State of the Great Lakes 2007 - Draft
            DRAFT - JUNE 2007

-------
Preface

The Governments of Canada and the United States are committed to providing public access to
environmental information that is reported through the State of the Great Lakes reporting process.
This commitment is integral to the mission to protect ecosystem health. To participate effectively
in managing risks to ecosystem health, all Great Lakes stakeholders (e.g., federal, provincial,
state and local governments; non-governmental organizations; industry; academia; private
citizens, Tribes and First Nations) should have access to accurate information of appropriate
quality and detail.

The information in this report, State of the Great Lakes 2007, has been assembled from various
sources with the participation of many people throughout the Great Lakes basin. The data are
based on indicator reports and presentations from the State of the Lakes Ecosystem Conference
(SOLEC), held in Milwaukee, Wisconsin, November 1-3, 2006. The sources of information are
acknowledged within each section.

Expanding upon previous State of the  Great Lakes reporting systems, the 2007 information is
presented in three different ways:

State of the Great Lakes 2007. This technical report contains the full indicator reports as
prepared by the primary authors, the indicator category assessments, and management challenges.
It also contains detailed references to data sources.

State of the Great Lakes 2007 Highlights. This report highlights key information presented in
the main report.

State of the Great Lakes Technical Summaries Series. These summaries provide information
from a variety of indicators such as: drinking water, swimming at the beaches, eating fish,  air
quality, aquatic invasive species, amphibians, birds, forests, coastal wetlands, the  Great Lakes
food web and special places such as islands, alvars and cobble beaches. In addition there is a
technical summary for each of the lakes, plus the St. Clair-Detroit River ecosystem and the St.
Lawrence River.

This approach of multiple reports addresses the needs of multiple audiences and also satisfies the
U.S. Guidelines for Ensuring and Maximizing the Quality, Objectivity, Utility, and Integrity of
Information Disseminated by Federal  Agencies, OMB, 2002, (67 FR 8452). The guidelines were
developed in response to U.S. Public Law  106-554: H.R. 5658, Section 515(a) of the Treasury
and General Government Appropriations Act for Fiscal Year 2001.

The State of the Lakes Ecosystem Conferences (SOLEC) and reports provide independent,
science-based reporting on the state of the health of the Great Lakes basin ecosystem. Four
objectives for the  SOLEC process include:
To assess the state of the Great Lakes  ecosystem based on accepted indicators
To strengthen decision-making and environmental management concerning the Great Lakes
To inform local decision makers of Great Lakes environmental issues
To provide a forum for communication and networking amongst all the Great Lakes stakeholders
                         Draft for Discussion at SOLEC 2006

-------
                                                                                  	
The role of SOLEC is to provide clear, compiled information to the Great Lakes community to
enable environmental managers to make better decisions. Although SOLEC is primarily a
reporting venue rather than a management program, many SOLEC participants are involved in
decision-making processes throughout the Great Lakes basin.

For more information about Great Lakes indicators and the State of the Lakes Ecosystem
Conference, visit: www.binational.net or www.epa.gov/glnpo/solec or
www.on.ec.gc.ca/solec.
1.0 Introduction

This State of the Great Lakes 2007 report presents the compilation, scientific analysis and
interpretation of data about the Great Lakes basin ecosystem. It represents the combined efforts of
many scientists and managers in the Great Lakes community representing federal, Tribal/First
Nations, state, provincial and municipal governments, non-government organizations, industry,
academia and private citizens.

The seventh in a series of reports beginning in 1995, the State of the Great Lakes 2007 provides
an assessment of the Great Lakes basin ecosystem components using a suite of ecosystem health
indicators. The Great Lakes indicator suite has been developed, and continues to be refined, by
experts as part of the State of the Lakes Ecosystem Conference (SOLEC) process.

The SOLEC process was established by the governments of Canada and the U.S. in response to
requirements of the Great Lakes Water Quality Agreement (GLWQA) for regular reporting on
progress toward Agreement goals and objectives.  Since the first conference in 1994, SOLEC has
evolved into a two-year cycle of data collection, assessment and reporting on conditions and the
major pressures in the Great Lakes basin. The year following each conference, a State of the
Great Lakes report is prepared, based on information presented and discussed at the conference
and post-conference comments. Additional information about SOLEC and the Great Lakes
indicators is available at www.binational.net.

The State of the Great Lakes 2007 provides assessments of 63 of approximately 80 ecosystem
indicators and overall assessments of the categories into which the indicators are grouped:
Contamination, Human Health,  Biotic Communities, Invasive Species, Coastal Zones and
Aquatic Habitats, Resource Utilization, Land Use-Land Cover, and Climate Change. Within most
of the main categories are sub-categories to further delineate issues or geographic areas.

Authors of the indicator reports assessed the status of ecosystem components in relation to
desired conditions or ecosystem objectives, if available. Five status categories were used (coded
by color in this report):

|     |  Good.  The state of the  ecosystem component is presently meeting ecosystem objectives
       or otherwise is in acceptable condition.
                         Draft for Discussion at SOLEC 2006

-------
I     1  Fair. The ecosystem component is currently exhibiting minimally acceptable conditions,
       but it is not meeting established ecosystem objectives, criteria, or other characteristics of
       fully acceptable conditions.

^^|  Poor.  The ecosystem component is severely negatively impacted and it does not display
       even minimally acceptable conditions.

|i	|  Mixed. The ecosystem component displays both good and degraded features.

|     |  Undetermined.  Data are not available or are insufficient to assess the status of the
       ecosystem component.

Four categories were also used to denote current trends of the ecosystem component (coded by
shape in this Highlights report):

       Improving. Information provided shows the ecosystem component to be changing
       toward more acceptable conditions.

       Unchanging. Information provided shows the ecosystem component to be neither
       getting better nor worse.

       Deteriorating.  Information provided shows the ecosystem component to be departing
       from acceptable conditions.

       Undetermined.  Data are not available to assess the ecosystem component over time, so
       no trend can be identified.
For many indicators, ecosystem objectives, endpoints, or benchmarks have not been established.
For these indicators, complete assessments are difficult to determine.

In 2006, the overall status of the Great Lakes ecosystem was assessed as mixed because some
conditions or areas were good while others were poor. The trends of Great Lakes ecosystem
conditions varied: some conditions were improving and some were worsening.

Some of the good features of the ecosystem leading to the Mixed conclusion include:
    •  Levels of most contaminants in herring gull eggs continue to decrease
    •  Phosphorus targets have been met in Lakes Ontario, Huron, Michigan and Superior.
    •  The Great Lakes are a good source for treated drinking water.
    •  Sustainable forestry programs throughout the Great Lakes basin are helping
       environmentally friendly management practices.
    •  Lake trout stocks in Lake Superior have remained self-sustaining, and some natural
       reproduction of lake trout is occurring in Lake Ontario and in Lake Huron.
    •  Mayfly (Hexagenia) populations have partially recovered in western Lake Erie.
Some of the negative features of the ecosystem leading to the Mixed conclusion include:
                         Draft for Discussion at SOLEC 2006

-------

    •  Concentrations of the flame retardant PBDEs are increasing in herring gull eggs
    •  Nuisance growth of the green alga Cladophora has reappeared along the shoreline in
       many places
    •  Phosphorus levels are still above guidelines in Lake Erie.
    •  Non-native species (aquatic and terrestrial) are pervasive throughout the Great Lakes
       basin, and they continue to exert impacts on native species and communities.
    •  Populations ofDiporeia, the dominant, native, bottom-dwelling invertebrate, continue to
       decline in Lake Michigan, Lake Huron, and Lake Ontario, and they may be extinct in
       Lake Erie.
    •  Groundwater withdrawals for municipal water supplies and irrigation, and the increased
       proportion of impervious surfaces in urban areas, have negatively impacted groundwater.
    •  Long range atmospheric transport is a continuing source of PCBs and other contaminants
       to the Great Lakes basin, and can be expected to be significant for decades.
    •  Land use changes in  favour of urbanization along the shoreline continue to threaten
       natural habitats in the Great Lakes and St. Lawrence River ecosystems.
    •  Some species of amphibians and wetland-dependent birds are showing declines in
       population numbers - in part due to wetland habitat conditions.

The listing of the State of the Great Lakes 2007 indicator reports, the categories, and the
indicator assessments for 2007, 2005, 2003, and 2001 are provided in the following summary
table. A complete listing of all indicators in the Great Lakes suite can be found in Section 6.0.

2.0 Assessing Data Quality

Through both the biennial Conferences and the State of the Great Lakes reports (Technical
Report, Highlights, Summary Series), SOLEC organizers seek to disseminate the highest quality
information available to a wide variety of environmental managers, policy officials, scientists and
other interested public.  The importance of this quality standard, including the availability of
reliable and useful data, is implicit in the main objectives of the SOLEC process.

To ensure that data and information made available to the public by federal  agencies adhere to a
basic standard of objectivity, utility, and integrity, the U.S. Office of Management and Budget
issued a set of Guidelines1 in 2002. Subsequently, other U.S. federal agencies have issued their
own guidelines for implementing the OMB  policies. According to the Guidelines issued  by the
U.S. Environmental Protection Agency2, information must be accurate, reliable, unbiased, useful
and uncompromised though corruption or falsification. The U.S. EPA further amplified its
Guidelines in 2003 with a review of "assessment factors" that the agency typically takes into
account when evaluating the  quality and relevance of scientific and technical information:3
    •  Soundness - The extent to which the scientific and technical procedures,  measures,
       methods or models employed to generate the information are reasonable for, and
       consistent with, the intended application
    •  Applicability and Utility - The extent to which the information is relevant for the
       Agency's intended use
                         Draft for Discussion at SOLEC 2006

-------


    •   Clarity and Completeness - The degree of clarity and completeness with which the data,
        assumptions, methods, quality assurance, sponsoring organizations and analyses
        employed to generate the information are documented
    •   Uncertainty and Variability - The extent to which the variability and uncertainty
        (quantitative and qualitative) in the information or in the procedures, measures, methods
        or models are evaluated and characterized
    •   Evaluation and Review - The extent of independent verification, validation and peer
        review of the information or of the procedures, measures, methods or models.

Recognizing the need to more formally integrate concerns about data quality into the SOLEC
process, SOLEC organizers developed a Quality Assurance Project Plan (QAPP) in 2004.  The
QAPP recognizes that SOLEC, as an entity, does not directly measure any environmental or
socioeconomic parameters. Existing data are contributed by cooperating federal, state and
provincial environmental and natural resource agencies, non-governmental environmental
agencies or other organizations engaged in  Great Lakes monitoring. Additional data sources may
include  local governments, planning agencies, and the published scientific literature.  Therefore,
SOLEC relies  on the quality of datasets reported by others. Characteristics of datasets that would
be acceptable for indicator reporting include:

    •   Data are documented, validated, or quality-assured by a recognized agency or
        organization.
    •   Data are traceable to original sources
    •   The source of the data is a known, reliable and respected generator of data.
    •   Geographic coverage and scale of data are appropriate to  the Great Lakes Basin.
    •   Data obtained from sources within the United States are comparable with those from
        Canada.
    •   Gaps in data availability are identified if data sets are unavailable for certain
        geographic regions and/or contain a level of detail insufficient to be useful in the
        evaluation of a particular indicator.
    •   Data are evaluated for feasibility of being incorporated into indicator reports.
        Considerations include budgetary constraints in acquiring data, type and format of data,
        time required to convert data to usable form, and the collection frequency for particular
        types of data.

SOLEC relies  on a distributed system of information in which the data reside with the original
providers. Although data reported through SOLEC are not centralized, clear links for
accessibility of the data and/or the indicator authors are provided. The authors hold the primary
responsibility for ensuring that the data used for indicator reporting meet criteria for objectivity,
usefulness and integrity. Users of the indicator information, however, are obliged to  evaluate  the
usefulness and appropriateness of the data for their own application, and they are encouraged to
contact the authors with any concerns or questions.

The SOLEC indicator reporting process is intended to be open and collaborative. Indicator
authors  are generally subject matter experts who are the primary generators of data, who have
direct access to the data, or who are able to obtain relevant data from one or more other sources
and who can assess the quality of data for objectivity, usefulness and integrity. In some cases,
authors  may serve as facilitators or leaders to coordinate a workgroup of experts who collectively


                         Draft for Discussion at SOLEC 2006                        5

-------
contribute their data and information, to arrange for data retrievals from agency or organization
databases, or to review published scientific literature or conduct online data searches from trusted
sources, e.g., U.S. census data or the National Land Cover Dataset.

Several opportunities are provided for knowledgeable people to review and comment on the
quality of the data and information provided.  These include:
    •  Coauthors - Most of the indicator reports are prepared by more than one author, and data
       are often obtained from more than one source. As the draft versions are prepared, the
       authors freely evaluate the data.
    •  Comments from the Author(s) - The section in each indicator report called "Comments
       from the Author(s)" provides an opportunity for the authors to describe any known
       limitations on the use or interpretation of the data that are being presented.
    •  Pre-SOLEC availability - The indicator reports  are prepared before each Conference, and
       they are made available online to SOLEC participants in advance. Participants are
       encouraged to provide comments and suggestions for improvements, including any data
       quality issues.
    •  During SOLEC discussions - The Conferences have been designed to encourage
       exchange of ideas and interpretations among the participants.  The indicator reports
       provide the framework for many of the discussions.
    •  Post-SOLEC review period - Following the Conferences, interested agencies,
       organizations and other stakeholders are encouraged to review and comment on the
       information and interpretations provided in the indicator reports.
    •  Preparation of State of the Great Lakes products - Prior to finalizing the Technical
       Report, Highlights, and Summary Series, any substantive comments on the indicator
       reports, including data quality issues, are referred back to the authors for resolution with
       the report editors.

The primary record and documentation of the indicator reports and assessments are the State of
the Great Lakes reports.  The Technical Report presents the full indicator reports as prepared by
the primary authors. It also contains detailed references to the data sources. A Highlights report
is also produced which refers to the detailed references and links. This approach of dual reports,
one summary version and one with details and references to data sources, also satisfies the
Guidelines for Ensuring and Maximizing the Quality, Utility, and Integrity of Information
Disseminated by Federal Agencies, OMB, 2002, (67 FR 8452). The guidelines were developed
in response to U.S. Public Law 106-554; H.R. 5658, Section 515 (a) of the Treasury and General
Government Appropriations Act for Fiscal Year 2001.
1 Guidelines for Ensuring and Maximizing the Quality, Objectivity,  Utility, and Integrity of
Information Disseminated by Federal Agencies, OMB, 2002, (61 FR 8452). The guidelines were
developed in response to U.S. Public Law 106-554: H.R. 5658, Section 515(a) of the Treasury
and General Government Appropriations Act for Fiscal Year 2001.

 Guidelines for Ensuring and Maximizing the Quality, Objectivity,  Utility, and Integrity, of
Information Disseminated by the Environmental Protection Agency. 2002. U.S. Environmental
Protection Agency EPA/260R-02-008, 62pp.
                         Draft for Discussion at SOLEC 2006

-------

 Assessment Factors. A Summary of General Assessment Factors for Evaluating the Quality of
Scientific and Technical Information. 2003. U.S. Environmental Protection Agency. EPA 100/B-
03/001, 18pp.
3.0 What is being done to improve conditions?

In an effort to restore and preserve the Great Lakes, legislators, managers, scientists, educators
and numerous others are responding to environmental challenges with multifaceted solutions.
The responses and actions referenced here are intended to serve as examples of positive strides
being taken in the Great Lakes basin to improve ecosystem conditions. Examples from both
Canada and the United States and from each of the Great Lakes are included. There are many,
many more actions that could have been recognized in this report. Each is an important part of
our collective commitment to a clean and healthy Great Lakes ecosystem.

Strategic planning occurs  at basin-wide, lake-wide and local scales. An example of strategic
planning is the Canada-Ontario Agreement, a federal-provincial agreement that supports the
restoration, protection, and conservation of the Great Lakes basin ecosystem. To achieve the
collective goals and results, Canada and Ontario work closely with local and regional
governments, industry, community and environmental groups. In the United States, more than
140 different federal programs help fund and implement environmental restoration and
management activities in the basin. The Great Lakes Water Quality Agreement, Great Lakes
Regional Collaboration and Federal Task Force, Great Lakes Binational Toxics Strategy,
Lakewide Management Plans, Binational Partnerships, and Remedial Action Plans are other
examples of strategic planning in the Great Lakes basin.

Research, monitoring and assessment efforts  operating at various geographic scales are the
backbone of management actions and decisions in the basin. Coordinated monitoring among
Canadian and United States federal, provincial,  state, and university groups began in 2003  to
focus on monitoring physical, biological, and chemical parameters with monitoring occurring on
a five-year rotation of one Great Lake per year.  The International Joint Commission maintains a
Great Lakes - St. Lawrence Research Inventory of the many funded projects that help increase
our knowledge about the structure and function of the Great Lakes ecosystem.

Canada and the United States implement numerous actions across the basin at national, regional
and local scales. For example, in Ontario, the City of Toronto is addressing water pollution
through the Wet Weather Flow Management Master Plan, a long-term solution to reduce
pollution from stormwater and combined sewer overflows.

Communities, states, the U.S. Environmental Protection Agency and local industry are working
together to remediate contaminated sediments in U.S. Areas of Concern (AOCs) with funding
provided through the U.S. Great Lakes Legacy Act. Since inception of the Act in 2002, sediment
remediation has been completed at three U.S. AOC sites (Ruddiman Creek and Ruddiman Pond
in Michigan, Black Lagoon in Michigan, and Newton Creek and Hog Island Inlet in Wisconsin).
                         Draft for Discussion at SOLEC 2006

-------
The Oswego River AOC on Lake Ontario was delisted in 2006, the first removal of an AOC
designation in the United States. In Canada, two AOCs have been delisted, both on Lake Huron
(Collingwood Harbour in 1994 and Severn Sound in 2003). Delisting of an Area of Concern
occurs when environmental monitoring has confirmed that the remedial actions taken have
restored the beneficial uses in the area and that locally derived goals and criteria have been met.

Effective actions are often based on collaborative work. In 2005, the Nature Conservancy, the
State of Michigan and The Forestland Group (a limited partnership), collaborated in a sale and
purchase agreement that created the largest conservation project in Michigan's history. This
purchase will protect more than 110,000 hectares (271,000 acres) through a working forest
easement on 100,362 hectares (248,000 acres) and acquisition of 9,445 hectares (23,338 acres) in
the Upper Peninsula of Michigan. By connecting approximately one million hectares (2.5 million
acres), the project curbs land fragmentation and incompatible development by establishing
buffers around conservation sites such as the Pictured Rocks National Lakeshore and Porcupine
Mountains Wilderness State Park.

Lake Superior communities have embraced a goal of zero discharge of critical pollutants by
engaging in a number of actions to remove contaminants. Efforts to reach this goal include
electronic and hazardous waste collection events run by Earth Keepers, a faith-based
environmental organization based in the Upper Peninsula of Michigan. On Earth Day 2006, over
272 metric tons (300 U.S. tons) of household hazardous waste, primarily household electronics,
were collected, disposed of, or recycled. In Canada, more than  11,500 mercury switches from
scrap automobiles were collected in 2005 through Ontario's mercury Switch Out program.

In many cases management and conservation actions are based on or supported by federal, state,
provincial, or local legislation. For example, Ontario's Greenbelt Act of 2005 enabled the
creation of a Greenbelt Plan to protect about 728,437 hectares (1.8 million acres) of
environmentally-sensitive and agricultural land in the Golden Horseshoe region from urban
development and sprawl. The Plan includes and builds upon approximately 324,000 hectares
(800,000 acres) of land within the Niagara Escarpment Plan and the Oak Ridges Moraine
Conservation Plan.

Proving that some legislation effectively crosses national borders, in December, 2005, the Great
Lakes Governors and Premiers signed the Annex 2001 Implementing Agreements at the Council
of Great Lakes Governors'  Leadership Summit that will provide unprecedented protection for the
Great Lakes-St. Lawrence River basin. The agreements detail how the states and provinces will
manage and protect the basin and provide a framework for each state and province to enact laws
for its protection, once the agreement is ratified.

Education and outreach about Great Lakes environmental issues are essential actions for
fostering both a scientifically-literate public as well as informed decision-makers. The Lake
Superior Invasive-Free Zone Project involves community groups in the inventorying and control
of non-native invasive terrestrial and emergent aquatic plants through education. The project
combines Canadian and United States programs at federal, state, provincial, municipal, and local
levels and has the goal of eliminating non-native plants within a designated 291 hectare (720
acre) area.
                         Draft for Discussion at SOLEC 2006

-------
A Shoreline Stewardship Manual developed for the Southeast shore of Lake Huron and promoted
through workshops and outreach programs encourages sustainable practices to improve and
maintain the quality of groundwater and surface water and the natural landscape features that
support them. The Shoreline Stewardship Manual is a collaborative effort by the Huron County
Planning Department, the University of Guelph, the Huron Stewardship Council, the Ausable
Bayfield Conservation Authority, the Lake Huron Centre for Coastal Conservation, and the
Friends of the Bayfield River, and a high level of community engagement has been instrumental
in its success.

The Great Lakes Conservation Initiative of the Shedd Aquarium in Chicago aims to draw public
attention to the value and vulnerabilities of the Great Lakes. With collaboration by Illinois-
Indiana Sea Grant and the U.S. Fish and Wildlife Service, the Shedd Aquarium opened a new
exhibit in 2006 which features many of the invasive species found in the Great Lakes. This
exhibit provides public audiences with the opportunity to see many of these live animals  and
plants, and is also highlighted in teacher workshops.

As these examples show, there is much planning, information gathering, research and education
occurring in the Great Lakes basin. Much more remains to be done to meet the goals of the
GLWQA, but progress is being made with the involvement of all Great Lakes stakeholders.

4.0 Indicator Category Assessments and Management Challenges
Contamination

The transfer of natural and human-made substances from air, sediments, groundwater,
wastewater, and runoff from non-point sources is constantly changing the chemical composition
of the Great Lakes. Over the last 30 years, concentrations of some chemicals or chemical groups
have declined significantly. There is a marked reduction in the levels of toxic chemicals in air,
water, biota, and sediments. Many remaining problems are associated with local regions such as
Areas of Concern. However, concentrations of several other chemicals that have been recently
detected in Great Lakes have been identified as chemicals of emerging concern.

Levels of most contaminants in herring gull eggs continue to decrease in all the Great Lakes
colonies monitored, although concentration levels vary from good in Lake Superior, to mixed in
Lake Michigan, Lake Erie and Lake Huron, to poor in Lake Ontario. While the frequency of
gross effects of contamination on wildlife has subsided, many subtle (mostly physiological and
genetic) effects that were not measured in earlier years of sampling remain in herring gulls.
Concentrations of flame-retardant polybrominated diphenyl ethers (PBDEs) are increasing in
herring gull eggs.

Concentrations of most organic contaminants in the offshore waters of the Great Lakes are low
and are declining, indicating progress in the reduction of persistent toxic  chemicals. Indirect
inputs of in-use organochlorine pesticides are most likely the current source of entry to the Great
Lakes. Continuing sources of entry of many organic contaminants to the Great Lakes include
indirect inputs such as atmospheric deposition, agricultural land runoff, and resuspension of
                         Draft for Discussion at SOLEC 2006

-------
                                              ! \ *$" mm     ~Z™«	j
                                              'SfSftfft&f-X?: Tfl-S b^fCV^fn^' -. "'M"«;iS'rV	
                                              ll^f >UiP?ilj»";'-i?'S W"  ?*•!';ft'-".'; jj1;' '""".- *-U .13 vj^'   j
contaminated sediments. Overall, mercury concentrations in offshore waters are well below water
quality guidelines. Mercury concentrations in waters near major urban areas and harbors,
however, exceed water quality criteria for protection of wildlife. Concentrations of poly cyclic
aromatic hydrocarbons (PAHs) and dioxins in offshore waters have declined below water quality
guidelines, largely due to the control of point sources.

The status of atmospheric deposition of toxic chemicals is mixed and improving for
polychlorinated biphenyls (PCBs), banned organochlorine pesticides, dioxins, and furans, but
mixed and unchanging or slightly improving for PAHs and mercury across the Great Lakes. For
Lake Superior, Lake Michigan, and Lake Huron, atmospheric inputs are the largest source of
toxic chemicals due to the large surface areas of these lakes. While atmospheric concentrations of
some substances are very low at rural sites, they may be much higher in some urban areas.

Juvenile spottail shiner, an important preyfish species in the Great Lakes, is a good indicator of
nearshore  contamination because the species limits its  distribution to localized, nearshore areas
during its first year of life.  Total dichlorodiphenyltrichloroethane (DDT) in juvenile spottail
shiner has declined over the last 30 years but still exceeds GLWQA criteria at most locations.
Concentrations of PCBs in juvenile spottail shiner have decreased below the GLWQA guideline
at many, but not all, sites in the Great Lakes.

The status of contaminants in lake trout, walleye and smelt as monitored annually in the open
waters of each of the Great Lakes is mixed and improving for PCBs, DDT, toxaphene, dieldrin,
mirex, chlordane, and mercury. Concentrations of PBDEs and other chemicals of emerging
concern such as perflourinated chemicals, however, are increasing. Both the United States and
Canada continue to monitor for these chemicals in whole fish tissues and have over 30 years of
data to support the status and trends information.

Phosphorus concentrations in the Great Lakes were a major concern in the 1960s and 1970s, but
private and government actions have reduced phosphorus loadings,  thus maintaining or reducing
phosphorus concentrations in open waters. However, high phosphorus concentrations are still
measured in some embayments, harbors,  and nearshore areas. Nuisance growth of the green alga
Cladophora has reappeared along the shoreline in many places and may be related, in part, to
increased availability of phosphorus.

Management Challenges:
Presently,  there are no standardized analytical monitoring methods and tissue residue guidelines
for new contaminants and chemicals of emerging concern, such as PBDEs.
PCBs from residual sources in the United States, Canada, and throughout the world enter the
atmosphere and are transported long distances. Therefore, atmospheric deposition of PCBs to the
Great Lakes will still be significant at least decades into the future.
Assessment of the capacity and operation of existing sewage treatment plants for phosphorus
removal, in the context of increasing human populations being served, is warranted.
Monitoring of tributary, point source, and urban and rural non-point source contributions of
phosphorus will allow tracking of various sources of phosphorus loadings.
Investigating the causes of Cladophora reappearances  will aid in the reduction of its impacts on
the ecosystem.
                         Draft for Discussion at SOLEC 2006

-------
Chemical Integrity - What the Experts are Saying

Chemical Integrity of the Great Lakes - What the Experts are Saying
In addition to the ecosystem information derived from indicators, six presentations on the theme
of "Chemical Integrity of the Great Lakes " were delivered at SOLEC 2006 by Great Lakes
experts. The definition of Chemical Integrity proposed by SOLEC is "the capacity to support and
maintain a balanced, integrated and adaptive biological system having the full range of elements
and processes expected in a region's natural habitat. " James R. Karr, 1991 (modified)

The presentations focused on the status of anthropogenic (man-made) contaminants and
imbalances in naturally-occurring chemicals in the Great Lakes basin. The key points of each
presentation are summarized here.

Anthropogenic Chemicals
Ron Kites, Indiana University: While concentrations of banned or regulated toxic substances such
as PCBs and PAHs have decreased over the past 30 years, the rate of decline has slowed
considerably over the past decade. Virtual elimination of most of these chemicals will not occur
for another 10 to 30 years despite restrictions or bans on their use. Further decreases in the
environmental concentrations of PCBs, PAHs, and some pesticides may well depend on emission
reductions in cities.

Derek Muir, Environment Canada: Some 70,000 commercial and industrial compounds are now
in use, and an estimated 1,000 new chemicals are introduced each year. Several chemical
categories have been identified as chemicals of emerging concern, including polybrominated
diphenyl ethers (flame retardants), perfluorooctanyl sulfonate (PFOS) and carboxylates,
chlorinated paraffins and naphthalenes, various pharmaceutical and personal care products,
phenolics, and approximately 20 currently-used pesticides. PBDEs, siloxanes and musks are now
widespread in the Great Lakes environment. Implementation of a more systematic program for
monitoring new persistent toxic substances in the Great Lakes will require significant investments
in instrumentation and researchers.

Joanne Parrot, Environment Canada: Some pharmaceuticals and personal care products appear to
cause negative effects in aquatic organisms at very low concentrations in laboratory experiments.
Some municipal waste water effluents within the Great Lakes discharge concentrations of these
products within these ranges. There is some evidence that fish and turtles show developmental
effects when exposed to municipal wastewater effluent in the laboratory. Whether these effects
appear in aquatic organisms including invertebrates, fish, frogs, and turtles, in environments
downstream of municipal wastewater effluent is not known, indicating the need for more research
in this area.

Naturally-occurring Chemicals
Harvey Bootsma, University of Wisconsin-Milwaukee: Changes in levels of nitrate, chloride and
phosphorus in Great Lakes waters are attributed to human activities, with potential effects on
phytoplankton and bottom-dwelling algae. Changes in lake chemistry, shown through variations
in calcium, alkalinity, and even chlorophyll, are linked to the biological activity of non-native
                         Draft for Discussion at SOLEC 2006

-------
species. Non-native species also appear to be altering nutrient cycling pathways in the Great
Lakes, by possibly intercepting nearshore nutrients before they can be exported offshore and
transferring them to the lake bottom.

Susan Watson, Environment Canada: The causes and occurrences of taste and odor impairments
in surface waters are widespread, erratic, and poorly characterized but are likely caused by
volatile organic compounds produced by species of plankton, benthic organisms, and
decomposing organic materials. In recent years, there has been an increase in the frequency and
severity of nuisance algae such as Cladophora outbreaks in the Great Lakes, particularly in the
lower Great Lakes. Type E botulism outbreaks and resulting waterbird deaths continue to occur in
Lake Michigan, Lake Erie and Lake Ontario.

David Lam, Environment Canada: Models and supporting monitoring data are used to predict
Great Lakes water quality. A post-audit of historical models for Great Lakes water quality
revealed the general success of setting target phosphorus loads to reduce open water phosphorus
concentrations.
Human Health

Levels ofPCBs in sportfish continue to decline, progress is being made to reduce air pollution,
beaches are better assessed and more frequently monitored for pathogens, and treated drinking
water quality continues to be assessed as good. Although concentrations of many organochlorine
chemicals in the Great Lakes have declined since the 1970s, sportfish consumption advisories
persist for all of the Great Lakes.

The quality of municipally-treated drinking water is considered good. The risk of human
exposure to chemicals and/or microbiological contaminants in treated drinking water is generally
low. However, improving and protecting source water quality (before treatment) is important to
ensure good drinking water quality.

In 2005, 74 percent of monitored Great Lakes beaches in the United States and Canada remained
open more than 95 percent of the swimming season. Postings, advisories or closures  were due to a
variety of reasons, including the presence of E.  coli bacteria, poor water quality, algae abundance,
or preemptive beach postings based on storm events and predictive models. Wildlife waste on
beaches can be more of a contributing factor towards bacterial contamination of water and
beaches than previously thought.

Concentrations of organochlorine contaminants in Great Lakes sportfish are generally decreasing.
However, in the United States, PCBs drive consumption advisories of Great Lakes sportfish. In
Ontario, most of the consumption advisories for Great Lakes sportfish are driven by  PCBs,
mercury, and dioxins. Toxaphene also contributes to consumption advisories of sportfish from
Lake Superior and Lake Huron. Monitoring for other contaminants, such as PBDEs,  has begun in
some locations.
                         Draft for Discussion at SOLEC 2006

-------

Overall, there has been significant progress in reducing air pollution in the Great Lakes basin.
However, regional pollutants, such as ground-level ozone and fine particulates, remain a concern,
especially in the Detroit-Windsor-Ottawa corridor, the Lake Michigan basin, and the Buffalo-
Niagara area. Air quality will be further impacted by population growth and climate change.

Management Challenges:
Maintenance of high-quality source water will reduce costs associated with treating water,
promote a healthier ecosystem, and lessen potential contaminant exposure to humans.
Although the quality of treated drinking water remains good, care must be taken to maintain
water treatment facilities.
One-fourth of monitored beaches still have beach postings or closures.
A decline in some contaminant concentrations has not eliminated the need for Great Lakes
sportfish consumption advisories.
Most urban and local air pollutant concentrations are decreasing. However, population growth
may impact future air pollution levels.
Biotic Communities

Despite improvements in levels of contaminants in the Great Lakes, many biological components
of the ecosystem are severely stressed. Populations of the native species near the base of the food
web such as Diporeia and species of zooplankton are in decline in some of the Great Lakes.
Native prey fish populations have declined in all lakes except Lake Superior. Significant natural
reproduction of lake trout is occurring in Lake Huron and Lake Superior only. Walleye harvests
have improved but are still below fishery target levels. Lake sturgeon are locally extinct in many
tributaries and waters where they once spawned and flourished. Habitat loss and deterioration
remain the predominant threat to Great Lakes amphibian and wetland-dependant bird
populations.

The aquatic food web is severely impaired in all the Great Lakes with the exception of Lake
Superior. Zooplankton populations have declined  dramatically in Lake Huron, and a similar
decline is occurring in Lake Michigan. Populations of Diporeia, the dominant native benthic
(bottom-dwelling) invertebrate in offshore waters, continue to decline in Lake Huron, Lake
Michigan and Lake Ontario, and they may be locally extinct in Lake Erie. The decline of
Diporeia coincides with the introduction of non-native zebra and quagga mussels. Both
zooplankton and Diporeia are crucial food sources for many other species, so their population
size and health impact the entire system.

The current mix of native and non-native (stocked and naturalized) prey and predator fish species
in the system has confounded the natural balance within most of the Great Lakes. In all but
Lake Superior, native preyfish populations have deteriorated. However, the recent decline of non-
native preyfish (alewife and smelt) abundance in all Great Lakes except Lake Superior could have
positive impacts on other preyfish populations. Preyfish populations  are important for their role in
supporting predator fish populations, so the potential effects of these changes will be a significant
factor to be considered in fisheries management decisions.
                         Draft for Discussion at SOLEC 2006

-------
Despite basin-wide efforts to restore lake trout populations that include stocking, harvest limits,
and sea lamprey management, lake trout have not established self-sustaining populations in Lake
Michigan, Lake Erie, and Lake Ontario. In Lake Huron, substantial and widespread natural
reproduction of lake trout was observed starting in 2004 following the near collapse of alewife
populations. This change may have been due to the reduced predation on juvenile lake trout by
adult alewives and the alleviation of a trout vitamin deficiency problem caused by trout
consuming alewives. In Lake Superior, lake trout stocks have recovered such that hatchery-reared
trout are no longer stocked.

Reductions in phosphorus loadings during the 1970s substantially improved spawning and
nursery habitat for many fish species in the Great Lakes. Walleye harvests have improved but are
still below target levels. Lake sturgeon are now locally extinct in many tributaries and waters
where they once spawned and flourished, although some remnant lake sturgeon populations exist
throughout the Great Lakes. Spawning and rearing habitats have been destroyed, altered or access
to them blocked. Habitat restoration  is required to help re-establish vigorous lake  sturgeon
populations.

From 1995 to 2005, the American toad, bullfrog, chorus frog, green frog and northern leopard
frog exhibited significantly declining population trends while the spring peeper was the only
amphibian species that exhibited a significantly increasing population trend in Great Lakes
coastal wetlands. For this same time  period, 14  species of wetland-dependant birds exhibited
significantly declining population trends, while only six species exhibited significantly increasing
population trends.

The  Great Lakes are now facing a challenge from viral hemorrhagic septicemia (VHS). This virus
has affected at least 37 fish species and is blamed for fish kills in Lake Huron, Lake St. Clair,
Lake Erie, Lake Ontario, and the St.  Lawrence River.

Management Challenges:
Populations ofDiporeia continue to  decline in Lake Michigan, Lake Huron, and Lake Ontario,
and may be locally extinct in Lake Erie. Management actions to address the declines may be
ineffective until the underlying causes of the declines are identified.
The  decline ofDiporeia coincides with the spread of non-native zebra and quagga mussels. Cause
and effect linkages between non-native species  in the Great Lakes and ecological  impacts are
essential, however, they may be difficult to establish.
Identification of remnant lake sturgeon spawning populations should assist the selection of
priority restoration activities to improve degraded lake sturgeon spawning and rearing habitats.
Protection of high-quality wetland habitats and adjacent upland areas  will help support
populations of wetland-dependent birds and amphibians.
Invasive Species

Activities associated with shipping are responsible for over one-third of the aquatic non-native
species introductions to the Great Lakes. Total numbers of non-native species introduced and
established in the Great Lakes have increased steadily since the 1830s. However, numbers of
                         Draft for Discussion at SOLEC 2006

-------

ship-introduced aquatic species have increased exponentially during the same time period. High
population density, high-volume transport of goods, and the degradation of native ecosystems
have also made the Great Lakes region vulnerable to invasions from terrestrial non-native
species. Introduction of these species is one of the greatest threats to the biodiversity and natural
resources of this region, second only to habitat destruction.

There are currently 183 known aquatic and 124 known terrestrial non-native species that have
become established in the Great Lakes basin. Non-native species are pervasive throughout the
Great Lakes basin, and they continue to exert impacts on native species and communities.
Approximately 10 percent of aquatic non-native species are considered invasive and have an
adverse effect, causing considerable ecological, social, and economic burdens.

Both aquatic and terrestrial wildlife habitats are adversely impacted by invasive  species. The
terrestrial non-native emerald ash borer, for example, is a tree-killing beetle that has killed more
than 15 million trees in the state of Michigan alone as of 2005. The emerald ash borer probably
arrived in the United States on solid wood packing material carried in cargo ships or airplanes
originating from its native Asia.

Introductions of non-native invasive species as a result of world trade and travel have increased
steadily since the 1830s and will continue to rise if prevention measures are not improved. The
Great Lakes basin is particularly vulnerable to non-native invasive species because it is a major
pathway of trade and is an area that is already disturbed.

Management Challenges:
A better understanding of the entry routes of non-native invasive species would aid in their
control and prevention.
Prevention and control require coordinated regulation and enforcement efforts to effectively limit
the introduction of non-native invasive species.
Prevention of unauthorized ballast water exchange by ships will eliminate one key pathway of
non-native aquatic species introductions to the Great Lakes.
The unauthorized release, transfer, and escape of introduced aquatic non-native species and
private sector activities related to aquaria, garden ponds, baitfish, and live food fish markets need
to be considered.
Coastal Zones and Aquatic Habitats

Coastal habitats are degraded due to development, shoreline hardening and establishment of
local populations of non-native invasive species. Wetlands continue to be lost and degraded. In
addition to providing habitat and feeding areas for many species of birds, amphibians and fish,
wetlands also serve as a refuge for native mussels and fish that are threatened by non-native
invasive species.

The Great Lakes coastline is more than 17,000 kilometers (10,563 miles) long. Unique habitats
include more than 30,000 islands, over 950 kilometers (590 miles) of cobble beaches, and over
30,000 hectares (74,131 acres) of sand dunes. Each coastal zone region is subject to a
combination of human and natural stressors such as agriculture, residential development, point
                          Draft for Discussion at SOLEC 2006

-------
                                                     r       -       .''!*-?''
                                             i S I«i< Js si >3J i ',, j!u"jiiV'1H!*iiri-i>'' •'•/"",*
                                             •ijjfe ?*«r ^tAwl-r *^~*
-------
An educated public is essential to ensuring wise decisions about the stewardship of the Great
Lakes basin ecosystem.
Protection of groundwater recharge areas, conservation of water resources, informed land use
planning, raising of public awareness, and improved monitoring are essential actions for
improving groundwater quality and quantity.
Resource Utilization

Although water withdrawals have decreased, overall energy consumption is increasing as
population and urban sprawl increase throughout the Great Lakes basin. Human population
growth will lead to an increase in the use of natural resources.

The population of the Great Lakes basin is approximately 42 million. Growth forecasts for the
western end of Lake Ontario (known as the Golden Horseshoe) predict that this portion of the
Canadian population will grow by an additional 3.7 million people by 2031. Population size,
distribution, and density are contributing factors to resource use in the basin, although many
trends have not been adequately assessed. In general, resource use is connected to economic
prosperity and consumptive behaviors.

Although the Great Lakes and their tributaries contain 20 percent of the world's supply of surface
freshwater, less than one percent of these waters is renewed annually through precipitation, run-
off and infiltration. The net basin water supply is estimated to  be 500 billion liters (132 billion
gallons) per day. In 2000, water from the Great Lakes was used at a rate equal to approximately
35 percent of the available daily supply. The majority of water withdrawn is returned to the basin
through discharge or run-off. However, approximately seven percent is lost through evapo-
transpiration or depleted by human activities. Due to the shutdown of nuclear power facilities and
improved water efficiency at thermal power plants, water use in Canada and the United States has
decreased since 1980. In the future, increased pressures on water resources are expected to come
from population growth in communities bordering the basin, and from climate change.

Population size, geography, climate, and trends in housing size and density all affect the amount
of energy consumed in the basin. Electricity generation was the largest energy consuming sector
in the Great Lakes basin.

Population growth and urban sprawl in the basin have led to an increase in the number of vehicles
on roads, fuel consumption, and kilometers/miles traveled. Over a ten year period (1994-2004)
fuel consumption increased by 17 percent in the U.S. states bordering the Great Lakes and by 24
percent in the province of Ontario. Kilometers/miles traveled within the same areas increased 20
percent for the United States and 56 percent for Canada. The increase in registered vehicles
continues to outpace the increase in licensed drivers.

Management Challenges:
Increasing requests for water from communities bordering the basin, where existing water
supplies are scarce or of poor quality will require careful evaluation.
Energy production and conservation need to be carefully managed to meet current and future
energy consumption demands.
                         Draft for Discussion at SOLEC 2006

-------
Population growth and urban sprawl are expected to challenge the current and future
transportation systems and infrastructures in the Great Lakes basin.
Land Use-Land Cover

The Great Lakes basin encompasses an area of more than 765,000 square kilometers (295,000
square miles). How land is used impacts not only water quality of the Great Lakes, but also
biological productivity, biodiversity, and the economy.

Data from 1992 and 2002 indicate that forested land covered 61 percent of the Great Lakes basin
and 70 percent of the land immediately buffering surface waters, known as riparian zones. The
greater the forest coverage in a riparian zone, the greater the capacity for the watershed to
maintain biodiversity, store water, regulate water temperatures, and limit excessive nutrient and
sediment loadings to the waterways. Urbanization, seasonal home construction, and increased
recreational use are among the general demands being placed on forest resources nationwide.
Additional disturbances caused by lumber removal and forest fires can also alter the structure of
Great Lakes basin forests. However, the area of forested lands certified under sustainable forestry
programs has significantly increased in recent years, exemplifying continued commitment from
forest industry professionals to practices that help protect local ecosystem  sustainability.
Continued growth in these practices will lead to improved soil and water resources and increased
timber productivity in areas of implementation.

Under the pressure of rapid population growth in the Great Lakes region, urban development has
undergone unprecedented growth. Sprawl  is increasing in rural and urban fringe areas of the
Great Lakes basin, placing a strain on infrastructure and consuming habitat in areas that tend to
have healthier environments than those that remain in urban areas. This trend  is expected to
continue, which will exacerbate other problems, such as longer commute times from residential to
work areas, increased consumption  of fossil fuels, and fragmentation of habitat. For example, at
current development rates in Ontario,  residential building projects are predicted to consume some
1,000 square kilometers (386 square miles) of the countryside, an area double the size of Toronto,
by 2031. Also, vehicle gridlock could increase commuting times by 45 percent, and air quality
could decline due to an estimated 40 percent increase in vehicle emissions.

In 2006, The Nature Conservancy Great Lakes Program and the Nature Conservancy of Canada
Ontario Region released the Binational Conservation Blueprint for the Great Lakes. The
Blueprint identified 501 areas across the Great Lakes that are a priority for biodiversity
conservation. The Blueprint was developed by scientifically and systematically identifying native
species, natural communities,  and aquatic  system characteristics of the region, and determining
the sites that need to be preserved to ensure their long-term survival.

Management Challenges:
As the volume of data on land use and land conversion grows, stakeholder discussions will assist
in identifying the associated pressures and management implications.
Comprehensive land use planning that incorporates "green" features, such as cluster development
and greenway areas, will help to alleviate the pressure from development.
                         Draft for Discussion at SOLEC 2006

-------

Managing forest lands in ways that protect the continuity of forest cover can allow for habitat
protection and wildlife species mobility, therefore maintaining natural biodiversity.
Policies that favor an economically viable forestry industry will motivate private and commercial
landowners to maintain land in forest cover versus conversion to alternative uses such as
development.
Climate Change

A qualitative assessment of the indicator category Climate Change could not be supported for
this report. Some observed effects in the Great Lakes region, however, have been attributed to
changes in climate. Winters are getting shorter; annual average temperatures are growing
warmer; extreme heat events are occurring more frequently; duration of lake ice cover is
decreasing as air and water temperatures are increasing; and heavy precipitation events, both
rain and snow, are becoming more common.

Continued declines in the duration and extent of ice cover on the Great Lakes and possible
declines in lake levels due to evaporation during the winter are expected to occur in future years.
If water levels decrease as predicted with increasing temperature, shipping revenue may decrease
and the need for dredging could increase. Northward migration of species  naturally found south
of the Great Lakes region and invasions by warm water, non-native aquatic species will likely
increase the stress on native species. A change in the distribution of forest types and an increase
in forest pests are expected. An increase in the frequency of winter run-off and intense storms
may deliver more non-point source pollutants to the lakes.

Management Challenges:
Increased modeling, monitoring and analysis of the effects of climate change on Great Lakes
ecosystems would aid in related management decisions.
Increased public awareness of the causes of climate change may lead to more environmentally-
friendly actions.
                         Draft for Discussion at SOLEC 2006

-------

Salmon and Trout
Indicator #8
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving
The number of stocked salmonines per year is decreasing due to
improvements in suppressing the abundance of the non-native preyfish,
alewife. Many of the introduced salmonines are also reproducing
successfully in the Great Lakes. The combined effect of a decrease in
the number of alewife, as well as the increased health and reproduction
of the salmonines is creating an improvement in the Great Lakes
ecosystem.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Fair
Improving
The number of stocked salmonines per year in Lake Superior is decreasing
at a steady rate.  Populations of salmon, rainbow trout and brown trout are
being stocked at suitable rates to restore and manage indigenous fish species
in Lake Superior.
Mixed
Slightly Improving
The number of salmonines stocked each year in Lake Michigan is slightly
declining.  The goal for Lake Michigan is to establish self-sustaining lake
trout populations. Currently, there are more salmon than lake trout stocked,
which suggests that the lake trout are beginning to meet the self-sustaining
goal for a balance in the ecosystem. This lake has the highest stocking rates
out of all the Great Lakes.
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Fair
Improving
The number of salmonines stocked each year in Lake Huron is declining.
This lake has the second highest number of stocked salmonines, but the
numbers are decreasing faster than Lake Superior, suggesting a larger
reproduction rate and a balance in the ecosystem.
Good
Improving
Lake Erie is one of the lowest stocked out of all the Great Lakes. The
objective for Lake Erie is to provide sustainable harvests of valued fish
including lake trout, rainbow trout, and other salmonoids. Fisheries
                        Draft for Discussion at SOLEC 2006

-------

Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   restoration programs in Ontario and New York State have established
                   regulations to conserve the harvest and increase fish populations for the
                   next five years.
Mixed
Unchanging
Lake Ontario has the second largest stocking rates (after Lake Michigan).
The number of stocked salmonines has slightly declined in the last couple
decades, but stocking numbers have been fairly constant in the last four
years. The main objective for Lake Ontario is to have a diversity of
naturally produced salmon and trout, with an abundance of rainbow trout
and the top predator to be Chinook salmon.  There is an abundance of
rainbow trout and Chinook salmon, but the salmon and trout are not being
naturally produced based on the high numbers of stocked fish each year.
Purpose
•To assess trends in populations of introduced salmon and trout species;
•To infer trends in species diversity in the Great Lakes basin; and
•To evaluate the resulting impact of introduced salmonines on native fish populations and the
  preyfish populations that supports them.

Ecosystem Objective
In order to manage Great Lakes fisheries, a common fish community goal was developed by
management agencies responsible for the Great Lakes fishery. The goal is:

"To secure fish communities, based on foundations of stable self-sustaining stocks, supplemented
by judicious plantings of hatchery-reared fish, and provide from these communities an optimum
contribution of fish, fishing opportunities and associated benefits to meet needs identified by
society for wholesome food, recreation, cultural heritage, employment and income, and a healthy
aquatic environment" (GLFC 1997).

Fish Community Objectives (FCOs) for each lake address introduced salmonines such as chinook
and coho salmon, rainbow and brown trout (see Table 1 for definitions offish terms). The
following objectives are used to establish stocking and harvest targets consistent with FCOs for
restoration of native salmonines such as lake trout, brook trout, and, in Lake Ontario, Atlantic
salmon:

Lake Ontario (1999):  Establish a diversity of salmon and trout with an abundant population of
rainbow trout and the chinook salmon as the top predator supported by a diverse preyfish
community with the alewife as an important species. Amounts of naturally produced (wild)
salmon and trout, especially rainbow trout that are consistent with fishery and watershed plans.

Lake Erie and Lake St. Clair (2003): Manage the eastern basin to provide sustainable harvests of
valued fish species, including.. .lake trout, rainbow trout, and other salmonids.
                         Draft for Discussion at SOLEC 2006

-------

Lake Huron (1995): Establish a diverse salmonine community that can sustain an annual harvest
of 2.4 million kg with lake trout the dominant species and stream-spawning species also having a
prominent place.

Lake Michigan (1995): Establish a diverse salmonine community capable of sustaining an annual
harvest of 2.7 to 6.8 million kg (6 to 15 million Ib), of which 20-25% is lake trout, and establish
self-sustaining lake trout populations.

Lake Superior (2003): Manage populations of Pacific salmon, rainbow trout, and brown trout that
are predominantly self-sustaining but may be supplemented by stocking that is compatible with
restoration and management goals established for indigenous fish species.
Term
Salmonine
Salmonid
Pelagic
Definition
Refers to salmon and trout species
Refers to any species of fish with an adipose fin,
whitefish, graying, and cisco
Living in open water, especially where the water
including trout,
is more than 20
salmon,
m deep
Table 1. Glossary of various terms used in this report

State of the Ecosystem
First introduced to the Great Lakes in the late 1870s, non-native salmonines have emerged as a
prominent component of the Great Lakes ecosystem and an important tool for Great Lakes
fisheries management. Fish managers stock non-native salmonines to suppress abundance of the
non-native preyfish, alewife, thereby reducing alewife predation and competition with native fish,
while seeking to avoid wild oscillations in salmomine-predator/alewife-prey ratios. In addition,
non-native salmonines are stocked to create recreational fishing opportunities with substantial
economic benefit (Rand and Stewart 1998).

After decimation of the native top predator (lake trout) by the non-native, predaceous sea
lamprey, stocking of non-native salmonines increased dramatically in the 1960s and 1970s. Based
on stocking data obtained from the Great Lakes Fishery Commission (GLFC),  approximately 922
million non-native salmonines were stocked in the Great Lakes basin between  1966 and 2005.
This estimate excludes the stocking of Atlantic salmon in Lake Ontario because they are native to
this lake. Non-native salmonines also reproduce in the Great Lakes. For example, many of the
chinook salmon in Lake Huron are wild and not stocked. This includes mostly Chinook salmon,
followed by Rainbow trout. Since 2002,  74 million non-native salmonines have been stocked in
the Great Lakes. Although, this is a large amount offish being stocked, the number of stocked
salmonines has actually decreased 32% from 2002 to 2004.

Of non-native salmonines, chinook salmon are the most heavily stocked, accounting for about
45% of all non-native salmonine releases (Figure 1). Rainbow trout are the second highest non-
native stocked species, accounting for 25% of all non-native salmonine releases. Chinook salmon,
which prey almost exclusively on alewife, are the least expensive of all non-native salmonines to
rear, thus making them the backbone of stocking programs in alewife-infested  lakes, such as
Lakes Michigan, Huron and Ontario (Bowlby and Daniels 2002). Like other salmonines, chinook
salmon are also stocked in order to provide an economically important sport fishery. While
                         Draft for Discussion at SOLEC 2006

-------
chinook salmon have the greatest prey demand of all non-native salmonines, an estimated 76,000
tonnes of alewife in Lake Michigan alone are consumed annually by all salmonine predators
(Kocik and Jones 1999).

Data are available for the total number of non-native salmonines stocked in each of the Great
Lakes from 1966-2005 (Figure 2).

Of the five major Great Lakes (excluding Lake St. Clair), Lake Michigan is the most heavily
stocked, with a maximum stocking level in 1998 greater than 16 million non-native salmonines.
In contrast, Lake Superior has the lowest rates of stocking, with a maximum greater than 5
million non-native salmonines in 1991. Lakes Huron and Erie both display a similar overall
downward trend in stocking, especially in recent years. Lake Ontario has a constant, yet slightly
declining trend in stocking. In Lake Ontario, this trend can be explained by stocking cuts
implemented in 1993 by fisheries managers to lower prey consumption by salmonine species by
50% over two years (Schaner et al. 2001). Since the late 1980s, the number of non-native
salmonines stocked in the Great Lakes has been nearly constant or slightly declining with the
exception  of a 1998 peak in Lakes Michigan and Huron.

Overall, the Great Lakes are improving based on a general trend of reduced numbers of stocked
salmonines. The goal of creating a balanced ecosystem within each lake is occurring at different
levels for each individual lake. Lakes Superior and Erie are improving at the fastest rates with the
lowest stocking levels, while Lake Ontario is improving at the slowest rate out of all of the Great
Lakes. Lake Michigan's stocking levels are declining slightly more than Lake Ontario's levels,
but it also  has the highest number of stocked salmon and trout. Lake Huron has higher stocking
rates than Lake Erie and Superior, but the levels have been decreasing faster each year than any
other lake.

The number of stocked salmonines per year in Lake Superior is decreasing at a steady rate.
Populations of salmon, rainbow trout and brown trout are being stocked at suitable rates to restore
and manage indigenous fish species in Lake Superior.  Stocking rates have decreased in the last 5
years suggesting successful reproduction rates and suitable conditions for an improvement
towards a balanced ecosystem in the near future.

The number of salmonines stocked each year in Lake Michigan is slightly declining. The goal for
Lake Michigan is to establish self-sustaining lake trout populations. Currently, there are more
salmon than lake trout stalked, which suggests that the lake trout are beginning to meet the self-
sustaining goal for a balance in the ecosystem.  This lake has the highest stocking rates out of all
the Great Lakes.

The goal for Lake Huron is to  make the lake trout the dominant species.  The lake trout is one of
the few native deepwater predators found in the Great Lakes. Their populations in Lake Huron
and Lake Michigan were decimated in the 1950's  by over-fishing and predation by the exotic sea
lamprey (US Fish and Wildlife Service, 2005).  The number  of lake trout has increased in the last
decade due to the decrease in the number of sea lampreys (Madenjian and Desorcie, 2004). This
lake has the second highest number of stocked salmonines suggesting a low reproduction rate, but
an improvement in the balance of the ecosystem since these stocking levels are decreasing.
                         Draft for Discussion at SOLEC 2006

-------
Lake Erie is one of the lowest stocked out of all the Great Lakes.  The objective for Lake Erie is
to provide sustainable harvests of valued fish including lake trout, rainbow trout, and other
salmonoids. Based on figure 1, the need for stalking has dropped dramatically over the last few
years, suggesting that sustainable harvests are occurring in Lake Erie. Fisheries restoration
programs in Ontario and New York State have established regulations to conserve the harvest and
increase fish populations for the next five years (Lake Erie Lamp, 2003). This program is well on
its way since there have already been improvements in the fish populations.

Lake Ontario has the second largest stocking rates, following Lake Michigan. The number of
stocked salmonines has slightly declined in the last couple decades, but stocking numbers have
been fairly constant in the last four years. The main objective for Lake Ontario is to have a
diversity of naturally produced salmon and trout, with an abundance of rainbow trout and the top
predator to be Chinook salmon. Rainbow trout are the second highest stocked fish in Lake
Ontario, following Chinook salmon. Therefore, part of this goal has been met since the Chinook
salmon are readily available as the top predator, and Rainbow trout are abundant in Lake Ontario
because of the high stocking levels. However, the objective of having naturally producing salmon
and trout has not been met due to the need for high stocking rates  in Lake Ontario. The salmon
and trout are not naturally producing based on the high numbers of stocking each year.  Lake
Ontario received a "mixed" rating rather than deteriorating rating  because, although the
objectives have not been met, there is still a need for high stalking levels. Salmon and trout are
stalked not only to create a balance in the ecosystem, but for a popular recreational activity.
Sport fishing has been a very popular activity in Lake Ontario for many years. Native lake trout
are at the top of the food chain and would have disappeared if they weren't being stocked for
sport fishing. Sport fishing is a $3.1 billion annual business, according to a recent industry study
(Edgecomb, 2006). High stocking rates are needed to keep up with the popularity of sport fishing
in Lake Ontario, which explains the increased need for higher stocking levels in Lake Ontario.

Pressures
The introduction of non-native salmonines into the Great Lakes basin, beginning in the late
1870s, has placed  pressures on both the introduced species and the Great Lakes ecosystem. The
effects of introduction on the non-native salmonine species include changes in rate of survival,
growth and development, dispersion and migration, reproduction, and alteration of life-history
characteristics (Crawford 2001).

The effects of non-native salmonine introductions on the Great Lakes ecosystem are numerous.
Some of the effects on native species are; 1) the risk of introducing and transferring pathogens
and parasites (e.g. furunculosis, whirling disease, bacterial kidney disease, and infectious
pancreatic necrosis), 2) the possibility of local decimation or extinction of native preyfish
populations through predation, 3) competition between introduced and native species for food,
stream position, and spawning habitat, and 4) genetic alteration due to the creation of sterile
hybrids (Crawford 2001). The introduction of non-native salmonines to the Great Lakes basin is a
significant departure from lake trout's historic dominance as key predator.

With few exceptions (such as kokanee salmon), introduced salmonines are now reproducing
successfully in portions of the basin, and they are  considered naturalized components of the  Great
Lakes ecosystem.  Therefore, the question is no longer whether non-native salmonines should be
                         Draft for Discussion at SOLEC 2006

-------
introduced, but rather how to determine the appropriate abundance of salmonine species in the
lakes.

Within any natural system there are limits to the level of stocking that can be maintained. The
limits to stocking are determined by the balance between lower and higher trophic level
populations (Kocik and Jones 1999). Rand and Stewart (1998) suggest that predatory salmonines
have the potential to create a situation where prey (alewife) is limiting and ultimately predator
survival is reduced. For example, during the 1990s, chinook salmon in Lake Michigan suffered
dramatic declines due to high mortality and high prevalence of Bacterial Kidney Disease (BKD)
when alewife were no longer as abundant in the preyfish community (Hansen and Holey 2002).
Salmonine predators could have been consuming as much as 53 percent of alewife biomass in
Lake Michigan annually (Brown et al. 1999). While suppressing alewife populations, managers
seek to avoid extreme "boom and bust" predator and prey populations, a condition not conducive
to biological integrity. Currently managers seek to produce a predator/prey balance by adhering to
stocking ceilings established for lakes such as Michigan and Ontario, based on assessment of
forage species and naturally produced salmonines.

Because of their importance as  a forage base for the salmonine sport fishery,  alewife are no
longer viewed as a nuisance by some managers (Kocik and Jones 1999). However, alewives prey
on the young of a variety of native fishes, including yellow perch and lake trout, and they
compete with native fishes for zooplankton. In addition, the enzyme thiaminase in alewives
causes Early Mortality Syndrome (EMS) in salmonines that consume alewife, threatening lake
trout rehabilitation in the lower four lakes and Atlantic salmon restoration in Lake Ontario. As
alewife populations increase, massive over-winter die-offs can occur, particularly in severe
winters, fouling local beaches that  are used for recreation and impacting the health of the
surrounding ecosystem.

Management Implications
In Lakes Michigan, Huron and  Ontario, many salmonine species are stocked in order to maintain
an adequate population to suppress non-native prey species (alewife) as well as to support
recreational fisheries. Determining stocking levels that will avoid oscillations in the forage base
of the ecosystem is an ongoing  challenge. Alewife populations, in terms of an adequate forage
base for introduced salmonines, are difficult to estimate as there is a delay before stocked salmon
become significant consumers of alewife; meanwhile, alewife can suffer severe die offs in
particularly severe winters.

Fisheries managers seek to improve their means of predicting appropriate stocking levels in the
Great Lakes basin based on the alewife population. Long-term data sets and models track the
population of salmonines and species with which they interact. However, more research is needed
to determine the optimal number of non-native salmonines, to estimate abundance of naturally
produced salmonines, to assess the abundance of forage species,  and to better understand the role
of non-native salmonines and non-native prey species in the Great Lakes ecosystem.
Chinook salmon will likely continue to be the most abundantly stocked salmonine species in
Lakes Michigan, Huron, and Ontario because they are inexpensive to rear, feed heavily on
alewife, and they are highly valued by recreational fishers. Fisheries managers should continue to
model, assess, and practice adaptive management with the ultimate objective being to support fish
                         Draft for Discussion at SOLEC 2006

-------
community goals and objectives that GLFC lake committees established for each of the Great
Lakes.

Comments from the author(s)
This indicator should be reported frequently as salmonine stocking is a complex and dynamic
management intervention in the Great Lakes ecosystem.

Acknowledgments
Authors: Tracie Greenberg, Environment Canada, Burlington, ON.
Contributors: Melissa Greenwood and Erin Clark, Environment Canada, Downsview, ON,
John M. Dettmers, Great Lakes Fishery Commission, Senior Fishery Biologist, Ann Arbor, MI.

Data Sources
Bowlby, J.N., and Daniels, M.E. 2002. Lake Ontario Pelagic Fish 2: Salmon and Trout. 2002
Annual Report. www.glfc.org/lakecom/loc/mgmt_unit/index.html, last accessed May 14, 2006.

Brown, E.H., Jr., Busiahn, T.R., Jones, M.L., and Argyle, R.L. 1999. Allocating Great Lakes
Forage Bases in Response to Multiple Demand. In Great Lakes Fisheries Policy and
Management: a Binational Perspective, eds. W.W. Taylor and C.P. Ferreri, pp. 355-394. East
Lansing, MI: Michigan State University Press.

Crawford, S.S. 2001. Salmonine Introductions to the Laurentian Great Lakes: An Historical
Review  and Evaluation of Ecological Effects. Canadian Special Publication of Fisheries and
Aquatic Sciences, pp. 132-205.

Edgecomb, M. 2006. Critters that sport fish feed on are dwindling - Number of invasive species
in lake is up. Rochester Democrat and Chronicle.  Article from June 20, 2006, last accessed at
http://www.democratandchronicle.com/apps/pbcs.dll/article7AID-/20060620/NEWS01/6062003
32/1002/RSS01
Great Lakes Fishery Commission (GLFC). 2001. Strategic Vision of the Great Lakes Fishery
Commission for the First Decade of the New Millennium, http://www.glfc.org, last accessed
April 30, 2006.

Great Lakes Fishery Commission (GLFC). 1997. A Joint Strategic Plan for Management of Great
Lakes Fisheries, http://www.glfc.org/fishmgmt/sglfmp97.htm, last accessed April 28, 2006.

Hansen, M.J., and Holey, M.E. 2002. Ecological factors affecting the sustainability of chinook
and coho salmon populations in the Great Lakes, especially Lake Michigan. In Sustaining North
American salmon: Perspectives across regions and discipline, eds. K.D. Lynch, M.L. Jones and
W.W. Taylor, pp.155-179. Bethesda, MD: American Fisheries Society Press.

Indiana Division of Fish and Wildlife, Great Lakes Sport Fishing Council. 1997. Alewife Die-
Offs Expected on Indiana Shores, http://www.great-lakes.org/5-05-97.html, last accessed May 4,
2006.
                         Draft for Discussion at SOLEC 2006

-------
                                                          «~'j(tffi,<«-5w%i!A=d;:*.» J\sBi&j!tH«£,j][*jj   ^,>-^«f-:^r'J^"w
Kocik, J.F., and Jones, M.L. 1999. Pacific Salmonines in the Great Lakes Basin. In Great Lakes
Fisheries Policy and Management: a Binational Perspective, eds. W.W. Taylor and C.P. Ferreri,
pp. 455-489. East Lansing, MI, Michigan State University Press.

Lake Erie Lamp. 2003. Lake Erie Lamp Update 2003. Lakewide Management Plan.
http://www.binational.net/pdfs/erie/update2003-e.pdf. last accessed May 10, 2006.

Madenjian, C., and Desorcie, T. 2004.  Lake Trout Rehabilitation in Lake Huron—2004
Progress Report on Coded-Wire Tag Returns. Lake Huron Committee. Proceedings from the
Great Lakes Fishery Commission Lake Huron Committee Meeting Ypsilanti, Michigan, March 21,
2005.

Rand, P.S., and Stewart, DJ. 1998. Prey fish exploitation, salmonine production, and pelagic
food web efficiency in Lake Ontario. Can. J. Fish. Aquat. Sci. 55(2):318-327.

Schaner, T., Bowlby, J.N., Daniels, M., and Lantry, B.F. 2001. Lake Ontario Offshore Pelagic
Fish Community. Lake Ontario Fish Communities and Fisheries: 2000 Annual Report of the Lake
Ontario Management Unit.

US Fish and Wildlife Service. 2005.  Lake Trout Restoration Program.
http://www.fws.gov/midwest/alpena/laketrout.htm. last accessed May 15, 2006
List of Tables
Table 1. Glossary of various terms used in this report.

List of Figures
Figure 1. Non-Native salmonine stocking by species in the Great Lakes, 1966-2004 excluding
Atlantic salmon in Lake Ontario and brook trout in all Great Lakes.
Source: Great Lakes Fishery Commission Fish Stocking Database (www.glfc.org/fishstocking)

Figure 2. Total number of non-native salmonines stocked in the Great Lakes, 1966-2005
excluding Atlantic salmon in Lake Ontario and brook trout in all Great Lakes.
Source: Great Lakes Fishery Commission Fish Stocking Database (www.glfc.org/fishstocking)

Last updated
SOLEC 2006
                         Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
     E

     _c



     i"
     ^
     u
     o

     C/J
     L.
     o>
     J2

     E
     3
                    Non-Native Salmonine Stocking by Species, 1966-2004

                         A<0  A
-------
                             State of the Great Lakes 2007 - Draft
               Number of Non-Native Salmonines Stocked per Lake 1966-2005
                                      Year
                     -ER
                             -Ml
                                      HU
                                              SU
                                                     -ON
                                                              -SC
Figure 2. Total number of non-native salmonines stocked in the Great Lakes, 1966-2005
excluding Atlantic salmon in Lake Ontario and brook trout in all Great Lakes.
Source: Great Lakes Fishery Commission Fish Stocking Database (www.glfc.org/fishstocking)
10
Draft for Discussion at SOLEC 2006

-------

Walleye
Indicator #9
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Fair
Unchanging
An exceptionally strong 2003 hatch has bolstered walleye abundance in
nearly all of the Great Lakes and should keep them at low to moderate
levels for the next several years. Low reproductive success post-2003
will not permit populations to increase in many areas. Fisheries
harvests have improved in recent years but remain below targets in
nearly all areas.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Assessed Since Last Report
Undetermined
Recent harvest estimates were not available for this report. Through 2003,
commercial yields were below the historical average while tribal harvest
was above average.
Fair
Undetermined
Recreational harvest was below historical levels in 2004-2005. Tribal
fishery yields were not available but were well-above average in the four
most recent years where data exist (2000-2003). Green Bay stocks appear
to be stable, perhaps improving. Fishery yields remain well below targets of
100-200 metric tons per year.
Fair
Unchanging
Fishery yields are at historical average levels but far below targets of 700
metric tons each year.  Commercial harvest trends continue to decline while
recreational harvest trends are flat or perhaps improving.  Reproductive
success has greatly improved between 2003 and 2005 in Saginaw Bay and
perhaps other parts of the lake, and is attributed to the decline of alewives.
Fair
Unchanging
The fisheries objective of sustainable harvests lake wide has not been
realized since the late-1990s but has improved recently with contributions
from the strong 2003 hatch.  Commercial harvest increased substantially in
2005 while recreational fisheries remained static due to size restrictions.
Harvest by both fisheries is expected to increase substantially in 2006.
                         Draft for Discussion at SOLEC 2006

-------

Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Below-average reproductive success in 2004-2005 will reduce adult
                   abundance over the next few years but the 2003 hatch should keep the
                   population at low to moderate levels of abundance.
Fair
Unchanging
After a decade long decline, walleye populations appear to have stabilized.
Fishery yields are roughly half of the average over the past 30 years.
Recent hatches should keep the population at current levels of abundance
for the next several years.
Purpose
•To show status and trends in walleye populations in various Great Lakes habitats;
•To infer changes in walleye health; and
•To infer ecosystem health, particularly in moderately productive (mesotrophic) areas of
  the Great Lakes.

Ecosystem Objective
Protection, enhancement, and restoration of historically important, mesotrophic habitats that
support natural stocks of walleye as the top fish predator are necessary for stable, balanced, and
productive elements of the Great Lakes ecosystem.

State of the Ecosystem
Reductions in phosphorus loadings during the 1970s substantially improved spawning and
nursery habitat for many fish species in the Great Lakes. Improved mesotrophic habitats (i.e.,
western Lake Erie, Bay of Quinte, Saginaw Bay and Green Bay) in the 1980s, along with
interagency fishery management programs that increased adult survival, led to a dramatic
recovery of walleyes in many areas of the Great Lakes, especially in Lake Erie. High water levels
also may have played a role in the recovery in some lakes or bays. Trends in annual assessments
of fishery harvests generally track walleye recovery in these areas, with peak harvests occurring
in the mid-1980s to early 1990s followed by declines from the mid-1990s through 2000, and
increases in most areas after 2000 (Figure 1). Total yields were highest in Lake Erie (annual
average of about 4,500 metric tons, 1975-2005), intermediate in Lakes Huron (average  of 90
metric tons) and Ontario (average of 224 metric tons), and lowest in Lakes Michigan (average of
14 metric tons) and Superior (average of 2 metric tons).  Declines after the mid-1990s were
possibly related to shifts in environmental states (i.e., from mesotrophic to less favorable
oligotrophic conditions), variable reproductive success, influences from invading species, and
changing fisheries.  Recent improvements in abundance are due to a strong 2003 hatch  across the
Great Lakes Basin, presumably due to ideal weather conditions. Reproductive success has
remained very strong since 2003 in Saginaw Bay, and perhaps other parts of Lake Huron, and is
attributed to the decline of alewives in that lake during the same time period.  In general, walleye
yields peaked under ideal environmental conditions and declined under less favorable (i.e., non-
mesotrophic) conditions. Overall, environmental conditions remain improved relative to the
                         Draft for Discussion at SOLEC 2006

-------

1960s and early 1970s but concerns about food web disruption, pathogens (e.g., botulism,
viruses), noxious algae, and watershed management practices persist.

Pressures
Natural, self-sustaining walleye populations require adequate spawning and nursery habitats. In
the Great Lakes, these habitats exist in tributary streams and nearshore reefs, wetlands, and
embayments, and they have been used by native walleye stocks for thousands of years.
Degradation or loss of these habitats is the primary concern for the health of walleye populations
and can result from both human causes, as well as from natural environmental variability.
Increased human use of nearshore and watershed environments continues to alter the natural
hydrologic regime, affecting water quality (i.e., sediment loads) and rate of flow. Environmental
factors that affect precipitation patterns ultimately alter water levels, water temperature, water
clarity and flow. Thus, global warming and its subsequent effects on temperature and
precipitation in the Great Lakes basin may become increasingly important determinants of
walleye health. Non-native invaders, like zebra and quagga mussels, ruffe,  and round gobies
continue to disrupt the efficiency of energy transfer through the food web, potentially affecting
growth and survival of walleye and other fishes through a reduced supply of food.  Recent
experience in Lake Huron has elevated the concern over the predatory and  competitive effects of
the non-native alewife on walleye. In their absence, walleye reproductive success has surged,
indicating that the deleterious effect of alewife predation on larval walleye  may have been much
greater than previously realized. Alterations in the food web can also affect environmental
characteristics (like water clarity), which can in turn affect fish behavior and fishery yields.
Pathogens, like viral hemorrhagic septicemia and botulism, may also be affecting walleye
populations in some areas of the Great Lakes.

Management Implications
To improve the health of Great Lakes walleye populations, managers must enhance walleye
reproduction, growth and survival rates. Most walleye populations are dependent on natural
reproduction, which is largely driven by uncontrollable environmental events (i.e., spring weather
patterns and alewife abundance).  However, a lack of suitable spawning and nursery habitat is
limiting walleye reproduction in some areas due to human activities and can be remedied through
such actions as dam removal, substrate enhancement or improvements to watersheds to reduce
siltation and restore natural flow conditions. Growth rates are dependent on weather (i.e., water
temperatures), quality of the prey base, and walleye density, most of which are not directly
manageable. Survival rates can be altered through fishery harvest strategies, which are generally
conservative across all of the Great Lakes. Continued interactions between land managers and
fisheries managers to protect and restore natural habitat conditions in mesotrophic areas of the
Great Lakes are essential for the long term health of walleye populations. Elimination of
additional introductions of invasive species and control of existing non-native species, where
possible, is also critical to future health of walleyes and other native species.

Comments from the author(s)
Fishery yields are appropriate indicators of walleye health but only in a general sense. Yield
assessments are lacking for some fisheries (recreational, commercial, or tribal) or in some years
for all of the areas. Moreover, measurement units are not standardized among fishery types (i.e.,
commercial fisheries are measured in pounds while recreational fisheries are typically measured
in numbers), which means additional conversions are necessary and may introduce errors.  Also,
                         Draft for Discussion at SOLEC 2006

-------
"zero" values are not differentiated from "missing" data in the figure. Therefore, trends in yields
across time (blocks of years) are probably better indicators than absolute values within any year,
assuming that any introduced bias is relatively constant over time. Given the above, I recommend
a 10-year reporting cycle on this indicator. Many agencies have developed, or are developing,
population estimates for many Great Lakes fishes. Walleye population estimates for selected
areas (i.e., Lake Erie,  Saginaw Bay, Green Bay, and Bay of Quinte) would probably be a better
assessment of walleye population health in the Great Lakes than harvest estimates across all lakes
and I recommend switching to them as they become available in all areas.

Acknowledgments
Author: Roger Knight, Ohio Department of Natural Resources.

Data Sources
Fishery harvest data were obtained from the following sources:
Lake Superior: Ken Cullis, OMNR, ken.cullis@mnr.gov.on.ca
Lake Superior/Michigan/Huron: Karen Wright, CORA, kwright@sault.com
Lake Michigan:  Kevin Kapuscinski, WDNR, Kevin.Kapuscinski@dnr.state.wi.us
Lake Huron: Lloyd Mohr, OMNR, lloyd.mohr@mnr.gov.on.ca
Lake Huron: David Fielder, MDNR, fielderd@michigan.gov
Lake Erie: Roger Knight, ODNR, roger.knight@dnr.state.oh.us
Lake Ontario: Jim Hoyle, OMNR. jim.hoyle@mnr.gov.on.ca
Lake Ontario: Steve Lapan, NYSDEC, srlapan@gw.dec.state.ny.us

Various annual Lake Erie fisheries reports from the Ontario Ministry of Natural Resources, Ohio
Department of Natural Resources, and the Great Lakes Fishery Commission commercial fishery
data base were used as data sources.

Fishery data should not be used for purposes outside of this document without first
contacting the agencies that collected them.

List of Figures
Figure 1.  Recreational, commercial, and tribal harvest of walleye from the Great Lakes. Fish
Community Goals and Objectives are: Lake Michigan, 100-200 metric tons; Lake Huron, 700
metric tons; Lake Erie, sustainable harvest in all basins.
Source: Chippewa Ottawa Resource Authority, Michigan Department of Natural Resources, New
York State Department of Environmental Conservation, Ontario Ministry of Natural Resources,
Ohio Department of Natural Resources , Wisconsin Department of Natural Resources

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes  2007 - Draft
              Lake Superior
  u

  I
                            n Tribal
                            D Recreational
                            • Commercial
                     Year

                Lake Huron
    400


    300
n Tribal
o Recreational
• Commercial
                                                Lake Michigan
                                                                D Tribal
                                                                D Recreational
                                                                • Commercial
                                             §»
                                                       Year

                                                  Lake Erie
                                                                          D Recreational
                                                                          • Commercial
                     Year
                                                                Year
              Lake Ontario
                               D Tribal
                               n Recreational
                                 Commercial
                     Year
Figure 1. Recreational, commercial, and tribal harvest of walleye from the Great Lakes. Fish

Community Goals and Objectives are: Lake Michigan, 100-200 metric tons; Lake Huron, 700

metric tons; Lake Erie, sustainable harvest in all basins.
Source: Chippewa Ottawa Resource Authority, Michigan Department of Natural Resources, New

York State Department of Environmental Conservation, Ontario Ministry of Natural Resources,

Ohio Department of Natural Resources , Wisconsin Department of Natural Resources
                        Draft for Discussion at SOLEC 2006

-------
Preyfish Populations
Indicator #17

Overall Assessment
           Status:  Mixed
           Trend:  Deteriorating

Lake-by-Lake Assessment
Lake Superior
           Status:  Mixed
           Trend:  Improving

Lake Michigan
           Status:  Mixed
           Trend:  Deteriorating
Lake Huron
Lake Erie
           Status:  Mixed
           Trend:  Deteriorating
           Status:  Mixed
           Trend:  Deteriorating
Lake Ontario
           Status:  Mixed
           Trend:  Deteriorating

Purpose
• To assess the abundance and diversity of preyfish populations; and
• To infer the stability of predator species necessary to maintain the biological integrity of each
lake.

Ecosystem Objective
The importance of preyfish populations to support healthy, productive populations of predator
fishes is recognized in the Fish Community Goals and Objectives for each lake. For example, the
fish community objectives for Lake Michigan specify that in order to restore an ecologically
balanced fish community, a diversity of prey species at population levels matched to primary
production and predator demands must be maintained. This indicator also relates to the 1997
Strategic Great Lakes Fisheries Management Plan Common Goal Statement for Great Lakes
fisheries agencies.

State of the Ecosystem
Background
The preyfish assemblage forms important trophic links in the aquatic ecosystem and constitutes
the majority of the fish production in the Great Lakes. Preyfish populations  in each of the lakes
are currently monitored on an annual basis in order to quantify the population dynamics of these
                         Draft for Discussion at SOLEC 2006

-------

important fish stocks leading to a better understanding of the processes that shape the fish
community and to identify those characteristics critical to each species. Populations of lake trout,
Pacific salmon, and other salmonids have been established as part of intensive programs designed
to rehabilitate (or develop new) game fish populations and commercial fisheries. These
economically valuable predator species sustain increasingly demanding and highly valued
fisheries, and information on their status is crucial. In turn, these apex predators are sustained by
preyfish populations. In addition, some preyfishes, such as the bloater and the lake herring, which
are native species, and the rainbow smelt, which is non native, are also directly important to the
commercial fishing industry. Therefore, it is very important that the current status and estimated
carrying capacity of the preyfish populations be fully understood in order to fully address (1) lake
trout restoration goals, (2) stocking projections, (3) present levels of salmonid abundance and (4)
commercial fishing interests.

The component of the Great Lakes' fish communities that we classify as preyfish comprises
species - including both pelagic and benthic species - that prey on invertebrates for their entire
life history. As adults, preyfish depend on diets of crustacean zooplankton and macroinvertebrates
Diporeia and Mysis. This convention also supports the recognition of particle-size distribution
theory and size-dependent ecological processes. Based on size-spectra theory, body size is an
indicator of trophic  level, and the smaller, short-lived fish that constitute the planktivorous fish
assemblage discussed here are a discernable trophic group of the food web. At present, bloaters
(Coregonus hoyi), lake herring (Coregonus artedi), rainbow smelt (Osmerus mordax), alewife
(Alosapseudoharengus}, and deepwater sculpins (Myoxocephalus thompsonif), and to a lesser
degree species like lake whitefish (Coregonus dupeaformis), ninespine stickleback (Pungitius
pungitius), round goby (Neogobius melanostomus) and slimy sculpin (Coitus  cognatus) constitute
the bulk of the preyfish communities (Figure 1). The successful colonization of Lakes Michigan,
Huron, Erie, and Ontario by non-native dreissenids, notably the zebra mussel  (Dreissena
polymorphd) in the  early 1990s and more recently the quagga mussel (Dreissena bugensis), has
had a significant impact on the trophic structure of those lakes by shunting pelagic planktonic
production to mussels, an energetic dead end in the food chain as few native fishes can eat the
mussels. As  a result of profound ongoing changes  in trophic structure in four Great Lakes, these
ecosystems will continue to change, and likely in unpredictable ways. In Lake Erie, the preyfish
community is unique among the Great Lakes in that it is characterized by relatively high species
diversity. The preyfish community comprises primarily gizzard shad (Dorosoma cepedianuni)
and alewife (grouped as clupeids); emerald (Notropis atherinoides) and spottail shiners (TV.
hudsonius), silver chubs (Hybopsis storeriand), trout-perch (Percopsis omiscomaycus), round
gobies, and rainbow smelt (grouped as soft-rayed); and age-0 yellow (Percaflavescens) and
white perch (Morone americand), and white bass (M. chrysops) (grouped as spiny-rayed).

State of Preyfish Populations
Lake Ontario: Mixed, deteriorating
The non-native alewife, and to a lesser degree non-native rainbow smelt, dominate the preyfish
community. Their populations remain  at levels well below that of the early 1980s.  Rainbow
smelt have an abbreviated age and size structure that suggests the population is under heavy
predation pressure.  Abundance of the  non-native round goby is increasing and round goby have
the potential to negatively impact native, bottom-dwelling, preyfishes such as slimy and
deepwater sculpins, and trout-perch.  Deepwater sculpin, not reported from the lake since 1972,
                         Draft for Discussion at SOLEC 2006

-------
were collected sporadically in 1996-2004.  During 2005-2006, catches of deepwater sculpin
increased and juveniles dominated the catches suggesting that the long-depressed population was
recovering. Deepwater ciscoes, however, have not been reported from the lake since 1983 and
the large area of the lake they once occupied is largely devoid offish for much of the year.

Lake Erie: Mixed, deteriorating
The preyfish community in all three basins of Lake Erie has shown declining trends. In the
eastern basin, rainbow smelt (part of soft-rayed group) have shown declines in abundance over
the past two decades. The declines have been attributed to lack of recruitment associated with
expanding Driessenid colonization and reductions in productivity. The western and central basins
also have shown declines in preyfish abundance associated with declines in abundance of age-0
white perch and rainbow smelt, although slight increases for white perch have been reported in
the past couple years. The clupeid component of the preyfish community is at the lowest level
observed since 1998 and well below the mean biomass during 1987-2005.. The biomass estimates
for western Lake Erie were based on data from bottom trawl catches, depth strata extrapolations
(0-6 m, and >6 m), and trawl net measurements using acoustic mensuration gear.

Lake Michigan: Mixed, deteriorating
Bloater abundance in Lake Michigan fluctuated greatly during 1973-2005, as the population
showed a strong recovery during the 1980s but rapidly declined during the late 1990s.  Bloaters
may be cycling in abundance with a period of about 30 years.  The substantial decline in alewife
abundance during the 1970s and early  1980s has been attributed to increased predation by salmon
and trout.  The deepwater sculpin population exhibited a strong recovery during the 1970s and
early 1980s, and this recovery has been attributed to the decline in alewife abundance.  Alewives
have been suspected of interfering with reproduction by deepwater sculpins by feeding upon
deepwater sculpin fry. Slimy sculpin abundance appeared to be primarily regulated by predation
by juvenile lake trout. Slimy sculpin is a favored prey of juvenile lake trout. Temporal trends in
abundance of rainbow smelt were difficult to interpret.  Yellow perch year-class strength in 2005
was the highest on record  dating back to 1973.  Thus, early signs of a recovery by the yellow
perch population in the main basin of Lake Michigan were evident. The first catch of round
gobies in our annual lakewide survey occurred in 2003, and round goby abundance in the main
basin of the lake has remained low through 2005.

Lake Huron: Mixed, deteriorating
The Lake Huron fish community changed dramatically during 2003-2006, primarily due a 99%
decline in alewife numbers. Loss of alewife appears due to heavy salmonid predation that resulted
from increased Chinook salmon abundance as a result of wild reproduction. Alewife decline was
followed immediately by increased reproduction of other fish species; record year classes of
walleye and yellow perch were produced in Saginaw Bay, while in the main basin increased
reproduction by bloaters (chubs), rainbow smelt, and deepwater sculpins was observed. In 2004,
USGS surveys captured 22 wild juvenile lake trout — more than had been captured in the 30 year
history of those surveys. However, despite increased reproduction by prey species, biomass
remains low because newly recruited fish are still small. No species has  taken the place of
alewife, and prey biomass has declined by over 65%. Salmon catch rates by anglers declined, as
did average size and condition of those fish. The situation is exacerbated by changes at lower
trophic levels. The deepwater amphipod Diporeia has declined throughout Lake Huron's main
basin, and the zooplankton community has grown so sparse that it resembles the assemblage
                         Draft for Discussion at SOLEC 2006

-------
found in Lake Superior. The reasons underlying these changes are not known, but the most
widely held hypothesis is that zebra and quagga mussels are shunting energy into pathways that
are no longer available to fish.

Lake Superior: Mixed, improving
Since 1994, biomass of the Lake Superior preyfish has declined compared to the peak years in
1986, 1990, and 1994, a period when lake herring was the dominant preyfish species and wild
lake trout populations were starting to recover. Since the early 1980s, dynamics in preyfish
biomass have been driven largely by variation in recruitment of age-1 lake herring. Strong year
classes in 1984, 1988-1990, 1998, and most recently 2003 were largely responsible for peaks in
lake herring biomass in 1986, 1990-1994, 1999, 2004-2005. Prior to 1984, the nonnative rainbow
smelt was the dominant preyfish, but fluctuating population levels and recovery of native
coregonids after 1984 resulted in reduced biomass and rank among preyfish species. During
2002-2004, rainbow smelt biomass declined to the lowest levels in the time series, though a
moderate recovery occurred in 2005. There is strong evidence that declines in rainbow smelt
biomass are tied to increased predation by recovered lake trout populations. Biomass of bloater
and lake whitefish has increased since the early 1980s, and  biomass for both species has been less
variable than that of lake herring. Other preyfish species, notably sculpins, burbot, and ninespine
stickleback have declined in abundance since the recovery of wild lake trout populations in the
mid-1980s. Thus, the current state of the Lake Superior preyfish community appears to be largely
the result of increased predation by recovered wild lake trout stocks and, to a lesser degree, the
resumption of human harvest of lake trout, lake herring, and lake whitefish.

Pressures
The influences of predation by salmon and lake trout on preyfish populations appear to be
common across all lakes. Additional pressures from Dreissena, which is linked to the collapse of
Diporeia are strong in all lakes save Superior. Bottom-up effects on the preyfishes have already
been observed in Lakes Ontario, Huron, and Michigan suggesting that dynamics of preyfish
populations in those lakes could be driven by bottom-up rather than top-down effects in future
years,  Moreover,  the effect of non-native zooplankters, Bythotrephes and Cercopagis, on
preyfish populations,  although not fully understood at present, has the potential to increase
bottom up pressure.

Management Implications
Recognition of significant predation effects on preyfish populations has resulted in recent salmon
stocking cutbacks in Lakes Michigan and Huron and only minor increases in Lake Ontario.
However, even with a reduced population, alewives have exhibited the ability to produce strong
year classes when climatic conditions are favorable such that the continued judicious use of
artificially propagated predators seems necessary to avoid domination by alewife. It should be
noted that this is not an option in Lake Superior because lake trout and salmon are almost entirely
lake-produced. Potential bottom-up effects on preyfishes would be difficult to mitigate owing to
our inability to affect changes. This scenario only reinforces the need to avoid further
introductions of exotics into the Great Lake ecosystems.
                         Draft for Discussion at SOLEC 2006

-------


Comments from the author(s)
It has been proposed that in order to restore an ecologically balanced fish community, a diversity
of prey species at population levels matched to primary production and predator demands must be
maintained. However, the current mix of native and naturalized prey and predator species, and the
contributions of artificially propagated predator species into the system confound any sense of
balance in lakes other than Superior. The metrics of ecological balance as the consequence of fish
community structure are best defined through food-web interactions. It is through understanding
the exchanges of trophic supply and demand that the fish community can be described
quantitatively and ecological attributes such as balance can be better defined and the limits
inherent to the ecosystem realized.

Continued monitoring of the fish communities and regular assessments of food habits of
predators and preyfish will be required to quantify the food-web dynamics in the Great Lakes.
This recommendation is especially supported by continued changes that are occurring not only in
the upper but also in the lower trophic levels. Recognized sampling limitations of traditional
capture techniques (bottom trawling) have prompted the application of acoustic techniques as
another means to estimate absolute abundance of preyfishes in the Great Lakes. Though not an
assessment panacea, hydro-acoustics have provided additional insights and have demonstrated
utility in the estimates of preyfish biomass.

Protecting or reestablishing rare or extirpated members of the once prominent native preyfishes,
most notably the various members of the whitefish family (Coregonus spp.), should be a priority
in all the Great Lakes but especially in Lake Ontario where vast areas of the lake once occupied
by extirpated deepwater ciscoes are devoid of fish for much of the year. This recommendation
should be reflected in future indicator reports. Lake Superior, whose preyfish assemblage is
dominated by indigenous species and retains a full complement of ciscoes, should be examined
more closely to better understand the trophic ecology of its more natural system.

With the continuous nature of changes that seems to  characterize the preyfishes, and the lower
trophic levels on which they depend, the appropriate frequency to review this indicator is on a 5-
year basis.

Acknowledgments
Author: Owen T. Gorman, U.S. Geological Survey, Great Lakes Science Center, Lake Superior
Biological Station, Ashland, WI.

Contributors: Robert O'Gorman and Maureen Walsh, U.S. Geological Survey (USGS) Great
Lakes Science Center, Lake Ontario Biological Station, Oswego, NY;
Charles Madenjian and  Jeff Schaeffer, USGS Great Lakes Science Center, Ann Arbor, MI;
Mike Bur and Marty Stapanian, USGS Great Lakes Science Center, Lake Erie Biological Station,
Sandusky, OH;
Jeffrey Tyson, Ohio Division of Wildlife, Sandusky Fish Research Unit, Sandusky, OH;
Steve LaPan, New York State Department of Environmental Conservation, Cape Vincent
Fisheries Research Station, Cape Vincent, NY

Data Sources
                         Draft for Discussion at SOLEC 2006

-------
Bur, M.T., Stapanian, M.A., Kocovsky, P.M., Edwards, W.H.,and Jones, J.M. 2006.
Surveillance and Status of Fish Stocks in Western Lake Erie, 2005.  U.S. Geological Survey,
Great Lakes Science Center, Lake Erie Biological Station, 6100 Columbus Avenue, Sandusky,
OH 44870, USA. Available online:
http://www.glsc.usgs.gov/files/reports/2005LakeErieReport.pdf

Lantry, B. F., O'Gorman, R., Walsh, M. G., Hoyle, J. A., Keir, M. J., and Lantry, J. R.
Reappearance of Deepwater Sculpin in Lake Ontario: Start of a Resurgence or Last Gasp of a
Doomed Population? Journal of Great Lakes Research: Submitted

Lantry, B. F., O'Gorman, R. 2006. Evaluation of Offshore Stocking to Mitigate Piscivore
Predation on Newly Stocked Lake Trout in Lake Ontario.  U.S. Geological Survey (USGS)
Lake Ontario Biological Station, 17 Lake St., Oswego, NY 13126. Available online:
http://www.glsc.usgs.gov/files/reports/2005NYSDECLakeOntarioReport.pdf

Madenjian, C. P., Fahnenstiel, G. L., Johengen, T. H., Nalepa, T. F., Vanderploeg, H. A.,
Fleischer, G. W., Schneeberger, P. J., Benjamin, D. M., Smith, E. B., Bence, J. R., Rutherford, E.
S., Lavis, D. S., Robertson, D. M., Jude, D. J., and Ebener, M. A. 2002. Dynamics of the Lake
Michigan food web, 1970-2000. Can. J. Fish. Aquat. Sci. 59: 736-753.

Madenjian, C. P., Hook, T.  O., Rutherford, E. S., Mason, D. M., Croley, T. E., II, Szalai, E. B.,
and Bence, J. R. 2005. Recruitment variability of alewives in Lake Michigan. Trans. Am. Fish.
Soc. 134:218-230.

Madenjian, C. P., Hondorp, D. W., Desorcie, T. J., and Holuszko, J. D.  2005. Sculpin
community dynamics in Lake Michigan.  J. Gt. Lakes Res. 31: 267-276.

Madenjian, C. P., Bunnell, D.  B., Desorcie, T. J., Holuszko, J. D., and Adams, J. V. 2006.  Status
and trends of preyfish populations in Lake Michigan, 2005. U. S. Geological Survey, Great
Lakes Science Center, Ann Arbor, Michigan. Available online:
http://www.glsc.usgs.gov/files/reports/2005LakeMichiganReport.pdf

Murray, C.,  R. Haas, M. Bur, J. Deller, D. Einhouse, T. Johnson, J. Markham, L. Rudstam, M.
Thomas, E. Trometer, J. Tyson, and L. Witzel.  2006. Report of the Forage Task Group to the
Standing Technical Committee of the Lake Erie Committee, Great Lakes Fishery Commission.
38pp.

O'Gorman, R., Gorman, O., and Bunnel, D.  2006.  Great Lakes Prey Fish Populations:
A Cross-Basin View of Status and Trends in 2005.  U.S. Geological Survey, Great Lakes Science
Center, Deepwater Science Group, 1451 Green Rd, Ann Arbor, MI  48105.  Available on line:
http://www.glsc.usgs.gov/files/reports/2005GreatLakesPrevfishReport.pdf

O'Gorman, R., Walsh, M. G., Strang, T., Adams, J. V., Prindle, S. E., and Schaner, T. 2006.
Status of alewife in the U.S. waters of Lake Ontario, 2005. Annual Report Bureau of Fisheries
Lake Ontario Unit and St. Lawrence River Unit to Great Lakes Fishery Commission's Lake
                        Draft for Discussion at SOLEC 2006

-------

Ontario Committee.  March 2006.  Section 12, 4-14.  Available online:
http://www.glsc.usgs.gov/files/reports/2005LakeOntarioPreyfishReport.pdf

Roseman, E.F., Schaeffer, J.S, French III, J.R.P., O'Brien, T.P., and Paul, C.S. 2006. Status and
Trends of the Lake Huron Deepwater Demersal Fish  Community, 2005.  U.S. Geological Survey,
Great Lakes Science Center, 1451 Green Rd, Ann Arbor, MI 48105. Available online:
http://www.glsc.usgs.gov/files/reports/2005LakeHuronDeepwaterReport.pdf

Schaeffer, J.S., O'Brien, T.P., Warner, D.M., and Roseman, E.F.  2006.  Status and Trends of
Pelagic Prey Fish in Lake Huron, 2005: Results From a Lake-Wide Acoustic Survey. U.S.
Geological Survey, Great Lakes Science Center, 1451 Green Rd, Ann Arbor, MI 48105.
Available online: http://www.glsc.usgs.gov/files/reports/2005LakeHuronPrevfishReport.pdf

Stockwell, J.D., Gorman, O.T., Yule, D.L., Evrard, L.M., and Cholwek, G.M.  2006. Status and
Trends of Prey Fish Populations in Lake Superior, 2005. U.S. Geological Survey, Great Lakes
Science Center, Lake Superior Biological Station, 2800 Lake Shore Dr. E., Ashland, WI 54806.
Available online: http://www.glsc.usgs.gov/files/reports/2005LakeSuperiorPreyfishReport.pdf

Strang, T., Maloy, A. and Lantry, B. F. 2006. Mid-lake assessment in the U.S. waters of Lake
Ontario, 2005.  Annual Report Bureau of Fisheries Lake Ontario Unit and St. Lawrence River
Unit to Great Lakes Fishery Commission's Lake Ontario Committee. March 2006.  Section 12,
32-34.  Available online:
http://www.glsc.usgs.gov/files/reports/2005LakeOntarioPrevfishReport.pdf

Walsh, M. G., O'Gorman, R., Maloy, A. P., and Strang, T. 2006. Status of rainbow smelt in the
U.S. waters of Lake Ontario, 2005.  Annual Report Bureau of Fisheries Lake Ontario Unit and St.
Lawrence River Unit to Great Lakes Fishery Commission's Lake Ontario Committee. March
2006. Section 12, 15-19. Available online:
http://www.glsc.usgs.gov/files/reports/2005LakeOntarioPrevfishReport.pdf

Walsh, M. G., O'Gorman, R., Lantry, B. F., Strang, T.,  and Maloy, A. P.  2006. Status of
sculpins and round goby in the U.S. waters of Lake Ontario, 2005. Annual Report Bureau of
Fisheries Lake Ontario Unit and St. Lawrence River Unit to Great Lakes Fishery Commission's
Lake Ontario Committee. March 2006.  Section 12, 20-31.  Available online:
http://www.glsc.usgs.gov/files/reports/2005LakeOntarioPrevfishReport.pdf

Warner, D.M., Randall M. Claramunt, R.M., and Paul, C.S. 2006. Status of Pelagic Prey Fishes
in Lake Michigan, 1992-2005.  Geological Survey, Great Lakes Science Center, 1451 Green Rd,
Ann Arbor, MI  48105. Available online:
http://www.glsc.usgs.gov/files/reports/2005LakeMichiganPreyfishReport.pdf

List of Figures
Figure  1. Preyfish trends based on annual bottom trawl  surveys. All trawl surveys were performed
by USGS - Great Lakes Science Center, except for Lake Erie, which was conducted by the
USGS, Ohio Division of Wildlife and the Ontario Ministry of Natural Resources (Lake Erie
Forage Task  Group), and Lake Ontario, which was conducted jointly by USGS and the New York
State Department of Environmental Conservation.
                        Draft for Discussion at SOLEC 2006

-------
Sources: U.S. Geological Survey - Great Lakes Science Center, Ohio Division of Wildlife,
Ontario Ministry of Natural Resources, and New York State Department of Environmental
Conservation.

Last updated
SOLEC 2006
                       Draft for Discussion at SOLEC 2006

-------
          State of the Great Lakes 2007 - Draft
                                                                     • Round Goby
                                                                     D Trout Fterch
                                                                     D Stickleback
                                                                     D Sculpin
                                                                     D Bloater
                                                                     • Rainbow Smelt
                                                                     D Alew if e
D Lake Wnitefisn
D Bloater
• Rainbow Smelt
D Lake Herrin
      1978 1981  1984 1987 1990  1993 1996 1999  2002 2005
                                                                    • Rainbow Srrelt
                                                                    D Alew if e
                                             1978 1981 1984 1987 1990 1993  1996 1999  2002 2005
                                                             Year
       • Slimy Sculpin
       D Deepw ater Sculpin
       D Bloater
       • Rainbow Smelt
       D Alew ife
                                        D Spiny-rayed
                                        • Soft-rayed
                                        D Clupeids
     1973 1976 1979 1982 1985 1988 1991 1994 1997 2000 2003
                      Year
                          1993  1995
                               Year
                                                                      2001 2003 2005
Figure 1. Preyfish trends based on annual bottom trawl surveys. All trawl surveys were
performed by USGS - Great Lakes Science Center, except for Lake Erie, which was conducted by
the USGS, Ohio Division of Wildlife and the Ontario Ministry of Natural Resources (Lake Erie
Forage Task Group),  and Lake Ontario, which was conducted jointly by USGS and the New York
State Department of Environmental Conservation.
Sources: U.S. Geological Survey - Great Lakes Science Center, Ohio Division of Wildlife,
Ontario Ministry of Natural Resources, and New York State Department of Environmental
Conservation.
                          Draft for Discussion at SOLEC 2006

-------
                STATE   OP
Itiinii ,f	
i«,{t»*';'<,l;
Sea Lamprey
Indicator #18

Assessment: Good/Fair, Improving

Purpose
  To estimate the abundance of sea lamprey as an indicator of
the status of this invasive species; and
  To infer the damage sea lamprey cause to the fish communi-
ties and aquatic ecosystems of the Great Lakes.

Ecosystem Objective
The  1955 Convention of Great Lakes Fisheries created the Great
Lakes Fishery Commission (GLFC) "to formulate and imple-
ment a comprehensive program for the purpose of eradicating or
minimizing the sea lamprey populations in the Convention area"
(GLFC 1955). Under the Joint Strategic Plan for Great Lakes
Fisheries, all fishery management agencies established Fish
Community Objectives (FCOs) for each of the lakes. These
FCOs call for suppressing sea lamprey populations to levels that
cause only insignificant mortality of fish in order to achieve
objectives for lake trout and other members of the fish commu-
nity (Horns et al. 2003, Eshenroder et al. 1995, DesJardin et al.
1995, Ryan et al. 2003., Stewart et al. 1999).

The GLFC and fishery management agencies have agreed on tar-
get abundance levels for sea lamprey populations that corre-
spond to the FCOs (Table 1). Targets were derived from avail-
able  estimates of the abundance of spawning-phase  sea lampreys
and from data on wounding rates on lake trout. Suppressing sea
lampreys to abundances within the target range is predicted to
result in tolerable mortality on lake trout and other fish species.
Lake
Superior
Michigan
Huron
Erie
Ontario
FCO Sea Lamprey
Abundance Targets
35,000
58,000
74,000
3,000
29,000
Target Range (+/- 95%
Confidence Interval)
18,000
13,000
20,000
1,000
4,000
Table 1 . Fish Community Objectives for sea lamprey
abundance targets.
Source: Great Lakes Fishery Commission
State of the Ecosystem
Background
Populations of the native top predator, lake trout, and other fish-
es are negatively affected by mortality caused by sea lamprey.
The first complete round of stream treatments with the lampri-
cide TFM, as early as 1960 in Lake Superior, successfully sup-
pressed sea lamprey to less than 10% of their pre-control abun-
dance in all of the Great Lakes.
        Mark and recapture estimates of the abundance of sea lamprey
        migrating up rivers to spawn are used as surrogates for the abun-
        dance of parasites feeding in the lakes during the previous year.
        Estimates of individual spawning runs in trappable streams are
        used to estimate lake-wide abundance using a new regression
        model that relates run size to stream characteristics (Mullett et
        al. 2003). Sea lamprey spend one year in the lake after metamor-
        phosing, so this indicator has a two-year lag in demonstrating
        the effects of control efforts.

        Status of Sea Lamprey
        Annual lake-wide estimates of sea lamprey abundance since
        1980, with 95%  confidence intervals, are presented in Figure 1.
        The FCO targets and ranges also are included for each lake.

        Lake Superior. During the past 20 years, populations have fluc-
        tuated but remain at levels less than 10% of peak abundance
        (Heinrich et al. 2003). Abundances were within the FCO target
        range during the late 1980s  and mid-1990s. Abundances have
        trended upward from a low  during 1994 and have been above
        the target range from 1999-2003. These recent increases in abun-
        dance have raised concern in all waters. Rates of sea lamprey
        markings on fish have shown the same pattern of increase. These
        increases appear to be most dramatic in the Nipigon  Bay and
        north-western portion of the lake  and in the Whitefish Bay area
        in the south-eastern portion of the lake.  Survival  objectives for
        lake trout continue to be met but lake trout populations could be
        threatened if these increases continue. In response to this
        increased abundance of sea  lampreys, stream  treatments with
        lampricides were increased beginning in 2001 through 2004. The
        effects of the increased treatments during 2001 may  have  con-
        tributed to the downward trend in the 2003 observation. The
        effects of additional stream  treatments in 2002 and beyond will
        be observed in the spawning-run estimates during 2004  and fol-
        lowing years.

        Lake Michigan: The population of sea lamprey has shown a con-
        tinuing, slow trend upward since  1980 (Lavis et al. 2003). The
        population was at or below the FCO target range until 2000. The
        marking rates on lake trout have shown the same upward trend
        past target levels during the recent years. Increases in abundance
        during the 1990s had been attributed to the St. Marys River. The
        continuing trend in recent years suggests sources of  sea lamprey
        in Lake Michigan itself.  Stream treatments were increased
        beginning in 2001 through 2004. This increase included treat-
        ment of newly discovered populations in lentic areas and treat-
        ment of the Manistique River, a large system  where  the  deterio-
        ration of a dam near the mouth allowed sea lamprey access to
        nursery habitat. The 2003 spawning-phase population estimate
        did not show any decrease as a result of the increased treatments
        during 2001.
                                                                                                                     83

-------
                                              OF   THE   GREAT
                             2007
                             Superior
_ 500 n
i 4°°-

o 300-
2 200
i
1 10°:
Q.
0^




/






\ T TJ
	 /41 .. . .. .V-K^-^-i- 	 -\--s- •f^,--/-'- 	
................................ .....^r....... .».. ............ |
                                                                                         Erie*
                                1990      1995
                               SpawningYear


                            Michigan
     |  80-
     ra
     en
     |  60-

     £  40 -
     a
     c
     3  .
ers (thousands)
w w ji en
0000
1 1 1
                                                                                  55555555 ; V"- ^ ~  "~ "	 •	 ^aViPi 5 i i 5*5
                                                                                1985      1990      1995
                                                                                        Spawning Year


                                                                                       Ontario
                                1990      1995
                               Spawning Year
                            Huron
       500;

     •| 400 -
     o)
     o 300 -

     |200-

     5 100-
500-
400-
300-

200-
100:



/
/
>'"*

I

n
\ / ~\ /-^- \ » *
1 I-i y VAV\ yv i
y^ "- i * ^_i \/

                                                                       0 -
                                                                        1980
                                                                                         1990      1995
                                                                                         Spawning Year
                                1990      1995
                               Spawning Year
 Figure 1. Total abundance of sea lampreys estimated during the spawning migration. Solid line and dashed line represent FCO tar-
 get abundance and ranges, respectively.
 *Note: the scale for Lake Erie is 1/5 that of the other four Lakes.
 Source: Great Lakes Fishery Commission
Lafe Huron: The first full round of stream treatments during the
late 1960s suppressed sea lamprey populations to levels less than
10% of those before control (Morse et al. 2003). During the
early 1980s, abundance increased in Lake Huron, particularly
the northern portion of the lake, peaking in 1993. Through the
1990s there were more sea lampreys in Lake Huron than all the
other lakes combined. FCOs were not being achieved. The dam-
age caused by this large population of parasites was so severe
that the Lake Huron Committee abandoned its lake trout restora-
tion objective in the northern portion of the lake during 1995.
The St. Marys River was identified as the source of the increas-
ing sea lamprey population. The size of this connecting channel
made traditional treatment with the lampricide TFM impractical.
A new integrated control  strategy, including targeted application
of a new formulation of a bottom-release lampricide,  enhanced
trapping of spawning animals, and sterile-male release, was initi-
ated in 1997 (Schleen et al. 2003). As predicted, the spawning-
phase abundance has been significantly lower since 2001 as a
result of the completion of the first full round of lampricide spot
treatments during 1999. However, the population shows consid-
erable variation and it increased during 2003. Wounding rates
and mortality estimates for lake trout have also declined during
the last three years. The full effect of the St. Marys River control
program will not be observed for another 2-4 years (Adams et
al. 2003). The GLFC has repeated lampricide treatments in lim-
ited areas with high densities of larvae during 2003 and 2004.
These additional treatments are aimed at continuing the decline
in sea lamprey in Lake Huron.

Lake Erie: Following the completion of the first full round of
84

-------
stream treatments in 1987, sea lamprey populations collapsed
(Sullivan et al. 2003). Marking rates on lake trout declined and
lake trout survival increased to levels sufficient to meet the reha-
bilitation objectives in the eastern basin. However, during the
mid-1990s, sea lamprey abundance increased to levels that
threatened the lake trout restoration effort. A major assessment
effort during  1998 indicated that the source of this increase was
several streams in which treatments had been deferred due to
low water flows or concerns for non-target organisms. These
critical streams were treated during 1999 and 2000. Sea lamprey
abundance was observed to decline to target levels in 2001
through 2003. Wounding rates on lake trout have also declined.

Lake Ontario: Abundance of spawning-phase sea lamprey has
shown a continuing declining  trend since the early 1980s
(Larson et al. 2003). The  abundance of sea lamprey has
remained stable in the FCO target range during 2000-2003.

Pressures
Since parasitic-phase sea lamprey are at the top of the aquatic
food chain and inflict high mortality on large  piscivores, popula-
tion control is essential for healthy fish communities. Increasing
abundance in Lake Erie demonstrates how short lapses in control
can result in rapid increases in abundance and that continued
effective stream treatments are necessary to overcome the repro-
ductive potential of this invading species. The potential for sea
lamprey to  colonize new locations is increased with improved
water quality and removal of dams. For example, the loss of
integrity of the dam on the Manistique River, and subsequent
production from this river, has contributed to the increase in sea
lamprey abundance in Lake Michigan. Any areas newly infested
with sea lamprey will require  some form of control to attain tar-
get abundance levels in the lakes.

As fish communities recover from the  effects of sea lamprey
predation or over-fishing, there is evidence that the survival of
parasitic sea lamprey may increase due to prey availability.
Better survival means that there will be more  residual sea lam-
prey to cause harm. Significant additional control efforts, like
those on the St. Marys River, may be necessary to maintain sup-
pression.

The GLFC has a goal of reducing reliance on lampricides and
increasing efforts to integrate other control techniques, such as
the sterile-male-release technique or the installation of barriers
to stop the upstream migration of adults. Pheromones that affect
migration and mating have been discovered and offer exciting
potential as new alternative controls. The use  of alternative con-
trols is consistent with sound practices of integrated pest man-
agement, but can put additional pressures on the ecosystem such
as limiting the passage of fish upstream of barriers. Care must be
taken in applying new alternatives or in reducing lampricide use
to not allow sea lamprey abundance to increase.

Management Implications
The GLFC has increased stream treatments and lampricide
applications in response to increasing abundances during 2001
through 2004. The GLFC has targeted these additional treat-
ments to maximize progress toward FCO targets. The GLFC
continues to focus on research and development of alternative
control strategies. Computer models, driven by empirical data,
are being used to best allocate treatment resources, and research
is being conducted to better understand and manage the variabil-
ity in sea lamprey populations.

Acknowledgments
Author: Gavin Christie, Great Lakes Fishery Commission, Ann
Arbor, MI.

Sources
Adams, J.V., Bergstedt, R.A., Christie, G.C., Cuddy, D.W.,
Fodale, M.F., Heinrich, J.W., Jones, M.L., McDonald, R.B.,
Mullett, K.M., and Young, RJ. 2003. Assessing assessment: can
we detect the expected effects of the St. Marys River sea lam-
prey control strategy? /. Great Lakes Res. 29 (Suppl. 1):717-
727.

DesJardine, R.L.,  Gorenflo, T.K., Payne, R.N., and Schrouder,
J.D.  1995. Fish-community objectives for Lake Huron. Great
Lakes Fish. Comm. Spec. Publ. 95-1.

Eshenroder, R.L., Holey, M.E., Gorenflo, T.K., and Clark, R.D.,
Jr. 1995. Fish-community objectives for Lake Michigan. Great
Lakes Fish. Comm. Spec. Publ. 95-3.

Great Lakes Fishery Commission (GLFC). 1955. Convention on
Great Lakes Fisheries, Ann Arbor, MI.

Heinrich, J.W., Mullett, K.M, Hansen, M.J., Adams, J.V., Klar,
G.T., Johnson, D.A., Christie, G.C., and Young, R.J. 2003. Sea
lamprey abundance and management in Lake Superior, 1957-
1999. /. Great Lakes Res. 29 (Suppl. l):566-583.

Horns, W.H., Bronte, C.R., Busiahn, T.R., Ebener, M.P,
Eshenroder, R.L., Greenfly, T., Kmiecik, N., Mattes, W., Peck,
J.W., Petzold, M., and Schneider, D.R. 2003. Fish-community
objectives for Lake Superior. Great Lakes Fish. Comm. Spec.
Publ. 03-01.

Larson, G.L., Christie, G.C.,  Johnson, D.A., Koonce, J.F.,
Mullett, K.M., and Sullivan, W.P 2003. The history of sea lam-
prey control in Lake Ontario  and updated estimates of suppres-
sion targets. /. Great Lakes Res. 29 (Suppl. l):637-654.
                                                                                                                       85

-------

Lavis, D.S., Hallett, A., Koon, E.M., and McAuley, T. 2003.
History of and advances in barriers as an alternative method to
suppress sea lampreys in the Great Lakes. /. Great Lakes Res. 29
(Suppl. l):584-598.

Morse, T.J., Ebener, M.P.,  Koon, E.M., Morkert, S.B., Johnson,
D.A., Cuddy, D.W., Weisser, J.W., Mullet, K.M., and Genovese,
J.H. 2003. A case history of sea lamprey control in Lake Huron:
1979-1999. J. Great Lakes Res.  29 (Suppl. 1):599-614.

Mullett, K M., Heinrich, J.W., Adams, J.V. Young, R. J., Henson,
M.P., McDonald, R.B., and Fodale, M.F. 2003. Estimating lake-
wide abundance of spawning-phase sea lampreys (Petromyzon
marinus) in the Great Lakes: extrapolating from sampled
streams using regression models. /. Great Lakes Res. 29 (Suppl.
l):240-253.

Ryan, P.S., Knight, R., MacGregor, R., Towns, G., Hoopes, R.,
and Culligan, W. 2003. Fish-community goals and objectives for
Lake Erie. Great Lakes Fish. Comm. Spec. Publ. 03-02.

Schleen, L.P., Christie, G.C., Heinrich, J.W., Bergstedt, R.A.,
Young, R.J., Morse, T.J., Lavis,  D.S., Bills, T.D., Johnson J., and
Ebener, M.P. 2003. In press. Development and implementation
of an integrated program for control of sea lampreys in the  St.
Marys River. J. Great Lakes Res. 29 (Suppl. l):677-693.

Stewart, T.J., Lange, R.E., Orsatti, S.D., Schneider, C.P.,
Mathers, A., and Daniels M.E. 1999. Fish-community objectives
for Lake Ontario. Great Lakes Fish. Comm. Spec. Publ. 99-1.

Sullivan, W.P., Christie, G.C., Cornelius, F.C., Fodale, M.F.,
Johnson, D.A.,  Koonce, J.F., Larson, G.L., McDonald, R.B.,
Mullet, K.M., Murray, C.K., and Ryan, PA. 2003. The sea lam-
prey in Lake Erie: a case history. /. Great Lakes Res. 29 (Suppl.
l):615-637.
                                                                                           2007
Last Updated
State of the Great Lakes 2005
Authors' Commentary
Targeted increases in lampricide treatments are predicted to
reduce sea lamprey abundance to acceptable levels. The effects
of increased treatments will be observed in this indicator two
years after they occur. Discrepancies among estimates of differ-
ent life-history stages need to be resolved. Efforts to identify all
sources of sea lamprey need to continue. In addition, research to
better understand lamprey/prey interactions, the population
dynamics of sea lamprey that survive control actions, and refine-
ment of alternative control methods are all key to maintaining
sea lamprey at tolerable levels.
86

-------
Native Freshwater Mussels
Indicator #68

Assessment: Not Assessed

Purpose
  To assess the location and status of freshwater mussel
(unionid) populations and their habitats throughout the Great
Lakes system, with emphasis on endangered and threatened
species; and
  To use this information to direct research aimed at identifying
the factors responsible  for mussel survival in refuge areas, which
in turn will be used to predict the locations of other natural sanc-
tuaries and guide their management for the protection and
restoration of Great Lakes mussels.

Ecosystem Objective
The objective is the restoration of the richness, distribution, and
abundance of mussels throughout the Great Lakes, which would
thereby reflect the  general health of the basin ecosystems. The
long-term goal is for mussel populations to be stable and self-
sustaining wherever possible throughout their historical range in
the Great Lakes, including the connecting channels and tributar-
ies.

State of the Ecosystem
Background
Freshwater mussels (Bivalvia: Unionacea) are of unique ecolog-
ical value as natural biological filters, food for fish and wildlife,
and indicators of good  water quality. In the United States, some
species are  commercially harvested for their shells and pearls.
These slow-growing, long-lived organisms can influence ecosys-
tem function such as phytoplankton ecology, water quality,  and
nutrient cycling. As our largest freshwater invertebrate, freshwa-
ter mussels may also constitute a significant proportion of the
freshwater invertebrate biomass where they occur. Because they
are sensitive to toxic chemicals, mussels may serve as an early-
warning system to  alert us of water quality problems. They are
also good indicators of environmental change due to their
longevity and sedentary nature.  Since mussels are parasitic  on
fish during  their larval  stage, they depend on healthy fish com-
munities for their survival.

The richness, distribution, and abundance of mussels reflect the
general health of the  aquatic ecosystems. Because their shells are
attractive and easy to find, they  were prized by amateur collec-
tors and naturalists in the past. As a result, many museums  have
extensive shell collections dating back 150 years or more that
provide us with an invaluable "window to the past" that is not
available for other  aquatic invertebrates.
Status of freshwater mussels
The abundance and number of species of freshwater mussels
have severely declined across North America, particularly in the
Great Lakes. Nearly 72% of the 300 species in North America
are vulnerable to extinction or already extinct. The decline of
unionids has been attributed to commercial exploitation, water
quality degradation (pollution, siltation), habitat destruction
(dams, dredging, channelization) riparian and wetland alter-
ations, changes in the distribution and/or abundance of host fish-
es, and competition with non-native species. In the Great Lakes
watershed, zebra mussels (Dreissena polymorphd) and, to a less-
er extent, quagga mussels (D. bugensis] have caused a severe
decline in unionid populations. Zebra mussels attach to a mus-
sel's shell, where they interfere with activities such  as feeding,
respiration and locomotion - effectively robbing it of the energy
reserves needed for survival and reproduction. Native mussels
are particularly sensitive to biofouling by zebra mussels and to
food competition with both zebra mussel and quagga mussels.

Many areas in the Great Lakes, such as Lake St. Clair and Lake
Erie, have lost over 99% of their native mussels of all species as
a result of the impacts of dreissenids. Although Lake Erie, Lake
St. Clair, and their connecting channels historically  supported a
rich mussel fauna of about 35 species, unionid mussels were
slowly declining in some areas even before the zebra mussel
invasion. For example, densities in the western basin of Lake
Erie decreased from 10 unionids/m2 in 1961 to 4/m2 in 1982,
probably due to poor water quality. In contrast, the impact of the
zebra mussel was swift and severe. Unionids were virtually
extirpated from the offshore waters of western Lake Erie by
1990 and from Lake St. Clair by 1994, with similar declines in
the connecting channels and many nearshore habitats. The aver-
age number of unionid species found in these areas  before the
zebra mussel invasion was 18 (Figure 1). After the invasion,
60% of surveyed sites had 3 or fewer species remaining, 40% of
sites had none left, and abundance had declined by 90-95%.

It was feared that unionid mussels would be extirpated from
Great Lakes waters by the zebra mussel. However, significant
communities were recently discovered in several nearshore areas
where zebra mussel infestation rates are low (Figure 1).

These remnant unionid populations, found in isolated habitats
such as river mouths and lake-connected wetlands, are at severe
risk.  Reproduction is occurring at some of these sites, but not all.
Further problems are associated with unionid species that were
in low numbers before the influx of the non-native dreissenids.
A number of species that are listed as endangered or threatened
in the United States or Canada are found in some of these isolat-
ed populations in the Great Lakes and in associated tributaries.
In the United States, these include the clubshell (Pleurobema
clava), fat pocketbook (Potamilus capax), northern riffleshell

                                                         87

-------
                                              OF   THE   GREAT
                                                                                          2007
                                                                                                 Port Maitland
                                   Lake St. Clair
   Grosse Point, Ml
        19911999
    Detroit River
                                         astern Snore
                                        Lake St. Clair
                                                   19™2001
                                                                      0   = no mussels

                                                                          = 10 species
                                                                                               hompson Bay Refuge
                                                                                         'resque Isle Bay
  1982-831992-94
Nearshore Western
Basin Refuge

      MetzgerMarsh
      Refuge
         /x
    Lake Erie SW Shore
           0
           1999    Sandusky Bay
                       2001
 Figure 1. Numbers of freshwater mussel species found before and after the zebra mussel invasion at 13 sites in Lake Erie, Lake St.
 Clair, and the Niagara and Detroit Rivers (no "before" data available for 4 sites), and the locations of the four known refuge sites
 (Thompson Bay, Metzger Marsh, Nearshore Western Basin, and St. Clair Delta).
 Source: Metcalfe-Smith, J.L., D.T. Zanatta, B.C. Masteller, H.L. Dunn, S J. Nichols, PJ. Marangelo, and D.W. Schloesser. 2002
(Epioblasma torulosa rangiana), and white catspaw
(Epioblasma obliquata perobliqua). In Canada, the northern rif-
fleshell, rayed bean (Villosafabalis), wavyrayed lampmussel
(Lampsilis fasciola), salamander mussel (Simpsonaias ambigua),
snuffbox (Epioblasma triquetra), round hickorynut (Obovaria
subrotunda), kidneyshell (Ptychobranchus fasciolaris) and round
pigtoe (Pleurobema sintoxia) are listed as endangered.

All of the refuge sites discovered to date have two characteristics
in common: they are very shallow (
-------
Pressures
Zebra mussel expansion is the main threat facing unionids in the
Great Lakes drainage basin. Zebra mussels are now found in all
of the Great Lakes and in many associated water bodies, includ-
ing at least 260 inland lakes and river systems such as the
Rideau River in Ontario and in two reservoirs in the Thames
River drainage in Ontario.

Other non-native species may also impact unionid survival
through the reduction or redistribution of native fishes. Non-
native fish  species such as the Eurasian ruffe (Gymnocephalus
cernuus] and round goby (Neogobius melanostomus) can com-
pletely displace native fish, thus causing the functional extirpa-
tion of local unionid populations.

Continuing changes in land use  (increasing urban sprawl, growth
of factory farms, etc.), elevated  use of herbicides to remove
aquatic vegetation from lakes for recreational purposes, climate
change and the associated lowering of water levels, and many
other factors will continue to have an impact on unionid popula-
tions in the future.

Management Implications
The long-term goal is for unionid mussel populations to be sta-
ble  and self-sustaining wherever possible throughout their histor-
ical range in the Great Lakes, including the connecting channels
and tributaries.  The most urgent activity is to prevent the  further
introduction of non-native species into the Great Lakes. A sec-
ond critical activity is to prevent the further expansion of non-
native species into the river systems and inland lakes of the
region where they may seriously harm the remaining healthy
populations of unionids that could be used to re-inoculate the
Great Lakes themselves  in the future.

To ensure the survival of remaining unionids in the  Great Lakes
basin, and to foster the restoration of their populations to  the
extent possible, the following actions are recommended:

      All existing information on the  status of freshwater mus-
    sels throughout the Great Lakes drainage basin should be
    compiled and reviewed. A complete analysis of trends over
    space and time is needed to properly assess the current
    health  of the fauna.
    comparability of data. The Freshwater Mollusk
    Conservation Society has prepared a peer-reviewed, state-
    of-the art protocol that should be consulted for guidance
    (Strayer and Smith 2003). Populations of endangered and
    threatened species should be specifically targeted.

      The locations of all existing refugia, both within and out-
    side of the influence of zebra mussels, should be document-
    ed,  and they must be protected by all possible means from
    future disturbance.

      Research is needed to determine the mechanisms respon-
    sible for  survival of unionids in the various refuge sites, and
    this knowledge should be used to predict the locations of
    other refugia and to guide their management.

      The environmental requirements of unionids need to be
    taken into account in wetland restoration projects.

      All avenues for educating the public about the plight of
    unionids  in the  Great Lakes should be pursued, as  well as
    legislation for their protection. This includes ensuring that
    all species that  should be listed are listed as quickly as pos-
    sible.

      The principles of the National Strategy for the
    Conservation of Native Freshwater Mussels (The National
    Native Mussel Conservation Committee 1998) should be
    applied to the conservation and protection of the Great
    Lakes unionid fauna.

Acknowledgments
Authors: Janice L. Smith, Biologist, Aquatic Ecosystem Impacts
Research Branch, National Water Research Institute, Burlington,
ON, Janice.Smith@ec.gc.ca; and
S. Jerrine Nichols, U.S. Geological Survey, Biological  Resources
Division, Ann Arbor, MI, sjerrine_nichols@usgs.gov.

Sources
Bowers, R., and De Szalay, F. 2003. Effects of hydrology on
unionids (Unionidae) and zebra mussels (Dreissenidae) in a
Lake Erie coastal wetland. American Midland Naturalist
151:286-300.
      To assist with the above exercise, and to guide future sur-
    veys, all data must be combined into a computerized, GIS-
    linked database (similar to the 8000-record Ontario database
    managed by the National Water Research Institute), accessi-
    ble to all relevant jurisdictions.

      Additional surveys are needed to fill data gaps, using
    standardized sampling designs and methods for optimum
Martel, A.L., Pathy, D.A., Madill, J.B., Renaud, C.B., Dean,
S.L., and Kerr, S.J. 2001. Decline and regional extirpation of
freshwater mussels (Unionidae) in a small river system invaded
by Dreissenapolymorpha: the Rideau River, 1993-2000. Can. J.
Zoo/. 79(12):2181-2191.

Metcalfe-Smith, J.L., Zanatta, D.T., Masteller, E.C., Dunn, H.L.,
Nichols, S.J., Marangelo, P.J., and Schloesser, D.W 2002. Some
                                                         89

-------

                                                                                           2007
near shore areas in Lake Erie and Lake St. Clair provide refuge
for native freshwater mussels (Unionidae) from the impacts of
invading zebra and quagga mussels (Dreissena). Presented at the
45th Conference on Great Lakes Research, June 2-6, 2002,
Winnipeg, MB, abstract on p. 87 of program.

Nichols, S.J., andAmberg, J. 1999. Co-existence of zebra mus-
sels and freshwater unionids; population dynamics ofLeptodea
fragilis in a coastal wetland infested with zebra mussels. Can. J.
Zoo/. 77(3):423-432.

Nichols, S.J., and Wilcox, D.A.  1997. Burrowing saves Lake
Erie clams. Nature 389:921.

Schloesser, D.W., Smithee, R.D., Longdon, G.D., and Kovalak,
W.P. 1997. Zebra mussel induced mortality of unionids in firm
substrata of western Lake Erie and a habitat for survival.
American Malacological Bulletin 14:67-74.

Strayer, D.L., and Smith, D.R. 2003. A guide to sampling fresh-
water mussel populations. American Fisheries Society,
Monograph 8, Bethesda, MD. 103 pp.

The National Native Mussel Conservation Committee. 1998.
National strategy for the conservation of native freshwater mus-
sels. J. Shellfish Res. 17(5):1419-1428.

Zanatta, D.T., Mackie, G.L., Metcalfe-Smith, J.L., and
Woolnough, D.A. 2002. A refuge for native freshwater mussels
(Bivalvia: Unionidae) from impacts of the non-native zebra
mussel (Dreissena polymorpha) in Lake St. Clair. /. Great Lakes
Res. 28(3):479- 489.

Last Updated
State of the Great Lakes 2005
90

-------

Lake Trout
Indicator #93

Overall Assessment
           Status:  Mixed
           Trend:  Unchanging

Lake-by-Lake Assessment
Lake Superior
           Status:  Good
           Trend:  Improving

Lake Michigan
           Status:  Poor
           Trend:  Declining
Lake Huron
Lake Erie
           Status:  Mixed
           Trend:  Improving
           Status:  Mixed
           Trend:  Unchanging
Lake Ontario
           Status:  Mixed
           Trend:  Declining

Purpose
•To track the status and trends in lake trout populations; and
•To infer the basic structure of the cold water predator community and the general health of the
ecosystem.

Ecosystem Objective
Self-sustaining, naturally reproducing populations that support target yields to fisheries are the
goal of the lake trout restoration program. Target yields approximate historical levels of lake trout
harvest or levels adjusted to accommodate stocked non-native predators such as Pacific salmon.
These targets are 4 million pounds (1.8 million kg) from Lake Superior, 2.5 million pounds (1.1
million kg) from Lake Michigan, 2.0 million pounds (0.9 million kg) from Lake Huron and 0.1
million pounds (0.05 million kg) from Lake Erie. Lake Ontario has no specific yield objective but
has a population objective of 0.5-1.0 million adult fish that produce 100,000 yearling recruits
annually through natural reproduction.

State of the Ecosystem
Background
Lake trout were historically the  principal salmonine predator in the coldwater communities of the
Great Lakes. By the late 1950s,  lake trout were extirpated throughout most of the Great Lakes
                         Draft for Discussion at SOLEC 2006

-------

mostly from the combined effects of sea lamprey predation and over fishing. Restoration efforts
began in the early 1960s with chemical control of sea lamprey, controls on exploitation, and
stocking of hatchery-reared fish to rebuild populations. Full restoration will not be achieved until
natural reproduction is established and maintained to sustain lakewide populations. To date, only
Lake Superior has that distinction.

Status of Lake Trout
Trends in the relative or absolute annual abundance of lake trout in each of the Great Lakes are
displayed in Figure 1. Lake trout abundance dramatically increased in all the Great Lakes after
initiation of sea lamprey control, stocking, and harvest control. Natural reproduction, from large
parental stocks of wild fish is occurring throughout Lake Superior, supports both onshore and
offshore populations, and it may be approaching historical levels. Stocking there has been
discontinued. Sustained natural reproduction, albeit at low levels, has also been occurring in Lake
Ontario since the early 1990s, and in some areas of Lake Huron, but has been largely absent
elsewhere in the Great Lakes. In Lake Huron  substantial and widespread natural reproduction was
seen starting in 2004 following near collapse  of alewife populations.  Abundance of hatchery-
reared adults was relatively high in Lake Ontario from 1986 - 1998, but declined by more than
30% in 1999 due to reduced stocking and poor survival of stocked yearlings since the early
1990s. Adult abundance again declined by 54% in 2006 likely due to ongoing poor recruitment
and mortality from sea lamprey predation. Parental stock sizes of hatchery-reared fish were
relatively high in some areas of Lakes Huron  and Michigan, but sea lamprey predation, fishery
extractions, and low stocking densities have limited population expansion elsewhere.

Pressures
Sea lamprey continues to limit population recovery, particularly in Lakes Michigan and Superior,
and parasitic adults are increasing basin-wide. Fishing pressures also continue to limit recovery.
More  stringent controls on fisheries are required to increase survival  of stocked fish. In northern
Lake Michigan parental  stock sizes are low and young in age due to low stocking densities,
moderate fishing mortality, and substantial sea lamprey mortality; hence egg deposition is low in
most historically important spawning areas. Fishing mortality has been reduced in recent years
but replaced by sea lamprey mortality. High biomass of alewives  and predators on lake trout
spawning reefs are thought to inhibit restoration through egg and  fry predation, although the
magnitude of this pressure is unclear. Recent  trends in Lake Huron suggest that alewife may need
to reach very low abundances to allow substantial natural reproduction.  A diet dominated by
alewives may be limiting fry survival (early mortality syndrome)  through thiamine  deficiencies.
The loss ofDiporeia and dramatic reductions in the abundance of slimy sculpins is reducing prey
for young lake trout and may be affecting survival. Current strains of lake trout stocked may not
be appropriate for offshore habitats, therefore limiting colonization potential.

Management Implications
Continued and enhanced sea lamprey control  is required basin-wide to increase survival of lake
trout to adulthood. New  sea lamprey control options, which include pheromone systems that
increase trapping efficiency and disrupt reproduction, are being researched and hold promise for
improved control. Continued and enhanced control on exploitation is being improved through
population modeling in the upper Great Lakes but needs to be applied throughout the basin.
Stocking densities need to be increased in some areas, especially in Lake Michigan. The use of
                         Draft for Discussion at SOLEC 2006

-------
alternate strains of lake trout from Lake Superior could be candidates for deep, offshore areas not
colonized by traditional strains used for restoration. Introduction of such strains has been initiated
in Lake Erie and hold promise. Direct stocking of eggs, fry, and yearling on or near traditional
spawning sites should be used where possible to enhance colonization.

Comments from the author(s)
Reporting frequency should be every 5 years. Monitoring systems are in place, but in most lakes
measures do not directly relate to stated harvest objectives. Population objectives may need to be
redefined as endpoints in units measured by the monitoring activities.

Acknowledgments
Authors: Charles R. Bronte, U.S. Fish and Wildlife Service, New Franken, WI;
James Markham, New York Department of Environmental Conservation;
Brian Lantry, U.S. Geological Survey, Oswego, NY;
Aaron Woldt, U.S. Fish and Wildlife Service, Alpena, MI; and
James Bence, Michigan State University, East Lansing, MI.

Data Sources
Bence, J.R., and Ebener, M.P. (eds.). 2002. Summary status of lake trout and lake whitefish
populations in 1936 treaty-ceded waters of Lakes Superior, Huron and Michigan in 2000, with
recommended yield and effort levels for 2001. Technical Fisheries Committee, 1836 Treaty-
Ceded Waters of Lakes Superior, Huron and Michigan.

Bronte, C.R., Ebener, M.P., Schreiner, D.R., DeVault, D.S., Petzold, M.M., Jensen, D.A.,
Richards, C., and Lozano, S.J. 2003a. Fish community change in Lake Superior, 1970-2000.  Can.
J. Fish. Aquat. Sci. 60:1552-1574.

Bronte, C. R., M. E. Holey, C. P. Madenjian, J. L. Jonas, R. M. Claramunt,  P.
C. McKee, M. L. Toneys, M. P. Ebener, B. Breidert, G. W. Fleischer, R. Hess, A. W. Martell, Jr.,
and E. J. Olsen.2006. Relative abundance, site fidelity, and survival of adult lake trout in Lake
Michigan from 1999-2001: implications for future restoration strategies. N. Am. J. Fish. Manage.
(in press).

Bronte, C.R., Jonas, J., Holey, M.E., Eshenroder, R.L., Toneys, M.L., McKee, P., Breidert, B.,
Claramunt, R.M., Ebener, M.P., Krueger, C.C., Wright, G., and Hess, R. 2003c. Possible
impediments to lake trout restoration in Lake Michigan. Lake Trout Task Group report to the
Lake Michigan Committee, Great Lakes Fishery Commission.

Bronte, C.R., Schram, S.T., Selgeby, J.H., and Swanson, B.L. 2002. Reestablishing a spawning
population of lake trout in Lake Superior with fertilized eggs in artificial turf incubators. N. Am.
J. Fish. Manage. 22:796- 805.

Cornelius, F.C., Muth, K.M., and Kenyon, R. 1995. Lake trout rehabilitation in Lake Erie: a case
of history. J. Great Lakes Res. 21 (Suppl. l):65-82.

DesJardine, R.L., Gorenflo, T.K., Payne, R.N., and Schrouder, J.D. 1995. Fish community
objectives for Lake Michigan. Great Lakes Fish. Comm.  Spec. Publ. 95-1. 38pp.
                         Draft for Discussion at SOLEC 2006

-------

Elrod, J.H., O'Gorman, R., Schneider, C.P., Eckert, T.H., Schaner, T., Bowlby, J.N., and Schleen,
L.P. 1995. Lake trout rehabilitation in Lake Ontario. J. Great Lakes Res. 21 (Suppl. 1): 83-107.

Eshenroder, R.L., Holey, M.E., Gorenflo, T.K., and Clark, R.D., Jr. 1995a. Fish community
objectives for Lake Michigan.  Great Lakes Fish. Comm. Spec. Publ. 95-3. 56pp.

Eshenroder, R.L., Payne, N.R., Johnson, J.E., Bowen II, C.A., and Ebener, M.P. 1995b. Lake
trout rehabilitation in Lake Huron. J. Great Lakes Res. 21 (Suppl. 1):108-127.

Hansen, M.J.  1999. Lake trout in the Great Lakes: basinwide stock collapse and binational
restoration. In Great Lakes Fisheries Policy and Management. Edited by W.W. Taylor and C.P.
Ferreri, Michigan State University Press, East Lansing, MI. pp. 417-454.

Hansen, M.J.  (ed.).  1996. A lake trout restoration plan for Lake Superior. Great Lakes Fishery
Commission,  34pp.

Holey, M.E., Rybicki, R.R., Eck, G.W., Brown,, E.H., Jr., Marsden, J.E., Lavis, D.S., Toneys,
M.L., Trudeau, T.N., and Horrall, R.M. 1995. Progress toward lake trout restoration in Lake
Michigan. J. Great Lakes Res.  21 (Suppl. 1):128-151.

Horns, W.H.,  Bronte, C.R., Busiahn, T.R., Ebener, M.P., Eshenroder, R.L., Gorenflo, T.,
Kmiecik, N., Mattes, W., Peck, J.W., Petzold, M., Schreiner, D.R. 2003. Fish community
objectives for Lake Superior. Great Lakes Fish. Comm. Spec. Publ. 03-01. 78pp.

Johnson, J.E., He, J.X., Woldt, A.P., Ebener, M.P., and Mohr, L.C. 2004.  Lessons in
rehabilitation  stocking  and management of lake trout in Lake Huron. In Propagated fish in
resource management.  Editors, M.J. Nickum, P.M. Mazik, J.G. Nickum and D.D. MacKinlay,
pp. 157-171.  , Bethesda, Maryland, American Fisheries Society, Symposium 44.

Lake Superior Lake Trout Technical Committee (LSLTTC). 1986. A lake trout restoration plan
for Lake Superior. In Minutes of the Lake Superior Committee (1986 annual minutes), Ann
Arbour, MI, Great Lakes Fishery Commission, March 20, 1986.
Lake Trout Task Group. 1985. A Strategic Plan for the rehabilitation of lake trout in eastern Lake
Erie. Lake Erie Committee. Ann Arbor, MI.

Lantry, B.F., Eckert, T.H., O'Gorman, R., and Owens, R.W. 2003. Lake trout rehabilitation in
Lake Ontario, 2002. In: NYDEC Annual Report to the Great Lakes Fishery Commission's Lake
Ontario  Committee, March, 2003.

Ryan, P.A., Knight, R., MacGregor, R., Towns, G., Hoopes, R., and Culligan, W. 2003. Fish-
community goals and objectives for Lake Erie. Great Lakes Fish. Comm. Spec. Publ. 03-02.
56pp.
                         Draft for Discussion at SOLEC 2006

-------

                                                      ^ ._
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''
Schneider, C.P., Schaner, T., Orsatti, S., Lary, S., and Busch, D. 1997. A Management Strategy
for Lake Ontario Lake Trout. Report to the Lake Ontario Committee, Great Lakes Fishery
Commission.

Wilberg, M. J., Hansen, M.J., and Bronte, C.R. 2003. Historic and modern density of wild lean
lake trout in Michigan waters of Lake Superior: implications for restoration goals. N. Am. J. Fish.
Manage. 23:100-108.

List of Figures
Figure 1. Relative or absolute abundance of lake trout in the Great Lakes.  The measurement
reported varies from lake to lake, as shown on the vertical scale, and comparisons between lakes
may be misleading. Overall trends over time provide information on relative abundances.
Source: U.S. Fish and Wildlife Service

Last updated
SOLEC 2006
                         Draft for Discussion at SOLEC 2006

-------
                               State of the Great Lakes 2007 - Draft
    80
       Lake Superior - U.S.
    60-
  I 40-

  m
  il 20-
                              Wild
                              Hatchery
1970  1975  1980   1985  1990  1995 2000
                Year
    80
       Lake Superior - Canada
60-

40-

20-
1975  1980  1985  1990  1995 2000
           Year
     1970


       Lake Michigan
10

 8

 6-

 4-

 2-
     1965 1970 1975 1980 1985 1990  1995 2000
                     Year
                                           30

                                         1 25-

                                         1 20-

                                         | 15-

                                         «i
                                         il  5-
                                              Lake Huron
                                                1975  1980
                                                         1985   1990
                                                             Year
                                                                       1995   2000
                                           10

                                            8-

                                            6-

                                            4-

                                            2-
                                                  Lake Erie
                                                                           — All Fish
                                                                           — Age 5+
                                                                           — Ages 1-3
                                            1986
                                                      1990
                                                              1994
                                                            Year
   1998
                                               25
                                              Lake Ontario
                                         E 20-
                                         •B
                                         * 15-

                                         I 10-

                                         I  5-
                                                                           — Females
                                                                           — Males
                                                                              Immature
                                            0-
                                            1980
                                                     1985
                                                            1990
                                                             Year
1995
                                                                      2000
Figure 1. Relative or absolute abundance of lake trout in the Great Lakes. The measurement
reported varies from lake to lake, as shown on the vertical scale, and comparisons between lakes
may be misleading. Overall trends over time provide information on relative abundances.
Source: U.S. Fish and Wildlife Service
                        Draft for Discussion at SOLEC 2006

-------
Benthos Diversity and Abundance - Aquatic Oligochaete Communities
Indicator #104
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Unchanging/ deteriorating
Some lakes or parts of lakes are good and unchanging, while other
lakes or parts of lakes are fair to poor and are either unchanging or
may be deteriorating.
Lake by Lake Assessment
Lake Superior
           Status:  Good
           Trend:
   Primary Factors
      Determining
  Status and Trend
Unchanging
All sites had index values that ranged from 0 to <0.5, indicating
oligotrophic conditions
Lake Michigan
Status:
 Trend:
Primary Factors
Determining
Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Unchanging, Deteriorating
Most sites had index values that ranged from 0 to <0.5, indicating
oligotrophic conditions. The two most southeastern, nearshore sites changed
from oligotrophic status in 2000, mesotrophic status in 2001,
mesotrophic/eutrophic status in 2002-2004, and back to mesotrophic in
2005. The most east-central, nearshore site changed from oligotrophic
(2000-2004) to mesotrophic (2005).
Mixed
Unchanging
Saginaw Bay remained mesotrophic throughout the six years. All other
sites were oligotrophic.
Lake Erie
Status:
 Trend:
Primary Factors
Determining
Status and Trend
Lake Ontario
Status:
 Trend:
Primary Factors
Mixed
Unchanging, Deteriorating
Most sites were mesotrophic to eutrophic. Two western sites were
oligotrophic mesotrophic due to reduced numbers of oligochaetes.
Eutrophic sites in the eastern part of the lake exhibited increasing index
values.
Mixed
Unchanging
Most sites were oligotrophic.  The three most southern, nearshore sites
                        Draft for Discussion at SOLEC 2006

-------
Determining        varied from oligotrophic to eutrophic on a year to year basis.
Status and Trend

Purpose
    •   To assess species diversity and abundance of aquatic oligochaete communities in order to
        determine the trophic status and relative health of benthic communities in the Great
        Lakes.

Ecosystem Objective
Benthic communities throughout the Great Lakes should retain species abundance and diversity
typical for benthos in similar unimpaired waters and substrates. A measure of biological response
to organic enrichment of sediments is based on Milbrink's (1983) Modified Environmental Index
(MEI).  This index was modified from Howmiller and Scott's (1977) Environmental Index. This
measure will have wide application in nearshore, profundal, riverine, and bay habitats of the
Great Lakes.  This indicator supports Annex 2 of the Great Lakes Water Quality Agreement.

State of the Ecosystem
Shortly after intensive urbanization and industrialization during the first half of the 20th century,
pollution abatement programs were initiated in the Great Lakes.  Degraded waters and substrates,
especially in shallow areas, began to slowly improve in quality. By the early 1980's, abatement
programs and natural biological processes changed habitats to the point where aquatic species
that were tolerant of heavy pollution began to be replaced by species that were intolerant of heavy
pollution.

The use of Milbrink's index values to characterize aquatic oligochaete communities provided one
of the earliest measures of habitat quality improvements (e.g., western Lake Erie). This index has
been used to measure changing productivity in waters of North America and Europe and, in
general, appears to be a reasonable measure of productivity in waters of all the Great Lakes
(Figure 1). The index values from sites in the upper lakes continue to be very low (<0.6),
indicating an oligotrophic status for these areas.  Index values from sites such as the nearshore
areas of southeastern and east-central Lake Michigan and Saginaw Bay in Lake Huron, which are
known to have higher productivity, exhibited higher index values that indicate mesotrophic (0.6-
1.0) to eutrophic (>1.0) conditions. Nearshore sites in southern Lake Ontario continued to be
classified as mesotrophic to eutrophic, while offshore sites  were oligotrophic. Sites in Lake Erie
exhibited the highest index values; nearly all of them fell within the mesotrophic or eutrophic
category (one site in western Lake Erie had low values characterized by low numbers of
oligochaetes). Over the last six years, a trend of increasing  index values was observed for eastern
Lake Erie.

Pressures
Future pressures that may change suitability of habitat for aquatic oligochaete communities
remain unknown. Pollution abatement programs and natural processes will assuredly continue to
improve water and substrate quality.  However, measurement of improvements could be
overshadowed by pressures such as zebra and quagga mussels, which were an unknown impact
only 10 years ago. Other possible pressures include non-point source pollution, regional
temperature and water level changes, and discharges of contaminants such as pharmaceuticals, as
                         Draft for Discussion at SOLEC 2006

-------

well as from other unforeseen sources.

Management Implications
Continued pollution abatement programs aimed at point source pollution will continue to reduce
undesirable productivity and past residual pollutants. As a result, substrate quality will improve.
Whatever future ecosystem changes occur in the Great Lakes, it is likely aquatic oligochaete
communities will respond early to such changes.

Comments from the authors
Biological responses of aquatic oligochaete communities are excellent indicators of substrate
quality, and when combined with a temporal component, they allow for the determination of
subtle changes in environmental quality, possibly decades before single species indicators.
However, it is only in the past several years that Milbrink's MEI has been applied to the open
waters of all the Great Lakes. Therefore, it is critical that routine monitoring of oligochaete
communities in the Great Lakes continue. Additionally, oligochaete taxonomy can be a
specialized and time-consuming discipline, and the taxonomic classification of species and their
responses to organic pollution is continually being updated. As future work progresses, it is
anticipated that the ecological relevance of existing and new species comprising the index will
increase.  Modifications to this index must be incorporated in future work, which includes the
assignment of index values to several taxa that are currently not included in the index, and the re-
evaluation of index values for a few of the species that are included in the index. It should be
noted that even though the index only addresses responses to organic enrichment in sediments, it
may be used with other indicators to assess the effects of other sediment pollutants.

Acknowledgments
Authors/Contributors: Kurt L. Schmude, Lake Superior Research Institute, University of
Wisconsin-Superior, Superior, WI; Don W. Schloesser, U.S. Geological Survey, Ann Arbor, MI;
Richard P. Barbiero, Computer Sciences Corporation, Chicago, IL; Mary Beth Giancario,
USEPA,  Great Lakes National Program Office, Chicago, IL.

Data Sources
U.S. Environmental Protection Agency, Great Lakes National Program Office, Biological Open
Water Surveillance Program of the Laurentian Great Lakes (2000-2005), through cooperative
agreement GL-96513791 with the University of Wisconsin-Superior.

Howmiller, R.P.,and Scott, M.A. 1977. An environmental index based on relative abundance of
oligochaete species. J. Wat. Poll. Cont. Fed. 49: 809-815.

Milbrink, G.  1983.  An improved environmental index based on the relative abundance of
oligochaete species. Hydrobiologia 102: 89-97.

List of Figures
Figure 1. Scatter plots of index values for Milbrink's (1983) Modified Environmental Index,
applied to data from GLNPO's 2000-2005 summer surveys.  Values ranging from 0-0.6 indicate
oligotrophic conditions; values from 0.6-1.0 indicate mesotrophic conditions (shaded area);
values above 1.0 indicate eutrophic conditions.  Index values for the taxa were taken from the
literature (Milbrink 1983, Howmiller and Scott 1977); immature specimens were not included in
                         Draft for Discussion at SOLEC 2006

-------
any calculations. Data points represent average of triplicate samples taken at each sampling site.
Source: U.S. Environmental Protection Agency, 2000-2005.

Figure 2. Map of the Great Lakes showing trophic status based on Milbrink's (1983) Modified
Environmental Index using the oligochaete worm community. Data taken from 2005. Gray
circles = oligotrophic; yellow squares = mesotrophic; red triangles = eutrophic.
Last updated
SOLEC 2006
           2.5
        TJ
        c
                                    Lake Superior
            1999   2000    2001    2002   2003   2004   2005    2006
                                    Year


                                      Lake Huron
          2.5

            2


          1.5
        -E   1
          O.E
                    X     A

X

            0              _     _     _
            1999    2000   2001    2002    2003    2004    2005    2006
                                     Year
                         Draft for Discussion at SOLEC 2006

-------
 State of the Great Lakes 2007 - Draft
                           Lake Michigan
CD
"ra
x
CD

^
2c
7 -
1 5 -

n f> -
n -




•

: i n ^ ^ n
	 * 	 — 	 — 	 7\ 	 _ 	 •• 	
                                                         * Northern Michigan
                                                         • Central Michigan
                                                          Southern Michigan
   1999    2000    2001    2002    2003    2004    2005    2006
                           Year
                     Lake Ontario
9 R -
Oj T
1 R -
X 1b
o
•o
n £
0-



» • * *
* ^ *
«• . A
^ A V
* • 1 * *
A A • X * A

# Western Ontario
• Eastern Ontario

1999   2000   2001   2002   2003   2004   2005   2006
                     Year
                Draft for Discussion at SOLEC 2006

-------
                                  Lake Erie
        1999
2000
203
2DC3
2004
2005
2006
                                 Year
Figure 1.  Scatter plots of index values for Milbrink's (1983) Modified Environmental
Index, applied to data from GLNPO's 2000-2005 summer surveys. Values ranging from 0-
0.6 indicate oligotrophic conditions; values from 0.6-1.0 indicate mesotrophic conditions (shaded
area); values above 1.0 indicate eutrophic conditions.  Index values for the taxa were taken from
the literature (Milbrink 1983, Howmiller and Scott 1977); immature specimens were not included
in any calculations. Data points represent average of triplicate samples taken at each sampling
site. Source: U.S. Environmental Protection Agency, 2000-2005.
                        Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
                                                                     A
                              0   50  100  150  200
Figure 2. Map of the Great Lakes showing trophic status based on Milbrink's (1983)
Modified Environmental Index using the oligochaete worm community. Data taken from
2005. Gray circles = oligotrophic; yellow squares = mesotrophic; red triangles = eutrophic.
                      Draft for Discussion at SOLEC 2006

-------
                                             OF   THE   GREAT
                                             2007
Phytoplankton Populations
Indicator #109

Assessment: Mixed, Trend Not Assessed
This assessment is based on historical conditions and expert
opinion. Specific objectives or criteria have not been deter-
mined.

Purpose
  To directly assess phytoplankton species composition, bio-
mass, and primary productivity in the Great Lakes; and
  To indirectly assess the impact of nutrient and contaminant
enrichment and invasive non-native predators on the microbial
food-web of the Great Lakes.

Ecosystem Objective
Desired objectives are phytoplankton biomass size and structure
indicative of oligotrophic conditions (i.e. a state of low biologi-
cal productivity, as is generally found in the cold open waters of
large lakes) for Lakes Superior, Huron and Michigan; and of
                mesotrophic conditions for Lakes Erie and Ontario. In addition.
                algal biomass should be maintained below that of a nuisance
                condition in Lakes Erie and Ontario, and in bays and in other
                areas wherever they occur. There are currently no guidelines in
                place to define what criteria should be used to assess whether or
                not these desired states have been achieved.

                State of the Ecosystem
                This indicator assumes that phytoplankton populations  respond
                in quantifiable ways to anthropogenic inputs of both nutrients
                and contaminants, permitting inferences to be made about sys-
                tem perturbations through the assessment of phytoplankton com-
                munity size, structure and productivity.

                Records for Lake Erie indicate that substantial reductions in
                summer phytoplankton populations occurred in the early 1990s
                in the western basin (Figure 1). The timing of this decline  sug-
                gests the possible impact of zebra mussels. In Lake Michigan, a
                significant increase in the size of summer diatom populations
                occurred during the  1990s. This was most likely due to the
     0)
     E
    _3
     O
     §
                                                                   Superior
                                                         Michigan

                                                               li

                                              83  85  87  89  91  93   95  97  99  83  85  87  89  91  93  95   97  99
0
                                                                      Huron
                                                           Ontario
  83  85  87  89   91  93  95  97  99   83  85  87  89  91   93  95  97  99
                                                         Erie Central  Basin
           83  85  87  89  91  93  95   97  99   83  85   87  89  91   93   95  97  99   83  85  87  89  91  93  95  97  99

                                                          Year

                    Q  Other            Q  Dinoflagellates    |  Cyanophytes     Q Cryptophytes
                    |  Chrysophytes    Q  Chlorophytes     Q  Diatoms
 Figure 1. Trends in phytoplankton biovolume (g/m3) and community composition in the Great Lakes 1983-1999. Samples were
 collected from offshore, surface waters during August.
 Source: U.S. Environmental Protection Agency, Great Lakes National Program Office
96

-------
                STATE   or
effects of phosphoras reductions on the silica mass balance in
this lake, and it suggests that diatom populations might be a sen-
sitive indicator of oligotrophication in Lake Michigan. No trends
are apparent in summer phytoplankton from Lakes Huron or
Ontario, while only three years of data exist for Lake Superior.
Data on primary productivity are no longer being collected. No
assessment of "ecosystem health" is currently possible on the
basis of phytoplankton community data, since reference criteria
and endpoints have yet to be developed.

It should be noted that these findings are at variance with those
reported for SOLEC 2000. This is due to problems with histori-
cal data comparability that were unrecognized during the previ-
ous reporting period. These problems continue to be worked on,
and as such, conclusions reported here should be regarded as
somewhat provisional.

Pressures
The two most important potential future pressures on the phyto-
plankton community are changes in nutrient loadings and contin-
ued introductions and expansions of non-native species.
Increases in nutrients can be expected to result in increases in
primary productivity and possibly also in increases in phyto-
plankton biomass. In addition, increases in phosphorus concen-
trations might result in shifts in phytoplankton community com-
position away from diatoms and towards other taxa. As seen in
Lake Michigan, reductions in phosphorus loading might be
expected to have the opposite effect. Continued expansion of
zebra mussel populations might be expected to result in reduc-
tions in overall phytoplankton biomass, and perhaps also in a
shift in species composition, although these potential effects are
not clearly understood. It is unclear what effects, if any, might be
brought about by changes in the zooplankton  community.

Management Implications
The effects of increases in nutrient concentrations tend to
become apparent in nearshore areas before offshore areas. The
addition of nearshore monitoring to the existing offshore moni-
toring program might therefore be advisable.  Given the greater
heterogeneity of the nearshore environment, any such sampling
program would need to be carefully thought out, and an ade-
quate number of sampling stations included to enable trends to
be discerned.

Acknowledgments
Authors: Richard P. Barbiero, DynCorp, A CSC company,
Chicago, IL, rick.barbiero@dyncorp.com;  and
Marc L. Tuchman, U.S. Environmental Protection Agency, Great
Lakes National Program Office, Chicago, IL,
tuchman.marc@epa.gov.
Sources
U.S. Environmental Protection Agency, Great Lakes National
Program Office. Unpublished data. Chicago, IL.
Authors' Commentary
A highly detailed record of phytoplankton biomass and commu-
nity structure has accumulated, and continues to be generated,
through regular monitoring efforts. However, problems exist
with internal comparability of this database. Efforts are currently
underway to rectify this situation, and it is essential that the
database continue to be refined and improved.

In spite of the existence of this database, its interpretation
remains problematic. While the use of phytoplankton data to
assess "ecosystem health" is conceptually attractive, there is cur-
rently no objective, quantitative mechanism for doing so.
Reliance upon literature values for nutrient tolerances or indica-
tor status of individual species is not recommended, since the
unusual physical regime of the Great Lakes makes it likely that
responses of individual species to their chemical environment in
the Great Lakes will vary in fundamental ways from those in
other lakes. Therefore, there is an urgent need for  the develop-
ment of an objective, quantifiable index specific to the Great
Lakes to permit use of phytoplankton data in the assessment of
"ecosystem health".

Last Updated
State of the Great Lakes 2003
                                                                                                                      97

-------
Phosphorus Concentrations and Loadings
Indicator #111
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Open Lake: Mixed  Nearshore: Poor
Open Lake: Undetermined Nearshore: Undetermined
Strong efforts begun in the 1970s to reduce phosphorus loadings have
been successful in maintaining or reducing nutrient concentrations in
the Lakes, although high concentrations still occur locally in some
embayments, harbors and nearshore areas.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Open Lake: Good Nearshore: Undetermined
Open Lake: Undetermined Nearshore: Undetermined
Average concentrations in the open waters are at or below expected levels.
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Open Lake: Good, Nearshore: Poor
Open: Improving Nearshore: Undetermined
Average concentrations in the open waters are at or below expected levels.
Phosphorus concentrations may exceed guidelines in nearshore waters for at
least part of the growing season.
Open Lake: Good Nearshore: Poor
Open Lake: Undetermined Nearshore: Undetermined
Average concentrations in the open waters are at or below expected levels.
Most offshore waters meet the desired guideline but some nearshore areas
and embayments experience elevated levels which likely contribute to
nuisance algae growths such as the attached green algae, Cladophora and
toxic cyanophytes such as Microcystis.
Open Lake: Fair-Poor  Nearshore: Poor
Open Lake: Undetermined Nearshore: Undetermined
Concentrations in the three basins of Lake Erie fluctuate from year to year
and frequently exceed target concentrations.  Extensive lawns of
Cladophora are common place over the nearshore lakebed in parts of
Eastern Lake Erie and are suggestive of phosphorus levels supportive of
nuisance levels of algal growth (Higgins et al. 2006 and Wilson et al.
2005).  Phosphorus levels in the nearshore (Canadian shores) of eastern
Lake Erie are periodically elevated above basin guideline value of 10 (ig/L,
however, the highly dynamic nature of water quality in the nearshore has
made it difficult to achieve either integrated nearshore assessments of
                        Draft for Discussion at SOLEC 2006

-------

Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   phosphorus levels, or to relate phosphorus levels to growth of Cladophora.
Open Lake: Good  Nearshore: Poor
Open Lake: Improving Nearshore: Undetermined
Average concentrations in the open lake are at or below expected levels.
Most offshore waters meet the desired guideline but some nearshore areas
and embayments experience elevated levels which likely  contribute to
nuisance algae growths such as the attached green algae, Cladophora and
toxic cyanophytes such as Microcystis. For example, in the Bay of Quinte,
control strategies at municipal sewage plants have reduced loadings by two
orders of magnitude since the early 1970s. In spite of these controls, mean
concentrations measured between May and October in the productive upper
bay have remained at 30-35 (ig/L in recent years. This  level of total
phosphorus is indicative of a eutrophic environment. Extensive lawns of
Cladophora are common place over the nearshore lakebed in parts of Lake
Ontario and are suggestive of phosphorus levels supportive of nuisance
levels of algal growth (Higgins et al. 2006 and Wilson et  al. 2005).
Phosphorus levels in the nearshore (Canadian shores) are  periodically
elevated above basin guideline value of 10 (ig/L, however, the highly
dynamic nature of water quality in the nearshore has made it difficult to
achieve either integrated nearshore assessments of phosphorus levels, or to
relate phosphorus levels to growth of Cladophora.
Purpose
This indicator assesses total phosphorus levels in the Great Lakes, and is used to support the
evaluation of trophic status and food web dynamics in the Great Lakes. Phosphorus is an essential
element for all organisms and is often the limiting factor for aquatic plant growth in the Great
Lakes. Although phosphorus occurs naturally, the historical problems caused by elevated levels
have originated from anthropogenic sources. Detergents, sewage treatment plant effluent,
agricultural and industrial sources have historically introduced large amounts into the Lakes.

Ecosystem Objective
The goals of phosphorus control are to maintain an oligotrophic state in Lakes Superior, Huron
and Michigan; to maintain algal biomass below that of a nuisance condition in Lakes Erie and
Ontario; and to eliminate algal nuisance growth in bays and in other areas wherever they occur
(GLWQA Annex 3). Maximum annual phosphorus loadings to the Great Lakes  that would allow
achievement of these objectives are listed in the  GLWQA. The expected concentrations of total
phosphorus in the open waters of the Great Lakes, if the maximum annual loads are maintained,
are listed in the following table: (insert Table 1: Phosphorus guidelines for the Great Lakes)

State of the Ecosystem
Strong efforts begun in the  1970s to reduce phosphorus loadings have been successful in
maintaining or reducing nutrient concentrations in the Lakes, although high concentrations still
occur locally in some embayments, harbors  and nearshore areas. Phosphorus loads have
                         Draft for Discussion at SOLEC 2006

-------


decreased in part due to changes in agricultural practices (e.g., conservation tillage and integrated
crop management), promotion of phosphorus-free detergents, and improvements made to sewage
treatment plants and sewer systems.

Average concentrations in the open waters of Lakes Superior, Michigan, Huron, and Ontario are
at or below expected levels. Concentrations in the three basins of Lake Erie fluctuate from year to
year (Figure 1) and frequently exceed target concentrations. In Lakes Ontario and Huron, most
offshore waters meet the desired guideline but some nearshore areas and embayments  experience
elevated levels which likely contribute to nuisance algae growths such as the attached  green
algae, Cladophora and toxic cyanophytes such as Microcystis. For example, in the Bay of
Quinte, Lake Ontario, control strategies at municipal sewage plants have reduced loadings by two
orders of magnitude since the early 1970's. In spite of these controls, mean concentrations
measured between May and October in the productive upper bay have remained at 30-35 (ig/L in
recent years. This level of total phosphorus is indicative of a eutrophic environment.  Typical of
other zebra mussel-infested and phosphorus enriched bays in the Great Lakes, toxic cyanophytes
such as Microcystis have increased in abundance in recent years with blooms occurring in late
August and early September.

Similarly, phosphorus concentrations may exceed phosphorus guidelines in nearshore  waters for
at least part of the growing season. Lake Michigan's eastern shoreline, when sampled in June,
2004, had a median concentration of 9 (ig/L. Summer sampling at the same locations  yielded a
median concentration of 6 (ig/L, with a number of sampling locations at or above the 7 (ig/L
guideline. By comparison, open water concentrations during  spring 2004 was 3.7 (ig/L.
Cladophora growth is a problem on much of this shoreline. In parts of Eastern Lake Erie and
Lake Ontario extensive lawns of Cladophora are common place and are suggestive of  phosphorus
levels supportive of nuisance levels of algal growth (Higgins et  al. 2006 and Wilson et al. 2005).
Phosphorus levels in the nearshore (Canadian shores) of eastern Lake Erie and Lake Ontario and
are periodically elevated above basin guideline value of 10 (ig/L, however, the highly  dynamic
nature of water quality in the nearshore has made it difficult  to achieve either integrated
nearshore assessments of phosphorus levels, or to relate phosphorus levels to growth  of
Cladophora. Phosphorus concentration in  nearshore  areas tend to be highly variable over time
and from point to point, at times on the scale of meters, due to influences of tributary and other
shore-based discharges, weather, biological activity and lake circulation.

Pressures
Even if current phosphorus controls are maintained, additional loadings can be expected.
Increasing numbers of people living along the Lakes will exert increasing demands on existing
sewage treatment facilities.  Even if current phosphorus concentration discharge limits are
maintained, increased populations may result in increased loads. Phosphorus management plans
with target loads need to be established for major municipalities. Recent research indicates that
even weather and climate changes may be  influencing the phosphorus loads to the lakes through
changes in snowmelt and storm patterns.

Management Implications
Because of its key role as the limiting nutrient for productivity and  food web dynamics of the
Great Lakes, vigilance must be exercised by water management agencies with respect  to
phosphorus loads to prevent a return to conditions observed in the 1960s. Future activities that
                         Draft for Discussion at SOLEC 2006

-------

are likely to be needed include:  1) Assess the capacity and operation of existing sewage treatment
plants in the context of increasing human populations being served. Utilization of state of the art
technology to lower effluent concentrations below current targets should be considered for
retrofits and upgrades to sewage treatment plants; 2) Conduct studies of the urban and rural
nonpoint contributions of phosphorus to better our understanding of their current overall
importance, especially with regards to nearshore eutrophication and Cladophora abundance, and
3) Conduct sufficient tributary and point source monitoring to track Phosphorus loadings and to
better understand the relative importance of various sources.

The surveillance of phosphorus  concentrations in the Great Lakes is ongoing and the data are
considered to be reliable. Plans are being formulated for an interagency laboratory comparison of
total phosphorus analysis. Enhanced monitoring of nearshore and embayment sites as well as
tributary monitoring may be accomplished with better coordination with existing state and
provincial environmental programs. Especially if they are tied to a framework, such as a
Lakewide Management Plan (LaMP) that recognizes the unique phosphorus related sensitivities
of the nearshore and also provides the means to interrelate nearshore and offshore nutrient
conditions and concerns. The recent reappearance of Cladophora in some areas of the Great
Lakes strengthens the importance of nearshore measurements.

The data needed to support loadings calculations have not been collected since 1991 in all lakes
except Lake Erie, which has loadings information up to 2002, and Lake Michigan with
information for 1994 and 1995.  Efforts to do so should be reinstated for at least Lake Erie, and
work is underway to accomplish this. For the other lakes,  the loadings component of this SOLEC
indicator will remain unreported, and changes in the different sources of phosphorus to these
Lakes may go undetected.

Acknowledgments
Authors: Alice Dove, Environment Canada, Burlington, ON & Glenn Warren, US EPA Chicago,
111
Additional contributions from: Scott Millard, Environment Canada, Burlington, ON & Todd
Howell, Ontario Ministry of Environment, Toronto, ON

Data Sources
Great Lakes Water Quality Agreement (GLWQA). 1978.  Revised Great Lakes Water Quality
Agreement of 1978. As amended by Protocol November 18, 1987. International Joint
Commission,  Windsor, Ontario.

Higgins, Scott N., E. Todd Howell, Robert E. Hecky, Stephanie J. Guildford and Ralph E. Smith
J., 2005. The Wall  of Green: The Status of Cladophora glomerata on the Northern Shores of
Lake Erie's Eastern Basin, 1995-2002.  Great Lakes Res. 31(4):547-563.

Howell, E. Todd and Duncan Boyd, Environmental Monitoring and Reporting Branch, Ontario
Ministry of Environment

Richardson, V. Environmental Conservation Branch, Environment Canada.
                         Draft for Discussion at SOLEC 2006

-------

                                                     ^ ._
                                  ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''
Warren, G. Great Lakes National Program Office, U.S. Environmental Protection Agency

Wilson, Karen A. E. Todd Howell and Donald A. Jackson, 2006.  Replacement of Zebra Mussels
by Quagga Mussels in the Canadian Nearshore of Lake Ontario: the Importance of Substrate,
Round Goby Abundance, and Upwelling Frequency
 J. Great Lakes Res. 32(l):ll-28.

List of Tables
Table 1: Phosphorus guidelines for the Great Lakes (GLWQA 1978)

List of Figures
Figure 1. Total Phosphorus Trends in the Great Lakes 1970 to 2005. Blanks indicate no
sampling.  Horizontal line on each graph represents the phosphorus guideline as listed in the
Great Lakes Water Quality Agreement for each Lake. Environment Canada data (white bars -
average of spring, surface measurements at open lake sites) are used for Lakes Ontario, Huron
and Superior, and are supplemented by US data for years in which no monitoring was conducted
on that lake. U.S. Environmental Protection Agency data (black bars - average of spring
measurements, all depths at open lake sites) are used for the three basins of Lake Erie and for
Lake Michigan, and are supplemented by Canadian data for years in which no US monitoring was
conducted on that lake.
Source: Science and Technology Branch, Environment Canada and Great Lakes National
Program Office, U.S. Environmental Protection Agency

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
                                                                           	
Lake
Superior
Huron
Michigan
Erie-
Western
Basin
Erie -
Central
Basin
Erie-
Eastern
Basin
Ontario
Phosphorus
Guideline (ug/L)
5
5
7
15


10


10


10
Table 1. Phosphorus guidelines for the Great Lakes (GLWQA 1978)
                      Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
                                               Total Open Lake Phosphorus Trends
                                   Lake Huron         in the Great Lakes (ug/L)
                                                           1970 to 2005
Figure 1. Total Phosphorus Trends in the Great Lakes 1970 to 2005. Blanks indicate no
sampling. Horizontal line on each graph represents the phosphorus guideline as listed in the
Great Lakes Water Quality Agreement for each Lake. Environment Canada data (white bars -
average of spring, surface measurements at open lake sites) are used for Lakes Ontario, Huron
and Superior, and are supplemented by US data for years in which no monitoring was conducted
on that lake. U.S. Environmental Protection Agency data (black bars - average of spring
measurements, all depths at open lake sites) are used for the three basins of Lake Erie and for
Lake Michigan, and are supplemented by Canadian data for years in which no US monitoring was
conducted on that lake.
Source: Science and Technology Branch, Environment Canada and Great Lakes National
Program Office, U.S. Environmental Protection Agency
                        Draft for Discussion at SOLEC 2006

-------


Contaminants in Young-of-the-Year Spottail Shiners
Indicator #114
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving
Although levels of polychlorinated biphenyls (PCBs) in forage fish
have decreased below the guideline at many sites around the Great
Lakes, PCB levels remain elevated. As well, dichloro-diphenyl-
trichloroethane (DDT) levels in forage fish have declined but remain
above the guideline at most Great Lakes' locations.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving
PCB concentrations in Lake Superior forage fish have declined over the
period of record and are currently below the guideline at all sample sites.
DDT has declined to levels near the guideline, except for Nipigon Bay,
where the most current levels (1990) are elevated.
N/A
N/A
N/A
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving
PCB levels in Lake Huron forage fish have remained static or declined over
the period of record and are currently at or below the guideline.  DDT
levels, however, remain elevated at Collingwood Harbour.
Mixed
Improving
PCB levels in Lake Erie forage fish have declined to levels at or below the
guideline.  DDT has also declined over the period of record but remains
above the guideline.
                        Draft for Discussion at SOLEC 2006

-------
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving
PCB levels in Lake Ontario forage fish have declined significantly over the
period of record and the most recent levels are at or below the guideline. At
some sites, DDT in forage fish has declined considerably, however, levels
remain at or above the guideline at all sites.  Mirex has also declined and
has remained below the detection limit in recent years.
Purpose
•To assess the levels of persistent bioaccumulative toxic (PBT) chemicals in young-of-the-year
spottail shiners;
•To infer local areas of elevated contaminant levels and potential harm to fish-eating wildlife; and
•To monitor contaminant trends over time for the nearshore waters of the Great Lakes.

Ecosystem Objective
Concentrations of toxic contaminants in juvenile forage fish should not pose a risk to fish-eating
wildlife. The Aquatic Life Guidelines in Annex 1 of the Great Lakes Water Quality Agreement
(United States and Canada, 1987), the New York State Department of Environmental
Conservation (NYSDEC) Fish Flesh Criteria (Newell et al,  1987) for the protection of
piscivorous wildlife, and the Canadian Environmental Quality Guidelines (CCME, 2001) are used
as acceptable guidelines for this indicator. Canadian Council of Ministers of the Environment
Contaminants monitored in forage fish and their respective guidelines are listed in Table 1.

State of the Ecosystem
Contaminant levels in fish are important indicators of contaminant levels in an aquatic ecosystem
due to the bioaccumulation of organochlorine chemicals in fish tissue. Contaminants that are
often undetectable in water may be detected in juvenile fish. Juvenile spottail shiner (Notropis
hudsonius} was originally selected by Suns  and Rees (1978) as the principal biomonitor for
assessing trends in contaminant levels in local or nearshore areas. It was chosen as the preferred
species because of its limited range in the first year of life; undifferentiated feeding habits in early
stages; importance as a forage fish; and its presence throughout the Great Lakes. The position it
holds in the food chain also creates an important link for contaminant transfer to higher trophic
levels. However, at some sites  along the Great Lakes spottail shiners are not as abundant as they
once were, and therefore can be difficult to  collect.  In this updated indicator report, bluntnose
minnow (Pimephales notatus) have been included in the Lake Huron/Georgian Bay dataset.

With the incorporation of the CCME guidelines, the total DDT tissue residue criterion is
exceeded at most locations. After total DDT, PCB is  the contaminant most frequently exceeding
the guideline. Mirex was historically detected and exceeded the guideline at Lake Ontario
locations.  However, mirex concentrations over the  past 10 years have been below detection.
Other contaminants listed in Table 1  are often not detected, or are present at levels well below the
guidelines.
                         Draft for Discussion at SOLEC 2006

-------

Lake Erie
Trends of contaminants in spottail shiners were examined for four locations in Lake Erie: Big
Creek, Thunder Bay Beach, Grand River and Leamington (Figure 1). Overall, the trends show
higher concentrations of PCBs in the early years (1970s) with a steady decline over time. At Big
Creek, PCB concentrations were elevated (>300 ng/g) until 1986. Since 1986, concentrations
have remained near the guideline (100 ng/g). At the Grand River and Thunder Bay beach
locations, PCB concentrations exceeded the guideline in the late 1970s, but have declined in
recent years and are currently below the IJC guideline (100 ng/g). At Leamington, PCB
concentrations were considerably higher than at the other Lake Erie sites. Although they declined
from 888 ng/g in 1975 to 204 ng/g in 2001, the concentrations exceeded the guideline in all years
except for a period in the early to mid-1990s. In the most recent collection (2004), levels have
declined to 136 ng/g, which only marginally exceeds the IJC guideline.

Total DDT concentrations at Lake Erie sites have also been declining. Concentrations of total
DDT  at Big Creek, Grand River and Thunder Bay Beach have declined considerably to levels
close to the guideline (14 ng/g). Maximum concentrations at these sites were found in the 1970s
and ranged from 38 ng/g  at Thunder Bay Beach to 75 ng/g at Big Creek. At Leamington,
however, total DDT levels peaked at 183 ng/g in 1986. Since then, levels have declined, but they
remain above the guideline.

Lake Huron
Trend data are available for two Lake Huron sites: Collingwood Harbour and Nottawasaga River
(Figure 2). At Collingwood Harbour the highest PCB concentrations were found when
sampling began in 1987 (206 ng/g). Since then, PCB  concentrations have remained near or just
below the guideline. At the Nottawasaga River the highest concentration of PCBs was observed
in 1977 (90 ng/g). Concentrations declined to less than the detection limit by 1987 and in 2002
were detected at very low levels.

Total DDT concentrations at Collingwood Harbour have remained near 40 ng/g since 1987. The
guideline of 14 ng/g was  exceeded in all years. At the Nottawasaga River site, there has been a
steady decline in total DDT since 1977 when concentrations peaked at 106 ng/g. In 2002, levels
were below the guideline.

Lake Superior
Trend data were examined for four locations in Lake  Superior: Mission River, Nipigon Bay,
Jackfish Bay and Kam River (Figure 3). Recent data are not available for the first three locations.

Generally PCB concentrations were low in all years and at all locations. The highest PCB
concentrations in Lake Superior were found at the Mission River in 1983 (139 ng/g). All other
analytical results were below the guideline (100 ng/g). The highest concentrations of PCBs at the
other three Lake Superior sites also occurred in 1983  and ranged from 51 ng/g at Nipigon Bay to
89 ng/g at Jackfish Bay.

At Mission River and Nipigon Bay, total DDT levels were high in the late 1970s but decreased
below the guideline (14 ng/g) by the mid-1980s. In 1990, the DDT level at Nipigon Bay was 66
ng/g, which is the highest concentration observed in juvenile fish from any Lake Superior site to
                         Draft for Discussion at SOLEC 2006

-------
date. At Jackfish Bay and the Kam River, total DDT levels were below the guideline each year,
except for the Kam River in 1991 when levels rose to 37 ng/g.

Lake Ontario
Contaminant concentrations from five sites were examined for trends: Twelve Mile Creek,
Burlington Beach, Bronte Creek, Credit River and the Humber River (Figure 4). PCBs, total DDT
and mirex were generally higher at these (and other Lake Ontario) locations than elsewhere in the
Great Lakes. Overall, PCBs at all locations tended to be higher in the early years, ranging from 3
to 30 times the guideline. The highest concentrations of PCBs were found at the Humber River in
1978 (2938 ng/g). In recent years PCBs at the five sites generally have ranged from 100 ng/g to
200 ng/g.

Total DDT concentrations at all five locations have declined considerably since the late 1970s
and early 1980s. However, at all of these locations, levels in juvenile fish still exceed the
guideline (14 ng/g). The maximum reported concentration was at the Humber River in 1978 (443
ng/g). Currently, the typical concentration of total DDT  at all five locations is approximately 50
ng/g. Mirex  has been detected intermittently at all five locations. The maximum concentration
was 37 ng/g at the Credit River in 1987. Since 1993, mirex has been below the detection limit at
all of these locations.

Lake Michigan
No spottail shiners were sampled from Lake Michigan.

Pressures
New and emerging contaminants, such as polybrominated diphenyl ethers, may apply new
pressures on Great Lakes' water quality. Analytical methods need to be developed and tissue
residue guidelines need to be established for these contaminants. Monitoring programs should
also be initiated.

Management Implications
For those contaminants that exceed the wildlife protection guidelines, additional remediation
efforts may be required. Continued monitoring is essential to determine the status  of
contaminants in forage fish from the Great Lakes.

Comments from the author(s)
Organochlorine contaminants have declined in juvenile fish throughout the Great Lakes.
However, regular monitoring should continue for all of these  areas to determine if levels are
below wildlife protection guidelines. Analytical methods should be improved to accommodate
revised guidelines and to include additional contaminants such as dioxins and furans, dioxin-like
PCBs and PBDEs. For Lake Superior, the historical data do not include toxaphene  concentrations.
Since this contaminant is responsible for some consumption restrictions on sport fish from this
lake (MOE,  2005), it is recommended that analysis of this contaminant be included in any future
biomonitoring studies in Lake Superior.

Spottail shiners have been a useful indicator of contaminant levels in the past. However, this
species is less abundant than it has been. Due to the difficulties in collecting this species in all
                         Draft for Discussion at SOLEC 2006

-------
areas of the Great Lakes, consideration should be given to adopting other forage fish species as
indicators when spottail shiners are not available. This year, bluntnose minnows were used for
one site in Georgian Bay. This will improve temporal and spatial trend data and result in a more
complete dataset for the Great Lakes.

Acknowledgments
Authors: Emily Awad, Sport Fish Contaminant Monitoring Program, Ontario Ministry of
Environment, Etobicoke, ON; and
Alan Hayton, Sport Fish Contaminant Monitoring Program, Ontario Ministry of Environment,
Etobicoke, ON.
Data: Sport Fish Contaminant Monitoring Program, Ontario Ministry of Environment.

Data Sources
CCME, 2001. Canadian Environmental Quality Guidelines. Canadian Council of Ministers of the
Environment.

MOE, 2005. Guide to eating Ontario sport fish, 2005-06 edition. PIBs 590B12.  Ontario
Ministry of the Environment, Etobicoke, ON.

Newell, A.J., Johnson, D.W., and Allen, L.K. 1987. Niagara River biota contamination project:
fish flesh criteria for piscivorous wildlife. Technical Report 87-3. New York State Department of
Environmental Conservation, Albany, NY.

Suns, K., and Rees, G. 1978. Organochlorine contaminant residues in young-of-the-year spottail
shiners from Lakes Ontario, Erie, and St. Clair. J. Great Lakes Res. 4:230-233.

United States and Canada. 1987. Great Lakes Water Quality Agreement of 1978, as amended by
Protocol signed November 18, 1987.  Ottawa and Washington.

List of Tables
Table 1. Tissue Residue Criteria for various organochlorine chemicals or chemical groups for the
protection of wildlife consumers of aquatic biota.

List of Figures
Figure 1. PCB and total DDT levels in juvenile spottail shiners from four locations in Lake Erie.
The figures show mean concentration plus standard deviation. The red line indicates the wildlife
protection guideline. When not detected, one half of the detection limit was used to calculate the
mean concentration.
Source: Ontario Ministry of the Environment

Figure 2. PCB and total DDT levels in juvenile spottail shiners from two locations in Lake Huron.
The figures show mean concentration plus standard deviation. The red line indicates the wildlife
protection guideline. When not detected, one half of the detection limit was used to calculate the
mean concentration.
Source: Ontario Ministry of the Environment
                         Draft for Discussion at SOLEC 2006

-------

Figure 3. PCB and total DDT levels in juvenile spottail shiners from four locations in Lake
Superior. The figures show mean concentration plus standard deviation. The red line indicates the
wildlife protection guideline. When not detected, one half of the detection limit was used to
calculate the mean concentration.
Source: Ontario Ministry of the Environment

Figure 4. PCB, mirex and total DDT levels in juvenile  spottail shiners from five locations in Lake
Ontario. The  figures show mean concentration plus standard deviation. The red line indicates the
wildlife protection guideline for PCBs and total DDT. For mirex, the red line indicates the
detection limit (5ng/g). When not detected, one half of the detection limit was used to calculate
the mean concentration.
Source: Ontario Ministry of the Environment

Last updated
SOLEC 2006
Contaminant
PCBs
DDT, ODD, DDE
Chlordane
Dioxin/Furans
Hexachlorobenzene
Hexachlorocyclohexane (BHC)
Mirex
Octachlorostyrene
Tissue Residue Criteria
(ng/g)
100*
14f (formerly 200)
500
0.000713 (formerly 0.003)
330
100
below detection*
20
               *IJC Aquatic Life Guideline for PCBs (IJC 1988); a Environment Canada, 2000 (CCME 2001);
               f Environment Canada, 1997  (CCME 2001). All othervalues from NYSDEC Fish Flesh
               Criteria (Newell et al. 1987).  Guidelines based on mammals and birds.

Table 1. Tissue Residue Criteria for various organochlorine chemicals or chemical groups for the
protection of wildlife consumers of aquatic biota.
                          Draft for Discussion at SOLEC 2006

-------
Mean PCB Levels in Juvenile Spottail
Shiners from Lake Erie at Big Creek
•3 800


o 4UU n
Q_ j-»*irt 1 1
A^ (


T IfT
ifx HT I
-•nfl— ntnrn — v-a TTTT 	 mr
Year
Mean PCB Levels in Juvenile Spottail
Shiners from Lake Erie at the Grand
River

O)
a? 400
o
o- 200 -L






-T — _ 	 — 	 — — -
Year
Mean PCB Levels in Juvenile Spottail
Shiners from Lake Erie at Thunder Bay
Beach

_. 800
» 400 1

°" 20° a -r
A* Q>
N^ N*




f-m-w •*•
Year
Mean PCB Levels in Juvenile Spottail
Shiners from Lake Erie at Leamington

1
=. eon —
£
o 400
200 1



r T

|][|[||||| Illrir TfirllTlT-n- 	 n — ff~
1975197819811984 198719901993199619992002
Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Erie at Big
Creek

51 150 	
O)
c
"" 100


Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Erie at the
Grand River
S1
rn -tern
£ 1
0
" 0 IT
i- Q ii nil n TT irnn Trnii
Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Erie at
Thunder Bay Beach

oj 150
o
5 50 - i 	
IT! I T T T T T
\J \J \J \J \J \J \J \J rpJ
Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Erie at
Leamington
^150 I

Q IUU
° 5° llwllirirl Tw irirlll IT J
'$'$><£'&<$•<$><$<$><$>•$'
Year
Figure 1. PCB and total DDT levels in juvenille spottail shiners from four locations in Lake Erie.
The figures show mean concentration plus standard deviation. The red line indicates the wildlife
protection guideline. When not detected, one half of the detection limit was used to calculate the
mean concentration.
Source: Ontario Ministry of the Environment

-------
Mean PCB Levels in Juvenile Spottail
Shiners and Bluntnose Minnows* from
Lake Huron at Collingwood Harbour

j"! 100 N T T Q D . w f n "
* :1 m n if
Year
Mean PCB Levels in Juvenile Spottail
Shiners from Lake Huron at the
Nottawasaga River
5
50 -mo T-
°- ill IT
n Is x
Year
Mean Total DDT Levels in Juvenile Spottail
Shiners and Bluntnose Minnows* from Lake
Huron at Collingwood Harbour
§ "3 80 -
— 0)
™ — 40
Q _


I 7TX T
p rat n f

Year

Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Huron at the
Nottawasaga River
— 1J3 Ff —
O) JL
o) n
& 8o J I
Q
° 40 U
s° F
"~ 0 P

I I T
h Hrri



u uu uu d
Year
Figure 2. PCB and DDT levels in juvenille spottail shiners from two locations in Lake Huron. The
figures show mean concentration plus standard deviation. The red line indicates the wildlife
protection guideline. When not detected, one half of the detection limit was used to calculate the
mean concentration.
Source: Ontario Ministry of the Environment

-------
IVI
SI
200
0) 150
| 100
CD
0 50
0
ean PCB Levels in Juvenile Spottail
liners from Lake Superior at Mission
River

m | | ,
Year
Mean PCB Levels in Juvenile Spottail
Shiners from Lake Superior at Nipigon
Bay
B 150 -
1"
•S. 100 -
(0
0 50
Q.



„ i
Year
Mean PCB Levels in Juvenile Spottail
Shiners from Lake Superior at Jackfish
Bay
O)
•s, 150 -
~ 100 -
o 50 -
0- n



i— i fi „ „
Aq><£)$>v
Year
Mean PCB Levels in Juvenile Spottail
Shiners from Lake Superior at Kam
River

i1
— 100
(0
8 50


fiff I
i \\\ _ n _ n
Q
Year
s
~ 80-
-?
Q 40-
Q
"o
H 0 -
N
Mean Total DDT Levels in Juvenile
pottail Shiners from Lake Superior at
Mission River


1 	
Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Superior at
Nipigon Bay
U>
Q 40 -
Q




A^ N 'b ^5 A Q)
Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Superior at
Jackfish Bay
? 60
Q 40
Q
3 20
"o
>- o




i i n n n
Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Superior at
Kam River
"6 60
^ 40
Q
Q 20 -
o 0 -

















\-
Year
Figure 3. PCB and total DDT levels in juvenille spottail shiners from four locations in Lake
Superior. The figures show mean concentration plus standard deviation. The red line indicates the
wildlife protection guideline. When not detected, one half of the detection limit was used to calculate
the mean concentration.
Source: Ontario Ministry of the Environment

-------
anile Spottail
rio at Twelve





nile Spottail
itario at
ch



rr n j

ile Spottail
> at Bronte



W w-, 	 »
? / /
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Ontario at
Twelve Mile Creek
O)
-g, 300 -
t- 200 -
o L
Q 100-0
3 o]


-r T I ¥ 1
IfrllT™ M MT T T T
/> ^ N#N ^ /" ^ ^ ^ ^ ^
Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Ontario at
Burlington Beach
•§> QOO
K 200 -
Q
Q loo-
's
£ n

c


1977 1981 1985 1989 1993 1997 2001
Year
Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Ontario at
Bronte Creek
-?
o) 3QQ - -
1— ooo - -
Q ^uu
Q 100 -
1 o-i




NA#\^N#\^\^N#V
Year
Mean Mirex Levels in Juvenile Spottail
Shiners from Lake Ontario at Twelve
Mile Creek
40
§30 	 T 	
f20 n n
I10 — II — Irr 1 	

1975 1979 1983 1987 1991 1995 1999 2003
Year
Mean Mirex Levels in Juvenile Spottail
Shiners from Lake Ontario at
Burlington Beach
50 -T

O)
c 30
(U on
'^ 10 -B 	 ^ _ir 	
xO'9?>9r>9?>9v)C^'6v:>d?>0N
& $> $> $> $> $> $> $> c^
Year
Mean Mirex Levels in Juvenile Spottail
Shiners from Lake Ontario at Bronte
Creek

O)
C QA 1
V
a °n 1
z. ^u
'£ 10 	 ?j 	
• •••-—•
1979 1982 1985 1988 1991 1994 1997 2000
Year

-------
inile Spottail
• at the Credit
 Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Ontario at
         the Credit River
                   — 400
                   -1
                      300

                      200

                      100

                        0
                         1976 1979 1982 1985 1988 1991 1994 1997

                                        Year
                                                  Mean Mirex Levels in Juvenile Spottail
                                                  Shiners from Lake Ontario at the Credit
                                                                 River
                                             1976 1979  1982 1985  1988 1991 1994 1997
                                                             Year
anile Spottail
tario at the
 Mean Total DDT Levels in Juvenile
Spottail Shiners from Lake Ontario at
        the Number River
                                                  Mean Mirex Levels in Juvenile Spottail
                                                     Shiners from Lake Ontario at the
                                                             Number River
_  400 -,

|>  300 -

t  200 -
                             .4—1443 ng/g |
                                           50

                                           40
                                                                 £.  30 -
                                                                 x
                                                                 £  20 -
                                                                 i  10-
                                                                     0
                                        Year
                                                                      1977  1980  1983  1986  1989  1992  1995
                                                                                      Year
3tal DDT levels in juvenille spottail shiners from five locations in Lake Ontario. The figures show mean concentration
: red line indicates the wildlife protection guideline for PCBs and total DDT. For mirex, the red line indicates the
hen not detected,  one half of the detection limit was used to calculate the mean concentration.
the Environment

-------
Contaminants in Colonial Nesting Waterbirds
Indicator #115
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving
The primary factors being used are: 1. the change in contaminant
concentrations in Herring Gull eggs between when they were first measured
(usually 1974) and currently, in 2005 (Jermyn-Gee et al. 2005; CWS,
unpubl.), 2. the overall ranking of contaminant concentrations at the  15
Great Lakes Herring Gull Egg Monitoring Sites (Weseloh et al. 2006) and
3. the direction and relative slope of the change-point regression line
calculated for each compound at each site. (Pekarik and Weseloh 1996;
Weseloh et al. 2003, 2005; CWS, unpubl.) Overall,  most contaminants have
declined substantially (>90%) since first measured.  Spatially, some sites in
2-3 of the lakes were much more contaminated than others. Temporally,
more than 70% of all contaminant concentrations at all colonies (N=105)
were currently declining as fast or faster than they did in the past.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
Good
Improving
For 6 contaminants that have been measured since the program started in
1974 (PCBs, DDE, HCB, HE, mirex and dieldrin), the two Herring Gull egg
monitoring sites in Lake Superior showed declines of 93.9 - 99.8% between
then and 2005. Both sites ranked among the lowest for concentrations of 7
major compounds (the above 6 + TCDD) among the 15 monitor sites. The
temporal pattern at the two sites showed 71% of colony-contaminant
comparisons declining as fast or faster than previously.
Mixed
Improving
For 6 contaminants that have been measured since the program began, the
two Herring Gull egg monitoring sites showed declines of 91.8 - 99.1%
between then and 2005. Eggs from one of the Lake Michigan sites ranked
as the 3rd most contaminated among the 15 monitor sites; eggs from the
other site ranked much lower (9th). The temporal pattern for the two sites
showed 86% of the colony-contaminant comparisons declining as fast or
faster than previously.
 Mixed
 Improving
Herring Gull eggs from two of three monitoring sites in Lake Huron were
relatively clean. The third site, in Saginaw Bay, had the most contaminated
                        Draft for Discussion at SOLEC 2006

-------

  Status and Trend
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
gull eggs among all sites tested and reduced the overall status of this
indicator in Lake Huron. The three sites showed contaminant declines of
68.9 - 99.7% in gull eggs in 2005. Two of three sites ranked among the
lowest for concentrations for 7 major compounds among 15 sites.
The temporal pattern at the three sites showed 86% of colony-contaminant
comparisons declining as fast or faster than previously.
Mixed
Improving
Of the two monitor sites in Lake Erie, the most easterly, at Port Colborne,
had the cleanest gull eggs of all 15 sites tested. Eggs from Middle Island, in
the Western Basin, were considerably more contaminated. The two sites
showed contaminant declines of 80.2 - 99.3% in gull eggs in 2005. Eggs
from Middle Island were in the mid-range and those from Port Colborne
were the lowest for contaminants. The temporal pattern at the two sites
showed 93% of colony-contaminant comparisons declining as fast or faster
than previously.
Poor
Improving
Eggs from the three Lake Ontario Herring Gull Monitoring Sites showed
declines of 88.6 - 99.0% in 2005. The three sites ranked among the top 8
for concentrations of contaminants in gull eggs. Temporally, 76% of
colony-contaminant comparisons were declining as fast or faster than
previously.
Purpose
•To assess current chemical concentrations and trends in representative colonial waterbirds (gulls,
terns, cormorants and/or herons) on the Great Lakes;
•To assess ecological and physiological endpoints in representative colonial waterbirds (gulls,
terns, cormorants and/or herons) on the Great Lakes; and
•To infer and measure the impact of contaminants on the health, i.e. the physiology and breeding
characteristics, of the waterbird populations.

Ecosystem Objective
One of the objectives of monitoring colonial waterbirds on the Great Lakes is to track progress
toward an environmental condition in which there is no difference in contaminant levels and
related biological endpoints between birds on and off the Great Lakes. Other objectives include
determining temporal and spatial trends in contaminant levels in colonial waterbirds and detecting
changes in their population levels on the Great Lakes. This includes monitoring contaminant
levels in Herring Gull eggs to ensure that the levels continue to decline and utilizing these data to
promote continued reductions of contaminants in the Great Lakes basin.
                         Draft for Discussion at SOLEC 2006

-------
State of the Ecosystem
Background
This indicator is important because colonial waterbirds are one of the top aquatic food web
predators in the Great Lakes ecosystem and they are very visible and well-known to the public.
They bioaccumulate contaminants to the greatest concentration of any trophic level organism and
they breed on all the Great Lakes. Thus, they are a very cost efficient monitoring system and
allow easy inter-lake comparisons. The current Herring Gull Egg Monitoring Program is the
longest continuously running annual wildlife contaminants monitoring program in the world
(1974-present). It determines concentrations of up to 20 organochlorines, 65 polychlorinated
biphenyls (PCB) congeners and 53 polychlorinated dibenzo-p-dioxin (PCDD) and
polychlorinated dibenzo furan (PCDF) congeners, as well as 16 brominated diphenyl ethers
BDEs) congeners (Braune et al. 2003).

Status of Contaminants in Colonial Waterbirds
The Herring Gull Egg Monitoring Program has provided researchers and managers with a
powerful tool (a 30-year database) to evaluate changes in contaminant concentrations in Great
Lakes wildlife (e.g., see Figure 1). The extreme longevity of the egg database makes it possible to
calculate temporal trends in contaminant concentrations in wildlife and to look for significant
changes within those trends. The database shows that most contaminants in gull eggs have
declined 90% or more since the program began in 1974 (Figure 2). In 2005, PCBs,
hexachlorobenzene (HCB), dichlorodiphenyl-dichloroethene (DDE), heptachlor epoxide (HE),
dieldrin, mirex and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) levels measured in eggs from the
15 Annual Monitor Colonies (Figure 3) were analysed for temporal trends (N=105 comparisons).
Analysis showed that in 83.8% of cases (88/105), the contaminants were decreasing as fast as or
faster in recent years than they had in the past. We interpreted that as a positive sign. In 9.5% of
cases (10/105), contaminants were decreasing more slowly than they had in the past (calculated
from Bishop et all992, Pettit et all994, Pekarik et all998 and Jermyn-Gee et al. 2005, as per
Pekarik and Weseloh 1998). This is  viewed as a negative sign. PCBs showed the most frequent
reduction in their rates of decline. The decline in contaminant concentrations in gull eggs,
however, may not be due wholly to a decrease in contaminants in the environment. Changes in
food web dynamics may be playing  a role in some of these declines, that is, contaminant exposure
at some colonies may have lessened because the birds  are now feeding on lower trophic level
prey.

The sole exception to these declining herring gull egg  contaminant concentrations appears to be
brominated diphenyl ethers (BDEs). These compounds, which are used as fire retardants in
plastics, furniture cushions, etc., increased dramatically in gull  eggs during 1981-2000 (Norstrom
et al. 2002). Recent data showed a combined 3.9% decline for the 15 monitor sites from 2000 to
2003 but a 25.3% increase from 2000 to 2005 (CWS, unpubl. data).

A comparison of concentrations of six contaminants (PCBs, HCB, DDE, HE, dieldrin and mirex)
at the 15 sites in 2003 and 2005 (N=90 comparisons) was made to show the variability in a short-
term (two year) assessment. TCDD was last measured in 2003, therefore for this short-term
assessment 2001 and 2003 data were used for an additional 15 comparisons. Of the total 105
comparisons, 89 (84.8%) decreased; only 16 (15.2%) increased. TCDD and PCBs were the most
frequently increasing contaminants (Canadian Wildlife Service (CWS) unpublished data). This is
illustrated for a single contaminant, PCBs, in Figure 4. Annual  fluctuations like these, including
                         Draft for Discussion at SOLEC 2006

-------
both short-term increases and decreases, are part of current contaminant patterns (Figures 1 and
4).

In terms of gross ecological effects of contaminants on colonial waterbirds, e.g. eggshell thinning,
failed reproductive success and population declines, most species appear to have recovered.
Populations of most species have increased over the past 25-30 years, e.g. see Figure 5 (Blokpoel
and Tessier 1993-1998; Austen et al. 1996; Scharf and Shugart 1998, Cuthbert et al.
2001,Weseloh et al. 2002; Morris et al. 2003, Havelka and Weseloh In review, Hebert et al. In
review, CWS unpubl. data). Although the gross effects appear to have subsided (but see Custer et
al. 1999), there are many other subtle, mostly physiological and genetic endpoints that are being
measured now that were not measured in earlier years (Fox et al. 1988, Fox 1993, Grasman et al.
1996, Yauk et al. 2000). A recent and ongoing study,  the Fish and Wildlife Health Effects and
Exposure Study, is assessing whether there are fish and wildlife health effects in Canadian Areas
of Concern (AOCs) similar to those reported for the human population (Environment Canada
2003). To date, the following abnormalities have been found in Herring Gulls in one or more
Canadian AOCs on the lower Great Lakes:  a male-biased sex ratio in hatchlings, elevated levels
of embryonic mortality, indications of feminization in more than 10% of adult males, a reduced
or suppressed ability to combat stress, an enlarged thyroid with reduced hormone production and
a suppressed immune system. Although there is little  question that Herring Gulls  and colonial
waterbirds on the Great Lakes are healthier now than  they were 30 years ago, these  findings show
that they are in a poorer state of health than are birds from clean reference sites in the Maritimes
(Environment Canada 2003).

Pressures
Future pressures for this indicator include all sources  of contaminants which reach the Great
Lakes. These include those sources that are already well-known, e.g., point sources, re-suspension
of sediments, and atmospheric inputs, as well as lesser known ones such as underground leaks
from landfill sites. There are also other, non-contaminant factors that regulate the stability of
populations, e.g. habitat modification (in the Detroit River), food availability (Lake Superior),
interspecific competition at breeding colonies (Lake Ontario) and predation (western Lake Erie).
Many of these factors pose much more tangible threats to our ability to collect eggs from these
colonies in the future.

Management Implications
Data from the Herring Gull Egg Monitoring Program suggest that, for the most part, contaminant
levels in wildlife are continuing to decline at a constant rate. However, even at current
contaminant levels, more physiological abnormalities in Herring Gulls occur at Great Lakes sites
than at cleaner, reference sites away from the Great Lakes basin. Also, with the noted increase in
concentrations of polybrominated diphenyl ethers  (PBDEs), steps should be taken to identify and
reduce sources of this compound to the Great Lakes. In short, although almost all contaminants
are decreasing and many biological impacts have lessened, we do not yet know the  full health
implications of the subtle effects and of newly monitored contaminants.

Future Activities
The annual collection and analysis of herring gull  eggs from 15 sites on both sides of the Great
Lakes and the assessment of this species' reproductive success is a permanent part of the CWS
                         Draft for Discussion at SOLEC 2006

-------

Great Lakes surveillance activities. Likewise, so is the regular monitoring of population levels of
most of the colonial waterbird species. The plan is to continue these procedures. Research on
improving and expanding the Herring Gull Egg Monitoring Program is done on a more
opportunistic, less predictable basis. A lake-by-lake intensive study of possible biological impacts
to herring gulls is currently underway in the lower lakes. Recently, ecological tracers (stable
isotopes and fatty acids) have been generated from archival eggs as part of the program and
provide insights into how food webs in the Great Lakes ecosystem are changing. This information
broadens the utility of the program from just examining contaminants to providing insights into
ecosystem change. Ecological tracer data are also directly relevant to the interpretation of
contaminant monitoring data.

Comments from the author(s)
We have learned much  about interpreting the Herring Gull egg contaminants data from associated
research studies.  However, much of this work is conducted on an opportunistic basis, when funds
are available. Several research activities should be incorporated into routine monitoring, e.g.
tracking of porphyria, vitamin A deficiencies, and evaluation of the avian immune system.
Likewise, more research should focus on new areas, e.g. the impact of endocrine disrupting
substances, factors regulating chemically induced genetic mutations and ecological tracers.

Acknowledgments
Authors: D.V. Chip Weseloh,  Canadian Wildlife Service, Environment Canada, Downsview, ON;
and Tania Havelka, Canadian Wildlife Service, Environment Canada, Downsview, ON.

Thanks to past and present staff at CWS-Ontario Region, including Glenn Barrett, Christine
Bishop, Birgit Braune, Neil Burgess, Rob Dobos, Pete Ewins, Craig Hebert, Kate Jermyn, Margie
Koster, Brian McHattie, Pierre Mineau, Cynthia Pekarik, Karen Pettit, Jamie Reid, Peter Ross,
Dave  Ryckman, John Struger and Stan Teeple as well as past and present staff at the CWS
National Wildlife Research Centre (Ottawa, ON), including Masresha Asrat, Glen Fox, Michael
Gilbertson, Andrew Oilman, Jim Learning, Rosalyn McNeil, Ross Norstrom, Laird Shutt, Mary
Simon, Suzanne Trudeau, Bryan Wakeford, Kim Williams and Henry Won and wildlife
biologists Ray Faber, Keith Grasman, Ralph Morris, Jim Quinn and Brian Ratcliff for egg
collections, preparation, analysis and data management over the 30 years of this project. We are
also grateful for the logistical and graphical support of the Technical Operations Division and the
Drafting Department at the Canada Centre for Inland Waters, Burlington, Ontario. Craig Hebert
reviewed an earlier version of this report.

Data  Sources
Austen, M.J., Blokpoel, H., and Tessier, G.D. 1996. Atlas of colonial waterbirds nesting on the
Canadian Great Lakes,  1989-1991. Part 4. Marsh-nesting terns on Lake Huron and the lower
Great Lakes system in 1991. Canadian Wildlife Service (CWS), Ontario Region, Technical
Report No. 217. 75pp.

Bishop, C.A., Weseloh, D.V.,  Burgess, N.B., Norstrom, R.J., Turle, R., and Logan, K.A. 1992.
An atlas of contaminants in eggs of colonial fish-eating birds of the Great Lakes (1970-1988).
Accounts by location and chemical. Volumes 1 & 2. Canadian Wildlife Service (CWS), Ontario
Region, Technical Report Nos. 152 and 153. 400 pp. and 300 pp.
                         Draft for Discussion at SOLEC 2006

-------
Blokpoel, H., and Tessier, G.D. 1993-1998. Atlas of colonial waterbirds nesting on the Canadian
Great Lakes, 1989-1991. Parts 1-3, 5. Canadian Wildlife Service (CWS), Ontario Region,
Technical Report Nos. 181, 225, 259, 272. 93 pp, 153 pp, 74 pp, 36 pp.

Braune B.M., Hebert, C.E., Benedetti, L.S., and Malone, BJ. 2003. An Assessment of Canadian
Wildlife Service Contaminant Monitoring Programs. Canadian Wildlife Service (CWS),
Technical Report No. 400. Headquarters. Ottawa, ON. 76 pp.

Custer, T.W., Custer, C.M., Hines, R.K., Gutreuter, S., Stromborg, K.L., Allen, P.D., and
Melancon, MJ. 1999. Organochlorine contaminants and reproductive success of Double-crested
Cormorants from Green Bay, Wisconsin, USA. Environ. Toxicol. & Chem. 18:1209-1217.

Cuthbert, F.J., McKearnan, J., and Joshi, A.R. 2001. Distribution and abundance of colonial
waterbirds in the U.S. Great Lakes, 1997 - 1999. Report to U.S. Fish and Wildlife Service, Twin
Cities, Minnesota.

Environment Canada. 2003. Fish and wildlife health effects in the Canadian Great Lakes Areas of
Concern. Great Lakes Fact Sheet.  Canadian Wildlife Service (CWS), Ontario Region,
Downsview, ON. Catalogue No. CW/66-223/2003E. ISBN 0-662-34076-0.

Fox, G.A. 1993. What have biomarkers told us about the effects of contaminants on the health of
fish-eating birds in the Great Lakes? The theory and a literature review. J. Great Lakes Res.
19:722-736.

Fox, G.A., Kennedy, S.W., Norstrom, R.J., and Wgfield, D.C.  1988. Porphyria in herring gulls: a
biochemical response to chemical contamination in Great Lakes food chains.  Environ. Toxicol. &
Chem. 7:831-839.

Grasman, K.A., Fox, G.A., Scanlon, P.P., and Ludwig, J.P. 1996. Organochlorine associated
immunosuppression in prefledging Caspian terns and herring gulls from the Great Lakes: an
ecoepidemiological study. Environmental Health Perspectives  104:829-842.

Havelka, T.,  and Weseloh, D.V. In review. Continued growth and expansion of the Double-
crested Cormorant (Phalacrocorax auritus) population on Lake Ontario, 1982-2002.

Hebert, C.E., Weseloh, D.V., Havelka, T., Pekarik, C., and Cuthbert, F. In review. Lake Erie
colonial waterbirds: Trends in populations, contaminant levels and diet. The State of Lake Erie.
In Ecovision World Monograph Series, Aquatic Ecosystem Health and Management Society, ed.
M. Munawar.

Jermyn-Gee, K., Pekarik, C., Havelka, T., Barrett, G., and Weseloh, D.V. 2005. An atlas of
contaminants in eggs of colonial fish-eating birds of the Great Lakes (1998-2001). Accounts by
location (Vol. I) & chemical (Vol. II). Technical Report No. 417. Canadian Wildlife Service
(CWS), Ontario Region, Downsview, ON. Catalogue No. CW69-5/417E-MRC. ISBN - 0-662-
37427-4.
                         Draft for Discussion at SOLEC 2006

-------


Morris, R.D., Weseloh, D.V., and Shutt, J.L. 2003. Distribution and abundance of nesting pairs of
Herring Gulls (Larus argentatus) on the North American Great Lakes. J. Great Lakes Res.
29:400-426.

Norstrom, R.J., Simon, M., Moisey, J., Wakeford, B., and Weseloh, D.V.C. 2002. Geographical
distribution (2000) and temporal trends (1981-2000) of brominated diphenyl ethers in Great
Lakes Gull eggs. Environ. Sci. & Technol. 36:4783-4789.

Pekarik, C., and Weseloh, D.V. 1998. Organochlorine contaminants in Herring Gull eggs from
the Great Lakes,  1974-1995: change point regression analysis and short term regression. Environ.
Monit. & Assess. 53:77-115.

Pekarik, C., Weseloh, D.V., Barrett, G.C., Simon, M., Bishop, C.A., and Pettit, K.E. 1998. An
atlas of contaminants in the eggs offish-eating colonial birds of the Great Lakes (1993-1997).
Accounts by location & chemical. Volumes I & 2. Canadian Wildlife Service (CWS), Ontario
Region, Technical Report Nos. 321 and 322. 245 pp. and 214 pp.

Pettit, K.E., Bishop, C.A., Weseloh, D.V., and Norstrom, RJ. 1994. An atlas of contaminants in
eggs of colonial fish-eating birds of the Great Lakes (1989-1992). Accounts by location &
chemical. Volumes 1 & 2. Canadian Wildlife Service (CWS), Ontario Region, Technical Report
Nos. 194 and 195. 319 pp. and 300 pp.

Scharf, W.C., and Shugart, G.W. 1998. Distribution and abundance of gull, tern and cormorant
nesting colonies of the U.S. Great Lakes, 1989 and 1990. In A.S. Publication No. 1 eds. W.W.
Bowerman and Roe, Sault Ste. Marie, MI: Gale Gleason Environmental Institute, Lake Superior
State University Press.

Weseloh, D.V.C, Faber, R.A., and Pekarik, C. 2005. Temporal and spatial trends of
organochlorine contaminants  in herring gull eggs from Lake Michigan.  Book Chapter In: Edsall,
T., and Munawar, M. (eds.) State of Lake Michigan: Ecology, Health, and management. Aquatic
Ecosystem Health & Management (AEHM), Ecovision World Monograph Series. Goodword
Books Pvt. Ltd., New Delhi, p. 393-417.

Weseloh, D.V., Joos, R.,  Pekarik, C., Farquhar, J., Shutt, L., Havelka, T., Mazzocchi, I., Barrett,
G., McCollough, R., Miller, R.L., Mathers, A. 2003. Monitoring Lake Ontario's waterbirds:
contaminants in Herring Gull eggs and population changes in the Lake's nearly 1,000,000
colonial waterbirds. Book Chapter In: Munawar, M. (ed.) State of Lake Ontario (SOLO) - Past,
Present and Future, Aquatic Ecosystem Health & Management (AEHM), Ecovision World
Monograph Series. Backhuys Publishers, Leiden, The Netherlands, p. 597-631.

Weseloh, D.V.C., Pekarik, C., Havelka, T., Barrett, G., and Reid, J. 2002. Population trends and
colony locations  of double-crested cormorants in the Canadian Great Lakes and immediately
adjacent areas, 1990-2000: a manager's guide. J. Great Lakes Res. 28:125-144.

Yauk, C.L., Fox, G.A., McCarry, B.E., and Quinn, J.S. 2000. Induced minisatellite germline
mutations in Herring Gulls (Larus argentatus) living near steel mills. Mutation Research 452:211-
218.
                         Draft for Discussion at SOLEC 2006

-------
List of Figures
Figure 1. Annual concentration of DDE in Herring Gull eggs, Toronto Harbour, 1974-2005.
Source: Environment Canada, Herring Gull Monitoring Program

Figure 2. Mean contaminant concentrations and percent decline of 7 contaminants in Herring Gull
eggs from year of first analysis to present, Middle Island, Lake Erie. Concentrations in [ig/g wet
weight except for dioxin in pg/g wet weight.
Source: Environment Canada, Herring Gull Monitoring Program

Figure 3. The distribution and locations of the 15 Herring Gull Annual Monitoring Colonies.
Source: Environment Canada, Herring Gull Monitoring Program and Canadian Wildlife Service

Figure 4. A comparison of PCB concentrations at all sites for 2003 and 2005. Note the between
year differences as well as the variation among sites.
Source: Environment Canada, Herring Gull Monitoring Program and Canadian Wildlife Service

Figure 5. Double-crested Cormorant nests (breeding pairs) on Lake Ontario, 1979-2005.
Source: Environment Canada, Canadian Wildlife Service

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------

    25
    20 --
  o
  5 1541
  c
  o

  Is 10
  i

  0)
  u
  c
  o
  o
     5 +
                                                         Unnnnn
HRrn
       1974  1977  1979  1981  1983  1985  1987  1989  1991  1993  1995  1997  1999  2001  2003  2005


                                        Year
Figure 1. Annual concentration of DDE in Herring Gull eggs, Toronto Harbour, 1974-2005.

Source: Environment Canada, Herring Gull Monitoring Program
                      Draft for Discussion at SOLEC 2006

-------
      100%-

  o,   75%-

  CD
       50%-
       25%-
                 D % remaining in 2005*
                 Q concentration measured in 1974* set to 100%
             PCB
% decline:     80%
X




x





^


•,,,,.
:.w.


14.4 1
r






/B


::22;3


0,192









f . . . . i
•.-.•.•.•
::7;02:
•.-.•.•.•
•.-.•.•.•
•.•.•.•.•
....
. . . .
•.-.•.•.•
•:•:•:•:•
•.•.•.•.•
^.004:


i
i

i
i
*


[•,,,,.i.
0.465:

	
	
0.008:
	








f- 	 1
	
	
: 0.580:
	
	
	
•'•'•'•'•'
	
	
	
	
	
	
	
	
	
: 0.003:










:0;1:55:


0.004
	
----------nja






^









:80,5:


27.6


f
»
i
i
»
i
»
i
//*
'
DDE     Mirex   Dieldrin   HCB
                            HE
 97%
99%     98%      99%      97%
                                                                       2378-
                                                                      dioxin*
                                                                        35%
 * dioxin first measured in 1984 and last measured in 2003
Figure 2. Mean contaminant concentrations and percent decline of 7 contaminants in Herring
Gull eggs from year of first analysis to present, Middle Island, Lake Erie. Concentrations in
wet weight except for dioxin in pg/g wet weight.
Source: Environment Canada, Herring Gull Monitoring Program
                        Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
                                                       1 Granite I.
                                                       2 AgiWiRlss.
                                                       4GullI.
                                                       5 Channel-Shelter I.
                                                       6 Double I.
                                                       7 Charily I.
SFightngl.
9 Middle I.
10 Port Colborre
11 HiagiraR.
12 Hamilton Hibr
13 Toronto Hrb r.
14 Snake I.
ISStrachanl
Figure 3. The distribution and locations of the 15 Herring Gull Annual Monitoring Colonies.
Source: Environment Canada, Herring Gull Monitoring Program and Canadian Wildlife Service
                        Draft for Discussion at SOLEC 2006
            11

-------

                                                  o°N
                                     Colony (west to east)

Figure 4. A comparison of PCB concentrations at all sites for 2003 and 2005. Note the between
year differences as well as the variation among sites.
Source: Environment Canada, Herring Gull Monitoring Program and Canadian Wildlife Service
                        Draft for Discussion at SOLEC 2006

-------
    30
    25
    15
    10 H
          m n  n  n a 111
       1979  1981   1983  1985  1987   1989  1991   1993  1995  1997   1999  2001   2003  2005
                                            Year

Figure 5. Double-crested Cormorant nests (breeding pairs) on Lake Ontario, 1979-2005.
Source: Environment Canada, Canadian Wildlife Service
                         Draft for Discussion at SOLEC 2006
13

-------

Zooplankton Populations
Indicator #116

Overall Assessment
           Status:   Mixed
           Trend:   Not Assessed

   Primary Factors   Changes in community structure are occurring in lakes Michigan,
      Determining   Huron, and Ontario due to declines in cyclopoid copepods and
  Status and Trend   cladocerans. Summer mean size has increased in these lakes
                   concurrent with the increase in the percent of calanoid copepods
Lake-by-Lake Assessment
Lake Superior
           Status:   Good
           Trend:
   Primary Factors
      Determining
  Status and Trend
Unchanging
Stable summer zooplankton community dominated by large calanoid
copepods.
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Assessed
Undetermined (changing)
Total summer biomass  has been declining since 2004 due to fewer
Daphnia and cyclopoid copepods. Summer mean size of zooplankton is
increasing.
Not Assessed
Undetermined (changing)
Total summer biomass  has declined dramatically since 2003 due to fewer
Daphnia, bosminids, and cyclopoid copepods. Summer mean size of
zooplankton is increasing.
Not Assessed
Undetermined
Variable biomass and composition of summer crustacean zooplankton
community in each basin. Most diverse zooplankton community in Great
Lakes. Very low biomass in Western Basin in August, 2001.
Not Assessed
Undetermined (changing)
Lowest percentage of calanoid copepods of all Great Lakes. Total summer
biomass has declined since 2004 due to a decline in cyclopoid copepods.
Summer mean size of zooplankton is increasing.
                        Draft for Discussion at SOLEC 2006

-------
                                                                                    	
Purpose
•To directly measure changes in community composition, mean individual size and biomass of
zooplankton populations in the Great Lakes basin;
•To indirectly measure zooplankton production; and
•To infer changes in food-web dynamics due to changes in vertebrate or invertebrate predation,
system productivity, the type and intensity of predation, and the energy transfer within a system.

Ecosystem Objective
Ultimately, analysis of this indicator should provide information on the biological integrity of the
Great Lakes, and lead to the support of a healthy and diverse fishery. Suggested metrics include
zooplankton mean length, the ratio of calanoid copepod abundance to that of cyclopoid copepods
plus cladocerans and zooplankton biomass. However, the relationship between these objectives
and the suggested metrics have not been fully worked out, and no specific criteria have yet been
identified for these metrics.

Planktivorous fish often feed size selectively, removing larger cladocerans and copepods.  High
densities of planktivores result in a reduction of the mean size of zooplankton in a community.  A
mean individual size of 0.8 mm has been suggested as "optimal" for zooplankton communities
sampled with a  153 (im mesh net, indicating a balance between planktivorous and piscivorous
fish. Declines in mean size of crustacean zooplankton between spring and late summer may
indicate increased predation by young fish or the presence of a greater proportion of immature
zooplankton. Interpretation of deviations from this average size objective, and the universality of
this objective remain unclear at this time. In particular, questions regarding its applicability to
systems impacted by predaceous cladocereans and dreissenids as well as planktivorous fish have
been raised.

Gannon and Stemberger (1978) found that cladocerans and cyclopoid copepods are more
abundant in nutrient enriched waters of the Great Lakes, while calanoid copepods dominate
oligotrophic communities. They reported that areas of the Great Lakes where the density of
calanoid copepods comprises over 50% of the summer crustacean zooplankton community (or the
ratio of calanoids/cyclopoids + cladocerans >1) could be classified  as oligotrophic. As with
individual mean size, though, clear objectives have not presently been defined.

State of the Ecosystem
Summer biomass of crustacean zooplankton communities in the offshore waters of Lake Superior
has remained at a relatively low but stable level for the past seven years (Figure 1). The plankton
community is dominated by large calanoid copepods (Leptodiaptomus sicilis and Limnocalanus
macrurus) that are characteristic of oligotrophic, cold water ecosystems. Biomass is generally
higher in the nutrient enriched lower lakes with more annual variation produced by seasonal
increases in cladocerans, primarily daphnids and bosminids.   Since 2003 the biomass of
cladocerans and cyclopoid copepods in Lake Huron has declined dramatically.  Data from 2005
suggests that a similar decline may now be occurring in Lake Michigan. Cyclopoid abundance
has also begun to decline in Lake Ontario. Mechanisms for these declines are not known at this
                         Draft for Discussion at SOLEC 2006

-------

time, but may be related to changes in nutrient levels, phytoplankton composition, exotic species
interactions, or fish predation pressure.

The proportion of calanoid copepods in Lake Superior has remained fairly stable at 70% (Figure
2) indicating oligotrophic conditions.  Summer zooplankton communities in Lakes Michigan and
Huron have shown an increasing proportion of calanoid copepods in recent years, suggesting an
improved trophic state.  Lake Ontario has the lowest proportion of calanoids, followed closely by
the nutrient enriched western basin of Lake Erie.  Values for the central and eastern basins of
Lake Erie are at intermediate levels and exhibit considerable annual variation.

Historical comparisons  of this metric are difficult to make because most historical data on
zooplankton populations in the Great Lakes seem to have been generated using shallow (20 m)
tows. Calanoid copepods tend to be deep living organisms; therefore the use of data generated
from shallow tows would tend to contribute a strong bias to this metric. This problem is largely
avoided in Lake Erie, particularly in the western and central basins, where most sites are
shallower than 20 m. Comparisons in those two basins have shown a statistically significant
increase in  the ratio  of calanoids to cladocerans and cyclopoids between 1970 and 1983-1987,
with this increase sustained throughout the 1990s. A similar increase was seen in the eastern
basin, although some of the data used to calculate the ratio were generated from shallow tows and
are therefore subject to  doubt.

Mean length of crustacean zooplankton in the  offshore waters of the Great Lakes is generally
greater in the spring than during the summer (Figure 3). In the spring, mean zooplankton size in
all of the Great Lakes is near the suggested level of 0.8 mm. Mean length in Lake Superior
declines during the summer due to the production of immature copepodids, but is still above the
criterion. Summer mean length in Lakes Huron and Michigan remain high and have begun to
show an increase in  recent years.  In Lakes Erie and Ontario, the mean length of zooplankton
declines considerably in the summer. Whether this decline is  due to predation pressure or to the
increased abundance of bosminids (0.4 mm mean length) and immature cyclopoids (0.65 mm
mean length) is unknown.

Historical data from the eastern basin of Lake  Erie, from 1985 to 1998, indicate a fair amount of
interannual variability in zooplankton mean length, with values from offshore sites ranging from
about 0.5 mm to  0.85 mm (Figure 4). As noted above, interpretation of these data are currently
problematic.

Pressures
The zooplankton community might be expected to respond to changes  in nutrient and
phytoplankton concentrations  in the lakes, although the potential magnitude of such "bottom  up"
effects is not well understood. The most immediate potential threat to the zooplankton
communities of the Great Lakes is posed by invasive species.  The continued proliferation of
dreissenid populations can be expected to impact zooplankton communities through the alteration
of the structure and abundance of the phytoplankton community, upon which many zooplankton
depend for  food.  Predation from the exotic cladocerans Bythotrephes longimanus and Cercopagis
pengoi may also  have an impact on zooplankton abundance and community composition.
Bythotrephes has been in the Great Lakes for approximately twenty years, and is suspected to
have had a  major impact on zooplankton community structure. Cercopagis pengoi was first noted
                         Draft for Discussion at SOLEC 2006

-------
                                                  *J^~~'...:...^...:...:....qHl.  ...J!!""""™
in Lake Ontario in 1998, and has now spread to the other lakes, although in much lower densities.
Continuing changes in predation pressure from planktivorous fish may also impact the system

Management Implications
Continued monitoring of the offshore zooplankton communities of the Great Lakes is critical,
particularly considering the current expansion of the range of the non-native cladoceran
Cercopagis and the probability of future invasive zooplankton and fish species.

Comments from the author(s)
Currently the most critical need is for the development of quantitative, objective criteria that can
be applied to the zooplankton indicator. The applicability of current metrics to the Great Lakes is
largely unknown, as are the limits that would correspond to acceptable ecosystem health.

The implementation of a long-term monitoring program on the Canadian side is also desirable to
expand both the spatial and the temporal coverage currently provided by American efforts. Since
the interpretation of various indices is dependent to a large extent upon the sampling methods
employed, coordination between these two programs, both with regard to sampling dates and
locations, and especially with regard to methods, would be highly recommended.

Acknowledgments

Authors and Contributors:

Mary Balcer, University of Wisconsin-Superior, Superior, WI mbalcer@uwsuper.edu;
Richard P. Barbiero, Computer Sciences Corporation, Chicago, IL, Chicago, IL;
Marc L. Tuchman, U.S. Environmental Protection Agency, Great Lakes National Program Office,
Chicago, IL; and
Ora Johannsson, Fisheries and Oceans Canada, Burlington, Ontario Canada.

Data Sources
Johannsson, O.E., Dumitru, C.,  and Graham, D.M. 1999. Examination of zooplankton mean
length for use in an index of fish community structure and application in Lake Erie. J. Great
Lakes Res. 25:179-186.

U.S. Environmental Protection Agency, Great Lakes National Program Office, Chicago, IL,
Biological Open Water Surveillance Program of the Laurentian Great Lakes, unpublished data
(2000-2005), produced through cooperative  agreement GL-96513791 with the University of
Wisconsin-Superior.

List of Figures
Figure  1. Average composition of crustacean zooplankton biomass at Great Lakes offshore
stations sampled in August of each year.  Samples were collected with 153(im mesh net tows to a
depth of 100 m or the bottom of the water column, whichever was shallower. Source: U.S.
Environmental Protection Agency, Great Lakes National Program Office.
                         Draft for Discussion at SOLEC 2006

-------

                                                      ^ ._
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''
Figure 2. Average percentage of calanoid copepods (by abundance) in crustacean zooplankton
communities from Great Lakes offshore stations sampled in August of each year. Samples were
collected with 153(un mesh net tows to a depth of 100 m or the bottom of the water column,
whichever was shallower. Line at 50% level is the suggested criterion for oligotrophic lakes.
Source: U.S. Environmental Protection Agency, Great Lakes National Program Office.

Figure 3. Average individual mean lengths of crustacean zooplankton in the Great Lakes in May
and August. Length estimates were generated from data collected with 153(im mesh net tows to a
depth of 100 m or the bottom of the water column, whichever was shallower. Values are the
indicate arithmetic averages of all sites sampled. Line at 0.8 mm is the suggested criterion for
balanced fish community. Source: U.S. Environmental Protection Agency, Great Lakes National
Program Office.

Figure 4. Trend in Jun27-Sep30 mean zooplankton length: NYDEC data (circles) collected with
153(un mesh net, DFP data (diamonds) converted from 64[un to 153(un mesh equivalent. Open
symbols = offshore, solid symbols = nearshore (<12m).  1985-1988 are means +/- 1 S.E.
Source: Johannsson et al. 1999.

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
                                 State of the Great Lakes 2007 - Draft
        1998|1999|2001|2002|2003|2004|2005

                   Superior
   IDaphnia BBosminids DOther Cladocerans D Immature Cyclopoids • Adult Cyclopoids DImmature Calanoids • Adult Calanoids
   IDaphnia BBosminids D Other Cladocerans D Immature Cyclopoids • Adult Cyclopoids D Immature Calanoids • Adult Calanoids
Figure 1. Average composition of crustacean zooplankton biomass at Great Lakes offshore
stations sampled in August of each year.  Samples were collected with 153um mesh net tows to a
depth of 100 m or the bottom of the water column, whichever was shallower. Source: U.S.
Environmental Protection Agency, Great Lakes National Program Office.
                         Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
Figure 2. Average percentage of calanoid copepods (by abundance) in crustacean zooplankton
communities from Great Lakes offshore stations sampled in August of each year.  Samples were
collected with 153(un mesh net tows to a depth of 100 m or the bottom of the water column,
whichever was shallower. Line at 50% level is the suggested criterion for oligotrophic lakes.
Source: U.S. Environmental Protection Agency, Great Lakes National Program Office.
                       Draft for Discussion at SOLEC 2006

-------
                              State of the Great Lakes 2007 - Draft
                                        Spring
Figure 3. Average individual mean lengths of crustacean zooplankton in the Great Lakes in May
and August. Length estimates were generated from data collected with 153(im mesh net tows to a
depth of 100 m or the bottom of the water column, whichever was shallower. Values are the
indicate arithmetic averages of all sites sampled. Line at 0.8 mm is the suggested criterion for
balanced fish community. Source: U.S. Environmental Protection Agency, Great Lakes National
Program Office.
                       Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
 E
 E
 O)
 C
 OJ
 TO
 0)
 C
 o
2
 Q.
 O
 O
N
       1 00 - Eastern Lake Erie
       0,80 -
       0.60 -
0.40
0.20
       0.00
                                            Objective (Mills etal. 1987}
                                          *".
           1984    1986    1988    1990    1992    1994     1996    1998


                                        Year

Figure 4. Trend in Jun27-Sep30 mean zooplankton length: NYDEC data (circles) collected with
153(un mesh net, DFP data (diamonds) converted from 64[un to 153(un mesh equivalent. Open
symbols = offshore, solid symbols = nearshore (<12m). 1985-1988 are means +/- 1 S.E.
Source: Johannsson et al. 1999.
                      Draft for Discussion at SOLEC 2006

-------

Atmospheric Deposition of Toxic Chemicals
Indicator #117
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving for polychlorinated biphenyls (PCBs), banned
organochlorine pesticides, and dioxins and furans
Unchanging or slightly improving for polycyclic aromatic
hydrocarbons (PAHs) and mercury
Mixed since different chemical groups have different trends over time;
levels in cities can be much higher than in rural areas
Lake-by-Lake Assessment
The indicator status is mixed for all Lakes. Levels of PBT chemicals in air tend to be lower over
Lakes Superior and Huron than over the other three Lakes (which are more impacted by human
activity), but their surface area is larger, resulting in a greater importance of atmospheric inputs.

While concentrations of some of these substances are very low at rural sites, they may be much
higher in "hotspots" such as urban areas. Lakes Michigan, Erie, and Ontario have greater inputs
from urban areas. The Lake Erie station tends to have higher levels than the other remote master
stations, most likely since it is located closer to an urban area (Buffalo, NY) than the other master
stations; it may also receive some influence from the East Coast of the U.S.

In general for PBT chemicals, atmospheric inputs dominate for Lakes Superior, Huron, and
Michigan due to their large surface areas (Strachan and Eisenreich 1991, Kreis 2005).
Connecting channel inputs dominate for Lakes Erie and Ontario, which have smaller surface
areas.

Purpose
•  To estimate the annual average loadings of persistent bioaccumulative toxic (PBT) chemicals
   from the atmosphere to the Great Lakes;
•  To determine trends over time in contaminant concentrations;
•  To infer potential impacts of toxic chemicals from atmospheric deposition on human health
   and the Great Lakes aquatic ecosystem; and
•  To track the progress of various Great Lakes programs toward virtual elimination of toxic
   chemicals to the Great Lakes.

Tracking atmospheric inputs is important since the air is a primary pathway by which PBTs reach
the Great Lakes. Once PBTs reach the Great Lakes, they can bioaccumulate in fish and other
wildlife and cause fish consumption advisories.

Ecosystem Objective
The Great Lakes Water Quality Agreement (GLWQA)  and the Binational Toxics Strategy both
state the virtual elimination of toxic substances in the Great Lakes as an objective. Additionally,
GLWQA General Objective (d) states that the Great Lakes should be free from materials entering
                        Draft for Discussion at SOLEC 2006

-------
                                                                                   	
the water as a result of human activity that will produce conditions that are toxic to human,
animal, or aquatic life.

State of the Ecosystem
 The Integrated Atmospheric Deposition Network (IADN) consists of five master sampling sites,
one near each of the Great Lakes, and several satellite stations. This joint United States-Canada
project has been in operation since 1990. Since that time, thousands of measurements of the
concentrations of PCBs, pesticides, PAHs and trace metals have been made at these sites.
Concentrations are measured in the atmospheric gas and particle phases and in precipitation.
Spatial and temporal trends in these concentrations and atmospheric loadings to the Great Lakes
can be examined. Data from other networks are used here to supplement the IADN data for
mercury, dioxins and furans.

PCBs. Concentrations of gas-phase PCBs (EPCB) have generally decreased  over time at the
master stations (Figure 1). EPCB is a suite of congeners that make up most of the PCB mass and
represent the full range of PCBs. Some increases are seen during the late 1990s for Lakes
Michigan and Erie and during 2000-2001 for Lake Superior. These increases  remain unexplained,
although there is some evidence of connections with atmospheric circulation phenomena such as
El Nino (Ma et al. 2004a). Levels decrease again by 2002. It is assumed that PCB concentrations
will continue to decrease slowly.  It should be noted that PCBs in precipitation samples at the
rural master stations are nearing levels of detection.

The Lake Erie site consistently shows relatively elevated EPCB concentrations compared to the
other master stations.  Back-trajectory analyses have shown that this is due to  possible influences
from upstate New York and the East Coast (Hafner and Kites 2003). Figure 2 shows that EPCB
concentrations at urban satellite stations in Chicago and Cleveland are about fifteen and ten times
higher, respectively, than at the remote master stations at Eagle Harbor (Superior) and Sleeping
Bear Dunes (Michigan).

Pesticides. In general, concentrations of banned or restricted pesticides measured by the IADN
(such as hexachlorocyclohexane [a-HCH] and DDT) are decreasing over time in air and
precipitation (Sun et al. 2006a, Sun et al. submitted).  Concentrations of chlordane are about ten
times higher at the urban stations than at the more remote master stations, most likely due to the
use of chlordane as a termiticide in buildings. Dieldrin shows a similar urban elevation; this
pesticide was also used as a termiticide until 1987, after all other uses were banned in  1974.
Current-use pesticide endosulfan shows mixed trends, with significant decreases at some sites in
some phases, but no trends at other sites.  Concentrations of endosulfan were  generally higher in
the summer, following application of this current-use pesticide (Sun et al. submitted).

PAHs. In general, concentrations of poly cyclic aromatic hydrocarbons can be roughly correlated
with population, with highest levels in Chicago and Cleveland, followed by the semi-urban site at
Sturgeon Point, and lower concentrations at the other remote master stations.  In general, PAH
concentrations in Chicago and Cleveland are about ten to one hundred times higher than at the
master stations.
                         Draft for Discussion at SOLEC 2006

-------


Concentrations of PAHs in the particle and gas phase are decreasing at Chicago, with half-lives
ranging from 3-10 years in the vapor phase and 5-15 years in the particle phase. At the other
sites, most gas phase PAH concentrations showed significant, but slow long-term decreasing
trends (>15 years).  For most PAHs, decreases on particles and in precipitation were only found
at Chicago (Sun et al. 2006b, Sun et al. submitted).

An example of a PAH is benzo[a]pyrene (BaP), a PAH, is produced by the incomplete
combustion of almost any fuel and is a probable human carcinogen. Figure 3 shows the annual
average particle-phase concentrations of BaP.

Dioxins and Furans. Concentrations  of dioxins and furans have decreased over time (Figure 4)
with the largest declines in areas with the highest concentrations (unpublished data, T. Dann,
Environment Canada 2006).

Mercury. Data from the Canadian Atmospheric Mercury Network (CAMNet) for the IADN
stations at Egbert, Point Petre,  and Burnt Island show decreases in total gaseous mercury (TGM)
concentrations between 1995 and 2004, with more of the decrease occurring in the 2000-2004
time period (Figure 5).  Median TGM  concentrations decreased by 7-19% from 2000 to 2004 for
those stations (Temme et al. 2006).

Data from the Mercury  Deposition Network (MDN) show that concentrations of mercury in
precipitation are decreasing for much of the U.S., but there is no trend for the stations in the upper
Midwest (Gay et al. 2006).

PBDEs.  Total PBDE concentrations during 2003-2004 were in the single pg/m3 range for the
rural master stations and in the 50-100 pg/m3 range  for the urban stations (Venier 2006).  This is
lower than total PCB levels, which are generally in the 10s to  100s of pg/m3 range. A meta-
analysis of PBDE concentrations in various environmental compartments and biota worldwide
revealed exponentially increasing concentrations with doubling times of about 4-6 years and
higher levels in North America than in Europe (Kites 2004).  US manufacturers of penta- and
octa-PBDEs phased out production in 2004, but deca-PBDEs are still being produced.  Future
data will  confirm whether PBDEs increase or decrease in the air of the Great Lakes.

Loadings. An atmospheric loading is  the amount of a pollutant entering a lake from the air,
which equals wet deposition (rain) plus dry deposition (falling particles) plus gas absorption into
the water minus volatilization out of the water. Absorption minus volatilization equals net gas
exchange, which is the most significant part of the loadings for many semi-volatile PBT
pollutants. For many banned or restricted substances that IADN monitors,  net atmospheric inputs
to the lake are headed toward equilibrium; that is, the amount going into the lake equals the
amount volatilizing out. Current-use pesticides, such as g-HCH (lindane) and endosulfan, as well
as PAHs  and trace metals, still have net deposition from the atmosphere to the Lakes.

A report on the atmospheric loadings of these compounds to the Great Lakes for data through
2004 will be published in late 2006 or  early 2007. It will be available online at:
http://www.epa.gov/glnpo/monitoring/air/iadn/iadn.html.
To receive a hardcopy, please contact one of the agencies listed at the end of this report.
                         Draft for Discussion at SOLEC 2006

-------
                                                                                    	
Pressures
Atmospheric deposition of toxic compounds to the Great Lakes is likely to continue into the
future. The amount of compounds no longer in use, such as most of the organochlorine pesticides,
may decrease to undetectable levels, especially if they are phased out in developing countries, as
is being called for in international agreements.

Residual sources of PCBs remain in the U.S. and throughout the world; therefore, atmospheric
deposition will still be significant at least decades into the future. PAHs and metals continue to be
emitted  and therefore concentrations of these substances may not decrease or will decrease very
slowly depending on further pollution reduction efforts or regulatory requirements. Even though
emissions from many sources of mercury and dioxin have been reduced over the past decade,
both pollutants are still seen at elevated levels in the environment. This problem will continue
unless the emissions of mercury and dioxin are reduced further.

Atmospheric deposition of chemicals of emerging concern, such as brominated flame retardants
and other compounds that may currently be under the radar, could also serve as a future stressor
on the Great Lakes. Efforts are being made to screen for other chemicals of potential concern,
with the intent of adding such chemicals to Great Lakes monitoring programs given available
methods and sufficient resources.

Management Implications
In terms of in-use agricultural chemicals, such as lindane, further restrictions on the use of these
compounds may be warranted. Transport of lindane to the Great Lakes following planting of
lindane-treated canola seeds in the Canadian prairies has been demonstrated by modellers (Ma et
al. 2004b).  On January 1, 2005, Canada withdrew registration of lindane for agricultural pest
control;  lindane is still registered for use in the U.S.

Controls on the emissions of combustion systems, such as those in factories and motor vehicles,
could decrease inputs of PAHs to the Great Lakes' atmosphere.

Although concentrations of PCBs continue to decline slowly, somewhat of a "leveling-off'  trend
seems to be occurring in air, fish, and other biota as shown  by various long-term monitoring
programs. Remaining sources of PCBs, such as contaminated sediments, sewage sludge, and in-
use electrical equipment, may need  to be  addressed more systematically through efforts like the
Canada-U.S. Binational Toxics Strategy and national regulatory programs in order to see more
significant declines. Many such sources are located in urban areas, which is reflected by the
higher levels of PCBs measured in Chicago and Cleveland by IADN, and by other researchers in
other areas  (Wethington and Hornbuckle  2005; Totten et al. 2001). Research to investigate the
significance of these remaining sources is underway. This is important since fish consumption
advisories for PCBs exist for all five Great Lakes.

Progress has been made in reducing emissions of dioxins and furans, particularly through
regulatory controls on incinerators.  Residential garbage burning (burn barrels) is now the largest
current source of dioxins and furans (Environment Canada  and U.S. Environmental Protection
Agency 2003). Basin- and nationwide efforts are underway to eliminate emissions from burn
barrels.
                         Draft for Discussion at SOLEC 2006

-------
Regulations on coal-fired electric power plants, the largest remaining source of anthropogenic
mercury air emissions, will help to decrease loadings of mercury to the Great Lakes.

Pollution prevention activities, technology-based pollution controls, screening of in-use and new
chemicals, and chemical substitution (for pesticides, household, and industrial chemicals) can aid
in reducing the amounts of toxic chemicals deposited to the Great Lakes. Efforts to achieve
reductions in use and emissions of toxic substances worldwide through international assistance
and negotiations should also be supported, since PBTs used in other countries can reach the  Great
Lakes through long-range transport.

Continued long-term monitoring of the atmosphere is necessary in order to measure progress
brought about by toxic reduction efforts. Environment Canada and USEPA are currently adding
dioxins and PBDEs to the IADN as funding allows. Mercury monitoring at Canadian stations is
being conducted through the CAMNet.  Additional urban monitoring is needed to better
characterize atmospheric deposition to the Great Lakes.

Acknowledgments
This report was prepared on behalf of the IADN Steering Committee by Melissa Hulting, IADN
Program Manager, U.S. Environmental Protection Agency, Great Lakes National Program Office.
Thanks to Tom Dann of Environment Canada's National Air Pollution Surveillance Network
(NAPS) for dioxin and furan information, David Gay of the Mercury Deposition Network for
mercury in precipitation information, and Ron Kites and Marta Venier of Indiana University for
PBDE data.

IADN Contacts
IADN Principal Investigator, Environment Canada, Science and Technology Branch, 4905
Dufferin Street, Toronto, Ontario,  M3H 5T4
Pierrette Blanchard, pierrette.blanchard@ec.gc.ca

IADN Program Manager, Great Lakes National Program Office, U.S. Environmental Protection
Agency, 77 West Jackson Boulevard (G-17J), Chicago, IL, 60604
Melissa Hulting, hulting.melissa@epa.gov

Data Sources
Environment Canada and U.S. Environmental Protection Agency. 2003. The Great Lakes
Binational Toxics Strategy 2002 Annual Progress Report.
http://binational.net/bns/2002/index.html last accessed 11.03.05.

Gay, D., Prestbo, E., Brunette, B., Sweet, C. 2006. Wet Deposition of Mercury in the U.S. and
Canada, 1996-2004: Results from the NADP Mercury Deposition Network (MDN).  Workshop:
What do we know about mercury deposition in the upper Midwest?  February 22, 2006.
Rosemont, IL.

Hafner, W.D., and Kites, R.A. 2003. Potential Sources of Pesticides, PCBs, and PAHs to the
Atmosphere of the Great Lakes. Environmental Science and Technology 37(17):3764-3773.
                         Draft for Discussion at SOLEC 2006

-------

Kites, R.A. 2004. Polybrominated Diphenyl Ethers in the Environment and in People: A Meta-
Analysis of Concentrations. Environmental Science and Technology 38(4):945-956.

Kreis, R. 2005. Lake Michigan Mass Balance Project: PCB Results.  October 28, 2005.  Grosse
He, MI.  Online at:  http://www.epa.gov/med/grosseile_site/LMMBP/

Ma, J., Hung, H., and Blanchard, P. 2004a. How Do Climate Fluctuations Affect Persistent
Organic Pollutant Distribution in North America? Evidence from a Decade of Air Monitoring.
Environmental Science and Technology 38(9):2538-2543

Ma, J., Daggupaty,  S., Harner, T., Blanchard, P., and Waite, D. 2004b. Impacts of Lindane Usage
in the Canadian Prairies on the Great Lakes Ecosystem: 2. Modeled Fluxes and Loadings to the
Great Lakes. Environmental Science and Technology 38(4):984-990.

Strachan, W. M. J.; Eisenreich, S. J.  1990. Mass Balance Accounting of Chemicals in the Great
Lakes. In Long Range Transport  of Pesticides, ed. D. A. Kurtz, pp. 291-301. Chelsea, Michigan:
Lewis Publishers.

Sun, P.,  Backus, S., Blanchard, P., Kites, R.A. 2006. Temporal and Spatial Trends of
Organochlorine Pesticides in Great Lakes Precipitation. Environmental Science and Technology
40(7): 2135-2141.

Sun, P.,  Blanchard, P., Brice, K.A., Kites, R.A.  Atmospheric Organochlorine Pesticide
Concentrations near the Great Lakes: Temporal and Spatial Trends.  Environmental Science and
Technology, submitted.

Sun, P.,  Backus, S., Blanchard, P., Kites, R.A. 2006. Annual Variation of Polycyclic
Aromatic Hydrocarbon Concentrations in Precipitation Collected near the Great Lakes.
Environmental Science and Technology 40(3): 696-701.

Sun, P.,  Blanchard, P., Brice, K.A. and Kites, R.A. Trends in Polycyclic Aromatic Hydrocarbon
Concentrations in the Great Lakes Atmosphere. Environmental Science and Technology,
submitted.

Temme, C., Blanchard P., Steffen, A., Banic, C., Beauchamp, S., Poissant, L., Tordon, R., Wiens
B. and Dastoor, A.  2006. Long-Term Trends of Total Gaseous Mercury Concentrations
from Selected CAMNet Sites (1995-2005). Great Lakes Binational Toxics Strategy Stakeholders
Forum.  May 17, 2006.  Toronto, Ontario.

Totten, L.A., Brunciak, P.A., Gigliotti, C.L., Dachs, J., Glenn, T.R., IV, Nelson, E.D., and
Eisenreich, SJ. 2001. Dynamic Air-Water Exchange of Poly chlorinated Biphenyls in the New
York-New Jersey Harbor Estuary. Environmental Science and Technology 35(19):3834-3840.

Venier, M., Hoh, E., and Kites, R.A.  2006. Atmospheric Brominated Flame Retardants and
Dioxins in the Great Lakes. 49th Annual Conference on Great Lakes Research. May 25, 2006.
University of Windsor, Windsor, Ontario, Canada.
                         Draft for Discussion at SOLEC 2006

-------



Wethington, D.M., III, and Hornbuckle, K.C. 2005. Milwaukee, WI as a Source of Atmospheric
PCBs to Lake Michigan. Environmental Science and Technology 39(l):57-63.

List of Figures
Figure 1. Annual Average Gas Phase Concentrations of Total PCBs (PCB Suite).
Source: Integrated Atmospheric Deposition Network (IADN) Steering Committee, unpublished,
2006.

Figure 2. Gas Phase PCB concentrations for rural sites versus urban areas.
Source:  IADN Steering Committee, unpublished, 2006.

Figure 3. Annual Average Particulate Concentrations of Benzo(a)pyrene.
Source:  IADN Steering Committee, unpublished, 2006.

Figure 4. Concentrations of dioxins and furans expressed as TEQ (Toxic Equivalent) in fg/m3 in
Windsor, Ontario.
Source: Environment Canada National Air Pollution Surveillance (NAPS) network, unpublished,
2006.

Figure 5. Trends from 2000 to 2004 for median concentrations of total gaseous mercury (ng/m3)
at CAMNet stations.
Source: Temme et al. 2006.

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
                             State of the Great Lakes  2007 - Draft
                                                                     Superior
                                                                     Michigan
                                                                     Huron
                                                                     Erie
                                                                     Ontario
         CV    ^0    V"   ^O    ^O    ^     Co    O5    O   *~~
         050505050505050500
         050505050505050500
Figure 1. Annual Average Gas Phase Concentrations of Total PCBs (PCB Suite).
Source: Integrated Atmospheric Deposition Network (IADN) Steering Committee, unpublished,
2006
                                                                • 2000
                                                                • 2001
                                                                • 2002
                                                                D2003
                                                                • 2004
                                                      «r
Figure 2. Gas Phase PCB concentrations for rural sites versus urban areas.
Source: IADN Steering Committee, unpublished, 2006
                       Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
                                                                    Superior
                                                                    Michigan
                                                                    Huron
                                                                    Erie
                                                                    Ontario
  O)
  a.
Figure 3. Annual Average Particulate Concentrations of Benzo(a)pyrene.
Source: IADN Steering Committee, unpublished, 2006
                       1995   1997  1999   2001   2003  2005
                                     Windsor
Figure 4. Concentrations of dioxins and furans expressed as TEQ (Toxic Equivalent) in fg/m3 in
Windsor, Ontario.
Source: Environment Canada National Air Pollution Surveillance (NAPS) network, unpublished,
2006
                       Draft for Discussion at SOLEC 2006

-------
                               State of the Great Lakes 2007 - Draft
                   Trend 2000 to 2004 for selected CAMNet stations/categories
                        Final trends for 2000 - 2004, based on difference of
                         annual median TGM concentrations (+ M-W U-test)
1 2 DO



               -•


                                      Stationi'Category
      iicfiff. Median (-)
                TT diff, Median (+)     - Median 2000
Median 2004
     Figure 5. Trends from 2000 to 2004 for median concentrations of total gaseous mercury (ng/m )
     at CAMNet stations.
     Source: Temme et al. 2006
     10
Draft for Discussion at SOLEC 2006

-------

                                  ^ffe*l'if?lif9"to r%"ilfS
                                  • fsiSsfs?^-? &9m I
                                    'as; "u. !'"»'
Toxic Chemical Concentrations in Offshore Waters
Indicator #118
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
Data for this indicator is not available system-wide for all chemicals.

Concentrations of most organic compounds are low and are declining in the
open waters of the Great Lakes, indicating progress in the reduction of
persistent toxic substances. Insufficient data are available at this time to
make a robust determination of the recent trend in concentrations of all
compounds.

Generally, organochlorine pesticide concentrations exhibit a north to south
gradient from lowest to highest (Superior
-------
                   Little or no information is currently available for some compounds, such as
                   dioxins, in offshore waters.  Concentrations of these compounds are
                   extremely low and difficult to detect in lake water samples. It may be more
                   appropriate to measure them in fish and/or sediment samples. Information
                   about compounds of new and emerging concern is being assessed and
                   information should be available for a future SOLEC update.
Lake-by-Lake Assessment
Lake Superior
           Status:  Fair
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Undetermined
Thirteen of a possible 21 organochlorines were detected in Lake Superior
and their concentrations were generally very low. Their presence is most
likely due to atmospheric deposition because the traditional sources (row-
crop agriculture and urban land uses) are low in this basin.  For example,
concentrations of the insecticide Dieldrin (Figure 2) reflect its usage in the
agricultural communities of the southern Great Lakes basin and are low in
Lake Superior (2005: open lake average = 0.11 ng/L). In contrast,
concentrations of Lindane (Figure 3), which was previously used in North
American agriculture, reflect greater atmospheric deposition in the north
(2005: open lake average = 0.31 ng/L).

Mercury concentrations in Lake Superior were very low offshore (2005
open lake average 0.41 ng/L), with higher concentrations near Thunder Bay
and Duluth.  With the exception of one station near Duluth, all samples met
the US EPA Great Lakes Initiative (GLI)  water quality criterion for
protection of wildlife of 1.3 ng/L.

PAHs are present throughout the Lake at extremely low concentrations.
Concentrations were many orders of magnitude below Ontario Water
Quality  Guidelines.  For example, the open lake average concentration of
Phenanthrene (Figure 4) was 0.03 ng/L and  the Ontario Guideline is 30
ng/L.

Fair
Undetermined
Preliminary data from 2004 indicate that concentrations of PCBs and
organochlorine pesticides have either decreased slightly or remained
constant since the mid-1990s, following a decrease in the 1970s through the
early 1990s. 2005 total mercury concentrations were all below the U.S.
EPA's Great Lakes Initiative (GLI) water quality criterion for protection of
wildlife  of 1.3 ng/L. Atrazine concentrations in the open lake waters were
consistent across Lake Michigan stations with an average concentration
ranging  from 33 to 48 ng/L between 1994 and 2000;  this is more than 50
times below the maximum concentration allowed for drinking water
                         Draft for Discussion at SOLEC 2006

-------

                                 ^ffe^'ifrlif^'to ^TilfS
                                 • fsiSsfs?^-? &9m I
                                   'as; "u. !'"»'
Lake Erie
          Status:
          Trend:
  Primary Factors
     Determining
 Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
 (Kannan et al 2006).

 Fair
 Undetermined
 In 2004, 16 of a possible 21 organochlorines were detected in Lake Huron,
 but only 11 were commonly found. Commonly found OCs included a-HCH,
 lindane, dieldrin, and g-chlordane. The concentrations were generally low,
 reflecting historical or diffuse sources.  For example, average
 concentrations of dieldrin in 2004 were 0.08 ng/L in Lake Huron and 0.07
 ng/L in Georgian Bay. These concentrations were lower than those found in
 the other Great Lakes and are well below the Ontario Water Quality
 Objective of 1.0 ng/L.

 Mercury concentrations in Lake Huron and Georgian Bay were low (2005
 open lake average: Lake Huron 0.58 ng/L, Georgian Bay 0.33 ng/L). The
 concentrations at all open lake stations were below the USEPA's Great
 Lakes Initiative (GLI) water quality criterion for protection of wildlife of
 1.3 ng/L (Figure 1), and only one nearshore station in Georgian Bay
 exceeded this level.

 PAH concentrations in Lake Huron and Georgian Bay are very low.  Of the
 20 and  19 PAH compounds found in Lake Huron and Georgian Bay,
 respectively, five were detected only within the North Channel
 (Dibenzo(a,h)antracene, Perylene, Benzo(a)pyrene, Anthracene, and 2-
 Chloronaphthalene). The open lake average concentration of Phenanthrene
 (Figure 4) was 0.08 ng/L in Lake Huron and 0.13 ng/L in Georgian Bay,
 well below the Ontario guideline of 30 ng/L.
Mixed
Undetermined
In 2004, Environment Canada's Great Lakes Surveillance Program detected
15 of a possible 21 organochlorine compunds in Lake Erie; 10 of these were
commonly found, including a-HCH, HCB, Lindane and Dieldrin.
Concentrations of most compounds were highest in the shallow western
basin and much lower in the central and eastern basins.  An exception is
Lindane, which showed similar concentrations in all three basins. Almost all
Canadian sources of Lindane to the Great Lakes are from the Canadian
prairies (Ma et al 2003).  Similar results were found in 1998 by Marvin et al.
(2004).  Between 1998 and 2004 average lakewide Lindane concentrations
fell (2004: 0.16 ng/1; 1998: 0.32 ng/1) indicating a possible downward trend.
Key contributors of hexachlorobenzene and octachlorostyrene were
identified in the St. Clair River (Marvin et al 2004).

The intensively-farmed agricultural and urban lands draining into Lake Erie
and Lake St. Clair are a major contributor of pesticides and other
contaminants to the Great Lakes. In these watersheds, approximately 75%
                         Draft for Discussion at SOLEC 2006

-------
                                                il^iflfetl1^!^!?!:;
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
of the land use is agriculture and about 40% of the Great Lakes population
resides here.   Pesticides were detected in every tributary monitored
between 1996 and 1998 (Kannan et al 2006). Some tributaries contained as
many as 18 different pesticides; among the highest counts for any watershed
monitored in North America.

Mercury concentrations in 2005 in Lake Erie were the highest of the Great
Lakes and reflected a decreasing concentration from west to east (average
concentrations 2.53 ng/L in the western basin,  0.52 ng/L in the central basin,
and 0.49 ng/L in the eastern basin). Higher concentrations (above 3.0 ng/L)
were found near the mouths of the Detroit and Maumee rivers.
Concentrations at all stations in the western basin, as well as some stations in
the central and eastern basins, exceeded the GLI mercury criterion of 1.3
ng/L.

PAH concentrations and distributions reflected urban source areas on the
Lake and upstream sources within the St. Clair River - Detroit River
corridor.  The highest concentrations of most PAHs were found in the
western basin,  and near the mouth of the Detroit River in particular. For
example the phenanthrene concentration (Figure 4) at the mouth of the
Detroit River was 2.5 ng/L, whereas the overall Lake average was 0.59 ng/L,
an almost 5-fold difference.

 Mixed
 Undetermined
 Seventeen  of a possible 21 OC pesticides were detected in Lake Ontario
 waters in 2005.  Dieldrin, lindane, and a-HCH were routinely found.
 Probable sources of these compounds include a combination of historical
 watershed uses, upstream loadings (e.g. the Niagara River) and atmospheric
 deposition.  Concentrations of many parameters were intermediate
 compared to the upper Great Lakes (which generally had lower
 concentrations) and Lake Erie (which generally had higher concentrations,
 especially in the western basin). Within Lake Ontario,  spatial trends were
 reflective of localized (predominantly urban) sources.

 Mercury concentrations in Lake Ontario were low in the offshore areas
 (average 0.48) and higher in the nearshore (average 0.80 ng/L). Spatial
 trends were reflective of localized sources (e.g. higher values in Toronto
 and Hamilton, Ontario and Rochester and Oswego, New York), but only
 samples taken from Hamilton Harbour exceeded the GLI objective of 1.3
 ng/L for mercury.

 PAH distribution and concentrations reflect urban source areas on the Lake
 (e.g., Rochester NY, Niagara River, and Hamilton, Ontario).  All offshore
 concentrations were below Ontario Water Quality Guidelines.
                         Draft for Discussion at SOLEC 2006

-------

Purpose
This indicator reports on concentrations of priority toxic chemicals in offshore waters, and, by
comparison to criteria for the protection for aquatic life and human health, infers the potential for
impacts to the health of the Great Lakes aquatic ecosystem. The indicator can be used to infer the
progress of virtual elimination programs as well.

Ecosystem Objective
The Great Lakes should be free from materials entering the water as a result of human activity
that will produce conditions that are toxic or harmful to human, animal, or aquatic life (GLWQA,
Article III(d)).

State of the Ecosystem
Many toxic chemicals are present in the Great Lakes and it is impractical to summarize the spatial
and temporal trends of them all within a few pages. For more information on spatial and
temporal trends in toxic contaminants in offshore waters, the reader is referred to Marvin et al.
(2004), Kannan et al. (2006), and Trends in Great Lakes Sediments and Surface Waters in
Chapter 8  of the Great Lakes Binational Toxics Strategy 2004 Progress report.

Surveys conducted between 1992 and 2000 (Marvin et al), and between 2004-2005  (Environment
Canada unpublished data) on Lakes Superior, Huron, Erie and Ontario showed that
concentrations of most organic compounds are low (i.e., below the most stringent water quality
guidelines) and declining in the open waters of the Great Lakes.  The decline in the  concentration
of banned organochlorine pesticides has leveled off since the mid-1980s and  current rates of
decline are slow.

Dieldrin, a-HCH, lindane (g-HCH), and heptachlor epoxide were the only organochlorine
pesticide compounds routinely detected in Lakes Superior, Erie and Ontario (Marvin et al. 2004).
The in-use herbicides atrazine and metolachlor were ubiquitous (Marvin et al. 2004). An
example of the spatial distribution of dieldrin using 2004/05 data is provided in Figure 2.

Many organic compounds (such as PCBs, hexachlorobenzene, octachlorostyrene, and DDT) show
a spatial pattern that indicates higher concentrations near historical, localized sources.  Currently
emitted compounds, such as PAHs and mercury, which are released during fossil fuel
combustion, also show spatial patterns that are indicative of sources. Concentrations of the
heavier PAHs, which are not as subject to atmospheric transport due to their partitioning to
particles, are highest in the lower Great Lakes, where human populations are greater.

Management Implications
Management efforts to control inputs of organochlorine pesticides have resulted in decreasing
concentrations in the Great Lakes; however, historical sources for some compounds still appear to
affect ambient concentrations in the environment. Further  reductions in the input of OC pesticides
are dependent, in part, on controlling indirect inputs such as atmospheric deposition and surface
runoff.  Monitoring programs should increase measurement of the major in-use pesticides, of
which currently only half are monitored. The additive  and synergetic effects of pesticide
mixtures should be examined more closely, since existing  water quality criteria have been
development for individual pesticides only (Kannan et al 2006).
                         Draft for Discussion at SOLEC 2006

-------
Beginning in 1986, Environment Canada has conducted toxic contaminant monitoring in the
shared waters of the Great Lakes.  Recently, Environment Canada has developed new
measurement techniques and has invested in an ultra-clean laboratory in order to more accurately
measure these trace concentrations of pollutants in the surface waters of the Great Lakes. The
data presented here represent the results of this new methodology.  Data is available for all of the
shared waters, although only partial coverage of Lake Ontario has been analyzed to date. The
analyte list includes PCBs (as congeners), organochlorines, polycyclic aromatic hydrocarbons
(PAHs), trace metals including mercury, as well as a limited number of in-use pesticides and
other compounds of emerging concern.

In 2003, USEPA initiated a monitoring program for toxics in offshore waters. EPA's spatial
coverage is more limited than the Canadian program, focusing mainly on Lake Michigan, but the
analyte list is more comprehensive and includes PCBs, organochlorine pesticides, toxaphene,
dioxins/furans, PBDEs, selected PAHs, mercury, and perfluorinated compounds. Information
from the USEPA is currently available for Lake Michigan for many organic compounds.
Different measurement and analytical techniques are used, but good agreement with Canadian
information is achieved for some parameters. Future efforts will need to focus on comparisons of
the  analytical methodologies used and  the results obtained. In 2006, some work to this end is
being initiated by the parties in Lake Michigan.

Efforts need to be maintained to identify and track the remaining sources and explore
opportunities to accelerate their elimination (e.g. The Great Lakes Binational Toxics  Strategy).
Targeted monitoring to identify and track  down local sources of LaMP critical pollutants is being
conducted in many Great Lakes tributaries. However, an expansion of the track down program
should be considered to include those chemicals whose distribution suggests localized influences.

Chemicals such as endocrine disrupting chemicals, in-use pesticides, and pharmaceuticals are
emerging issues. The agencies' environmental researchers are working with the monitoring
groups to include compounds of emerging concern in Great Lakes Surveillance cruises.  For
example, in-use pesticides and a suite of pharmaceuticals are being measured in each of the Great
Lakes between 2005 and 2007.

Comments from the author(s)
Data for Lakes Superior, Huron, Erie and  Ontario  are from Environment Canada's Great Lakes
Water Quality Monitoring and Surveillance Program. Data for Lake Michigan are from the US
EPA's Great Lakes Aquatic Contaminant  Surveillance (GLACS) program (Principal
Investigators: Dr. Matt Simcik of the University of Minnesota and Dr. Jeff Jeremiason of
Gustavus Adolphus College).

Lake Ontario 2005 data for PAHs and  OC pesticides reflects sampling conducted in the western
half of the lake only.

Acknowledgments
Authors:  Jennifer Vincent and Alice Dove, Environment Canada, Burlington, ON,  Melissa
Hulting, Great Lakes National Program Office, USEPA,  Chicago, IL.
                         Draft for Discussion at SOLEC 2006

-------
Data Sources
Great Lakes Binational Toxics Strategy. 2002 Progress Report. Environment Canada and US
Environmental Protection Agency.

Great Lakes Water Quality Agreement (GLWQA). 1978. Revised Great Lakes Water Quality
Agreement of 1978. As amended by Protocol November 18, 1987. International Joint
Commission, Windsor, Ontario.

Kannan, K, J. Ridal, J. Struger. 2006. Pesticides in the Great Lakes. In Persistent Organic
Pollutants in the Great Lakes ed. R. Kites, pp. 151-199. Germany: Springer.

Ma, J., S.M. Daggupaty, T. Harner, and Y.F. Li, 2003. Impacts of lindane usage in the Canadian
prairies to the Great Lakes ecosystem - Part 1: coupled atmospheric transport model and modeled
concentrations  in air and soil. Environmental Science and Technology 37:3774-3781.

Marvin, C., S. Painter, D. Williams, V. Richardson, R. Rossmann, P. Van Hoof. 2004. Spatial and
temporal trends in surface water and sediment contamination in the Laurentian Great Lakes.
Environmental Pollution. 129(2004): 131-144.

Rutherford, G., DJ. Spry, W. Scheider, and J. Ralston, 1999. Provincial Water Quality Standards.
Standards Development Branch and Program Development Branch, Ontario Ministry of
Environment and Energy. 31 pp.

Struger J., 1988. Organophosphorous insecticides and endosulfan in surface waters of the
Niagara fruit belt, Ontario, Canada.  Presented at the Society of Environmental Toxicology and
Chemistry meeting, Charlotte, North Carolina.

United States Environmental Protection Agency 2006. National Recommended Water Quality
Criteria for Priority Toxic Pollutants. Office of Water Science and Technology. 24pp.

Williams, DJ.  and M.L O'Shea. 2003. Niagara River Toxics Management Plan (NRTMP)
Progress Report and Work Plan. Prepared for the Niagara River Secretariat. Environment Canada,
US Environmental Protection Agency, Ontario Ministry of Environment and New York State
Department of Environmental Conservation.

List  of Figures
Figure  1. Great Lakes 2003-2005 Open Lake, Spring Cruise, Concentrations of Total Mercury
(ng/L).
Source: Environment Canada's Great Lakes Water Quality Surveillance Program, Burlington,
Ontario and U.S. Environmental Protection Agency's Great  Lakes National Program Office,
Chicago, Illinois.

Figure 2. Great Lakes 2004/05 Open Lake,  Spring Cruise, Concentrations of Dieldrin (ng/L).
Lake Ontario data for western half of the lake only.
Source: Environment Canada's Great Lakes Water Quality Surveillance Program, Burlington,
Ontario.
                         Draft for Discussion at SOLEC 2006

-------
Figure 3. Great Lakes 2004/05 Open Lake, Spring Cruise, Concentrations of Lindane (ng/L).
Lake Ontario data for western half of the lake only.
Source: Environment Canada's Great Lakes Water Quality Surveillance Program, Burlington,
Ontario.

Figure 4. Great Lakes 2004/05 Open Lake, Spring Cruise, Concentrations of Phenanthrene
(ng/L).
Source: Environment Canada's Great Lakes Water Quality Surveillance Program, Burlington,
Ontario.

Last updated
SOLEC 2006
Figure 1. Great Lakes 2003-2005 Open Lake, Spring Cruise, Concentrations of Total Mercury
(ng/L). Source: Environment Canada's Great Lakes Water Quality Surveillance Program,
Burlington, Ontario and U.S. Environmental Protection Agency's Great Lakes National Program
Office, Chicago, Illinois
                        Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
        Dieldrin (ng/L)

         o   I]-0.1 [I
         O  0.10-0.15
         O  0.15-0.20
         •  0.20+
Figure 2. Great Lakes 2004/05 Open Lake, Spring Cruise, Concentrations of Dieldrin (ng/L).
Lake Ontario data for western half of the lake only.
Source: Environment Canada's Great Lakes Water Quality Surveillance Program, Burlington,
Ontario and U.S. Environmental Protection Agency's Great Lakes National Program Office,
Chicago, Illinois
                        Draft for Discussion at SOLEC 2006

-------
         Lindane (ng/L)
           o    0-0.10
           O  0.10-0.15
           O  0.15-0.20
           •  0.20-0.30
           •  0.30 +
Figure 3. Great Lakes 2004/05 Open Lake, Spring Cruise, Concentrations of Lindane (ng/L).
Lake Ontario data for western half of the lake only.
Source: Environment Canada's Great Lakes Water Quality Surveillance Program, Burlington,
Ontario
10
Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
Figure 4. Great Lakes 2004/05 Open Lake, Spring Cruise, Concentrations of Phenanthrene
(ng/L).
Source: Environment Canada's Great Lakes Water Quality Surveillance Program, Burlington,
Ontario
                      Draft for Discussion at SOLEC 2006
11

-------

Concentrations of Contaminants in Sediment Cores
Indicator #119
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving/Undetermined
There have been significant declines over the past three decades in
concentrations of many contaminants including PCBs, DDT, lead, and
mercury. Knowledge is lacking regarding the occurrence of many new
contaminants including BFRs and fluorinated surfactants.
Lake-by-Lake Assessment
Lake Superior
           Status:   Mixed
           Trend:   Improving/Undetermined

Lake Michigan
           Status:   Mixed
           Trend:   Improving/Undetermined
Lake Huron
Lake Erie
           Status:   Mixed
           Trend:   Improving/Undetermined
           Status:   Mixed
           Trend:   Improving/Undetermined
Lake Ontario
           Status:   Mixed
           Trend:   Improving/Undetermined

Purpose
•To infer potential harm to aquatic ecosystems from contaminated sediments by comparing
contaminant concentrations to available sediment quality guidelines;
•To infer progress towards virtual elimination of toxic substances in the Great Lakes by assessing
surficial sediment contamination and contaminant concentration profiles in sediment cores from
open lake and, where appropriate, Areas of Concern index stations, and;
•To determine the occurrence, distribution, and fate of new chemicals in Great Lakes sediments.

Ecosystem Objective
The Great Lakes should be free from materials entering the water as a result of human activity
that will produce conditions that are toxic or harmful to human health, animal, or aquatic life
(Great Lakes Water Quality Agreement (GLWQA), Article III(d)). The GLWQA and the Great
Lakes Binational Toxics Strategy both state the virtual elimination of toxic substances to the
Great Lakes as an objective.
                        Draft for Discussion at SOLEC 2006

-------
State of the Ecosystem
Sediment Quality Index
A sediment quality index (SQI) has been developed that incorporates three elements: scope - the
percent of variables that did not meet guidelines; frequency - the percent of failed tests relative to
the total number of tests in a group of sites; and amplitude - the magnitude by which the failed
variables exceeded guidelines. A full explanation of the SQI derivation process and a possible
classification scheme based on the SQI score (0 - 100, poor to excellent) is provided in
Grapentine et al. (2002). Generally, the Canadian federal probable effect level (PEL) guideline
(CCME 2001) was used when available, otherwise the Ontario lowest effect level (LEL)
guideline (Persaud et al.  1992) was used. Application of the SQI to Lakes Erie and Ontario was
reported in Marvin et al. (2004). The SQI ranged from fair in Lake Ontario to excellent in eastern
Lake Erie. Spatial trends in sediment quality in Lakes Erie and Ontario reflected overall trends for
individual contaminant classes such as mercury and polychlorinated biphenyls (PCBs).

Environment Canada and USEPA integrated available data from the open waters of each of the
Great Lakes. To date, data on lead, zinc, copper, cadmium, and mercury have been integrated.
The site by site  SQIs for Great Lakes sediments based on these metals are illustrated in Figure 1.
The general trend in sediment quality across the Great lakes basin for the five metals is generally
indicative of trends for a wide range of persistent toxics. Areas of Lakes Erie, Ontario and
Michigan show the poorest sediment quality as a result of historical urban and industrial
activities.

Application of the SQI has been expanded to include contaminants in streambed and riverine
sediments for whole-watershed assessments. The SQI map for the Lake Erie - Lake St. Clair
drainages is shown in Figure 2. Poorest sediment quality is primarily associated with Areas of
Concern (AOC) where existing multi-stakeholder programs (e.g., Remedial Action Plans) are in
place to address environmental impairments related to toxic chemicals.

Pressures
Management efforts to control inputs of historical contaminants have resulted in decreasing
contaminant concentrations in the Great Lakes open-water sediments for the standard list of
chemicals. However,  additional chemicals such  as brominated flame retardants (BFRs) and
current-use pesticides (CUPs) may represent emerging issues and potential future stressors to the
ecosystem.

The distribution of hexabromocyclododecane  (HBCD) in Detroit River suspended sediments is
shown in Figure 3. This compound is the primary flame retardant used in polystyrene foams, and
is the third-most heavily produced BFR. Elevated levels of HBCD were  associated with heavily
urbanized/industrialized areas of the watershed.  The HBCD distribution differs from PCBs,
which are primarily associated with areas of contaminated sediment resulting from historical
industrial activities including steel manufacturing and chlor-alkali production. These results
corroborate observations made globally, which indicate that large urban  centers act as diffuse
sources of chemicals that are heavily used to support our modern societal lifestyle.

The temporal trend in the Niagara River of another class of BFRs, polybrominated diphenyl
ethers (PBDEs), is shown in Figure 4. Prior to 1988, PBDEs were generally detected at low
                         Draft for Discussion at SOLEC 2006

-------
(parts per billion, ppb) concentrations, but showed a trend toward increasing concentrations over
the period 1980 - 1988. After 1988, PBDE concentrations in the Niagara River showed a more
rapidly increasing trend. PBDE concentrations in suspended sediments of the Niagara River are
comparable to, or lower than, concentrations in sediments in other industrialized/urbanized areas
of the world. The Niagara River watershed does not appear to be a significant source of PBDEs to
Lake Ontario, and concentrations appear to be indicative of general contamination from a
combination of local, regional, and continental sources.

Management Implications
•The Great Lakes Binational Toxics Strategy needs to be maintained to identify and track the
remaining sources of contamination and to explore opportunities to accelerate their elimination.
•Targeted monitoring to identify and track down local sources of pollution should be considered
for those chemicals whose distribution in the ambient environment suggests local or sub-regional
sources.
•Ongoing monitoring programs in the Connecting Channels provide invaluable information on
the success of binational management actions to reduce/eliminate discharges of toxics to the
Great Lakes. These programs also provide important insights into pathways of new chemicals
entering the Great Lakes.

Acknowledgments
Authors: Scott Painter, Environment Canada, Burlington, ON; and
Chris Marvin, Environment Canada, Burlington, ON.

Data Sources
Canadian Council of Ministers of the Environment (CCME). 1999, updated 2001. Canadian
Environmental Quality Guidelines. Canadian Council of Ministers of the Environment.
Winnipeg, MB, Canada.

Grapentine, L., Marvin, C., and Painter, S. 2002. Development and evaluation of a sediment
quality index for the Great Lakes and associated Areas of Concern. Human and Ecological Risk
Assess. 8(7):1549-1567.

Marvin, C., Grapentine, L., and Painter, S. 2004. Application of a sediment quality index to the
lower Laurentian Great Lakes.  Environ. Monit. Assess. 91:1-16.

Marvin, C., Tomy, G.T., Alaee, M., and Maclnnis, G. 2006. Distribution of
hexabromocyclododecane in Detroit River suspended sediments. Chemosphere. 64:268-275.

Persaud, D., Jaagumagi, R., and Hayton, A. 1992. Guidelines for the protection and  management
of aquatic sediment quality in Ontario. Water Resources Branch, Ontario Ministry of the
Environment and Energy. June 1992.

U.S. Geological Survey (USGS). 2000. Areal distribution and concentrations of contaminants of
concern in surficial streambed and lakebed sediments, Lake Erie - Lake St. Clair drainages,
1990-97. Water Resources Investigations Report 00-4200.
                         Draft for Discussion at SOLEC 2006

-------
                              State of the Great Lakes 2007 - Draft
List of Figures
Figure 1. Site Sediment Quality Index (SQI) based on lead, zinc, copper, cadmium and mercury.
Source: Chris Marvin, Environment Canada (1997-2001 data for all lakes except Michigan); and
Ronald Rossmann, U.S. Environmental Protection Agency (1994-1996 data for Lake Michigan)

Figure 2. Sediment Quality Index (SQI) for the Lake Erie-Lake St. Clair drainages. More detailed
information on contaminants in sediments in the Lake Erie-Lake St. Clair drainages has been
reported by the USGS (2000).
Source: Dan Button, U.S. Geological Survey

Figure 3. Distribution of hexabromocyclododecane (HBCD) and PCBs in suspended sediments in
the Detroit River.
Source: Marvin et al. (2006).

Figure 4. Temporal trend in polybrominatd diphenyl ethers (PBDEs) in Niagara River suspended
sediments.
Source: Marvin et al. (2006).
Last updated
SOLEC 2006
                                   Great Lakes SQI PEL
             /»
             0-39 (Poor)
            40 - 59 (Marginal)
            60-79 (Fair)
            80-94 (Good)
            95-100 (Excellent)
Figure 1. Site Sediment Quality Index (SQI) based on lead, zinc, copper, cadmium and mercury.
Source: Chris Marvin, Environment Canada (1997-2001 data for all lakes except Michigan); and
Ronald Rossmann, U.S. Environmental Protection Agency (1994-1996 data for Lake Michigan)
                       Draft for Discussion at SOLEC 2006

-------
      State of the Great Lakes 2007 - Draft
    EXPLANATION
    StHJimwn Quality Index Basad on Probable Effect Levels (PEL)
     • 0 • 2S Poor QttaUf
       SO - W
     * 16 • 100 Good Quality
i
                          ».
                   '&£
                    &  *
                         •;"
                                 ^*v?
   ...--%-Vj    X-'-^.r-"
         gr*-J  c---  .   c^
         IT       .•.•••:"
.   w^=---;<^
             V  •
Figure 2. Sediment Quality Index (SQI) for the Lake Erie-Lake St. Clair drainages. More
detailed information on contaminants in sediments in the Lake Erie-Lake St. Clair drainages has
been reported by the USGS (2000).
Source: Dan Button, U.S. Geological Survey
                 Draft for Discussion at SOLEC 2006

-------
                             State of the Great Lakes  2007 - Draft
   Trenton Channel
      Elizabeth Park 1159
                                                             Lake
                                                             St. Clair
                                                           1160
                                                         1000
                                                          500
                                                               2500 n
                                                               2000-
                                                               1500-
                                                               1000-
                                                                500-
                                                                 o-l
                                                        PCBs(ng/g) HBCD(pg/g)
                  Lake Erie
Figure 3. Distribution of hexabromocyclododecane (HBCD) and PCBs in suspended sediments

in the Detroit River.

Source: Marvin et al. (2006).
                       Draft for Discussion at SOLEC 2006

-------

                                Total PBDEs
                                           Year
Figure 4. Temporal trend in polybrominatd diphenyl ethers (PBDEs) in Niagara River suspended
sediments.
Source: Marvin et al. (2006).
                     Draft for Discussion at SOLEC 2006

-------

Concentrations of Contaminants in Sediment Cores
Indicator #119
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving/Undetermined
There have been significant declines over the past three decades in
concentrations of many contaminants including PCBs, DDT, lead, and
mercury. Knowledge is lacking regarding the occurrence of many new
contaminants including BFRs and fluorinated surfactants.
Lake-by-Lake Assessment
Lake Superior
           Status:   Mixed
           Trend:   Improving/Undetermined

Lake Michigan
           Status:   Mixed
           Trend:   Improving/Undetermined
Lake Huron
Lake Erie
           Status:   Mixed
           Trend:   Improving/Undetermined
           Status:   Mixed
           Trend:   Improving/Undetermined
Lake Ontario
           Status:   Mixed
           Trend:   Improving/Undetermined

Purpose
•To infer potential harm to aquatic ecosystems from contaminated sediments by comparing
contaminant concentrations to available sediment quality guidelines;
•To infer progress towards virtual elimination of toxic substances in the Great Lakes by assessing
surficial sediment contamination and contaminant concentration profiles in sediment cores from
open lake and, where appropriate, Areas of Concern index stations, and;
•To determine the occurrence, distribution, and fate of new chemicals in Great Lakes sediments.

Ecosystem Objective
The Great Lakes should be free from materials entering the water as a result of human activity
that will produce conditions that are toxic or harmful to human health, animal, or aquatic life
(Great Lakes Water Quality Agreement (GLWQA), Article III(d)). The GLWQA and the Great
Lakes Binational Toxics Strategy both state the virtual elimination of toxic substances to the
Great Lakes as an objective.
                        Draft for Discussion at SOLEC 2006

-------
State of the Ecosystem
Sediment Quality Index
A sediment quality index (SQI) has been developed that incorporates three elements: scope - the
percent of variables that did not meet guidelines; frequency - the percent of failed tests relative to
the total number of tests in a group of sites; and amplitude - the magnitude by which the failed
variables exceeded guidelines. A full explanation of the SQI derivation process and a possible
classification scheme based on the SQI score (0 - 100, poor to excellent) is provided in
Grapentine et al. (2002). Generally, the Canadian federal probable effect level (PEL) guideline
(CCME 2001) was used when available, otherwise the Ontario lowest effect level (LEL)
guideline (Persaud et al.  1992) was used. Application of the SQI to Lakes Erie and Ontario was
reported in Marvin et al. (2004). The SQI ranged from fair in Lake Ontario to excellent in eastern
Lake Erie. Spatial trends in sediment quality in Lakes Erie and Ontario reflected overall trends for
individual contaminant classes such as mercury and polychlorinated biphenyls (PCBs).

Environment Canada and USEPA integrated available data from the open waters of each of the
Great Lakes. To date, data on lead, zinc, copper, cadmium, and mercury have been integrated.
The site by site  SQIs for Great Lakes sediments based on these metals are illustrated in Figure 1.
The general trend in sediment quality across the Great lakes basin for the five metals is generally
indicative of trends for a wide range of persistent toxics. Areas of Lakes Erie, Ontario and
Michigan show the poorest sediment quality as a result of historical urban and industrial
activities.

Application of the SQI has been expanded to include contaminants in streambed and riverine
sediments for whole-watershed assessments. The SQI map for the Lake Erie - Lake St. Clair
drainages is shown in Figure 2. Poorest sediment quality is primarily associated with Areas of
Concern (AOC) where existing multi-stakeholder programs (e.g., Remedial Action Plans) are in
place to address environmental impairments related to toxic chemicals.

Pressures
Management efforts to control inputs of historical contaminants have resulted in decreasing
contaminant concentrations in the Great Lakes open-water sediments for the standard list of
chemicals. However,  additional chemicals such  as brominated flame retardants (BFRs) and
current-use pesticides (CUPs) may represent emerging issues and potential future stressors to the
ecosystem.

The distribution of hexabromocyclododecane  (HBCD) in Detroit River suspended sediments is
shown in Figure 3. This compound is the primary flame retardant used in polystyrene foams, and
is the third-most heavily produced BFR. Elevated levels of HBCD were  associated with heavily
urbanized/industrialized areas of the watershed.  The HBCD distribution differs from PCBs,
which are primarily associated with areas of contaminated sediment resulting from historical
industrial activities including steel manufacturing and chlor-alkali production. These results
corroborate observations made globally, which indicate that large urban  centers act as diffuse
sources of chemicals that are heavily used to support our modern societal lifestyle.

The temporal trend in the Niagara River of another class of BFRs, polybrominated diphenyl
ethers (PBDEs), is shown in Figure 4. Prior to 1988, PBDEs were generally detected at low
                         Draft for Discussion at SOLEC 2006

-------
(parts per billion, ppb) concentrations, but showed a trend toward increasing concentrations over
the period 1980 - 1988. After 1988, PBDE concentrations in the Niagara River showed a more
rapidly increasing trend. PBDE concentrations in suspended sediments of the Niagara River are
comparable to, or lower than, concentrations in sediments in other industrialized/urbanized areas
of the world. The Niagara River watershed does not appear to be a significant source of PBDEs to
Lake Ontario, and concentrations appear to be indicative of general contamination from a
combination of local, regional, and continental sources.

Management Implications
•The Great Lakes Binational Toxics Strategy needs to be maintained to identify and track the
remaining sources of contamination and to explore opportunities to accelerate their elimination.
•Targeted monitoring to identify and track down local sources of pollution should be considered
for those chemicals whose distribution in the ambient environment suggests local or sub-regional
sources.
•Ongoing monitoring programs in the Connecting Channels provide invaluable information on
the success of binational management actions to reduce/eliminate discharges of toxics to the
Great Lakes. These programs also provide important insights into pathways of new chemicals
entering the Great Lakes.

Acknowledgments
Authors: Scott Painter, Environment Canada, Burlington, ON; and
Chris Marvin, Environment Canada, Burlington, ON.

Data Sources
Canadian Council of Ministers of the Environment (CCME). 1999, updated 2001. Canadian
Environmental Quality Guidelines. Canadian Council of Ministers of the Environment.
Winnipeg, MB, Canada.

Grapentine, L., Marvin, C., and Painter, S. 2002. Development and evaluation of a sediment
quality index for the Great Lakes and associated Areas of Concern. Human and Ecological Risk
Assess. 8(7):1549-1567.

Marvin, C., Grapentine, L., and Painter, S. 2004. Application of a sediment quality index to the
lower Laurentian Great Lakes.  Environ. Monit. Assess. 91:1-16.

Marvin, C., Tomy, G.T., Alaee, M., and Maclnnis, G. 2006. Distribution of
hexabromocyclododecane in Detroit River suspended sediments. Chemosphere. 64:268-275.

Persaud, D., Jaagumagi, R., and Hayton, A. 1992. Guidelines for the protection and  management
of aquatic sediment quality in Ontario. Water Resources Branch, Ontario Ministry of the
Environment and Energy. June 1992.

U.S. Geological Survey (USGS). 2000. Areal distribution and concentrations of contaminants of
concern in surficial streambed and lakebed sediments, Lake Erie - Lake St. Clair drainages,
1990-97. Water Resources Investigations Report 00-4200.
                         Draft for Discussion at SOLEC 2006

-------
                              State of the Great Lakes 2007 - Draft
List of Figures
Figure 1. Site Sediment Quality Index (SQI) based on lead, zinc, copper, cadmium and mercury.
Source: Chris Marvin, Environment Canada (1997-2001 data for all lakes except Michigan); and
Ronald Rossmann, U.S. Environmental Protection Agency (1994-1996 data for Lake Michigan)

Figure 2. Sediment Quality Index (SQI) for the Lake Erie-Lake St. Clair drainages. More detailed
information on contaminants in sediments in the Lake Erie-Lake St. Clair drainages has been
reported by the USGS (2000).
Source: Dan Button, U.S. Geological Survey

Figure 3. Distribution of hexabromocyclododecane (HBCD) and PCBs in suspended sediments in
the Detroit River.
Source: Marvin et al. (2006).

Figure 4. Temporal trend in polybrominatd diphenyl ethers (PBDEs) in Niagara River suspended
sediments.
Source: Marvin et al. (2006).
Last updated
SOLEC 2006
                                   Great Lakes SQI PEL
             /»
             0-39 (Poor)
            40 - 59 (Marginal)
            60-79 (Fair)
            80-94 (Good)
            95-100 (Excellent)
Figure 1. Site Sediment Quality Index (SQI) based on lead, zinc, copper, cadmium and mercury.
Source: Chris Marvin, Environment Canada (1997-2001 data for all lakes except Michigan); and
Ronald Rossmann, U.S. Environmental Protection Agency (1994-1996 data for Lake Michigan)
                       Draft for Discussion at SOLEC 2006

-------
      State of the Great Lakes 2007 - Draft
    EXPLANATION
    StHJimwn Quality Index Basad on Probable Effect Levels (PEL)
     • 0 • 2S Poor QttaUf
       SO - W
     * 16 • 100 Good Quality
i
                          ».
                   '&£
                    &  *
                         •;"
                                 ^*v?
   ...--%-Vj    X-'-^.r-"
         gr*-J  c---  .   c^
         IT       .•.•••:"
.   w^=---;<^
             V  •
Figure 2. Sediment Quality Index (SQI) for the Lake Erie-Lake St. Clair drainages. More
detailed information on contaminants in sediments in the Lake Erie-Lake St. Clair drainages has
been reported by the USGS (2000).
Source: Dan Button, U.S. Geological Survey
                 Draft for Discussion at SOLEC 2006

-------
                             State of the Great Lakes  2007 - Draft
   Trenton Channel
      Elizabeth Park 1159
                                                             Lake
                                                             St. Clair
                                                           1160
                                                         1000
                                                          500
                                                               2500 n
                                                               2000-
                                                               1500-
                                                               1000-
                                                                500-
                                                                 o-l
                                                        PCBs(ng/g) HBCD(pg/g)
                  Lake Erie
Figure 3. Distribution of hexabromocyclododecane (HBCD) and PCBs in suspended sediments

in the Detroit River.

Source: Marvin et al. (2006).
                       Draft for Discussion at SOLEC 2006

-------

                                Total PBDEs
                                           Year
Figure 4. Temporal trend in polybrominatd diphenyl ethers (PBDEs) in Niagara River suspended
sediments.
Source: Marvin et al. (2006).
                     Draft for Discussion at SOLEC 2006

-------
Contaminants in Whole Fish
Indicator #121
Overall Assessment
           Status:   Mixed
           Trend:   Improving
   Primary Factors
      Determining
  Status and Trend
                   Whole fish are monitored by both EPA GLNPO and Environment
                   Canada** to determine the effects of contaminant concentrations on
                   wildlife and monitor trends. Both governments collect and analyze
                   whole fish independently from a variety of locations within each Great
                   Lake using different methods.  The differences between the two
                   programs, collection sites in all 5 Great Lakes, and differences in
                   species yield a mixed status for the basin as a whole.
** In the spring of 2006, Environment Canada assumed the responsibilities of the
Department of Fisheries and Ocean (DFO) Fish Contaminant Surveillance Program. All
data included in this indicator report were produced by DFO.

Lake-by-Lake Assessment    PCB and DDT levels are measured in lake trout and walleye
                            while only smelt samples have recent Hg trend data available.

                  Fair
                  Improving
                   Concentrations of Total PCBs show little change and Total DDT show
                   fluctuating concentrations while mercury concentrations continue to
                   decline. Total PCB concentrations remain above GLWQA criteria while
                   Total DDT and mercury remain below. Contaminants in Lake Superior are
                   typically atmospherically derived.  The dynamics of Lake Superior allow
                   for the retention of contaminants much longer than any other lake.
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Fair
                   Improving
                   Concentrations of Total PCBs and Total DDT are both declining. Total
                   PCBs remain above GLWQA criteria and Total DDT remains below.  Food
                   web changes are critical to Lake Michigan contaminant concentrations, as
                   indicated by the failure of the alewife population in the 1980's and the
                   presence of the round goby. Aquatic invasive species, such as asian carp,
                   are also of major concern to the lake due to the connection of Chicago
                   Sanitary and Ship canal and the danger they pose to the food web.
                   Fair
                   Improving
                   Both Total PCBs and DDT show general declines in concentrations while
                   mercury displays flux in concentration. Total PCB concentrations remain
                   above GLWQA criteria while Total DDT and mercury remain below.
                        Draft for Discussion at SOLEC 2006

-------

Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Contaminant loading to Saginaw Bay continues to be reflected in fish
                   tissue.
Fair
Improving
Total PCBs and DDT show a pattern of annual concentration increases
linked to changes in invasive species populations, such as zebra and guagga
mussels. Aquatic invasive species are of major concern to Lake Erie
because the pathways and fate of persistent toxic substances will be altered
resulting in differing accumulation patterns, particularly near the top of the
food chain. Mercury concentrations are the highest ever recorded in Lake
Erie. Total PCB concentrations remain above GLWQA criteria while Total
DDT and mercury remain below.
Fair
Improving
 Both Total PCBs and DDT show a pattern of decline while mercury
 concentrations show little change.  Total PCB concentrations remain above
 GLWQA criteria while Total DDT and mercury remain below. Historic
 point sources of mirex and OCS in Lake Ontario have resulted in the
 highest concentration of these contaminants in any of the Great Lakes. The
 presence of contaminants of emerging concern, such as PBDEs and PFOS,
 continue to raise alarm in Lake Ontario, due to their continuing increases in
 concentration over time.
Purpose
•To describe temporal and spatial trends of bioavailable contaminants in representative open
water fish species from throughout the Great Lakes;
•To infer the effectiveness of remedial actions related to the management of critical pollutants;
and "To identify the nature and severity of emerging problems".

Ecosystem Objective
Great Lakes waters should be free of toxic substances that are harmful to fish and wildlife
populations and the consumers of this biota. Data on status and trends of contaminant conditions,
using fish as biological indicators, support the requirements of the Great Lakes Water Quality
Agreement (GLWQA, United States and Canada. 1987) Annexes 1 (Specific Objectives), 2
(Remedial Action Plans and Lakewide Management Plans), 11 (Surveillance and Monitoring),
and Annex  12 (Persistent Toxic Substances).

State of the Ecosystem
Background
Long-term (>25 yrs), basin-wide monitoring programs that measure whole body concentrations
of contaminants in top predator fish (lake trout and/or walleye) and in forage fish (smelt) are
conducted by the U.S. Environmental Protection Agency (USEPA) Great Lakes National
                         Draft for Discussion at SOLEC 2006

-------



Program Office (GLNPO) through the Great Lakes Fish Monitoring Program and Environment
Canada (EC), formerly DFO, through the Fish Contaminants Surveillance Program. Canada
reports annually on contaminant burdens in similarly aged lake trout (4+ - 6+ year range),
walleye (Lake Erie), and in smelt. GLNPO annually monitors contaminant burdens in similarly
sized lake trout (600-700 mm total length) and walleye (Lake Erie, 400-500 mm total length)
from alternating locations by year in each lake.

Chemical Concentrations in Whole Fish Great Lakes Fish:
Since the late 1970s, concentrations of historically regulated contaminants such as
polychlorinated biphenyls (PCBs), dichlorodiphenyl-trichloroethane (DDT) and mercury have
generally declined in most monitored fish species. The concentrations of other contaminants, both
currently regulated and unregulated, have demonstrated either slowing declines or, in some cases,
increases in selected fish communities. The changes are often lake-specific and relate both to the
characteristics of the substances involved and the biological composition of the fish community.

The GLWQA, first signed in 1972 and renewed in 1978, expresses the commitment of Canada
and the United States to restore and maintain the chemical, physical and biological integrity of the
Great Lakes basin ecosystem. When applicable, contaminant concentrations are compared to
GLWQA criteria.

E PCBs -Total PCB concentrations in Great Lakes top predator fish have continuously declined
since their phase out in the 1970s.  However, rapid declines are no longer observed and
concentrations in fish remain above the EPA wildlife protection value of 0.16 ppm and the
GLWQA criteria of 0.1 ppm. Concentrations remain high in top predator fish due to the
continued release of uncontrolled sources and their persistent and bioaccumulative nature.

E DDT - Total DDT concentrations in Great Lakes top predator fish have continuously declined
since the chemical was banned in 1972. However, large declines are no longer observed.  But
rather,  very small annual percent declines indicating near steady state conditions.  It is important
to note that the concentrations of this contaminant remain below the GLWQA criteria of 1.0
ppm. There is no EPA wildlife protection value for total DDT because the PCB value is more
protective.

Mercury - Concentrations of mercury are similar across all fish in all lakes.  It is assumed that
concentrations of mercury in top predator fish are atmospherically driven. It is important to note
that current concentrations in GLNPO top predator fish in all lakes remain above the GLWQA
criteria of .5 ppm and that Canadian smelt have never been observed to be above the GLWQA
criteria.

E Chlordane - Concentrations of total chlordane have consistently declined in whole top predator
fish since the  EPA banned it in 1988.  Total Chlordane is composed of cis and trans-chlordane,
cis and trans-nonachlor, and oxy chlordane, with trans-nonachlor being the most prevalent of the
compounds.  While trans-nonachlor was the minor component of the total chlordane mixture, it is
the least metabolized and predominates within the food web (Swackhamer, 2006).

Mirex - Concentrations of mirex are highest in Lake Ontario top predator fish due to its
continued release from uncontrolled historic sources near the Niagara River.
                         Draft for Discussion at SOLEC 2006

-------
                                                                                   	
Dieldrin - Concentrations of dieldrin in lake trout appear to be declining in all Lakes and are
lowest in Lake Superior and highest in Lake Michigan. Concentrations in Lake Erie walleye
were the lowest of all lakes. Aldrin is readily converted to dieldrin in the environment. For this
reason, these two closely related compounds (aldrin and dieldrin) are considered together by
regulatory bodies.

Toxaphene - Decreases in toxaphene concentrations have been observed throughout the Great
Lakes in all media following its ban in the mid- 1980's. However, concentrations have remained
the highest in Lake Superior due to its longer retention time, cold temperatures, and slow
sedimentation rate. It is assumed that concentrations of toxaphene in top predator fish are
atmospherically driven (Kites, 2006).

PBDEs - Both the US and Canada analyze for PBDEs in whole top predator fish.  Retrospective
analyses of archived samples have demonstrated the continuing increase in concentrations of
polybrominated diphenyl ethers (PBDE) and are confirmed by present day concentrations in top
predator fish.   It is important to note that the concentration of most other persistent organic
pollutants in top predator fish have declined, while PBDEs continue to increase.

Other Contaminants of Emerging Interest:
One of the most widely used BFRs is hexabromocyclododecane (HBCD). Based on its use
pattern as an additive BFR, it has the potential to migrate into the environment from its
application site. Recent studies have confirmed that HBCD isomers do bioaccumulate  in aquatic
ecosystem and do biomagnify as they move up the food chain. Recent studies by Tomy et al.
(2004) confirmed the food web biomagnification of HBCD isomers in Lake Ontario (Table 4).

Perfluoroctanesulfonate (PFOS) has also been detected in fish throughout the Great Lakes and
has also demonstrated the capacity for biomagnification in food webs. PFOS is used in
surfactants such as water repellent coatings  (i.e. Scotchguard ™) and fire suppressing foams. It
has been identified in whole lake trout samples from all the Great Lakes at concentrations from 3
to 139 ng/g wet weight (Stock et al. 2003). In addition, retrospective analyses of archived lake
trout samples from Lake Ontario have identified a 4.25-fold increase (43-180 ng/g wet weight,
whole fish) from 1980 to 2001 (Martin et al, 2004).

Pressures
Current - The impact of invasive nuisance species on toxic chemical cycling in the Great Lakes is
still being investigated. The number of non-native invertebrates and fish species proliferating in
the Great Lakes basin continues to increase, and they continue to spread more widely.  Changes
imposed on the native fish communities by non-native species will subsequently alter ecosystem
energy flows. As a consequence, the pathways and fate of persistent toxic  substances will be
altered, resulting in different accumulation patterns, particularly at the top of the food web. Each
of the Great Lakes is currently experiencing changes in the structure of the aquatic community,
and hence there may be periods of increases in contaminant burdens of some fish species.

A recently published, 15 year retrospective Great Lakes study showed that lake trout embryos and
sac fry are very sensitive to toxicity associated with maternal exposures to 2,3,7,8-
                         Draft for Discussion at SOLEC 2006

-------


tetrachlorodibenzo-p-dioxin (TCDD) and structurally related chemicals (Cook et al. 2003). The
increase in contaminant load of TCDD may be responsible for declining lake trout populations in
Lake Ontario. The models used in this study can be used in the other Great Lakes.

Future - Additional stressors in the future will include climate change, with the potential for
regional warming to change the availability of Great Lakes critical habitats, change the
productivity of some biological communities, accelerate the movement of contaminants from
abiotic sources into the biological communities, and effect the composition of biological
communities. Associated changes in the concentration of contaminants in the water, critical
habitat availability and reproductive success of native and non-native species are also factors that
will influence trends in the quantity of toxic contaminants in the Great Lakes basin ecosystem.

Management Implications
Much of the current, basin-wide, persistent toxic substance data that is reported focuses on legacy
chemicals whose use has been previously restricted through various forms of legislation. There
are also a variety of other potentially harmful contaminants at various locations throughout the
Great Lakes that are reported in literature. A comprehensive, basin-wide assessment program is
needed to monitor the presence and concentrations of these recently identified compounds in the
Great Lakes basin. The existence of long-term specimen archives (>25 yrs) in both Canada and
the United States could allow retrospective analyses of the samples to determine if concentrations
of recently detected contaminants are changing. Further control legislation might be needed for
the management of specific chemicals.

Acknowledgments
Authors: Elizabeth Murphy, U.S. Environmental Protection Agency, Great Lakes National
Program Office;
Cameron MacEachen, Environment Canada; D. Michael Whittle, Emeritus, Great Lakes
Laboratory for Fisheries  and Aquatic Sciences, Michael J. Keir, Environment Canada, and J.
Fraser Gorrie, Bio-Software Environmental Data.

Data Sources
Carlson, D.L., and Swackhamer D.L, Results from the U.S. Great Lakes Fish Monitoring
Program and Effects of Lake Processes on Contaminant Concentrations. Journal of Great Lakes
Research.  32 (2): 370 - 385.

Cook, P.M., Robbins, J.A., Endicott, D.D., Lodge, K.B., Guiney, P.O., Walker, M.K, Zabel,
E.W., and Peterson, R.E. 2003. Effects of Aryl Hydrocarbon Receptor-Mediated Early Life Stage
Toxicity on Lake Trout Populations in Lake Ontario during the 20th Century. Environ. Sci.
Technol.37(17):3878-3884.

Kites R.A, editor. 2006. Persistent Organic Pollutants in the Great Lakes. Heidelberg, Germany:
Springer.

Martin, J.W., Whittle, D.M., Muir, D.C.G., and Mabury, S.A. 2004. Perfluoroalkyl Contaminants
in the Lake Ontario Food Web. Environ. Sci. Technol. 38(20):5379-5385.
                         Draft for Discussion at SOLEC 2006

-------

Stock, N.L., Benin J., Whittle, D.M., Muir, D.C.G., and Mabury, S.A. 2003. Perfluoronated
Acids in the Great Lakes. SET AC Europe 13th Annual Meeting, Hamburg, Germany.

Tomy, G.T., Budakowski, W., Halldorson T., Whittle, D.M., Keir, M., Marvin, C., Maclnnis, G.,
and Alaee, M. 2004. Biomagnification of a and y-Hexabomocyclododecane in a Lake Ontario
Food Web. Environ. Sci. Technol 38:2298-2303.

United States and Canada. 1987. Great Lakes Water Quality Agreement of 1978, as amended by
Protocol signed November 18, 1987. Ottawa and Washington.
http://www.ijc.org/rel/agree/quality.html.

Whittle, D.M.,  MacEachen, D.C., Sergeant, D.B., Keir, M.J., and Moles, M.D. Food Web
Biomagnification of Brominated Diphenyl Ethers in Two Great Lakes Fish Communities, pp 83-
86. 3rd International Workshop on Brominated Flame Retardants, Toronto, Ontario.

List of Figures -

Figure 1. Total PCBs levels in Even Year whole Lake Trout (Walleye in Lake Erie), 1972 - 2002
jag/g wet weight +/- 95% C.I., composite samples. Lake Trout = 600 - 700 mm size range.  *Fish
collected between 1972 and 1982 were collected at even year sites only. Walleye = 450 - 550
mm size range. Source: U.S. Environmental Protection Agency

Figure 2. Total PCBs levels in Odd Year whole Lake Trout (Walleye in Lake Erie), 1991 - 2003
jag/g wet weight +/- 95% C.I., composite samples. Lake Trout = 600 - 700 mm size range.
Walleye = 450  - 550 mm size range. Source:  U.S. Environmental Protection Agency

Figure 3. Total PCBs in 4 to 6 year old individual whole Lake Trout collected 1977 through
2005, jag/g wet weight. Source: Fisheries and Oceans Canada.

Figure 4. Total PCBs in composite rainbow smelt collected 1977 through 2005, jag/g wet weight.
Source: Fisheries and Oceans  Canada.

Figure 5. DDT  levels in Even Year whole Lake Trout (Walleye in Lake Erie), 1972 - 2000. jag/g
wet weight +/-  95% C.I., composite samples. Lake Trout = 600 - 700 mm size range. *Fish
collected between 1972 and 1982 were collected at even year sites only. Walleye = 450 - 550
mm size range. Source: U.S. Environmental Protection Agency

Figure 6. DDT  levels in Odd Year whole Lake  Trout (Walleye in Lake Erie), 1991 - 2001. jag/g
wet weight +/-  95% C.I., composite samples. Lake Trout = 600 - 700 mm size range. Walleye =
450 - 550 mm size range. Source: U.S. Environmental Protection Agency

Figure 7. Total DDT in 4 to 6 year old individual whole Lake Trout collected 1977 through 2005,
jag/g wet weight. Source: Fisheries and Oceans Canada.

Figure 8. Total DDT in composite rainbow smelt collected 1977 through 2005, jag/g wet weight.
Source: Fisheries and Oceans  Canada.
                        Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
Figure 9. Interactive GIS map of basin and web link

Last updated
SOLEC 2006
              Total PCBs in Great Lakes Top Predator Fish, Even
                                     Year Sites
                             Lake Trout (Walleye in Lake Erie)
                                                              Superior
                                                              Michigan
                                                              Huron
                                                              B-ie
                                                              Ontario
                                        Year
Figure 1. Total PCBs levels in Even Year whole Lake Trout (Walleye in Lake Erie), 1972 - 2002
jag/g wet weight +/- 95% C.I., composite samples. Lake Trout = 600 - 700 mm size range. *Fish
collected between 1972 and 1982 were collected at even year sites only.  Walleye = 450 - 550
mm size range.
Source: U.S. Environmental Protection Agency
                      Draft for Discussion at SOLEC 2006

-------
                           State of the Great Lakes 2007 - Draft
   E
   a.
   m
   o
              Total PCBs in Great Lakes Top Predator Fish, Odd
                                    Year Sites
                             Lake Trout (Walleye in Lake Erie)
                                                            Superior
                                                            Michigan
                                                            Huron
                                                            Erie
                                                            Ontario
                                       Year
Figure 2. Total PCBs levels in Odd Year whole Lake Trout (Walleye in Lake Erie), 1991 - 2003
jag/g wet weight +/- 95% C.I., composite samples. Lake Trout = 600 - 700 mm size range.
Walleye = 450 - 550 mm size range.
Source: U.S. Environmental Protection Agency
                     Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
  V)
  m
  o
  o.
10
 9 -
 8 -
 7
 6 -
 5 -
 4 -
 3 -
 2
 1 -
 0
                    Total PCB in Great Lakes Lake Trout
                                        Year
                                                                 •Ontario
                                                                 -Erie
                                                                  Huron
                                                                  Superior
Figure 3.  Total PCBs in 4 to 6 year old individual whole Lake Trout collected 1977 through
2005, jj,g/g wet weight.
Source: Fisheries and Oceans Canada
                      Draft for Discussion at SOLEC 2006

-------
                           State of the Great Lakes 2007 - Draft
                     Total PCB in Great Lakes Smelt
                                      Year
Figure 4. Total PCBs in composite rainbow smelt collected 1977 through 2005, jag/g wet
weight.
Source: Fisheries and Oceans Canada
10
Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
              Total DDT in Great Lakes Top Predator Fish, Even
                                     Year Sites
                             Lake Trout (Walleye in Lake Erie)
                                                             Superior
                                                             Michigan
                                                             Huron
                                                             &ie
                                                             Ontario
                                        Year

Figure 5. DDT levels in Even Year whole Lake Trout (Walleye in Lake Erie), 1972 - 2000.
wet weight +/- 95% C.I., composite samples. Lake Trout = 600 - 700 mm size range. *Fish
collected between 1972 and 1982 were collected at even year sites only. Walleye = 450 - 550
mm size range.
Source: U.S. Environmental Protection Agency
                      Draft for Discussion at SOLEC 2006
11

-------
                           State of the Great Lakes 2007 - Draft
   E
   Q.
   O.
   VI
   m
   o
   a.
               Total DDT in Great Lakes Top Predator Fish, Odd
                                    Year Sites
                            Lake Trout (Walleye in Lake Erie)
                                                           Superior
                                                           Michigan
                                                           Huron
                                                           Erie
                                                           Ontario
                                       Year
Figure 6. DDT levels in Odd Year whole Lake Trout (Walleye in Lake Erie), 1991 - 2001. jag/g
wet weight +/- 95% C.I., composite samples. Lake Trout = 600 - 700 mm size range. Walleye =
450 - 550 mm size range.
Source: U.S. Environmental Protection Agency
12
Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
   O)
   0>
   D

  Q
  Q
4.5
  4
3.5
  3
2.5
  2
1.5
  1
0.5
  0
                   Total DDT in Great Lakes Lake Trout
                                 Year
Figure 7. Total DDT in 4 to 6 year old individual whole Lake Trout collected 1977 through
2005, jj,g/g wet weight.
Source: Fisheries and Oceans Canada
                     Draft for Discussion at SOLEC 2006
                                                                    13

-------
                          State of the Great Lakes 2007 - Draft
                     Total DDT in Great Lakes Smelt
                                Year
Figure 8. Total DDT in composite rainbow smelt collected 1977 through 2005, |ag/g wet weight.
Source: Fisheries and Oceans Canada.
14
Draft for Discussion at SOLEC 2006

-------


Hexagenia
Indicator # 122
Overall Assessment
           Status:   Mixed
           Trend:   Improving
   Primary Factors
      Determining
  Status and Trend
Lack of time-series and historical information.
To date, only one area (western Lake Erie) has exhibited any
substantial recovery of Hexagenia despite anecdotal reports of recovery
for many areas in the Great Lakes in the mid to early 1990s. After an
absence of 50 years, emerging Hexagenia were observed in open water
of western Lake Erie in 1992 (Figure 1). Studies confirmed the return
of nymphs to sediments between 1995 and 2005 (Figure 2). Between
1995 and 2005, the annual average density of nymphs was
approximately 300 nymphs/m2, a density similar to known historical
abundances of nymphs in the basin. The return of this taxon may be
entering the final stage of its recovery (i.e., stable annual abundances).
However, large decreases in density (1997 to 1998 and 2001 to 2002,
Figure 2) and poor young-of-year recruitment into the population (3 of
6 years, Figure 3) indicate that 'restoration' of nymphs has not been
totally successful. The cause(s) for population decreases and failed
recruitment is not known but it is suspected that it is related to residual
pollution. Effects of residual pollution will likely  decrease as pollution-
abatement programs continue. Continued work in western Lake Erie
will allow us to define a quantitative goal for successful 'restoration'  of
Hexagenia in mesotrophic waters in western Lake Erie and throughout
the Great Lakes (Figure 4).
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Poor
Undetermined
Lack of time-series and historical information.
Baseline (2001) information on the abundance of Hexagenia has been
obtained for Duluth Harbor, Minnesota and Wisconsin (Edsall et al. 2004).
Poor
Undetermined
Lack of time-series and historical studies.
There have been no scientific conformations of anecdotal reports of
Hexagenia except for sporadic accounts of adults near the Fox River, Green
Bay, Wisconsin.

The absence of Hexagenia was confirmed in Green Bay, Wisconsin in 2001
(Edsall et al. 2005).
                        Draft for Discussion at SOLEC 2006

-------

                                                                         _'l
                                                       l-^^-*J - *- '^\r  * -'te *•*•*»
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
Poor
Undetermined
Lack of time-series and historical information.
There have been no scientific conformations of anecdotal reports of
  Status and Trend  Hexagenia adults.
Lake Erie
           Status:

           Trend:

   Primary Factors
      Determining
  Status and Trend
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   The absence of Hexagenia was confirmed in Saginaw Bay in 2001 (Edsall
                   etal. 2005).
Good for western Lake Erie; Mixed for the southwest shore of central Lake
Erie
Improving for western Lake Erie; Mixed for southwest shore of central
Lake Erie
To date, western Lake Erie is the only place where Hexagenia has been
documented to be recovering in the Great Lakes (Krieger et al. 1996;
Madenjian et al. 1998, Schloesser et al. 2000).
Initial signs of recovery of Hexagenia (i.e., evidence of adults) along the
south shore of central Lake Erie (i.e., appearance and increasing
distribution) occurred 1997-2000. However, since that time reports have
decreased and intensive lake sampling (2001-2003) have not been able to
confirm Hexagenia recovery.
Not Assessed
Undetermined
Lack of baseline studies and historical information.
There have been no scientific conformations of anecdotal reports of
mayflies near Presque Isle, Pennsylvania and Bay of Quinte, Ontario.
Purpose
To assess the distribution and abundance of burrowing mayflies (Hexagenia) in the Great Lakes.
To establish a quantitative goal for the restoration of Hexagenia nymphs in mesotrophic waters of
the Great Lakes.

Ecosystem Objective
Historical mesotrophic habitats should be restored and maintained as balanced, stable, and
productive elements of the Great Lakes ecosystem with Hexagenia as the key benthic invertebrate
organism in the food chain. (Paraphrased from Final Report of the Ecosystem Objectives
Subcommittee, 1990, to the IJC Great Lakes Science Advisory Board). In addition, this indicator
supports Annex 2 of the GLWQA.
                         Draft for Discussion at SOLEC 2006

-------

State of the Ecosystem
In the early 20th century, mesotrophic ecosystems in the Great Lakes had unique faunal
communities that included commercially valuable fishes and associated benthic invertebrates.
The primary invertebrate taxon associated with mesotrophic habitats was Hexagenia. Hexagenia
was chosen by the scientific community to be a mesotrophic indicator because it is important to
fishes, is relatively long lived, lives in sediments where pollution often accumulates, and is
relatively sensitive to habitat changes brought on by urban and industrial pollution associated
with changes as mesotrophic systems deteriorate to eutrophic systems (Schloesser and Hiltunen
1984; Schloesser 1988; Reynoldson et al. 1989). For example, Hexagenia was very abundant and
important to yellow perch and walleye in the 1930s and 1940s. Then in the mid-1950s,
Hexagenia was eliminated by low oxygen and resulting anoxic conditions created by urban and
industrial pollution and growth of yellow perch declined (Beeton 1969;  Burns 1985).

Initiation of pollution-abatement programs in the 1970s improved water and sediment quality in
Hexagenia habitat throughout the Great Lakes, but the recovery of Hexagenia populations has
been elusive (Krieger et al. 1996; Schloesser et al. 2000). Then in the early 1990s, soon after the
invasion of exotic dreissenid mussels, anecdotal reports of adult Hexagenia (winged dun ans
spinner) occurred in many bays and interconnecting rivers of the Great Lakes after absences of
30-60 years (Figure 1).

The first sign of the potential recovery of Hexagenia in western Lake Erie began with  an
anecdotal report of adult mayflies in open waters of the basin by scientists on the research vessel
Limnos  (Kreiger et al. 1996; Madenjian et al. 1998; Schloesser et al. 2000). Nymphs were
confirmed in sediments at very low densities (ca. 9 nymphs/m2) in 1993 and intensive  studies
began in 1995 (Figure 2) (Kreiger et al. 1996; Schloesser, unpublished data). Densities of nymphs
increased between 1995 and 1997 and then decreased between 1997 and 1998. This pattern of
increasing densities followed by a large decrease occurred again between 2001 and 2002. A
population study of Hexagenia revealed that sharp declines in densities were partly attributable to
failed young-of-year (YOY) recruitment (Figure 3) (Bridgeman et al. 2002). No YOY nymphs
were found in 1997, which corresponded to the largest observed decline in Hexagenia density
during the last decade. A similar decline occurred between 2001 and 2002 when few YOY
nymphs were produced. However, a slight increase occurred between 2002 and 2003 even though
relatively few YOY nymphs were recruited into the population indicating that some other
factor(s) contributes to density fluctuations observed in western Lake Erie in the 1990s and
2000s.

Anecdotal reports of winged Hexagenia mayflies in the 1990s also included the south  shore of
Lake Michigan, Chicago, Illinois, the Fox River near Green Bay, Lake Michigan,  Saginaw Bay
near Standish, Michigan, the south shore of central Lake Erie near Sandusky, Ohio, Presque Isle
of eastern Lake Erie, Pennsylvania, and the northern shore in the Bay of Quinte, eastern Lake
Ontario, Picton, Ontario. To date, only the possible recovery of Hexagenia along the south shore
of central Lake Erie has been investigated (K. Kreiger, personal communication).  An initial
recovery of nymphs occurred along the south shore between 1997 and 2000. However, intensive
scientific surveys between 2001 and 2003 indicate that a sustained recovery of Hexagenia along
the shore of south central Lake Erie has not occurred.
                         Draft for Discussion at SOLEC 2006

-------
Pressures
Hexagenia are extirpated at moderate levels of pollution and may even show a graded response to
the degree of pollution (Edsall et al. 1991; Schloesser et al. 1991). High Hexagenia abundance is
strongly indicative of adequate levels of dissolved oxygen in overlying waters and
uncontaminated surficial sediments. Probable causative agents of impaired Hexagenia
populations include excess nutrients, oil, heavy metals, and various other pollutants in surficial
sediments.

A portion of the general public has developed a negative perception of en masse swarms of adult
Hexagenia because they can disrupt recreational use of shorelines and this perception has been
incorporated into management goals for the recovery of Hexagenia in western Lake Erie (see
Management Implications below). Such perceptions may create pressures for management to
implement actions that manage lake systems below the natural carrying capacity of Hexagenia in
mesotrophic waters of the Great Lakes.

Management Implications
Management entities in both Europe and North America desire some level of abundance of
burrowing mayflies, such as Hexagenia, in mesotrophic habitats (Fremling and Johnson 1990; Bij
de Vaate et al. 1992; Ohio Lake  Erie Commission 1998). Recoveries of burrowing mayflies, such
as Hexagenia spp., in rivers in Europe and North America and now in western Lake Erie clearly
show how properly implemented pollution controls can bring about the recovery of large
mesotrophic ecosystems. With recovery, Hexagenia in the Great Lakes will probably reclaim its
functional status as a major trophic link between detrital energy pools and economically valuable
fishes such as yellow perch and walleye.

The recovery of Hexagenia in western Lake Erie reminds us of an outstanding feature associated
with using Hexagenia as an indicator of ecosystem health — the massive swarms of winged
adults that are typical of healthy, productive Hexagenia populations. These swarms are highly
visible to the public who use them to judge success of pollution-abatement programs by seeing a
'real' species that signifies the return of a 'real' habitat to a desirable condition in the Great Lakes.
This public perception has influenced target values set by management for the recovery of
Hexagenia in western Lake Erie (i.e., imperiled and good above excellent, Figure 4). However,
values above excellent are based on societies' perception of excessive en masse emergences of
winged Hexagenia which affect  electrical power generation, vehicle traffic, and outdoor
activities. These values may not represent the best scientific information for the historic/natural
carrying capacity of Hexagenia in mesotrophic waters. For example, the target value of excellent
is based on historical densities, a desire to return the system to an earlier more 'pristine' condition,
and provide prey for valuable fishes. Yet, there is no scientific information that indicates densities
of nymphs above 'excellent' would be in conflict with historical data, previous system conditions,
and prey availability to fishes.

Comments from the author(s)
In the early 20th century, Hexagenia were believed to be abundant in all mesotrophic waters of the
Great Lakes including Green Bay (Lake Michigan), Saginaw Bay (Lake Huron), Lake St. Clair,
western Lake Erie, Bay of Quinte (Lake Ontario), and portions of interconnecting rivers and
harbors. Thirty years of pollution-abatement programs may have allowed Hexagenia to return to
                         Draft for Discussion at SOLEC 2006

-------


other areas of the Great Lakes besides western Lake Erie as evidenced by anecdotal sightings of
winged mayflies in the 1990s. However, anecdotal reports have slowed and only one scientific
study (K. Kreiger, personal communication) has been performed to confirm anecdotal reports and
that study in central Lake Erie could not verify any Hexagenia recovery.

The only sustained recovery of Hexagenia in the Great Lakes (i.e., western Lake Erie) should be
monitored for another 4-6 years to determine annual variability and the carrying capacity of this
taxon in mesotrophic waters. If scientifically measured, the recovery will provide management
agencies with a quantitative endpoint of Hexagenia density which can be used to measure
recovery to a mesotrophic state in waters throughout the Great Lakes. In addition, a scientifically
determined carrying capacity of Hexagenia may also be useful as a benthic indicator for
remediation of contaminated sediments and as a guide for acceptable levels for food for valuable
percid communities. Contaminant levels in sediments that meet USEPA and OMOE guidelines
(i.e., "clean dredged sediment") and IJC criterion for oil and hydrocarbons (i.e., "sediment not
polluted") will not impair Hexagenia populations. There will be a graded response to
concentrations of metals  and oil in sediment exceeding these guidelines for clean sediment.
Reductions in phosphorus levels in formerly eutrophic habitats are likely to be accompanied by
colonization of Hexagenia, if surficial sediments are otherwise uncontaminated. Since Hexagenia
can be  one of the largest  and most abundant prey for percid  fishes such as yellow perch and
young walleye the reestablishment of Hexagenia in nearshore  waters of Great Lakes should be
encouraged.

Acknowledgments
Authors: Don W Schloesser, USGS,  Great Lakes Science Center, Ann Arbor, Michigan 48105,
dschloesser@usgs.gov

Data Sources
Beeton, A. M.  1969. Changes in the environment and biota of the Great Lakes. Pages 150-187 in
Eutrophication: causes, consequences, correctives. Proceedings of a Symposium. National
Academy of Sciences, Washington, D.C.

Bij de Vaate, A., A. Klink, and F. Oosterbroek, 1992. The mayfly, Ephoron virgo (Olivier), back
in the Dutch Parts of the  rivers Rhine and Meuse. Hydrobiological Bulletin 25:237-240.

Bridgeman, T.B., D.W. Schloesser, and A.E. Krause. 2005.  Recruitment of Hexagenia mayfly
nymphs in western Lake  Erie linked to environmental variability. Ecological Applications
16(2):0000-0000.

Burns,  N.M. 1985.  Erie:  The Lake That Survived. Rowman & Allanheld Publishers, Totowa,
Illinois. 320 pp.

Dermott, R. personal communication. Canadian Center for Inland Waters, Burlington, Ontario).

Edsall, T. A., B. A. Manny, D. W. Schloesser, S. J. Nichols, and A. M. Frank. 1991. Production
of Hexagenia limbata nymphs in contaminated sediments in the upper Great Lakes connecting
channels. Hydrobiologia  219:353-361.
                         Draft for Discussion at SOLEC 2006

-------
Edsall, T.A., O.T. Gorman, and L.M. Evrard. 2004. Burrowing mayflies as indicators of
ecosystem health: status of populations in two western Lake Superior embayments. Aquatic
Ecosystem Health & Management 7(4):507-513.

Edsall, T.A., M. Bur, O.T. Gorman, and J.S. Schaeffer. 2005. Burrowing mayflies as indicators of
ecosystem health: status of populations in western Lake Erie, Saginaw Bay, and Green Bay.
Aquatic Ecosystem Health & Management 8(2):107-116.

Fremling, C.R. andD.K. Johnson. 1990. Recurrence of Hexagenia mayflies demonstrates
improved water quality in Pool 2 and Lake Pepin, Upper Mississippi River, 243-248. -In:
Mayflies and stoneflies. Campbell, I. (ed.). Kluwer Academic Publication.

Kolar, C. S., P. L. Hudson, and J. F. Savino. 1997. Conditions for the return and simulation of the
recovery of burrowing mayflies in western Lake Erie. Ecological Applications 7:665-676.

Kreiger, K. personal communication. Heidelberg College, Tiffin, Ohio.

Krieger, K. A., D. W. Schloesser, B. A. Manny, C. E. Trisler, S. E. Heady, J. J. H. Ciborowski,
and K. M. Muth. 1996. Recovery of burrowing mayflies (Ephemeroptera: Ephemeridae:
Hexagenia) in western Lake Erie. Journal of Great Lakes Research 22:254-263.

Madenjian, C.P., D.W. Schloesser, andK.A. Krieger. 1998: Population models of burrowing
mayfly recolonization in western Lake Erie. Ecological Applications 8(4): 1206-1212.

Ohio Lake Erie Commission. 1998. State of Ohio 1998: state of the Lake Report. Ohio Lake Erie
Commission, Toledo, Ohio. 88 pp. (Available from Ohio Lake Erie Commission, One Maritime
Plaza, 4th Floor, Toledo, Ohio 43604-1866, USA).

Ohio Lake Erie Commission. 2004. State of the Lake Report 2004; Lake Erie Quality Index. Ohio
Lake Erie Commission, Toledo, Ohio. 79 pp. (Available from Ohio Lake Erie Commission, One
Maritime Plaza, 4th  Floor, Toledo, Ohio 43604-1866, USA).

Reynoldson, T. B., D. W. Schloesser, and B. A. Manny. 1989. Development of a benthic
invertebrate objective for mesotrophic Great Lakes waters. Journal of Great Lakes Research
15:669-686.

Schloesser, D.W., T.A. Edsall, B.A. Manny and S.J. Nichols. 1991. Distribution of Hexagenia
nymphs and visible oil in sediments of the upper Great Lakes connecting channels. Hydrobiologia
219: 345-352.

Schloesser, D.W. and J.K. Hiltunen. 1984. Life cycle of the mayfly Hexagenia limbata in the st.
Marys River between Lake Superior and Huron. Journal of Great Lakes Research 10:435-439.

Schloesser, D.W. 1988.  Zonation of mayfly nymphs and caddisfly larvae in the St. Marys River.
Journal of Great Lakes Research 14:227-233.
                        Draft for Discussion at SOLEC 2006

-------



Schloesser, D. W., K. A. Krieger, J. J. H. Ciborowski, and L. D. Corkum. 2000. Recolonization
and possible recovery of burrowing mayflies (Ephemeroptera: Ephemeridae: Hexagenia spp.) in
Lake Erie of the Laurentian Great Lakes. Journal of Aquatic Ecosystem Stress and Recovery
8:125-141.

Schloesser, D.W and T.F. Nalepa. 2001. Changing abundance of Hexagenia mayfly nymphs in
western Lake Erie of the Laurentian Great Lakes: impediments to assessment of lake recovery?
International Review Hydrobiologia 86(1):87-103.

List of Figures
Figure 1. Typical life-cycle of a burrowing mayfly such as Hexagenia found in the Great Lakes.
Source: Drawn by Martha Thierry, courtesy of the Detroit Free Press.

Figure 2. Densities (number/m ) of Hexagenia obtained in three studies (colored markers) in
western Lake Erie 1995-2005. Line of abundance fit by eye.
Source: Unpublished data, DWS)

Figure 3. Recruitment of young-of-year Hexagenia in western Lake Erie 1997-2002 Source:
Schloesser and Nalepa 2001; Bridgeman et al. 2005.

Figure 4. Densities (number/m2) of Hexagenia, three-year running average of densities, and
subjective target-reference values of desired abundance (i.e., poor, fair, good, etc.) in western
Lake Erie.
Source: After Ohio Lake Erie Commission 2004.

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
                            State of the Great Lakes 2007 - Draft
Figure 1. Typical life-cycle of a burrowing mayfly such as Hexagenia found in the Great Lakes.
Source: Drawn by Martha Thierry, courtesy of the Detroit Free Press.
                      Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
  CL



  I

  .5

  0)
  on
  «
  X
  09
     600 i
     500 -
400 -
300 -
  o
  >•  200
  to
  c
  m
  a
     100 •
                     1997
                                          2001
                                                                2005
        1994
             1996
1998
2000
2002
2004
Figure 2. Densities (number/m2) offfexagenia obtained in three studies (colored markers) in

western Lake Erie 1995-2005. Line of abundance fit by eye.

Source: Unpublished data, DWS)
                       Draft for Discussion at SOLEC 2006

-------
      1200 i
               YOY Recruitment in
                Western Lake Erie
          0
               1997    1998
1999   2000
    Year
2001    2002
Figure 3. Recruitment of young-of-year Hexagenia in western Lake Erie 1997-2002 Source:
Schloesser and Nalepa 2001; Bridgeman et al. 2005.
                     Draft for Discussion at SOLEC 2006

-------

          1995  1996  1997   1998  1999  2000  2001  2002  2003   2004  2005
Figure 4. Densities (number/m ) offfexagenia, three-year running average of densities, and
subjective target-reference values of desired abundance (i.e., poor, fair, good, etc.) in western
Lake Erie.
Source: After Ohio Lake Erie Commission 2004.
                         Draft for Discussion at SOLEC 2006

-------


Abundances of the Benthic Amphipod Diporeia spp.
Indicator #123
Overall Assessment
           Status:   Mixed
           Trend:   Deteriorating
   Primary Factors
      Determining
  Status and Trend
Abundances of the benthic amphipod Diporeia spp. continue to decline
in Lakes Michigan, Huron, and Ontario. While it is presently gone or
rare in shallow waters in each of these lakes, it is also declining in
deeper, offshore waters. The decline in the latter regions is temporally
linked to the expansion and increase of quagga mussels. Studies on
trends in Lake Superior are conflicting, but the general opinion of
researchers is that declines are not occurring.  Diporeia are currently
gone or very rare in Lake Erie.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Unchanging
Data sets are conflicting on current trends of Diporeia populations in Lake
Superior. One long-term monitoring program shows that Diporeia
abundances are declining in offshore areas (> 90 m), but abundances in
nearshore areas (< 65 m) remain unchanged. Other long and short-term
sampling programs show no overall trend in either offshore or nearshore
areas.
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Huron
Poor
Deteriorating
Diporeia abundances continue to decline in Lake Michigan.  A recent
lakewide survey (in 2005) indicated abundances were lower by 84 %
compared to abundances found in 2000 (Figure 1).  Diporeia are now
completely gone from depths < 80 m over most of the lake and abundances
are in the state of decline at depths > 80 m.
                        Draft for Discussion at SOLEC 2006

-------
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
Poor
Deteriorating
Diporeia abundances continue to decline in Lake Huron.  The most recent
lakewide survey in the main basin (in 2003) indicated abundances were
lower by 57 % compared to abundances found in 2000. Diporeia are now
completely gone from depths < 60 m except in the northeastern end and
continue to decline at depths > 60 m. Annual monitoring at 11 sites
indicated that, in 2005, Diporeia were gone from 5 sites and abundances
were lower compared to 2004 at the other 6 sites. Because of insufficient
data, trends in Georgian Bay and North Channel are not known.  However,
limited temporal and spatial data from the southern end of Georgian Bay
showed that Diporeia have been declining since 2000 and are  now
completely gone at depths < 93 m.
Poor
Deteriorating
Because of shallow, warm waters, Diporeia are naturally not present in the
western and central basins. Diporeia declined in the eastern basin
  Status and Trend  beginning in the early 1990s and have not been found since 1998.
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Poor
Deteriorating
Based on several limited surveys in 2005, Diporeia continue to decline in
Lake Ontario.  In one survey of 11  sites, Diporeia declined at 2 sites and
increased slightly at 2 sites compared to 2004. It was not found at 6 sites in
both years. In another survey of 14 sites, Diporeia declined at sites < 140
m, but abundances increased slightly at sites > 190 m. It was not found at
sites < 90 m over most of the lake.
Purpose
To provide a measure of the biological integrity of the offshore regions of the Great Lakes by
assessing the abundance of the benthic macroinvertebrate Diporeia.

Ecosystem Objective
The ecosystem goal is to maintain a healthy, stable population of Diporeia in offshore regions of
the main basins of the Great Lakes, and to maintain at least a presence in nearshore regions.

State of the Ecosystem
Background
This glacial-marine relic was once the most abundant benthic organism in cold, offshore regions
(> 30 m) of each of the lakes. It was present, but less abundant in nearshore regions of the open
lake basins, but naturally absent from shallow, warm bays, basins, and river mouths. Diporeia
occurs in the upper few centimetres of bottom sediment and feeds on algal material that freshly
                         Draft for Discussion at SOLEC 2006

-------

                                        "   '' ^"4&lllll
settles to the bottom from the water column (i.e., mostly diatoms). In turn, it is fed upon by most
species of fish, in particular by many forage fish species which serve as prey for the larger
piscivores such as trout and salmon. For example, sculpin feed almost exclusively upon Diporeia,
and sculpin are fed upon by lake trout. Also, lake whitefish, an important commercial species,
feeds heavily on Diporeia. Thus, Diporeia was an important pathway by which energy was
cycled through the ecosystem, and a key component in the food web of offshore regions. The
importance of this organism is recognized in the Great Lakes Water Quality Agreement
(Supplement to Annex 1 - Specific Objectives).

On a broad scale, abundances are directly related to the amount of food settling to the bottom, and
population trends reflect the overall productivity of the ecosystem. Abundances can also vary
somewhat relative to shifts in predation pressure from changing fish populations. In nearshore
regions, this species is sensitive to local sources of pollution.

Status of Diporeia
Diporeia populations are currently in the state of dramatic decline in Lakes Michigan, Ontario,
and Huron, and are completely gone or very rare in Lake Erie. Results are conflicting for Lake
Superior. One data set shows a trend of declining abundances in offshore waters, but other data
sets show no trend. In all the lakes except Superior, abundances have decreased progressively
from shallow to deeper areas.  Initial declines were first observed in all lake areas within 2-3
years of when zebra mussels (Dreissena polymorphd) or quagga mussel (Dreissena bugensis)
first became established. These two species were introduced into the Great Lakes in the late
1980s via the ballast water of ocean-going ships. Reasons for the negative response of Diporeia
to these mussel species are not entirely clear. One hypothesis is that dreissenid mussels are out
competing Diporeia for available food. That is,  large mussel populations were filtering food
material before it reached the bottom, thereby decreasing amounts available to Diporeia.
However, evidence suggests that the reason for the decline is more complex than a simple decline
in food because Diporeia have completely disappeared from areas where food is still settling to
the bottom and where there are no local populations of mussels. Also, individual Diporeia show
no signs of starvation before or during population declines. Further, Diporeia and Dreissena
apparently coexist in some lakes outside of the Great Lakes (i. e., Finger Lakes in New York).

Pressures
As populations of dreissenid mussels continue to expand, it may be expected that declines in
Diporeia will become more extensive. In the open waters of Lakes Michigan, Huron, and
Ontario, zebra mussels are most  abundant at depths less than 50 m, and Diporeia are now gone or
rare from lake areas as deep as 90 m. Recently, quagga mussel populations have increased
dramatically in each of these lakes and are occurring at deeper depths than zebra mussels.  The
decline of Diporeia at depths > 90 m can be attributed to the expansion of quagga mussels  to
these depths.

Management Implications
The continuing decline of Diporeia has strong implications to the Great Lakes food web. As
noted, many fish species rely on Diporeia as a major prey item, and the loss of Diporeia will
likely have  an impact on these species. Responses may include changes in diet, movement to
areas with more food, or a reduction in weight or energy content. Implications to populations
include changes in distribution, abundance, growth, recruitment, and condition. Recent evidence
                         Draft for Discussion at SOLEC 2006

-------
                                                                           sutS"
suggests that fish are already being affected. For instance, growth and condition of an important
commercial species, lake whitefish, has declined significantly in areas where Diporeia
abundances are low in Lakes Michigan, Huron, and Ontario. Also, studies show that other species
such as alewife, slimy sculpin, and bloater have been affected. Management agencies must know
the extent and implications of these changes when assessing the current state and future trends of
the fishery. Any proposed rehabilitation of native fish species, such as the re-introduction of
deepwater ciscoes in Lake Ontario, requires knowledge that adequate food, especially Diporeia,
is present.
Comments from the author(s)
Because of the rapid rate at which Diporeia populations are declining and their significance to the
food web, agencies committed to documenting trends should report data in a timely manner. The
population decline has a defined natural pattern, and studies of food web impacts should be
spatially well coordinated. Also, studies to define the cause of the negative response of Diporeia
to Dreissena should continue and build upon existing information. With an understanding of
exactly why Diporeia populations are declining, we may better predict what additional areas of
the lakes  are at risk. Also, by better understanding the cause, we may better assess the potential
for population recovery if and when dreissenid populations stabilize or decline.

Acknowledgments
Authors:  T.F. Nalepa, Great Lakes Environmental Research Laboratory, National Oceanic and
Atmospheric Administration, Ann Arbor, ML; and
R. Dermott, Great Lakes Laboratory for Fisheries and Aquatic Sciences, Fisheries and Oceans
Canada, Burlington, ON.

The authors thank the Great Lakes National Program Office, EPA for providing some data used
in this report.

Data Sources
Dermott,  R. 2001. Sudden disappearance of the amphipod Diporeia from eastern Lake Ontario,
1993-1995. J. Great Lakes Res. 27:423-433.

Dermott,  R., and Kerec, D. 1997. Changes in the deepwater benthos of eastern Lake Erie since
the invasion of Dreissena: 1979- 1993.  Can. J. Fish. Aquat. Sci. 54:922-930.

Hondorp, D. W., Pothoven, S. A., and Brandt, S. B.  2005. Influence of Diporeia density on the
diet composition, relative abundance, and energy density of planktivorous fishes in southeast
Lake Michigan. Trans. Am. Fish. Soc.  134:  588-601.

Lozano, S.J.,  Scharold, J.V., and Nalepa, T.F. 2001. Recent declines in benthic macroinvertebrate
densities  in Lake Ontario. Can. J. Fish.  Aquat. Sci. 58:518-529.
                         Draft for Discussion at SOLEC 2006

-------

                                                  .'.:,;"". ^.~,:' ^'V';'#S^|
Mohr, L. C. andNalepa, T. F. 2005. Proceedings of a workshop on the dynamics of lake
whitefish (Coregonis dupeaformis) and the amphipod Diporeia spp. in the Great Lakes.  Great
Lakes Fish. Comm. Tech. Rep. 66.

Nalepa, T. F., Rockwell, D. C., and Schloesser, D. W.  2006. Disappearance of Diporeia spp. in
the Great Lakes: workshop summary, discussion, and recommedations. NOAA Technical
Memorandum GLERL-136, Great Lakes Environmental Research Laboratory, Ann Arbor, MI.

Nalepa, T. F., Fanslow, D. L., Foley, A. J., Ill, Lang, G. A., Eadie, B. J., and Quigley, M. A.
2006. Continued disappearance of the benthic amphipof Diporeia spp. in Lake Michigan: is there
evidence for food limitation?  Can. J. Fish. Aquat. Sci. 63:  872-890.

Owens, R.W., and Dittman, D.E. 2003. Shifts in the diets of slimy sculpin (Cottus cognatus) and
lake whitefish (Coregonus dupeaformis) in Lake Ontario following the collapse of the burrowing
amphipod Diporeia. Aquat. Ecosys. Health Manag. 6:311-323.
List of Figures
Figure 1.  Distribution and abundance (number per square meter) of the amphipod Diporeia spp.
in Lake Michigan in 1994-1995, 2000, and 2005.  Small crosses indicate location of sampling
stations.
Source: National Oceanic & Atmospheric Administration (NOAA) Great Lakes Environmental
Research Laboratory

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
               1994/95
2000
                                                                   2005
             Density (No. m' x
                                       Density (No.
                                                                Density {No. m' xlO')
Figure 1. Distribution and abundance (No. m"2) of the amphipod Diporeia spp. in Lake
Michigan in 1994/1995, 2000, and 2005.  Small crosses indicate location of sampling stations.
                        Draft for Discussion at SOLEC 2006

-------

 1995
 Amphipod Diporeia

< No Sample
• 0     •  401 • 800 in'2
• 1-90  •  801 - 1500
• 91 -200 •  1501 -3000
• 201-400 •3001-8000
Figure 2. Distribution and abundance (No. m") of the amphipod Diporeia spp. in Lake
Ontario in 1995, 2003, and 2005.  Small crosses indicate a site where no sample was taken.
                       Draft for Discussion at SOLEC 2006

-------



                                          •s '•: >:•-. >..
External Anomaly Prevalence Index for Nearshore Fish
Indicator #124

Overall Assessment
           Status:   Poor
           Trend:   Unchanging
   Primary Factors
      Determining
  Status and Trend
Lake-by-Lake Assessment
Lake Superior
           Status:   Not Assessed
           Trend:   Undetermined
Lake Michigan
           Status:
           Trend:
Not Assessed
Undetermined
Lake Huron
           Status:
           Trend:
Lake Erie
Not Assessed
Undetermined
           Status:   Poor
           Trend:   Unchanging
Lake Ontario
           Status:   Poor
           Trend:   Unchanging

Purpose
1) To assess select external anomalies in nearshore fish;
2) To identify nearshore areas that have populations of benthic fish exposed to contaminated -
sediments; and
3) To help assess the recovery of Areas of Concern (AOCs) following remedial activities
Insert Purpose text

Ecosystem Objective
The objective is to help restoration and protection of beneficial uses in Areas of Concern or in
open Great Lakes waters, including beneficial use (iv) Fish tumors or other deformities (Great
Lakes Water Quality Agreement (GLWQA), Annex 2). This indicator also supports Annex  12 of
the GLWQA.
                        Draft for Discussion at SOLEC 2006

-------
State of the Ecosystem
Background

The presence of contaminated sediments at AOCs has been correlated with an increased
incidence of external and internal anomalies in benthic fish species (brown bullhead and white
suckers) that may be associated with specific groups of chemicals. Elevated incidence of liver
tumors (histopathologically verified pre-neoplastic or neoplastic growths) were frequently
identified during the past two decades. These elevated frequencies of liver tumours have been
shown to be useful indicators of beneficial use  impairment of Great Lakes aquatic habitat.
External raised growths (histopathologically verified tumors on the body and lips), such as lip
papillomas, have also been useful indicators. Raised growths may not have a single etiology; but,
they have been produced experimentally by direct application of polynuclear aromatic
hydrocarbons (PAH) carcinogens to brown bullhead skin. Field and laboratory studies have
correlated verified liver and external raised growths with chemical contaminants found in
sediments at some AOCs in Lake Erie, Lake Michigan, Lake Ontario and Lake Huron.  Other
external anomalies may also be used to assess beneficial use impairment. The external anomaly
prevalence index (EAPI) will provide a tool for following trends in fish population health that can
be used by resource  managers and community-based monitoring programs.

The EAPI has been developed for mature (> 3 years  of age) fish as  a marker of both contaminant
exposure and of internal pathology. Brown bullhead have been used to develop the index. They
are the most frequently used benthic indicator species in the southern Great Lakes and have been
recommended by the International Joint Commission (IJC) as a  key indicator species (IJC 1989).
The most common external anomalies found in brown bullhead over the last twenty years from
Lake Erie are: 1) abnormal barbels (BA); 2) focal discoloration  (FD); and 3) raised growths (RG)
- on the body and lips (Figurel). Initial statistical analysis of sediments and external anomalies at
different  locations indicates that variations in the chemical mixtures (Total, priority and
carcinogenic PAHs;  DDT metabolites; organochlorine chemicals (OC); and total metals) show a
statistically significant relation with a differing prevalence of individual external anomalies
(raised growths and  barbell abnormalities). Age and  external anomalies indicate a positive
correlation (Figure 2). Impairment determinations should be based on age comparisons of the
prevalence of external anomalies at contaminated sites with the  prevalence at "reference" (least
impacted) sites (Figure 3). Preliminary data indicate  that if the prevalence of raised growths on
the body  and lip combined is > 5%, barbell abnormalities  >10% and focal discoloration
(melanistic alterations) > 5% in brown bullhead, the  population should be considered impaired.

Surveys conducted in 1999 and 2000 in the Detroit, Ottawa, Black, Cuyahoga, Ashtabula,
Buffalo, and Niagara Rivers and at Old Woman Creek in Lake Erie demonstrated that external
raised growths are positively associated with both PAH metabolites in bile and in PAH
concentrations in sediment. The association with PAH metabolites in bile (Figure 4) is  stronger
than that  with total PAH concentrations in sediments (Figure 5). Bile metabolite concentrations
may be a better estimate of potential exposure of PAHs to individual fish than concentrations in
sediments. The EAPI indicates the impacts from the  exposure to individual fish from the PAHs as
well as other compounds in the mixtures of compounds that may be present in sediments. Barbel
deformities (Figure 5) also showed a positive correlation with total PAH levels in sediment. In
addition to the locations  listed above, the Huron River and Presque Isle Bay sites all showed a
                         Draft for Discussion at SOLEC 2006

-------

statistically significant correlation between external raised growths and concentration of heavy
metals in sediment (Figure 6).

Pressures
Many Great Lakes AOCs and their tributaries remain in a degraded condition. Exposure of the
fish populations to contaminated sediment continues and the elevated evidence of external
anomalies still persist. The human population in the Great Lakes is expected to increase and
urbanization along Great Lakes tributaries and shorelines will likely expand in the future.
Therefore, some locations impacted by land use changes may continue to deteriorate even as
control and remediation actions improve conditions at the older contaminated sites. As
recommended for delisting, listed AOCs continue the gain knowledge in order to achieve a low
EAPI to help the delisting process of the BUI for fish tumors and other deformities. A single
common data base must be implemented for international brown bullhead data sets to evaluate
AOC and reference conditions in each of the Great Lakes.

Management Implications
The EAPI provides managers and researchers with a tool to monitor contaminant impacts to the
fish populations in Great Lakes AOCs. Additional remediation to clean up contaminated
sediments at Great Lakes AOCs will help to reduce rates of external anomalies.  The EAPI,
particularly for brown bullheads and white suckers and the inclusion of a single common data
base will help environmental managers to follow trends in fish population health and to
determine the status of AOCs that may be considered for delisting (IJC Delisting Criteria, see
IJC 1996).

Comments from the author(s)
This external anomaly index for benthic species has potential for defining habitats that may  or
may not be impacted from contaminants. Collaborative U.S. and Canadian studies investigating
the etiology and prevalence of external anomalies in benthic fishes over a gradient of polluted to
pristine Great Lakes habitats are desperately needed. These studies would create a common  index
that could be used as an indicator of ecosystem health. The establishment of single data base to
house all lake wide data for each Great Lake is necessary to enable managers and decision makers
to gain an understanding of the health of individual fish (e.g. brown bullhead) and their
populations. Unless this takes place, understanding of health conditions at AOCs compared to the
least impacted (reference) sites will remain unknown and the delisting process will not advance.

Acknowledgments
Authors:

Stephen B. Smith, U.S. Geological Survey, Biological Resources, Reston, VA;

Paul C. Baumann, U.S. Geological Survey, Biological Resources, Columbus, OH; and

* Scott Brown, Environment Canada, National Water Research Institute, Burlington, ON.

*Dedicated to our friend and collogue Scott Brown, whose untimely passing has saddened all
who knew him.
                         Draft for Discussion at SOLEC 2006

-------

Data Sources
International Joint Commission (IJC). 1996. Indicators to evaluate progress under the Great
Lakes Water Quality Agreement.
Indicators for Evaluation Task Force. ISBN 1-895058-85-3.

International Joint Commission (IJC). 1989. Guidance on characterization of toxic substances
problems in areas of concern in the Great Lakes Basin. Report of the Great Lakes Water Quality
Board, Windsor, ON.

Smith, S.B., Reader, D.R.P., Baumann, P.C., Nelson, S.R., Adams, J.A., Smith, K.A., Powers,
M.M., Hudson, P.L., Rosolofson, A.J., Rowan, M., Peterson, D., Blazer, V.S., Hickey, J.T., and
Karwowski, K. 2003. Lake Erie Ecological Investigation; Summary of findings: Part 1;
Sediments, Invertebrate communities, and Fish Communities: Part 2; Indicators, anomalies,
Histopathology, and ecological risk assessment. U.S.Geological Survey, Mimeo.
List of Figures
Figure 1. External Anomalies on brown bullhead collected from Lake Erie from the 1980's
through 2000.  BA- barbel abnormalities, RG- raised growth (body and lip), FD-focal
discoloration, LE-lesion (total ca. 2400 fish). Source: Great Lakes Science Center, Ann Arbor,
MI.

Figure 2. Age of brown bullhead at Lake Erie sites from 1986-87 and 1998-2000 collections in
relation to combined external anomalies. Age groups; age 3, ages 4&5, ages 6&7.  Source: S.B.
Smith, unpublished data.

Figure 3. External anomalies (Melanoma. Raised Growth on body and lips, and Barbell
abnormalities) in relation to  sites classified for sediment contaminants  and BB morphology from
all collections in the 1980's  and 1990's. Source: S. B. Smith, unpublished data.

Figure 4. Prevalence of external raised growths in brown bullhead from Lake Erie tributaries
compared to PAH metabolite concentrations in bile (B[P]  and NAPH-type unit are (ig/mg protein.
Source: Yang and Baumann, unpublished data.

Figure 5. Prevalence of external raised growths and barbel deformities in brown bullhead from
Lake Erie tributaries compared to PAH concentrations in sediment. Source: Yang and Baumann,
unpublished data.

Figure 6. Prevalence of external raised growths in brown bullhead from Lake Erie tributaries
compared to concentrations  of heavy metals in sediment. Source: Yang and Baumann,
unpublished data.

Last updated
SOLEC 2006
                         Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
                        LE 5%
      FD 20%
    RG-L 15%
A 37%
                                       Eye 11%
              RC-B 7%
                      Fin 3%  Gill 2%
Figure 1. External Anomalies on brown bullhead collected from Lake Erie from the 1980's
through 2000. BA- barbel abnormalities, RG- raised growth (body and lip), FD-focal
discoloration, LE-lesion (total ca. 2400 fish).
Source: Great Lakes Science Center, Ann Arbor, MI.
                      Draft for Discussion at SOLEC 2006

-------
                  CO
                     o
                                                                    6&7YR
                                                                   SYR
                              Sites
Figure 2. Age of brown bullhead at Lake Erie sites from 1986-87 and 1998-2000 collections
in relation to combined external anomalies. Age groups; age 3, ages 4&5, ages 6&7.
Source: S.B. Smith, unpublished data.
                       Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
          4 and 5
                6 and 7
                                4and5   Band?

                                 Age Groups
4 and 5
      6 and 7
Figure 3. External anomalies (Melanoma, Raised Growth on body and lips, and Barbell
abnormalities) in relation to sites classified for sediment contaminants and BB morphology
from all collections in the 1980's and 1990's.
Source: S. B. Smith, unpublished data.
                      Draft for Discussion at SOLEC 2006

-------
^_^
^
CO
c
o
1
"ro
L_
2
X
0>
«•_
o
0
o
0
03
0
QL


100 -•
90-
80 •
70-
60-
50-

40-

30-

20-
10 -

C -
0 B[a]P-type, Spearman r = 0.60, p = 0.12
• NAPH-type, Spearman r = 0.69, p = 0.06

o m


o m


o •
o •
o •
0 


-------
        State of the Great Lakes 2007 - Draft

w
t_
2
1
1
^
X
0)
"o
8
1
03
1
c


100 -I
90 -
80-
70-
60-
50-

40 -

30-
20-

10-


o External lesbns, Spearman r = 0.41, p = 0.24
• Barbel def ormfties. Spearman r = 0.53, p = 0.12


• m


o
m m
O f>
o •
-
0 0
o
0 5 10 15 20 25
Concentration of select sediment PAHs (pg/g dry weight)
Figure 5. Prevalence of external raised growths and barbel deformities in brown bullhead
from Lake Erie tributaries compared to PAH concentrations in sediment.
Source: Yang and Baumann, unpublished data.
                     Draft for Discussion at SOLEC 2006

-------

?
i
i
"re
E
0)
"x
0)
**-*
o

HI
c
01
re

2!


100 -I
90-
80 -
70-
60-

50-
40-
30-


20-

10-

0 -


* Spearman r = 0.71 , p = 0.02


"


•

• A
+w



* *
J
T r T ^^^^ ^^
0 200 400 600 800 1000 1200 1400
Mean concentration of heavy metals in sediment ((jg/g dry weight)
Figure 6. Prevalence of external raised growths in brown bullhead from Lake Erie
tributaries compared to concentrations of heavy metals in sediment.
Source: Yang and Baumann, unpublished data.
10
Draft for Discussion at SOLEC 2006

-------
Status of Lake Sturgeon in the Great Lakes
Indicator #125
Overall Assessment
           Status:  Mixed
           Trend:  Improving
   Primary Factors
      Determining
  Status and Trend
There are remnant populations in each basin of the Great Lakes, but few of
these populations are large. Much progress has been made in recent years
learning about population status in many tributaries.  Confirmed
observations and captures of lake sturgeon are increasing in all lakes.
Stocking is contributing to increased abundance in some areas. There
remains a need for information on some remnant spawning populations.
Little is known about the juvenile life stage. In many areas habitat
restoration is needed as spawning and rearing habitat has been destroyed,
altered or access is blocked.
Lake by Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving or Undetermined
Lake sturgeon abundance shows an increasing trend in a few remnant
populations and where stocked in the Ontonagon and St. Louis rivers. Lake
sturgeons currently reproduce in at least 10 of 21 known historic spawning
tributaries.
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving and Undetermined
Remnant populations persist in at least 8 tributaries having unimpeded
connections to Lake Michigan. Successful reproduction has been
documented in six rivers and abundance has increased in a few in recent
years. Active rehabilitation has been initiated through rearing assistance in
1 remnant population and reintroductions have been initiated in three rivers.
Mixed
Improving and Undetermined
Current lake sturgeon spawning activity is limited to five tributaries, four in
Georgian Bay and the North Channel and one in Saginaw Bay. Abundant
stocks of mixed sizes are consistently captured in the North Channel,
Georgian Bay, southern Lake Huron and Saginaw Bay.
Poor
Undetermined
Current lake sturgeon spawning activity is unknown except for three
spawning areas identified in the Detroit and St. Clair Rivers. The western
basin of Lake Erie, the North Channel of the St. Clair River and Anchor
Bay in Lake St. Clair appear to be nursery areas for juveniles. In the central
and eastern basins lake sturgeon are scarcer.

-------
Lake Ontario
           Status:   Mixed
           Trend:   Improving
   Primary Factors   Lakewide incidental catches since 1995 indicate a possible improvement in
      Determining   their status. Spawning occurs in the Niagara River, Trent River, and
  Status and Trend   possibly the Black River. There are sizeable populations within the St.
                    Lawrence River system. Stocking for restoration began in 1995 in New
                    York.

Purpose
• Lake sturgeon was a key component of the nearshore benthivore fish community and their
presence and abundance indicates the health and status of that component of the Great Lakes
ecosystem.

Ecosystem Objective
Lake sturgeon is identified as an important species in the Fish Community Objectives for each of
the Great Lakes. Lake Superior has a lake sturgeon rehabilitation plan, and many of the Great
Lakes States have lake sturgeon recovery/rehabilitation plans which call for increasing numbers
of lake sturgeon beyond current levels. [Conserve, enhance or rehabilitate self-sustaining
populations of lake sturgeon where the species historically occurred and at a level that will permit
all State, Provincial and Federal delistings.]

State of the Ecosystem
Background
Lake sturgeon, Acipenser fulvescens, were historically abundant in the Great Lakes with
spawning populations using many of the major tributaries, connecting waters, and shoal areas
across the basin. Prior to European settlement of the region, they were a dominant component of
the nearshore benthivore fish community, with populations estimated in the millions in each of
the Great Lakes (Baldwin et al. 1979). In the mid- to late-1800s, they contributed significantly as
a commercial species ranking among the  five most abundant species in the commercial catch
(Baldwin et al. 1979, Figure 1).

The decline of lake sturgeon populations  in the Great Lakes was rapid and commensurate with
habitat destruction, degraded water quality, and intensive fishing associated with settlement and
development of the region. Sturgeon were initially considered a nuisance species of little value by
European settlers, but by the mid-1800s, their value as a commercial species began to be
recognized and a lucrative fishery  developed. In less than 50 years, their abundance had declined
sharply, and since 1900, they have remained a highly depleted species of little consequence to the
commercial fishery. Sturgeon are now extirpated from many tributaries and waters where they
once spawned and flourished (Figure 2 and Figure 3). They are considered rare, endangered,
threatened, or of watch  or special concern status by the various Great Lakes fisheries
management agencies. Their harvest is currently prohibited or highly regulated in most U.S. and
Canadian waters of the  Great Lakes.

Status of Lake Sturgeon
Efforts are continuing by many agencies and organizations to gather information on remnant
spawning populations in the Great Lakes. Most sturgeon populations continue to sustain
themselves at a small fraction of their historical abundance. In many systems, access to spawning
habitat has been blocked, and other habitats have been altered. However, there are remnant
populations in each basin of the Great Lakes, and some of these populations are large in number

-------
(10's of thousands offish, Figure 3).  Genetic analysis has shown that Great Lakes populations
are regionally structured and show significant diversity within and among lakes.

Lake Superior: The fish community of Lake Superior remains relatively intact in comparison to
the other Great Lakes (Bronte et al. 2003). Historic and current information indicate that at least
21 Lake Superior tributaries supported spawning lake sturgeon populations (Harkness and
Dymond 1961; Auer 2003; Holey et al. 2000). Lake sturgeons currently reproduce in at least 10
of these tributaries. Sturgeon populations in Lake Superior continue to sustain themselves at a
small fraction of their historical abundance.

Current populations in Lake Superior are reduced from historic levels and none meet  all
rehabilitation targets. The number of lake sturgeon in annual spawning runs has been estimated
over a multi-year period to range from 200-375 adults in the Sturgeon River, (Hay-Chmielewski
and Whelan 1997; Holey et al. 2000), 200-350 adults in the Bad River in 1997 and 1998 (U.S.
Fish and Wildlife Service, Ashland Fishery Resource Office, USFWS, 2800 Lake Shore Drive,
Ashland, Wisconsin, 54806, unpublished data), and 140 adults in the Kaministiquia River,
Ontario (Stephenson 1998). Estimates of lakewide abundance are available from  the period
during or after targeted commercial harvests in the 1880s. Using data from Baldwin et al. (1979),
Hay-Chmielewski and Whelan (1997) estimated that historic lake sturgeon abundance in Lake
Superior was 870,000 individuals of all ages. If the Rehabilitation Plan target of 1,500 adults
were met in all 21 tributaries, the minimum lakewide abundance of adult fish would be  31,500.

Radio telemetry studies suggest that a river resident population inhabits  the Kaministiquia River
(Mike Friday, OMNR, Upper Great Lakes Management Unit-Lake Superior, 435 James St.
South,  Thunder Bay, Ontario P7E 6S8, personal communication). The Pic River also  has the
potential to support a river resident population. Juvenile lake sturgeon index surveys conducted
by the Great Lakes Indian Fish and Wildlife Commission and U.S. Fish  and Wildlife  Service in
Wisconsin waters show a gradually increasing trend in catch per unit effort from  1994-2002
(Table 1). Since 2001, sturgeon spawning surveys have been conducted  for the first time in 8
tributaries. Genetic analysis has shown that lake sturgeon populations in Lake Superior  are
significantly different from those in the other Great Lakes. Currently, there is no  commercial
harvest of lake sturgeon allowed in Lake Superior. Regulation of recreational and
subsistence/home use harvest in Lake Superior varies by agency.

Lake Michigan:  Sturgeon populations in Lake Michigan continue to sustain themselves at a small
fraction of their historical abundance. An optimistic estimate of the lakewide adult abundance is
less than 5,000 fish, well below 1% of the most conservative estimates of historic abundance
(Hay-Chmielewski and Whelan 1997). Remnant populations currently are known to spawn in
waters of at least 8 tributaries having unimpeded connections to Lake Michigan (Schneeberger et
al 2005). Two rivers, the Menominee and Peshtigo, appear to support annual spawning runs of
200 or more adults, and four rivers, the Manistee, Muskegon, Fox and Oconto, appear to support
annual spawning runs of between 25 and 75 adults. Successful reproduction has been documented
in all six of these rivers, although actual recruitment levels remain unknown. However,
abundance in some of these rivers appears to be increasing in recent years. Two other rivers, the
Manistique and Kalamazoo, appear to have annual spawning runs of less than 25 fish, and
reproductive success remains unknown. Lake sturgeon have been observed during spawning
times in a few other Lake Michigan tributaries such as the St. Joseph, Grand and  Millecoquins,
and near some shoal areas where sturgeon are thought to have spawned historically. It is not
known if spawning occurs regularly in these systems, however, and their status is uncertain.

-------
Lake Huron:  Lake sturgeon populations continue to be well below historical levels.  Spawning
has been identified in the Garden, Mississaugi and Spanish rivers in the North Channel, in the
Nottawasaga River in Georgian Bay and in the Rifle River in Saginaw Bay. Adult spawning
populations for each of these river systems are estimated to be in the ten's and are well below
rehabilitation targets (Hay-Chmielewski and Whelan 1997; Holey et al. 2000).  Barriers on
Michigan tributaries to Lake Huron continue to limit successful rehabilitation.  Stocks of lake
sturgeon in Lake Huron are monitored primarily through the volunteer efforts of commercial
fishers cooperating with the various resource management agencies. To date the combined efforts
of researchers in U.S. and Canadian waters has resulted in over 6,600 sturgeon tagged in Saginaw
Bay, southern Lake Huron, Georgian Bay and the North Channel, with relatively large stocks of
mixed sizes being captured at each of these general locations. Tag recoveries and telemetry
studies indicate that lake sturgeon are moving within and between jurisdictional boundaries and
between lake  basins, supporting the need for more cooperative management between the states
and between the U.S. and Canada.  The Saginaw River watershed and the St. Mary's River
systems are being assessed for spawning, both projects are ongoing and will continue through
2007. Similar research is being planned for the Thunder and Rifle Rivers in Michigan.

Lake Erie: Lake sturgeon populations continue to be well below historical levels.  Spawning has
been identified at two locations in the St. Clair River and at one location in the Detroit River
(Manny and Kennedy 2002). Tag recovery data and telemetry research indicates that a robust
lake sturgeon stock (> 45,000 fish) reside in the North Channel of the St. Clair River and Lake St.
Clair (Thomas and Haas 2002). The North Channel, Anchor Bay and the western basin of Lake
Erie have been identified as nursery areas as indicated by consistent catches in commercial and
survey fishing gears. In the central and eastern basins of Lake Erie  lake sturgeon are scarcer with
only occasional catches of sub-adult or adult lake sturgeon in commercial fishing nets and none in
research nets.  A botulism-related die off in 2001 and 2002, and declines in sightings by anglers
and others near Buffalo indicate a possible decline in population abundance of lake sturgeon in
Lake Erie. Research is scheduled in 2007 to identify if spawning stocks of sturgeon are using
reputed historic spawning sites in the lower Detroit River  and the Maumee River. Research
efforts will continue to focus on identifying new spawning locations, genetic difference between
stocks, habitat requirements, and migration patterns.

Lake Ontario: Lake Ontario has lake sturgeon spawning activity documented in two major
tributaries (Niagara River and Trent River) and suspected in at least one more (Black River) on an
infrequent basis.  There is no targeted assessment of lake sturgeon in Lake Ontario, but incidental
catches in research nets have occurred since 1997 (Ontario Ministry of Natural Resources 2004)
and 1995 (Eckert 2004), indicating a possible improvement in population status. Age analysis of
lake sturgeon captured in the lower Niagara River indicates successful reproduction in the mid-
1990s. New York State Department of Environmental Conservation initiated a stocking program
in 1995 to recover lake sturgeon populations. Lake  sturgeon have been stocked in the St.
Lawrence River and some of its tributaries, inland lakes in New York, and the Genesee River.
There are sizeable populations within the St.  Lawrence River system, most notably the Des
Prairies River, Lac St. Pierre and the St.  Maurice River. However, access is inhibited for many of
the historical  spawning grounds in tributaries by small dams and within the St. Lawrence River
by the Moses-Saunders Dam.

Pressures
Low numbers or lack of fish (where extirpated) is itself is a significant impediment to recovery in
many spawning areas. Barriers that prevent lake sturgeon  from moving into tributaries to spawn
are a major problem. Predation on eggs and newly hatched lake sturgeon by non-native predators
may also be a problem. The genetic structure of remaining populations is being studied by

-------
university researchers and fishery managers, and this information will be used to guide future
management decisions. With the collapse of the Caspian Sea sturgeon populations, black market
demand for sturgeon caviar could put tremendous pressure on Great Lakes lake sturgeon
populations. An additional concern for lake sturgeon in Lake Erie and Lake Ontario is the
presence of high densities of round gobies and the spread of Botulism Type E, which produced a
die-off of lake sturgeon in Lake Erie in 2001 and 2002. Botulism may also have been the cause of
similar mortalities observed in Lake Ontario in 2003 and in Green Bay of Lake Michigan.

Management Implications
Lake sturgeon are an important native species that are listed in the Fish Community Objectives
for all of the Great Lakes. Many of the Great Lakes states and provinces either have or are
developing lake sturgeon management plans promoting the need to inventory, protect and restore
the species to greater levels of abundance.

While overexploitation removed millions of adult fish, habitat degradation and alteration
eliminated traditional spawning grounds. Current work is underway by state, federal, tribal,
provincial and private groups to document active spawning sites, assess habitat condition and
availability of good habitat, and determine the genetics of remnant Great Lakes lake sturgeon
populations.

Several meetings and workshops have been held focusing on identifying the research and
assessment needs to further rehabilitation of lake sturgeon in the Great Lakes (Holey et al. 2000),
and a significant amount of research and assessment directed towards these needs has occurred in
the last 10 years. Among these is the research to better define the genetic structuring of Great
Lakes lake sturgeon populations, and genetics-based rehabilitation plans are being developed to
help guide reintroduction and rehabilitation efforts being implemented across the Great Lakes.
Research into new fish passage technologies that will allow safe upstream and downstream
passage around barriers to migration also have been underway for several years. Many groups  are
continuing to work to identify current lake sturgeon spawning locations in the Great Lakes, and
studies are being initiated to identify habitat preferences for juvenile lake sturgeon (ages 0-2).

Comments from the author(s)
Research and development is needed to determine ways to pass lake  sturgeon at man-made
barriers on rivers. In addition, there are significant, legal, logistical,  and financial hurdles to
overcome in order to restore degraded spawning habitats in connecting waterways and tributaries
to the Great Lakes. More monitoring is needed to determine the current status of Great Lakes
lake sturgeon populations, particularly the juvenile life stage. Cooperative effort between law
enforcement and fishery managers is required as  world pressure on sturgeon stocks will result  in
the need to protect large adult lake sturgeon in the Great Lakes.

Acknowledgments
Authors: Betsy Trometer and Emily Zollweg, U.S. Fish and Wildlife Service, Lower Great Lakes
Fishery Resources Office, 405 N. French Rd., Suite 120A, Amherst,  NY 14228;
Robert Elliott, U.S. Fish and Wildlife Service, Green Bay Fishery Resources Office, 2661 Scott
Tower Drive, New Franken, WI 54229;
Henry Quinlan, U.S. Fish and Wildlife Service, Ashland Fishery Resources Office, 2800
Lakeshore Drive E., Ashland, WI 54806; and
James Boase, U.S. Fish and Wildlife Service, Alpena Fishery Resources Office, 145 Water Street,
Alpena, MI, 49707..

-------
Data Sources
Auer, N.A. (ed.). 2003. A lake sturgeon rehabilitation plan for Lake Superior. Great Lakes
Fishery Commission Misc. Publ. 2003-02.

Baldwin, N.S., Saalfeld, R.W., Ross, M.A., and Buettner, HJ. 1979. Commercial fish production
in the Great Lakes 1867-1977. Great Lakes Fishery Commission Technical Report 3.

Bronte, C.R., Ebener, M.P., Schreiner, D.R., DeVault, D.S., Petzold, M.M., Jensen, D.A.,
Richards, C., and Lozano, SJ. 2003. Fish community changes in Lake Superior, 1970-2000. Can.
J. Fish. Aquat. Sci. 60: 1552-1574.

Eckert, T.H.  2004. Summary of 1976-2003 Warm Water Assessment. In New York State
Department of Environmental Conservation. Lake Ontario Annual Report 2003. Bureau of
Fisheries, Lake Ontario Unit and St. Lawrence River Unit. Cape Vincent and Watertown, NY.

Harkness, W.J., and Dymond, J.R. 1961. The lake sturgeon: The history of its fishery and
problems of conservation. Ontario Dept. of Lands and Forests, Fish and Wildl. Branch. 120 pp.

Hay-Chmielewski, E.M., and Whelan, G.E. 1997. Lake sturgeon rehabilitation strategy. Michigan
Department of Natural Resources Fisheries Division, Special Report Number 18, Ann Arbor, MI.

Hill T.D., and McClain, J.R. (eds.). 2004. 2002 Activities of the Central Great Lakes Binational
Lake Sturgeon Group. U.S. Fish and Wildlife Service, Alpena, MI.

Holey,  M.E., Baker, E.A., Thuemler, T.F., and Elliott, R.F. 2000. Research and assessment needs
to restore Lake Sturgeon in the Great Lakes: results of a workshop sponsored by the Great Lakes
Fishery Trust. Lansing, MI.

Manny, B.A., and Kennedy, G.W. 2002. Known lake sturgeon (Acipenser fulvescens) spawning
habitat in the channel between Lakes Huron and Erie in the Laurentian Great Lakes. J. Applied
Ichthyology  18:486-490.

Ontario Ministry of Natural Resources. 2004. Lake Ontario Fish Communities and Fisheries:
2003 Annual Report of the Lake Ontario Management Unit. Ontario Ministry of Natural
Resources, Picton, ON.

Schneeberger, P.J. Elliott, R.F., Jonas, J.L. and Hart, S.  2005. Benthivores. In The state of Lake
Michigan in  2000. Edited by M.E. Holey and T.N. Trudeau. Great Lakes Fish. Comm. Spec. Pub.
05-01, pp. 25-32.

-------



Thomas, M.V., and Haas, R.C. 2002. Abundance, age structure, and spatial distribution of lake
sturgeon, Acipenser fulvescens, in the St. Clair system. J. Applied Ichthyology 18: 495-501.

Zollweg, B.C., Elliott, R.F., Hill, T.D., Quinlan, H.R., Trometer, E., and Weisser, J.W. (eds.).
2003. Great Lakes Lake Sturgeon Coordination Meeting. In Proceedings of the December 11-12,
2002 Workshop, Sault Ste. Marie, MI.

List of Tables
Table 1. Trends in juvenile lake sturgeon CPE during June in Lake Superior near the mouth of the
Bad River.

List of Figures
Figure 1. Historic lake sturgeon harvest from each of the Great Lakes.
Source: Baldwinsal. 1979

Figure 2. Historic distribution of lake sturgeon.
Source: Zollweg et al. 2003

Figure 3. Current distribution of lake sturgeon.
Source: Zollweg et al. 2003

Last updated
SOLEC 2006
                         Draft for Discussion at SOLEC 2006

-------
                             State of the Great Lakes 2007 - Draft
     ilKKI
       1870
                                                                  1970
Figure 1. Historic lake sturgeon harvest from each of the Great Lakes.
Source: Baldwin*??a/. 1979
Year
1994
1995
1996
1997
1998
1999
2000
Month
6
6
6
6
6
6
CPE
0.333333
1
0.714286
1.142857
1.769231
2.5
6 | 2.25
2001 6 4.5
2002 6 5.5
Table 1. Trends in juvenile lake sturgeon CPE during June in Lake Superior near the mouth of
the Bad River.
                      Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
  0 62 5125  250   375   500
Figure 2. Historic distribution of lake sturgeon.
Source: Zollweg et al. 2003
                                                                   population status
                                                                    • • rfnt
                                                                    • bigg
                                                                    • runn
                                                                      unSmwnftnwd
Figure 3. Current distribution of lake sturgeon.
Source: Zollweg et al. 2003
                       Draft for Discussion at SOLEC 2006

-------
Commercial/Industrial Eco-Efficiency Measures
Indicator #3 514

Assessment: Not Assessed

Purpose
  To assess the institutionalized response of the commercial/
industrial sector to pressures imposed on the ecosystem as a
result of production processes and service delivery.

Ecosystem Objective
The goal of eco-efficiency is to deliver competitively priced
goods and services that satisfy human needs and increase quality
of life, while progressively reducing ecological impacts and
resource intensity throughout the lifecycle, to a level at least in
line with the earth's estimated carrying capacity (WBCSD
1996). In quantitative terms, the goal is to increase the ratio of
the value of output(s) produced by a firm to the sum of the
environmental pressures generated by the firm (OECD  et al.
1998).

State of the Ecosystem
Background
This indicator report for eco-efficiency is based upon the public
documents produced by the 24 largest employers in the basin
which report eco-efficiency measures and implement eco-effi-
ciency strategies. The 24 largest employers were selected as
industry leaders and as a proxy for assessing commercial/indus-
trial  eco-efficiency measures. This indicator should not be con-
sidered a comprehensive evaluation of all the activities of the
commercial/industrial sector, particularly small-scale organiza-
tions, though it is presumed that many other industrial/commer-
cial organizations are implementing and reporting on similar
strategies.

Efforts to track eco-efficiency in the Great Lakes basin and in
North America are still in the infancy stage. This is the first
assessment of its kind in the Great Lakes region. It includes 24
of the largest private employers, from a variety of sectors, oper-
ating in the basin. Participation in eco-efficiency was tabulated
from publicly available environmental reporting data from 10
Canadian companies and 14 American companies based in (or
with major operations in) the Great Lakes basin.

Tracking of eco-efficiency indicators is based on the notion that
what is measured is what gets done. The  evaluation of this indi-
cator is conducted by recording presence/absence of reporting
related to performance in seven eco-efficiency reporting cate-
gories (net sales, quantity of goods produced, material consump-
tion, energy consumption, water consumption, greenhouse gas
emissions, emissions of ozone depleting  substances (WBCSD
2002)). In addition, the evaluation includes an enumeration of
146
                                                                                           2007
  specific initiatives that are targeted toward one or more of the
  elements of eco-efficiency success (material intensity, energy
  intensity, toxic dispersion, recyclability and product durability
  (WBCSD 2002)).

  State of Eco-Efficiency
  Of the 24 companies surveyed, 10 reported publicly (available
  online or through customer service inquiry) on at least some
  measures  of eco-efficiency. Energy consumption and, to some
  extent, material consumption were the most commonly reported
  measures. Of the 10 firms that reported on some elements of
  eco-efficiency, three reported on all seven measures.
  Of the 24 companies surveyed, 19 (or 79%) reported on imple-
  mentation of specific eco-efficiency related initiatives. Two corn-
     Energy Consumption    Materials     Water Consumption   GHG Emissions
                   Consumption
                Eco-Efficiency Measure (based on WBCSD measures)
Ozone depleting
  emissions
Figure 1. Number of the 24 largest employers in the Great Lakes basin
that publicly report eco-efficiency measures. GHG = green house gas.
Source: WBCSD = World Business Council for Sustainable
Development
  panies reported activities related to all five success areas.
  Reported initiatives were most commonly targeted toward
  improved recycling and improved energy efficiency.

  Overall, companies in the manufacturing sector tended to pro-
  vide more public information on environmental performance
  than the retail or financial sectors. At the same time, nearly all
  firms expressed a commitment to reducing the environmental
  impact of their operations. A select number of companies, such
  as Steelcase Inc. and General Motors in the U.S. and Nortel
  Networks in Canada, have shown  strong leadership in compre-
  hensive,  easily accessed, public reporting on environmental per-
  formance. Others, such as Haworth Inc. and Quad/Graphics,
  have shown distinct creativity and innovation  in implementing
  measures to reduce their environmental impact.

-------
The concept of eco-efficiency was defined in 1990 but was not
widely accepted until several years later. Specific data on com-
      Material intensity    Energy intensity   Toxic dispersion    Recyclability

                   Sucess Criteria (as defined by WBCSD)
                                                 Product durability
Figure 2. Number of the 24 largest employers in the Great Lakes
basin that publicly report initiatives related to eco-efficiency
success criteria.
Source: WBCSD = World Business Council for Sustainable
Development
mercial/industrial measures are only just being implemented,
therefore it is not yet possible to determine trends in eco-effi-
ciency reporting. In general, firms appear to be working to
improve the efficiency of their goods and service delivery. This
is an important trend as it indicates the growing ability of firms
to increase the quantity/number of goods and services produced
for the same or a lesser quantity of resources per unit of output.

While one or more eco-efficiency measures are often included in
environmental reporting, only a few firms recognize the com-
plete eco-efficiency concept. Many firms recognize the need for
more  environmentally sensitive delivery of goods and services;
however, the implementation of more environmentally efficient
processes appears narrow in scope. These observations indicate
that more could be done toward more sustainable goods and
services delivery.

Pressures
Eco-efficiency per unit of production will undoubtedly increase
over time, given the economic, environmental and public rela-
tions incentives for doing so. However,  as Great Lakes popula-
tions and economies grow, quantity of goods and services pro-
duced will likely increase.  If production increases by a greater
margin than eco-efficiency improvements, then the overall com-
mercial / industrial environmental impact will continue to rise.
Absolute reductions in the sum of environmental pressures are
necessary to deliver goods and services  within the earth's carry-
ing capacity.
Management Implications
The potential for improving the environmental and economic
efficiency of goods and services delivery is unlimited. To meet
the ecosystem objective, more firms in the commercial / indus-
trial sector need to recognize the value of eco-efficiency and
need to monitor and reduce the environmental impacts of pro-
duction.

Acknowledgments
Author: Laurie Payne, LURA Consulting, Oakville, ON.
Contributors: Christina Forst, Oak Ridge Institute for Science
and Education, on appointment to U.S. Environmental Protection
Agency, Great Lakes National Program Office; and Dale
Phenicie & George Kuper, Council of Great Lakes Industries.
Tom Van Camp and Nicolas Dion of Industry Canada provided
several data resources.
Many of the firms surveyed in this report also contributed envi-
ronmental reports and other corporate information. Chambers of
commerce in many states and provinces around the Great Lakes
provided employment data.

Sources
InfoUSA®, Omaha, NE. Largest Employers Database. 2001
www.acinet.org. employers.database@infoUSA.com.

Organization for Economic Cooperation and Development
(OECD), Environment Policy Committee, Environment
Directorate.  1998. Eco-Efficiency: Environment Ministerial
Steering Group Report. Paris, France.

Report on Business Magazine. 2002. The TOP 1000 2002: 50
Largest Employers, http://top 1000.robmagazine.com, last
accessed July 1, 2002.

Stratos: Strategies to Sustainability in collaboration with Alan
Willis and Associates and SustainAbility. 2001. Stepping
Forward: Corporate Sustainability Reporting in Canada.

Vrooman Environmental Inc. and Legwork Environmental Inc.
for Industry Canada. 2001. The Status of Eco-Efficiency and
Indicator Development in Canadian Industry. A Report on
Industry Perceptions and Practices.

World  Business Council on Sustainable Development
(WBCSD). 2000. Eco-efficiency: creating more value with less
impact.

World  Business Council on Sustainable Development
(WBCSD). 2000. Measuring eco-efficiency: A guide to reporting
company performance.
World Business Council on Sustainable Development
                                                                                                                     147

-------

(WBCSD). 1996. Eco-efficient Leadership for Improved
Economic and Environmental Performance. Geneva,
Switzerland.

National Round Table on Environment and Economy, Ottawa.
1999. Measuring eco-efficiency in business: feasibility of a core
set of indicators.
Authors' Commentary
By repeating this evaluation at a regular interval (i.e. every 2 or
4 years), trends in industrial / commercial eco-efficiency can be
determined. The sustainability of goods and service delivery in
the Great Lakes basin can only be determined if social justice
measures are also included in commercial/industrial sector
assessments. The difficulty in assessing the impacts of social jus-
tice issues precludes them from being included in this report,
however, such social welfare impacts should be included in
future indicator assessment.

Last Updated
State of the Great Lakes 2003
                                                                                          2007
148

-------

Drinking Water Quality
Indicator #4175
Overall Assessment
           Status:  Good
           Trend:  Unchanging
   Primary Factors
      Determining
  Status and Trend
Based on the information provided in the annual CC/WQRs and the Ontario
annual reports from the DWSs, the overall quality of the finished drinking
water in the Great Lakes Basin can be considered good. Because very few
violations of federally, provincially, or state regulated MCLs, MACs, or
treatment techniques occurred, the WTPs/DWSs are, in fact, employing
treatment techniques that are successfully treating water. As such, the
potential risk of human exposure to the noted chemical and/or
microbiological continents, and any associated health effects, is generally
low.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Assessed
Undetermined
Not available at this time.
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Assessed
Undetermined
Not available at this time.
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Assessed
Undetermined
Not available at this time.
Not Assessed
Undetermined
Not available at this time.
Lake Ontario
           Status:
Not Assessed
                        Draft for Discussion at SOLEC 2006

-------

           Trend:   Undetermined
   Primary Factors   Not available at this time.
      Determining
  Status and Trend

Purpose
• To evaluate the chemical and microbial contaminant levels in source water and in treated water;
and
• To assess the potential for human exposure to drinking water contaminants and the
effectiveness of policies and technologies to ensure safe drinking water.

Ecosystem Objective
The ultimate goal of this indicator is to ensure that all drinking water provided to the residents of
the Great Lakes basin is protected at its source, and treated in such a way that it is safe to drink
without reservations. As such, the treated water should be free from harmful chemical and
microbiological contaminants. This indicator supports Great Lakes Quality Agreement Annexes
1,2, 12, and 16.

State of the Ecosystem
Background
The information provided by the United States for this report focuses mainly on finished, or
treated, drinking water. This format was chosen as the focus for U.S. reporting in order to adapt
to the recommendations of the Environmental Health Indicator Project
(www.cdc.gov/nceh/indicators/default.htm). Additionally, the U.S. is in the process of
establishing an inclusive national drinking water database, which will include raw, or source
water data, thus providing an extensive array of information to all WTPs/D WSs, researchers,  and
the general public.  The information provided by Canada focuses on both finished and raw, or
source, water.

In the U.S., the Safe-Drinking Water Act Re-authorization of 1996 requires  all drinking water
utilities to provide yearly water quality information to their consumers. To satisfy this obligation,
U.S. Water Treatment Plants (WTPs) produce an annual Consumer Confidence/Water Quality
Report (CC/WQR). These reports provide information regarding: source water type (i.e. lake,
river or groundwater), the water treatment process, contaminants detected in the finished water,
any violations  that occurred, and other relevant information. For this indicator report the
CC/WQRs were collected from 59 WTPs for the operational year 2004 (2005 when available).
Furthermore, the U.S. based  Safe Drinking Water Information System (SDWIS) was also used as
a means to verify information presented in the reports and to provide any other relevant
information, where CC/WQRs were not yet available.

The data used  for the Canadian component of the report were provided by the Ontario Ministry of
the Environment and include results from two program areas. Data collected as part of the
Drinking Water Surveillance Program (DWSP) was provided for the period 2001/2002. DWSP  is
a voluntary partnership program with municipalities that monitors drinking water quality.
Ontario's Drinking Water Systems Regulation (O. Reg. 170/03), made under the Safe Drinking
Water Act, 2002, requires that the  owner of a Drinking Water Systems (DWS) prepare an annual
                         Draft for Discussion at SOLEC 2006

-------


report on the operation of the system and the quality of its water. DWSs must provide the Ontario
Ministry of the Environment (OMOE) with their drinking water quality data. Data from January
to June 2004, collected as part of this regulatory framework from 74 DWSs, were also provided
for analysis.

There are several sources of drinking water within the Great Lakes basin which include; the Great
Lakes themselves, smaller lakes/reservoirs, rivers, streams, ponds, and groundwater i.e. springs
and wells.  However, these systems are vulnerable to contamination from several sources
(chemical, biological, and radioactive).  Substances that may be present in the source water
include: microbial contaminants, such as viruses and bacteria; inorganic contaminants, such as
salts and metals; pesticides and herbicides; organic chemical contaminants, including synthetic
and volatile organic chemicals; and radioactive contaminants. After collection, the raw water
undergoes a detailed treatment process prior to being sent to the distribution system where it is
then dispersed to consumer taps. The treatment process involves several basic steps, which are
often varied and repeated depending on the condition of the source water.  It is important to note
that raw water can also affect the finished water that is consumed. Good quality raw water is an
important part of a multi-barrier approach to assuring the safety and quality of drinking water.

Status of Drinking Water in the Great Lakes Basin
Ten drinking water parameters were chosen to provide the best assessment of drinking water
quality in the Great Lakes Basin, which include several chemical parameters, microbiological
parameters, and other indicators of potential health hazards. These parameters are regulated by
an established standard, which when exceeded, has the potential to have serious affects on human
health. The U.S. Environmental Protection Agency (USEPA) defines this regulated standard as
the Maximum Contaminant Level  (MCL), or the highest level of a contaminant that is allowed in
drinking water. The Ontario drinking water standards are described by the Maximum Acceptable
Concentration (MAC), which is established for parameters that when present above a certain
concentration, have known or suspected health effects, and the Interim Maximum Acceptable
Concentration (IMAC), which is established for parameters either when there is insufficient
toxicological data to establish a MAC with reasonable certainty, or when it is not feasible, for
practical reasons, to establish a MAC at the desired level.

Chemical Contaminants
The chemical contaminants of concern include; atrazine, nitrate, and nitrite.  Exposure to these
contaminants above the regulated standards has the potential to negatively affect human health.

Atrazine-Atrazine, which has been widely used as an organic herbicide, can enter source water
though agricultural runoff and/or wastewater from manufacturing facilities. Consumption of
drinking water that contains atrazine in excess of the regulated standard, for extended periods  of
time, can potentially lead to health complications. The USEPA has set the MCL for atrazine at 3
parts per billion (ppb) and the Ontario Drinking water standards specify the IMAC to be 5 ppb,
which is the lowest level at which  WTPs/DWSs could reasonably be required to remove this
contaminant given the present technology and resources.

In the U.S., atrazine was infrequently detected in finished water supplies, and was only found  in
finished water originating from Lake Erie, rivers, and small lakes/reservoirs. However when
detected, it was found at levels that did not exceed the MCL. Violations of monitoring
                         Draft for Discussion at SOLEC 2006

-------

requirements were reported for two WTPs for failure to monitor atrazine and other contaminants
between February and June 2004 and during July 2004, respectively. Therefore, as indicated by
the annual CC/WQRs there is a low risk of human exposure to atrazine.

In Ontario, data from the 2003/2004 DWSP indicated that 22 percent of the water samples
collected had trace amount of atrazine present. However, the highest level detected was only 0.59
ppb (about one order of magnitude less than the IMAC), which was identified from a raw water
source located within an agricultural watershed.

Nitrogen-Nitrogen is a naturally occurring nutrient that is also used in many agricultural
applications. However, in natural waters most nitrogenous material tends to be converted into
nitrates, which when ingested at levels exceeding the MCL or MAC can cause serious health
effects, particularly to infants.  The USEPA has set the MCL for nitrate at 10 parts per million
(ppm) and nitrite at 1 ppm and the province of Ontario has set the MAC for nitrate at 10 ppm and
nitrite at 1 ppm.

In the U.S., nitrate was detected in over 70 percent of the finished water supplies which
originated from WTPs using  all sources of water except Lake Huron. However, it was never
found at levels that exceeded the MCL and therefore, while there is some risk of exposure to
nitrate, it is not likely to lead to serious health complications.

In Ontario, over 90 percent of the of the water samples contained nitrates; however, the highest
level detected was 9.11 ppm, from a raw ground water sample.  As such, there is a risk of
exposure to nitrates, especially in agricultural areas, but it is not likely to cause health
complications as detected levels never exceeded the Ontario contamination standard.

In the U.S., nitrite was rarely detected in finished water supplies.  It was only found in finished
water for WTPs which use rivers and small lakes/reservoirs as source water. As such, there is
only a small potential for human exposure to nitrite from drinking water. No MCL or monitoring
regulation violations were reported for nitrites.

Over fifty percent of the water samples contained a measurable amount of nitrite according to the
Ontario drinking water system reports. However, the highest value for this contaminant only
reached 0.365 ppm, which is  lower than the Ontario MAC and the highest value detected last year
(0.434 ppm).

Microbiological Parameters
The microbiological parameters evaluated include total coliform, Escherichia coli (E. coli),
Giardia, and Cryptosporidium. These microbial contaminants are included as indicators of water
quality, but also as an indication of the presence of hazardous and possibly fatal pathogens in the
water.

Total Co///brw-Coliforms are a broad class of bacteria that are ubiquitous in the environment and
in the feces of humans and animals.  The USEPA has set a MCL for total coliform at 5% of the
total monthly samples (e.g. for water systems that collect fewer than 40 routine samples per
month, no more than one sample can be total coliform-positive per month). Canada has set an
                         Draft for Discussion at SOLEC 2006

-------

MCL of 0 colony forming units (CPU) for DWSs.  Both Canada and the U.S. require additional
analysis of positive total coliform samples to determine if specific types of coliform, such as fecal
coliform or E coli, are present.

Escherichia coli (E. coli)-E. coli is a type of thermo tolerant (fecal) coliform bacteria that is
generally found in the intestines, and fecal waste, of all animals, including humans. This type of
bacteria commonly enters source water through contaminated runoff, which is often the result of
precipitation. Detection of E. coli in water strongly indicates recent contamination of sewage  or
animal waste, which may contain many types of disease-causing organisms. It is mandatory for
all WTPs to inform consumers if E. coli is present in their drinking and/or recreational water
(U.S. waters only).

In the U.S., the presence of total coliform was detected in finished water from WTPs using all
source water types, except Lake Superior.  It was repeatedly detected in finished water from
WTPs using Lake Michigan, groundwater, rivers, and small lakes/reservoirs as source water.
Between July 2004 and October 2005, there were four violations with regard to total coliform
levels exceeding the MCL. As such, repeat samples were collected at the same  locations as the
positive total coliform bacteria sample and at nearby locations to determine if the original positive
sample indicated a localized water problem, or a sampling or testing error.  However, samples
from two of these WTPs tested positive for either fecal coliform or E. coli.  Additionally,
violations of monitoring requirements of USEPA's Total Coliform Rule (TCR)  were reported in
one WTP, for not collecting enough repeat samples after coliform bacteria was detected in the
monthly routine  samples. Although there is a potential for human exposure to total coliform, it is
not likely to be a human health hazard in itself.  However, the presence of coliform bacteria,
especially at levels exceeding the MCL, indicates the possibility that microbial pathogens may be
present, and this can be hazardous to human health.

In Ontario, total  coliform was detected in many of the raw water samples; however only a few
treated water samples contained this contaminant.  Furthermore, E.  coli was identified in raw
water samples, which originating mostly from small lakes and rivers, in small amounts.
However,  the presence of E. coli was not identified in finished water supply, indicating that the
treatment facilities are working adequately to remove both of these microbiological parameters.

Giardia and Cryptosporidium- Giardia and Cryptosporidium are parasites that exist in water and
when ingested may cause gastrointestinal illness in humans.  The U.S. treated water standards,
which controls the presence of these microorganisms in the treated water, dictate that 99% of
Cryptosporidium should be physically removed by filtration. In addition, Giardia must be 99.9%
removed and/or inactivated by filtration and disinfection.  These regulations are confirmed by the
levels of post treatment turbidity and disinfectant residual levels. Ontario has also  adopted
removal/inactivation for Giardia and Cryptosporidium, however, there is no data to report at this
time.

In the U.S., neither Giardia nor Cryptosporidium were detected in finished water supplies from
any of the WTPs. However,  several of the CC/WQRs discussed the presence of these
microorganisms  in the source waters (Lake Erie, Lake Huron, Lake Michigan, Lake Ontario,
small lakes/reservoirs). The presence of these organisms in raw water but not in finished water
indicates that current treatment techniques are effective at removing these parasites from drinking
                         Draft for Discussion at SOLEC 2006

-------
water.  Nevertheless, implementing measures to prevent or reduce microbial contamination from
source waters should remain a priority. Even a well-operated WTP cannot ensure that drinking
water will be completely free of Cryptosporidium. Furthermore, very low levels of
Cryptosporidium may be of concern for the severely immuno-compromised because exposure can
compound their illness.

The annual CC/WQRs indicate that there is a potential for consumers to be exposed to the
aforementioned microbiological contaminants.  However, total coliform was the most common
microbiological contaminant detected.  Furthermore, there were very few if any confirmed
detections of the more serious contaminants including, E. coli,  Giardia, and Cryptosporidium, in
the finished water of the U.S.. As a result, it is not likely that consumption of drinking water
containing these contaminants will lead to any serious health complications.

Treatment Technique Parameters
The treatment technique parameters evaluated include turbidity, total organic carbon (TOC) in the
U.S. and dissolved organic carbon (DOC) in Canada.  These parameters do not pose a direct
danger to human health but often indicate other health hazards.

7wr&/W/^-Turbidity is a measure of the cloudiness of water and can be used to indicate water
quality and filtration efficiency. Higher turbidity levels, which  can inhibit the effectiveness of the
disinfection/filtration process and/or provide a medium for microbial growth, are associated with
higher levels of disease-causing microorganisms such as viruses, parasites and some bacteria. A
significant relationship has been demonstrated between increased turbidity and the number of
Giardia cysts and Cryptosporidium oocysts breaking through filters. USEPA's surface water
treatment rules require WTPs using surface water or ground water under the direct influence of
surface water must disinfect and filter their water.  In the U.S.,  turbidity levels must not exceed 5
Nephelolometric Turbidity Units (NTU) at any time, while WTPs that filter must ensure that the
turbidity go no higher than 1 NTU, and must not exceed 0.3 NTU in 95% of daily samples in any
month.  Ontario has set the aesthetic objective for turbidity at 5.0 NTU, at which point turbidity
becomes visible to the naked eye.

In the U.S., turbidity data is difficult to assess due to the different requirements and regulations
for WTPs  depending on the source water and treatment technique used. However, there were no
MCL or monitoring regulations violations reported from January 2004 to October 2005.

In Ontario, the 2003/2004 DWSP report indicated that 78 raw water samples, many of which
originated from Lake St. Clair and the Detroit River, exceeded  the aesthetic objective.
Furthermore, one treated water sample  exceed the aesthetic objective with a turbidity level of
11.1  NTU.

Total Organic Car&cw-Although the presence of total organic carbon (TOC) in water does not
directly imply a health hazard, the organic  carbon can react with chemical disinfectants to form
harmful byproducts. WTPs remove TOC from the water by using treatment techniques such as
enhanced coagulation or enhanced softening. Conventional WTPs with excess TOC in the raw
water are required to remove a certain percentage of the TOC depending upon the TOC and the
alkalinity level of the raw water. The USEPA does not have a MCL for TOC.
                         Draft for Discussion at SOLEC 2006

-------


In the U.S., TOC was detected in finished water from WTPs using all source water types, except
Lake Superior.  However, TOC data was difficult to assess due to the varying formats of
CC/WQRs and the way data was presented.  As such, it was difficult to quantitatively evaluate
and compare the TOC levels reported by each WTP.  Violations of monitoring requirements
and/or failure to report the results were reported for one WTP from July to September 2005.

Dissolved Organic Car&ow-Dissolved organic carbon (DOC) can indicate the potential possibility
of water deterioration during storage and distribution. Acting as a growth nutrient, increased
levels of carbon can aid in the proliferation of biofilm, or microbial cells that attach to the surface
of pipes and multiply to form a layer of film or slime on the pipes, which can harbor and protect
coliform bacteria from disinfectants.  High DOC levels can also indicate the potential of
chlorination by-products problems. The use of coagulant treatment or high pressure membrane
treatment can be used to reduce DOC. The aesthetic  objective for DOC in Ontario's drinking
water is 5 ppm.

In Ontario, there were 110 DOC violations, 11.4 ppm being the highest level, identified  from raw
water sample; however, no treated water sample contained DOC levels exceeding the  aesthetic
objective.  Most of the high DOC results came from raw water originating from small rivers and
lakes.

Taste and Odor
While taste and odor do not necessarily reflect any health hazards, these water characteristics
affect the consumer perception of the drinking water  quality.

In the U.S., there were no reports of offensive taste or odors associated with the finished drinking
water as indicated by the 2005 CC/WQRs.

In Ontario, there has  been an increase in the number of reports associated with offensive taste and
odor over the past several years; however, specific data is unavailable as it is difficult to
quantitatively evaluate and compare results. Many drinking-water systems have now installed
granular activated carbon filters to decrease the effect and intensity of these taste and odor events,
which are due, in part, to the increased decomposition of blue-green algae in the Great Lakes
(Ministry of Environment, 2004).

Summary
Based on the information provided in the annual CC/WQRs and the Ontario annual reports from
the DWSs, the overall quality of the finished drinking water can be considered good. However,
over the past several  years there has been an increase in the quantity of contaminants found in
raw source water in the Great Lakes Basin. The overall potential risk of human exposure to the
noted chemical and/or microbiological continents, and any associated health effects, is generally
low as very few violations of federally, provincially,  or state regulated MCLs, MACs, or
treatment techniques occurred. This indicates that the WTPs/DWSs are employing treatment
techniques that are successfully treating water

Pressures
                         Draft for Discussion at SOLEC 2006

-------
The greatest pressure to the quality of drinking water within the Great Lakes Basin would be
degraded runoff. Several causes for this reduction in quality would including; the increasing rate
of industrial development on or near water bodies, low-density urban sprawl, and agriculture -
both crop and livestock operations. Point source pollution, from wastewater treatment plants for
example, can also contribute to the contamination of raw water supplies and therefore can be
considered an important pressure as well.  Additionally, there is an emerging  set of pressures such
as newly introduced chemicals, chemicals of emerging concern (i.e. pharmaceuticals and personal
care products (PPCPs), endocrine disrupters, antibiotics and antibacterial agents) and invasive
species which might affect water quality; however to what extent is still unknown.

Management Implications
A more standardized, updated approach to monitoring contaminants and reporting data for
drinking water needs to be established. Even though the USEPA has established an extensive list
of contaminants, and their MCLs, newer parameters of concern might not be listed due to
available resources or technology. Additionally, state monitoring requirements  may differ;
requiring only a portion of this list to be monitored.  This would make trend analysis easier, and
thus provide a more effective assessment of the potential health hazards associated with drinking
water.

Furthermore, a more extensive monitoring program must be implemented in order to successfully
correlate drinking water quality with the status of the Great Lakes Basin. Although the
CC/WQRs provide useful information regarding the quality of finished drinking water, they
merely depict the efficiency of the WTP,  rather than the overall quality of the region.
Additionally, by solely focusing on treated water, WTPs that rely on several type of source water
will not provide accurate data with regard to contaminant origin.  Therefore, in order to properly
assess the  state of the ecosystem, source water data would need to be reviewed.

Another concern for future efforts would be the adherence of a consistent guideline when
identifying usable data; a guideline that obtains sufficient data while also providing adequate
geographical coverage. In the U.S., data  from WTPs serving a population of 50,000 or great was
used, while data from all DWSs in Ontario serving a population of 10,000 or greater was
analyzed.  Furthermore, focusing on this criterion for WTPs only provides a fragmented view of
the drinking water patterns in the Great Lakes Basin; however by sporadically including
additional WTPs to expand the geographical coverage area, bias results may be  introduced.

In addition to raw and treated water, some effort should also be made to analyze distributed
water.  Even though there are numerous precautions in place to ensure the quality of finished
water, contamination is also possible during the distribution stage. Corrosion of copper or lead
pipes and/or bacterial growth within these pipes could affect the overall quality  of drinking water.
Even though WTPs/DWSs are implementing actions to prevent or hinder such contamination,
without sufficient data from distributed water supplies it is impossible to determine whether these
efforts are effective or need to be altered.

Acknowledgments
Authors: Jeffrey C. May, U.S. Environmental Protection Agency, GLNPO Intern.  77 W.
Jackson Blvd (G-17J) Chicago, Illinois 60604, May.Jeffrey@epa.gov
                         Draft for Discussion at SOLEC 2006

-------

Tracie Greenberg, Environment Canada, Intern. Burlington, Ontario

Data Sources
Guillarte, A., and Makdisi, M. 2003 Implementing Indicators 2003, A Technical Report.
Environmental Canada and U.S. Environmental Protection Agency.
http://binational.net/sogl2003/sogl03  tech  eng.pdf

Fellowes, D. Personal Communication. Ontario 2003/2004 Drinking Water Surveillance
Program (DWSP) Data, Environmental Monitoring and Reporting Branch, Environmental
Sciences and Standards Division, Ontario Ministry of the Environment.

Ontario  Ministry of the Environment. 2006 (revised from 2003). Technical Support Document for
Ontario Drinking  Water: standards, objectives, and guidelines. Ontario Ministry of the
Environment.

Ontario  Ministry of the Environment. 2004. Drinking Water Surveillance Program Summary
Report. Taken on August 25, 2006 from
http://www.ene.gov.on.ca/envision/water/dwsp/0002/index.htm

Consumer Confidence Reports
Akron Public Utilities Bureau - Annual Drinking Water Quality Report for 2005
Alpena Water Treatment Plant - 2005 Annual Drinking Water Quality Report
Aqua Ohio, Inc. PWS - 2005 Water Quality Report
Aqua Ohio - Mentor - 2005 Water Quality Report
Buffalo  Water Authority - 2005 Annual Water Quality Report
City of Ann Arbor Water Utilities - 2005 Annual Report on Drinking Water
City of Battle Creek Public Works - 2005 Annual Water Quality Report
City of Cleveland Division of Water - 2006 Water Quality Report
City of Duluth Public Works and Utilities Department - 2005 Guide to Drinking Water Quality
City of Evanston - 2005 Water Quality Report
City of Kalamazoo - 2005 Water Quality Report
City of Kenosha Water Utility - 2005 Annual Drinking Water Quality Report
City of Marquette Water Filtration Plant - 2005 Annual Drinking Water Quality Report
City of Muskegon Water Filtration Plant - 2005 Annual Water Quality Report
City of Oshkosh - Drinking  Water Quality Report 2005
City of Rochester - Water Quality Report 2005
City of Syracuse Department of Water - Annual Drinking Water Quality Report for 2005
City of Toledo Water Treatment Plant - 2005 Drinking Water Quality Report
City of Warren - 2005 Water Quality Report
City of Waukegan - 2006 Water Quality Report
City of Wyoming - 2005 Water Quality Report
Department of Utilities Appleton Water Treatment Facility - 2005 Annual Water Quality Report
to our Community
Detroit Water & Sewer Department - 2005 Water Quality Report
Elmira Water Board - Annual Drinking Water Quality Report 2005
Elyria Water Department - 2005 Annual Water Quality Report
Erie County Water Authority - 2005 Water Quality Report
                         Draft for Discussion at SOLEC 2006

-------
Erie Water Works (EWW) - Water Quality Report for Year 2005
Fort Wayne City Utilities - 2006 Annual Drinking Water Quality Report
Grand Rapids Water System - Water Quality Report 2005
Green Bay Water Utility - 2006 Annual Drinking Water Quality Report
Indiana-American Water Company, Inc.  (Northwest Operations) - 2005 Annual Water Quality
Report
Lansing Board of Water & Light - 2005 Annual Water Quality Report
Michael C. O'Laughlin Municipal Water Plant - Annual Drinking Water Quality Report for 2005
Milwaukee Water Works - 2005 Water Quality Report
Mohawk Valley Water Authority - 2005 Water Quality Report
Monroe County Water Authority (MCWA) - 2005 Annual Water Quality Report
Onondaga County Water Authority (OCWA) - 2005 Consumer Confidence Report & Annual
Water Supply Statement
Port Huron Water Treatment Plant - 2005 Annual Drinking Water Quality Report
Saginaw Water Treatment Plant - Drinking Water Quality Report for 2005
South Bend Water Works -  Water Quality Report 2005
The City of Chicago - Water 2005 Quality Report
Waterford Township - 2005 Annual Water Quality Report
Waukesha Water Utility - 2005 Consumer Confidence Report
Waukesha Water Utility - 2006 Consumer Confidence Report

Last updated
SOLEC 2006
1C)
Draft for Discussion at SOLEC 2006

-------

Biological Markers of Human Exposure to Persistent Chemicals
Indicator #4177
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
At present, no routine Great Lakes human biomonitoring programs
exist to monitor biological markers of human exposure to persistent
chemicals.  Individual epidemiological studies have been conducted or
are on going in the Great Lakes to monitor specific populations. For
this reason, the status is mixed and no trends can be determined
regarding biological markers of human exposure.
Lake-by-Lake Assessment    No lake by lake assessments can be determined for this indicator.
                            Instead, a list of ongoing research funded by ATSDR's Great
                            Lakes Human Health Effects Research Program is provided
                            according to the institution conducting the research.
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
Mixed
Undetermined
No ATSDR studies are currently being funded by any institution in the
Lake Superior basin. However, basin wide studies do incorporate Lake
Superior information.
Mixed
Undetermined
•   Health Effects of PCB Exposure from Contaminated Fish (Susan L.
    Schantz, PhD  University of Illinois at Urbana-Champaign)
•   Organo-chlorides and Sex Steroids in two Michigan Cohorts (Janet
    Osuch, M.D., Michigan State University)
•   A Pilot Program to Educate Vulnerable Populations about Fish
    Advisories in Upper Peninsula of Michigan (Rick Haverkate, MPH,
    Inter-Tribal Council of Michigan, Inc.
Mixed
Undetermined
No ATSDR studies are currently being funded by any institution in the
Lake Huron basin. However, basin wide studies do incorporate Lake Huron
information.
Mixed
Undetermined
No ATSDR studies are currently being funded by any institution in the
Lake Erie basin. However, basin wide studies do incorporate Lake Erie
                        Draft for Discussion at SOLEC 2006

-------
  Status and Trend  information.

Lake Ontario
           Status:  Mixed
           Trend:  Undetermined
   Primary Factors  •  Neuropsychological and Thyroid Effects of PDBEs (Edward Fitzgerald,
      Determining     PhD, State University of New York at Albany)
  Status and Trend  •  PCB Congener and Metabolite Patterns in Adult Mohawks: Biomarkers
                      of Exposure and Individual Toxicokinetics (Anthony DeCaprio, PhD
                      State University of New York at Albany)
                   •  Neurobehavioral Effects of Environmental Toxics - Oswego Children's
                      Study: Prenatal PCB Exposure and Cognitive Development (Paul
                      Stewart, PhD., State University of New York at Oswego)

Purpose
•To assess the levels of persistent toxic substances such as methyl mercury, poly chlorinated
biphenyls (PCBs), and dichlorodiphenyl dichloroethenes (DDEs) in the human tissue of citizens
of the Great Lakes basin; and
•To infer the efficacy of policies and technology to reduce these persistent bioaccumulating toxic
chemicals in the Great Lakes ecosystem.

Ecosystem Objective
Citizens of the Great Lakes basin should be safe from exposure to harmful bioaccumulating toxic
chemicals found in the environment. Data on the status and trends of these chemicals should be
gathered to help understand how human health is affected by multimedia exposure and the
interactive effects of toxic  substances.  Collection of such data supports the requirement of the
Great Lakes Water Quality Agreement Annex 1 (Specific Objectives), Annex  12 (Persistent
Toxic Substances), and Annex 17 (Research and Development).

State of the Ecosystem
Women and Infant Child Study
Data presented for this indicator are solely based upon one biomonitoring study that Wisconsin
Department of Public Health (WiDPH) conducted in the basin. However, information on previous
biomonitoring studies has been collected and is highlighted as a way to support the results of the
WiDPH study and to illustrate previous and other ongoing efforts.

In the study conducted by WiDPH, the level of bioaccumulating toxic chemicals was analyzed in
women of childbearing age 18-45 years of age. Hair and blood samples were collected from
women who visited one of six participating Women Infant and Child (WIC) clinics located along
Lake Michigan and Lake Superior. Levels of mercury were measured in hair samples, and
mercury, PCBs,  and DDEs were measured in blood serum. Awareness offish consumption
advisories was assessed through a survey.

There was greater awareness of fish consumption advisories in households in which someone
fished compared to those in which no one did (Figure  1), and there was greater awareness of
advisories from individuals with at least a high school education compared to those with only
                        Draft for Discussion at SOLEC 2006

-------

some high school or less education (Figure 2). More women in the 36-45 age category were
aware of advisories than those of other ages, but there was less than 50% awareness in all age
classes (Figure 3). More Asian women were aware of advisories that those of other races, and
Hispanic women were least aware of the advisories (Figure 4).

Sixty-five hair samples were analyzed for mercury levels. The average mercury concentration in
hair from fish-eating women was greater than that from non-fish eaters, ranging from 128%
increase in women who ate few fish meals to 443% increase in those who ate several meals of
sport-caught fish (Table 1).

Five samples of blood were drawn and analyzed for PCBs, DDEs and mercury levels. Although
the small sample precludes definitive findings, the woman consuming the most fish (at least 1
sport-caught fish meal per week) had the highest concentration of DDE and the only positive
finding of PCB in her serum. The woman consuming the fewest fish per year (6-18 fish meals)
had the lowest concentration of DDE in her serum, and no PCBs were detected (Table 2).

Effects on Aboriginals of the Great Lakes (EAGLE) Project
A similar study was conducted by a partnership between the Assembly of First Nations, Health
Canada and First Nations in the Great Lakes basin between 1990 and 2000 to examine the effects
of contaminants on the health of the Great Lakes Aboriginal population. The Contaminants in
Human Tissues Program (CHT),  a major component of the EAGLE Project, identified three main
goals: To determine the levels of environmental contaminants in the tissues of First Nations
people in the Great Lakes basin; To correlate these levels with freshwater fish and wild game
consumption; and, To provide information and advice to First Nations people on the levels of
environmental contaminants found in their tissues.

The EAGLE project also analyzed hair samples for levels of mercury and blood serum for levels
of PCBs and DDEs. A survey was also used to identify frequency offish and wildlife
consumption. However, the EAGLE project analyzed both male and female voluntary
participants from 26 First Nations in the Great Lakes basin. The participants were volunteers, not
selected on a random basis, and the project did not specifically target only fish eaters.

Key findings of the study included:
•Males consumed more fish than females and carried greater contaminant levels;
•No significant relationship was found between total fish or wild game consumption and the
contaminant levels in the body;
•Levels of mercury in hair from First Nations people in the Canadian portion of the Great Lakes
basin suggest the levels have decreased since 1970;
•PCBs and DDE were the most frequently appearing contaminants in the serum samples;
•Increased age of participants correlated with increased contaminant concentrations;
•Mean levels of PCBs reported in the EAGLE CHT Program were lower than or within the
similar range of PCBs in fish-eaters in other Canadian health studies (Great Lakes,  Lake
Michigan, and St. Lawrence);
•Most people have levels of contaminants that were within Health Canada's guidelines for PCBs
in serum and mercury in hair;
•Levels of DDE were similar to levels found in other Canadian health studies; and
•There was little difference between serum levels of DDE in male and female participants.
                         Draft for Discussion at SOLEC 2006

-------
^4 TSDR-sponsored Studies
The Agency for Toxic Substances and Disease Registry (ATSDR) and the U.S. Environmental
Protection Agency (USEPA) established the Great Lakes Human Health Effects Research
Program through legislative mandate in September 1992 to "assess the adverse effects of water
pollutants in the Great Lakes system on the health of persons in the Great Lakes States" (ATSDR,
http://www.atsdr.cdc.gov/grtlakes/historical-background.html). This program assesses critical
pollutants of concern, identifies vulnerable and sensitive populations, prioritizes areas of research,
and funds research projects. Results from several recent Great Lakes biomonitoring research
projects are summarized here.

Data collected from 1980 to 1995 from Great Lakes sport fish eaters showed a decline in serum
PCB levels from a mean of 24 parts per billion (ppb) in 1980 to 12 ppb in  1995. This decline was
associated with an 83% decrease in the number offish meals consumed (Tee et al. 2003).

A large number of infants (2716) born between 1986 and 1991 to participants of the New York
State Angler Cohort Study were studied with respect to duration of maternal consumption of
contaminated fish and potential effects on gestational  age and birth size. The data indicated no
significant correlations gestational age or birth size in these infants and their mother's lifetime
consumption of fish. The researchers noted that biological determinants such as parity, and
placental infarction and maternal smoking were significant determinants of birth size (Buck et al.
2003).

The relationship between prenatal exposure to PCBs and methylmercury and performance  on the
McCarthy Scales of Children's Abilities was assessed in 212 children. Negative associations
between prenatal exposure to methylmercury and McCarthy performance were found in  subjects
with higher levels of prenatal PCB exposure at 38 months. However, no relationship between
PCBs and methylmercury and McCarthy performance was observed when the  children were
reassessed at 54 months. These results partially replicated the findings of others and suggest that
functional recovery may occur. The researchers concluded that the interaction between PCBs and
methylmercury can not be considered conclusive until it has been replicated in subsequent
investigations (Steward et al. 2003b).

Response inhibition in preschool children exposed parentally to PCBs may be  due to incomplete
development of their nervous system. One hundred and eighty-nine children in the Oswego study
were tested using a continuous performance test. The researchers measured the splenium of the
corpus callosum, a pathway in the brain implicated in  the regulation of response inhibition, in
these children by magnetic resonance imaging. The results indicated the smaller the splenium, the
larger the association between PCBs and the increased number of errors the children made  on the
continuous performance test. The researchers suggest  if the association between PCBs and
response inhibition is indeed causal, then children with suboptimal development of the splenium
may be particularly vulnerable to these effects (Stewart et al. 2003a).

Long term consumption of fish, even at low levels, contributes significantly to body burden levels
(Bloom et al. 2005).
                         Draft for Discussion at SOLEC 2006

-------


•American Indians were assessed for their exposure to PCBs via fish consumption by analysis of
blood samples and the Caffeine Breath Test (CBT). Serum levels of PCB congers #153, #170 and
#180 were significantly correlated with CBT values. CBT values may be a marker for early
biological effects of exposure to PCBs (Fitzgerald et al. 2005).

•Maternal exposure via fish consumption to dichlorodiphenyl dichloroethylene (DDE) and PCBs
indicated that only DDE was associated with reduced birth weight in infants (Weisskopf et al.
2005).

•The association between maternal fish consumption and the risk of major birth defects among
infants was assessed in the New York State Angler Cohort Study. The results indicated mothers
who consumed 2 or more fish meals per month had a significantly elevated risk for male children
being born with a birth defect (males: Odds Ratio = 3.01, in comparison to female children: Odds
Ratio = 0.73) (Mendola et al. 2005).

Pressures
Contaminants of emerging concern, such as certain brominated flame-retardants, are increasing in
the environment and may have negative health impacts. According to a recent study conducted by
Environment Canada, worldwide exposure to polybrominated diphenyl ethers (PBDEs, penta) is
highest in North America with lesser amounts in Europe and Asia. Food consumption is a
significant vector for PBDE exposure in addition to other sources. The survey analyzed PBDE
concentration in human milk by region in Canada in 1992 and in 2002 and showed a tenfold
increase in concentration in Ontario (Ryan 2004).

The health effects of contaminants  such as  endocrine disrupters are somewhat understood.
However, there  is little known about the synergistic or additive effects of bioaccumulating toxic
chemicals. Additional information about toxicity and interactions of a larger suite of chemicals,
with special attention paid to how bioaccumulating toxic chemicals work in concert, is needed to
better assess threats to human health from contaminants in the Great Lakes basin ecosystem.
ATSDR has developed 5 interaction toxicological profiles for mixtures of Volatile Organic
Compounds, metals, pesticides and for contaminants found in breast milk and fish.

Management Implications
There have been many small-scale  studies regarding human biomarkers and bioaccumulating
toxic chemicals. However, to this date, there have been no  large-scale or basin-wide studies that
can provide a larger picture of the issues facing the citizens of the basin. It is important that those
in management positions in Federal, State,  Provincial, and Tribal governments and universities
foster cooperation and collaboration to identify gaps in existing biomonitoring data and to
implement larger, basin-wide monitoring efforts. A Great Lakes environmental health tracking
program, similar to the Center for Disease Control  (CDC) Environmental Health Tracking
Program, should be established by key Great Lakes partners.

Comments from the author(s)
A region-specific biomonitoring program, similar to the CDC's National Health and Nutrition
Examination Survey (NHANES) project could provide needed biomonitoring  information and fill
in data gaps.
                         Draft for Discussion at SOLEC 2006

-------
It is important that additional studies assessing the levels of bioaccumulative toxic chemicals
through biomarkers be conducted on a much larger scale throughout the basin. In order to build
up on the WIC study it would be important for a question about fish consumption from
restaurants be included in future surveys. Because all states have WIC clinics, or something
similar, the WiDPH monitoring tool could be implemented basin-wide.

In the future, ATSDR's Great Lakes Human Health Effects Research Program plans to continue
to provide research findings to public health officials to improve their ability to assess and
evaluate chemical exposure in vulnerable populations. ATSDR also plans to focus on research
priorities of children's health, endocrine disrupters, mixtures, surveillance, and identification of
biomarkers, i.e., exposure, effect, and susceptibility. In addition, the program will use established
cohorts to monitor changes in body burdens of persistent toxic substances and specified health
outcomes, and develop and evaluate new health promotion strategies and risk communication
tools.

Acknowledgments
Authors: Elizabeth Murphy, U.S. Environmental Protection Agency, Great Lakes National
Program Office;
Jacqueline Fisher, U.S. Environmental Protection Agency, Great Lakes National Program Office;
Henry A. Anderson, Wisconsin Department of Health and Family Services;
Dyan Steenport, Wisconsin Division of Public Health;
Kate Cave, Environment Canada; and
Heraline E. Hicks, Agency for Toxic Substance and Disease Registry.

Data Sources
Anderson, H. 2004. SOLEC  Health Indicator Refinement and Implementation Progress Report.
Wisconsin Department of Health and Family Services. March, 22, 2004.

Bloom M.S., Vena, J.E., Swanson, M.K., Moysich, K.B., and Olsen, J.R. 2005. Profiles of ortho-
polychlorinated biphenyl congeners, dichlorodiphenyl dichloroethylene, hexachlorobenzene, and
mirex among male Lake Ontario sport fish consumers: the New York state angler cohort study.
Environ. Res. 97(2): 177-193.

Buck, G.M., Grace, P.T., Fitzgerald, E.F., Vena, J.E., Weiner, J.M., Swanson, M., and Msall,
M.E. 2003. Maternal fish consumption and infant birth size and gestation: New York state angler
cohort study. Environ. Health 2:7-15.

Davies, K., and Phil, D. 2001. EAGLE Project: Contaminants in human tissue. Health Canada,
Ottawa, ON.

Fitzgerald, E.F., Hwang, S.A., Lambert, G., Gomez, M., and Tarbell, A. 2005. PCB exposure and
in vivo CYP1A2 activity among Native Americans. Environ. Health Perspect. 113(3):l-6.

Mendola, P., Robinson, L.K., Buck, G.M., Druschel,  C.M., Fitzgerald, E.F., Sever, L.E., and
Vena, J.E. 2005. Birth defects associated with maternal sport fish consumption: potential effect
modification by sex of offspring. Environ. Res. 97:133-140.
                         Draft for Discussion at SOLEC 2006

-------
Ryan, JJ. 2004. Polybrominated Diphenyl Ethers (PBDEs) in Human Milk; Occurrence
Worldwide. Prepared for the 2004 BFR (Brominated Flame Retardants) conference. Health
Products and Food Branch, Health Canada, Toronto, ON.

Stewart P.W., Fitzgerald, S., Rehiman, J., Gump, B., Lonky, E., Darvill, T.J., Pagano, J., and
Hauser, P. 2003 a. Prenatal PCB exposure, the corpus callosum, and response inhibition. Environ.
Health Perspective 111:1670-1677.

Stewart, P.W., Reihman, J., Lonky, E.I., Daravill, T.J., and Pagano, J. 2003b. Cognitive
development in preschool children parentally exposed to PCBs and MeHg. Neurotoxicol. Teratol.
25:11-22.

Tee, P.O., Sweeney, A.M., Symanski, E., Gardiner, J.C., Gasior, D.M., and Schantz, S. 2003. A
longitudinal examination of factors related to changes in serum polychlorinated biphenyl levels.
Environ. Health. Perspect. 111(5):720-707.

Weisskopf, M.G., Anderson, H.A., Hanrahan, L.P., Kanarek, M.S., Falk, C.M., Steenport, D.M.,
Draheim, L.A., and the Great Lakes Consortium. 2005. Maternal exposure to Great Lakes  sport-
caught fish and dichlorodiphenyl dichloroethylene, but not polychlorinated biphenyls is
associated with reduced birth weight. Environ. Res. 97:149-162.

List of Tables
Table 1. Concentration of mercury in hair samples from women who consumed sport-caught or
not sport-caught fish during the previous three months. Source: Wisconsin Department of Health
and Family Services

Table 2. Number offish meals consumed and concentration of PCBs, DDE and mercury in blood
serum of 5 women who participated in the WIC study. Source: Wisconsin Department of Health
and Family Services

List of Figures
Figure 1. Percent of responders to the survey who are (red) or are not (yellow) aware offish
consumption advisories and who do (yes) or do not (no) have someone in the household who
fishes. Source: Wisconsin Department  of Health and Family Services

Figure 2. Percent of responders to the survey who are (red) or are not (yellow) aware offish
consumption advisories according to level of education. Source: Wisconsin Department of Health
and Family Services

Figure 3. Percent of responders to the survey who are (red) or are not (yellow) aware offish
consumption advisories according to age group. Source: Wisconsin Department of Health  and
Family Service

Figure 4. Percent of responders to the survey who are (red) or are not (yellow) aware offish
consumption advisories according to race.
Source: Wisconsin Department of Health and Family Service
                        Draft for Discussion at SOLEC 2006

-------
Last updated
SOLEC 2006
Fish meals/3 months
Sport-caught (Y/N)
0
1-9 (N)
1-9 (Y)
10+ (N)
10+ (Y)
Min
(UG/G)
0.00
0.04
0.03
0.04
0.09
Ave
(UG/G)
0.07
0.16
0.30
0.33
0.38
Max
(UG/G)
0.24
0.59
0.99
1.23
1.53
N
14
28
7
7
9
Ave no. fish
meals
0
2.3
2.4
12.8
8.11
Table 1. Concentration of mercury in hair samples from women who consumed sport-caught or
not sport-caught fish during the previous three months.
Source: Wisconsin Department of Health and Family Services
ID
100 Sheb
100 Sup
100AGB
105GB
101AGB
Fish Meals
Commercial = I/week
Sport Caught = none
Commercial = 5/month
Sport Caught = 30/year
Commercial =<6/Year
Sport Caught = 6-12/Year
Commercial = I/week
Sport Caught = I/week
Commercial = 4/month
Sport Caught = 2/month
PCB
0.0
0.0
0.0
0.4
0.0
DDE
0.34
0.40
0.25
1.20
0.49
Mercury
<5 mcg/L
<5 mcg/L
<5 mcg/L
<5 mcg/L
<5 mcg/L
Table 2. Number offish meals consumed and concentration of PCBs, DDE and mercury in blood
serum of 5 women who participated in the WIC study.
Source: Wisconsin Department of Health and Family Services
                       Draft for Discussion at SOLEC 2006

-------
                        Yes
                                     Fishing in Household
Figure 1. Percent of responders to the survey who are (red) or are not (yellow) aware offish
consumption advisories and who do (yes) or do not (no) have someone in the household who
fishes.
Source: Wisconsin Department of Health and Family Services
                        Draft for Discussion at SOLEC 2006

-------

           Elementary      Some HS       HS Grad       Coll/Tech
                                          Education
                                                              Grad
                                                                         Unknown
Figure 2. Percent of responders to the survey who are (red) or are not (yellow) aware offish
consumption advisories according to level of education.
Source: Wisconsin Department of Health and Family Services
                         Draft for Discussion at SOLEC 2006

-------
               18-25
                                 26-35
                                                   36^5
                                                                    Unknown
                                       Age Category
Figure 3. Percent of responders to the survey who are (red) or are not (yellow) aware offish
consumption advisories according to age group.
Source: Wisconsin Department of Health and Family Service
                         Draft for Discussion at SOLEC 2006

-------
                                                                        ^v5
    0.1
           White
                                                                       Other
Figure 4. Percent of responders to the survey who are (red) or are not (yellow) aware offish
consumption advisories according to race.
Source: Wisconsin Department of Health and Family Service
                        Draft for Discussion at SOLEC 2006

-------
Beach Advisories, Postings and Closures
Indicator #4200

*Previous beach reports for the Canadian side included inland beach data.  All data for inland
beaches has been removed for this 2006 report, which has skewed the results of the doughnut for
previous years.
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Static
While there's been an increase in monitoring and in the number of beaches
reporting, the percentage of beaches open during beach season over the last
8 years remains constant in the U.S. - at roughly 70% and slightly declining
conditions at 52% in Canada (see Figure 1). The percentage of beaches
posted more than 10% of the beach season averaged 13% in the U.S. and
38% in Canada since 2000. The significant difference in the number of
open beaches in the U.S. and Canada may be due to the difference in
posting criteria. The Ontario standard is a geometric mean of 100 E. coli
colony-forming units per 100ml of water, while in the U.S., beaches are
typically posted using a single sample maximum of 235 E. coli cfu per
100ml.
Lake-by-Lake Assessment
Lake Superior
           Status:  Good
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Undetermined (due to vast increase in number of reported beaches)
During 2004 and 2005, 90% or more of Lake Superior beaches (green &
blue - Figure 2a) were open more than 95% of the time in the U.S.  This
meets the key objective of the 2002 U.S. Great Lakes Strategy goal:  By
2010, 90% of monitored, high priority Great Lakes beaches will meet
bacteria standards more than 95% of the swimming season. In Canada,
during 2005, 5 of 9 beaches were open more than 95% of the time (green &
blue - Figure 2b).
Fair
Undetermined (due to vast increase in number of reported beaches)
Since 2000, on average, 77% of Lake Michigan beaches were open more
than 95% of the time (green & blue - Figure 3). Increased monitoring has
resulted in approximately twice as many postings since 2000 (yellow & red
- Figure 3). Several groups are collaborating to identify and remediate
sources  of beach contamination in Lake Michigan.
Lake Huron
           Status:  Good
           Trend:  U.S.: Static; Canada: Undetermined
   Primary Factors  Since 1998, on average, 94% of U.S. Lake Huron beaches are open more
                        Draft for Discussion at SOLEC 2006

-------

      Determining  than 95% of the beach season. This meets the key objective of the 2002
  Status and Trend  U.S. Great Lakes Strategy goal. However, in Ontario, an average of 49% of
                   Lake Huron beaches were open more than 95% during 1999 through 2005
                   beach seasons (green & blue - Figures 4a & 4b).
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Fair
Undetermined
From 1998 to 2005, on average, 76% of U.S. Lake Erie beaches were open
more than 95% of the beach season. From 1999 through 2005, in Ontario,
an average of 55% of Lake Erie beaches were open more than 95% of the
beach seasons (green & blue - Figures 5a & 5b). Contamination source
identification work is being conducted at Lake Erie beaches.
Fair
Undetermined
From 1998 to 2005, on average, 84% of Lake Ontario beaches in the U.S.
were open more than 95% of the beach season. From 1999 through 2005, in
Ontario, an average of 46% of Lake Ontario beaches were open more than
95% of the beach season (green & blue - Figures 6a & 6b).
Purpose
Assess the number of health-related swimming posting days for freshwater recreational areas
(beaches) in the Great Lakes basin.

Ecosystem Objective
Waters used for recreational activities involving body contact should be substantially free from
pathogens that may harm human health, including bacteria, parasites, and viruses. As the
surrogate indicator, E. coli levels should not exceed national, state or provincial standards set for
recreational waters. This indicator supports Annexes 1, 2 and 13 of the Great Lakes Water
Quality Agreement (United States and Canada 1978).

State of the Ecosystem
Background
A health-related posting day is one that is based upon elevated levels of E. coli, or other indicator
organisms,  as reported by county or municipal health departments in the Great Lakes basin. E.
coli and other indicator organisms are measured in order to infer potential harm to human health
through body contact with nearshore recreational waters because they act as indicators for
potential pathogens.

The Ontario provincial standard is 100 E. coli cfu per 100 mL, based on the geometric mean of a
minimum of one sample per week from each sampling site (minimum of 5 sampling sites per
beach) (Ministry of Health 1998). It is recommended by the Ontario Ministry of Health and
Long-Term Care that beaches of 1000 metres of length or greater require one sampling site per
200 metres. In some cases local Health Units in Ontario have implemented a more frequent
                         Draft for Discussion at SOLEC 2006

-------


sampling procedure than is outlined by the provincial government. When E. coli levels exceed the
limit, the beach is posted as unsafe for the health of bathers. Each beach in Ontario has a
different swimming season length, although the average swimming season for Ontario beaches
begins in early June and continues until the first weekend in September.  The difference in the
swimming season length may skew the final result of the % of beaches posted throughout the
season

The bacteria criteria recommendations for E. coli from the U.S. Environmental Protection
Agency (USEPA) are a single sample maximum value of 235 cfu per 100 ml. For enterococci,
another indicator bacterium, USEPA's recommendations are a single sample maximum value of
62/100 ml (USEPA 1986). When levels of these indicator organisms exceed water quality
standards, swimming at beaches is prohibited or advisories are issued to inform beachgoers that it
may not be safe to swim.

One of the most important factors in nearshore recreational water quality determination is that
indicator bacterial counts are at a level that is safe for bathers. Recreational waters may become
contaminated with animal and human feces from sources and conditions such as combined sewer
overflows (CSOs) and sanitary sewer overflows (SSOs), malfunctioning septic systems and poor
livestock management practices. This pollutant input can become further emphasized in certain
areas after heavy rains. The trends provided by this indicator will aid in beach management and in
the prediction of episodes of poor water quality. In addition, states, provinces, and municipalities
are continuing to identify point and non-point sources of pollution at their beaches, which will
determine why beach areas are becoming impaired. As some sources of contamination are
identified, improved remediation measures can be taken to reduce the number of postings at
beaches.

Status of Great Lakes Beach Advisories, Postings and Closures
Figure 1, shows that as the  frequency of monitoring and reporting increases in the U.S. and
Canada, more postings are also observed, especially after 1999. In fact, both countries
experienced  an approximate percentage doubling of beaches that had postings for more than 10%
of the season in 2000 due to increases in monitoring and reporting. The number of U.S. beaches
being included in the monitoring and reporting program in 2005 has expanded significantly (more
than double since 2002) due to funding from USEPA through the BEACH Act, however, the
percentage of U.S. beaches open all season and the percentage of beaches posted more than 10%
of the season in 2005 are virtually unchanged when compared to 2000-2004.

While the number of beaches reporting in 2004 and 2005 in Canada decreased, the number of
postings each swimming season is fairly constant at about 49% over the last 8 years, excluding
2002 and 2003 (Figure 1).  Although, Lakes Ontario, Huron, and Erie have not met the key
objective of the Great Lakes Strategy 2002, there are measures being taken to improve the
beaches on these lakes. A new version of the Guideline for Canadian Recreational Water Quality
will be out this year, focusing on implementing measures to reduce the risk of contamination
(Robertson, 2006).  Beach surveys, barriers, and preventive weather measures are some of the
actions that will be taken to assist in improving beach quality for the Canadian Great Lakes.

U.S. beaches in Lakes  Superior and Huron are meeting the key objective of the U.S. Great Lakes
Strategy 2002 (http ://www. epa. gov/glnpo/gls/index.html). The Great Lakes Strategy envisions
                         Draft for Discussion at SOLEC 2006

-------
that all Great Lakes beaches will be swimmable and sets a goal that by 2010, 90% of monitored,
high priority Great Lakes beaches will meet bacteria standards more than 95% of the swimming
season (Figures 2a & 4a - except for Lake Huron in 2002).  To help meet this goal, USEPA will
build local capacity in monitoring, assessment and information dissemination to help beach
managers and public health officials comply with USEPA's National Beach Guidance (USEPA
2002b) at 95% of high priority coastal beaches.

Further analysis of the data may show seasonal and local trends in recreational water quality. It
has been observed in the Great Lakes  basin that unless contaminant sources are removed or new
sources introduced, beach sample results contain similar bacteria levels after events with similar
meteorological conditions (primarily wind direction and volume and duration of rainfall). If
episodes of poor recreational water quality can be associated with specific events (such as
meteorological events of a certain threshold), then forecasting for episodes of elevated bacterial
counts may become more accurate.

Pressures
Future pressures: There may be new indicators and new detection methods available through
current research efforts occurring binationally in both public and private sectors and academia.
Although currently a concern in recreational waters, viruses and parasites are difficult to  isolate
and quantify, and feasible measurement techniques have yet to be developed. Comparisons of the
frequency of beach postings are typically limited due to the use of different water quality criteria
in different localities. In the U.S., all coastal states (including those along the Great Lakes) have
criteria as protective as USEPA's recommended bacteriological criteria (use of E. coli or
enterococci indicators) applied to their coastal waters. Conditions required to post Ontario
beaches as unsafe have become more  standardized due to the 1998 Beach Management Protocol,
but the conditions required to remove  the postings remain variable.

Current pressures: Additional point and non-point source pollution at coastal areas due to
population growth and increased land use may result in additional beach postings, particularly
during wet weather conditions. In addition, due to the nature of the laboratory analysis, each set
of beach water samples requires an average of one to two days before the results are
communicated to the beach manager.  Therefore, a lag time in posting exists in addition to the
lifting of any restrictions from the beach when safe levels are again reached. The delay in
developing a rapid test protocol for E. coli is lending support to advanced models to predict when
to post beaches.

Management Implications
Continued BEACH Act funding for beach monitoring and notification programs should be
encouraged as well as funding for beach water contaminant source identification and remediation,
rapid test methods research, and development of predictive models.

In Canada, a partnership between Environment Canada (Ontario Region) and the Ontario
Ministry of Health and Long-Term Care have created the Seasonal Water Monitoring and
Reporting System (SWMRS). This web-based application will provide local Health Units with a
tool to manage beach sampling data, as well as link to the meteorological data archives of
                         Draft for Discussion at SOLEC 2006

-------

Environment Canada. The result will be a system that potentially can be evolved to have some
predictive modeling capability.

Comments from the author(s)
Wet weather sources of pollution have the potential to carry pathogenic organisms to waters used
for recreation and contaminate them beyond the point of safe use. There is a need to begin
identifying beach water contamination sources and implement remediation measures to reduce
contaminant loading.

Many municipalities are in the process of developing long-term control plans that will result in
the selection of CSO controls to meet water quality standards. The City of Toronto has an
advanced Wet Weather Flow Management Master Plan, which could serve as a model to other
urban areas. Information on this initiative can be obtained at:
http://www.city.toronto.on.ca/wes/techservices/involved/wws/wwfmmp/index.htm.

Environment Canada (Ontario Region), in conjunction with the Ontario Ministry of Health and
Long-Term Care and other potential partners, will work to implement the SWMRS reporting
system. Future work will include a predictive modeling capability as well as improving the
interface for public use. The system, once running, will help identify areas of chronic beach
postings and, as a result, will aid in improved targeting of programs to address the sources of
bacterial contamination.

Creating wetlands around rivers, or areas that are wet weather sources of pollution, may help
lower the levels of bacteria that cause beaches to be posted. The wetland area may reduce high
bacterial levels that are typical after storm events by detaining and treating water in surface  areas
rather than releasing  the bacteria-rich waters into the  local lakes and recreational areas. Studies by
the Lake Michigan Ecological Research Station show that wetlands could lower bacterial levels at
state park beaches, but more work is needed (Mitchell 2002).

Variability in the data from year to year may result due to changing seasonal weather conditions,
the process of monitoring and variations in reporting, and may not be solely attributable to actual
increases or decreases in levels of microbial contaminants. At this time, most of the beaches in
the Great Lakes basin are monitored and have quality public notification programs in place. In
addition, state beach managers are submitting their beach monitoring and advisory/closure data to
the USEPA annually. The state of Michigan has an online site (http://www.glin.net/beachcast)
where beach monitoring data is posted by Michigan beach managers. In Ontario, the  SWMRS
program will increase the efficiency and accuracy of the data collection and reporting.

To ensure accurate and timely posting of Great Lake beaches, methods must be developed to
deliver quicker results that focus not just on indicator organism levels but on water quality in
general. This issue is being addressed. The BEACH Act requires  EPA to initiate studies for use in
developing appropriate and effective indicators for improving detection in a timely manner in
Coastal Recreation Waters.  In connection with this requirement,  the USEPA and the Centers  for
Disease Control and Prevention are conducting the National Epidemiological and Environmental
Assessment of Recreation Waters study at various coastal freshwater and marine beaches across
the country to evaluate new rapid and specific indicators of recreational water quality and to
determine their relationships to health effects. Until new indicators are available, predictive
                         Draft for Discussion at SOLEC 2006

-------
                                                  IW^"'^  ^T• ••••••^	N	h.L..............................................rr......
models and/or the experience of knowledgeable environmental or public health officers (who
regularly collect the samples) can be used on both sides of the border. Each method takes a
variety of factors into account, such as amount of rainfall, cloud coverage, wind (direction and
speed), current, point and non-point source pollution inputs, and the presence of wildlife, to
predict whether it is likely that E.  coli levels will likely exceed established limits in recreational
waters.

Acknowledgments
Authors: Tracie Greenberg, Environment Canada Intern, Ontario Region, Burlington, ON;
David Rockwell, U.S. Environmental Protection Agency, Great Lakes National Program Office,
Chicago, IL;
Holiday Wirick, U.S. Environmental Protection Agency, Region 5, Water Division, Chicago, IL;

Data Sources
Canadian data obtained from Ontario Health Units along the Great Lakes.

Health Canada. 1999. Guidelines  for Canadian Recreational Water Quality, 1992. http://www.hc-
sc.gc.ca/ehp/ehd/catalogue/bch_pubs/recreational_water.htm, last accessed July 12, 2002.

Ministry of Health, Algoma Health Unit. 1998. Beach management protocol - safe  water
program.
http://www.ahu.on.ca/health_info/enviro_health/enviro_water/enviro_water_BeachManagement
%20Protocol.htm, last accessed July 12, 2002.

Mitchell, D. 2002. E.  coli testing may have outlived usefulness. The Times Online.
http://www.thetimesonline.com/index.pl/articlesubpc?id=23927743, last accessed July 17, 2002.

Robertson, J.  2006. Evolution of the Guidelines for Canadian Recreational Water Quality. In
Proceedings from Great Lakes Beaches Symposium. June 19, 2006, Toronto, Ontario, Canada.

United States and Canada. 1987. Great Lakes Water Quality Agreement of 1978, as amended by
Protocol signed November 18, 1987. Ottawa and Washington, http://www.on.ec.gc.ca/glwqa/.

U.S. Environmental Protection Agency (USEPA). 2002a. Great Lakes strategy 2002.
http://www.epa.gov/glnpo/gls/index.html, last accessed March 14, 2005.

U.S. Environmental Protection Agency (USEPA). 2002b. National beach guidance and required
performance criteria for grants. www.epa.gov/OST/beaches, last accessed March 14, 2005.

U.S. Environmental Protection Agency (USEPA). 2002c. National health protection survey of
beaches for swimming (1998 to 2001). http://www.epa.gov/waterscience/beaches, last accessed
March 14, 2005.

U.S. Environmental Protection Agency (USEPA). 1986. Ambient water quality criteria for
bacteria - 1986. www.epa.gov/OST/beaches, last accessed March 14,  2005.
                         Draft for Discussion at SOLEC 2006

-------

List of Figures
Figure 1. Proportion of Great Lakes beaches with postings in the United States and Canada for
the 1998-2005 bathing seasons.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units

Figure 2. Proportion of Great Lakes beaches with postings for Lake Superior.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units

Figure 3. Proportion of Great Lakes beaches with postings for Lake Michigan.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003

Figure 4. Proportion of Great Lakes beaches with postings for Lake Huron.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units

Figure 5. Proportion of Great Lakes beaches with postings for Lake Erie.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units

Figure 6. Proportion of Great Lakes beaches with postings for Lake Ontario.
Note: The Ontario standard is 100 E. coli colony-forming units per 100ml of water, while the
U.S. standard is a single sample maximum of 235 E. coli cfu per 100ml.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
                       Proportion of U.S. Great Lake Basin Beaches
                    with Postings for the 1998 - 2005 Bathing Seasons
                                                          • 0% posted
                                                          • 1%-4% posted
                                                          D 5%-9% posted
                                                          • >10% posted
                    Proportion of Canadian Great Lakes Beaches
               with Beach Postings for the 1998-2005 Bathing Season
Number of Great Lake Basin
    Beaches reported
  Canada      U.S.
    194-2005-892
    161 -2004-787
    270 - 2003 - 649 *
    272-2002-381
    304-2001 -304
    293-2000-333
    238- 1999-320
    218- 1998-303
  * Data Source NRDC
                                                                 • 0% posted

                                                                 • 1% -4% posted

                                                                 a 5% - 9% posted

                                                                 • >10% posted
Figure 1. Proportion of Great Lakes beaches with postings in the United States and Canada for
the 1998-2005 bathing seasons.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units
                        Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
                           Proportion of U.S. Lake Superior Beaches
                       with Beach Postings for the 1998-2005 Bathing Seasons
\
\ \
\ \
\ \
I



/
i




/ /






/ /
/ /
/ /
7
• 0% posting
• 1% -4% posting
D5% -9% posting
• >10% posting

Number of Lake
Superior Beaches
reported each year:
2005 -194 beaches
2004 -167 beaches
2002- 14 beaches
2001 - 10 beaches
2000 - 9 beaches
1999 - 9 beaches
1998- 7 beaches
                              Lake Superior - Canada
                                     2005 45%

                                     	^
                                     2000  50%
                                                                     2005-
                                                                     2004-
                                                                     2003-
                                                                     2002-
                                                                     2001 -
                                                                     2000-
                                                                     1999
 9 Beaches
 0 Beaches
 0 Beaches
 0 Beaches
 0 Beaches
 4 Beaches
- 4 Beaches
                                                                        D0%

                                                                        • 1-4%

                                                                        D5-<10%

                                                                        • >or=10%
Figure 2. Proportion of Great Lakes beaches with postings for Lake Superior.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units
                         Draft for Discussion at SOLEC 2006

-------
                              State of the Great Lakes 2007 - Draft
                             Proportion of Lake Michigan Beaches
                       with Beach Postings for the 1998-2005 Bathing Seasons
V
\
\
1
1

1

72%


/
/
/
• 0% posting
• 1%-4% posting
D5%-9% posting
• >10% posting

Number of Lake
Michigan Beaches
reported each year:
2005 - 445 beaches
2004 - 428 beaches
2002 - 204 beaches
2001 - 157 beaches
2000- 177 beaches
1999- 173 beaches
1998- 158 beaches
Figure 3. Proportion of Great Lakes beaches with postings for Lake Michigan.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003
10
Draft for Discussion at SOLEC 2006

-------
          State of the Great Lakes 2007 - Draft
                               Proportion of Lake Huron Beaches
                        with Beach Postings for the 1998-2005 Bathing Seasons
                                                                       Number of Lake
                                                                       Huron Beaches
                                                                      reported each year:

                                                                      2005-137 beaches
                                                                      2004 - 93 beaches
                                                                      2002-43 beaches
                                                                      2001 -28 beaches
                                                                      2000 - 36 beaches
                                                                      1999 - 34 beaches
                                                                      1998 - 30 beaches
                                  Lake Huron - Canada
                                                                        2005-
                                                                        2004-
                                                                        2003-
                                                                        2002-
                                                                        2001 -
                                                                        2000-
                                                                        1999
 58 Beaches
- 45 Beaches
- 74 Beaches
- 75 Beaches
 55 Beaches
 44 Beaches
-41 Beaches
Figure 4. Proportion of Great Lakes beaches with postings for Lake Huron.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units
                          Draft for Discussion at SOLEC 2006
               11

-------
                                  State of the Great Lakes 2007 - Draft
                               Proportion of Lake Erie Beaches
                       with Beach Advisories for the 1998-2005 Bathing Seasons
                                                                   Number of Lake
                                                                    Erie Beaches
                                                                  reported each year:

                                                                  2005 - 85 beaches
                                                                  2004 - 69 beaches
                                                                  2002 - 79 beaches
                                                                  2001-71 beaches
                                                                  2000 - 75 beaches
                                                                  1999-74 beaches
                                                                  1998-78 beaches
                                  Lake Erie - Canada
                                                                        2005-
                                                                        2004-
                                                                        2003-
                                                                        2002-
                                                                        2001 -
                                                                        2000-
                                                                        1999
                                                  57 Beaches
                                                  49 Beaches
                                                  76 Beaches
                                                  77 Beaches
                                                  64 Beaches
                                                  60 Beaches
                                                  - 26 Beaches
Figure 5. Proportion of Great Lakes beaches with postings for Lake Erie.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units
12
Draft for Discussion at SOLEC 2006

-------
          State  of the Great Lakes 2007 - Draft
                                   Proportion of Lake Ontario Beaches
                             with Beach Postings for the 1998-2005 Bathing Seasons
                                                                          Number of Lake
                                                                          Ontario Beaches
                                                                         reported each year:

                                                                         2005-19 beaches
                                                                         2004-19 beaches
                                                                         2002 - 21 beaches
                                                                         2001 -21 beaches
                                                                         2000-19 beaches
                                                                         1999-13 beaches
                                                                         1998-13 beaches
                                      Lake Ontario - Canada
                                                                             2005-
                                                                             2004-
                                                                             2003-
                                                                             2002-
                                                                             2001 -
                                                                             2000-
                                                                             1999
 75 Beaches
 72 Beaches
 90 Beaches
 90 Beaches
 67 Beaches
 69 Beaches
- 61 Beaches
Figure 6. Proportion of Great Lakes beaches with postings for Lake Ontario.
Note: The Ontario standard is 100 E. coli colony-forming units per 100ml of water, while the
U.S. standard is a single sample maximum of 235 E. coli cfu per 100ml.
Source: U.S. data: U.S. Environmental Protection Agency, Great Lakes National Programs Office
and the National Resource Defense Council for 2003; Canadian data compiled by Environment
Canada from Ontario Health Units
                          Draft for Discussion at SOLEC 2006
         13

-------
Contaminants in Sport Fish
Indicator #4201
Overall Assessment
          Status:  Mixed
          Trend:  Improving
   Primary Factors
      Determining
  Status and Trend
The Great Lakes Fish Monitoring Program (GLNPO) and the Sport
Fish Contaminant Monitoring Program (Ontario Ministry of the
Environment, OMOE) have been monitoring contaminant levels in
Great Lakes fish for over three decades. To demonstrate trends in
organic contaminant levels, average-size (60cm) lake trout were chosen
by OMOE as the representative fish species due to their presence in all
of the Great Lakes, their potential for exploitation by anglers and their
high accumulation rates for organic contaminants. To demonstrate
trends in mercury levels, average-size (45cm) walleye were chosen by
OMOE due to high mercury accumulation rates. The GLNPO
program was not designed to determine trends in levels of
contaminants in sport fish, and it relies on individual Great Lakes
States and Tribes to issue consumption advice. Rather, the GLNPO
program can compare mean concentration levels to a set standard, the
Protocol for a Uniform Great Lakes Sport Fish Consumption Advisory,
by year. Other important differences between the GLNPO and OMOE
programs include composite analysis versus individual analysis, skin on
versus skin off, and whole fillet analysis versus dorsal plug analysis
respectively. For this reason, only general comparisons between
GLNPO and OMOE data should be made.
Lake-by-Lake Assessment
 Lake Superior
          Status:
          Trend:
   Primary Factors
      Determining
  Status and Trend
        EPA - GLNPOs data can not be used for statistical trend analysis.
        Any trend discussions in the lake assessments below are based on
        OMOE data.

        Contaminant concentrations for both EPA - GLNPO and OMOE
        can be compared to meal category advice. OMOE calculates its
        own advice and EPA - G LNPO compares its contaminant
        concentrations to the Protocol for a Uniform Great Lakes Sport
        Fish Consumption Advisory categories.
Mixed
Improving
PCB concentrations in Lake Superior lake trout have declined considerably
over the period of record. In the late 1970s, PCB concentrations exceeded
the current OMOE  "do not eat" consumption limit. Since 1990,
concentrations have generally fluctuated between 0.153 and 0.610 ppm,
which would permit the consumption of 2 to 4 meals per month.  Current
EPA - GLNPO concentrations range between the one meal per week and
the one meal per month categories.
                       Draft for Discussion at SOLEC 2006

-------
                                                   , $/'*? (ft •* ipt-iln"!"* v «S ^WJw iJ*w 'to^S?'   *-i i jr--« -'-Mi«*«s
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Mercury levels in 45cm walleye from Lake Superior have ranged from 0.62
                   to 0.30 ppm between 1973 and 2002. With the exception of a maximum
                   level reached in 1989 (0.84 ppm), levels of mercury in walleye have
                   declined over the last few decades. In the last 5 years of the period of
                   record, levels of mercury in 45cm walleye have been around 0.30 ppm,
                   permitting the consumption of 4 meals per month for the sensitive
                   population.  These mercury levels are similar to those found in fish from
                   other Ontario lakes and rivers.

                   Toxaphene has historically been high in fish from Lake Superior due to
                   atmospheric deposition.  In 60cm lake trout from Lake Superior, toxaphene
                   has ranged from 0.810 to 0.214 ppm between 1984 and 2003. In 1993,
                   levels of toxaphene in lake trout exceeded 1 ppm. The most current
                   concentration is below the consumption limits and does not result in any
                   fish consumption advisories.
Mixed
Improving
EPA - GLNPO data can be used to discern general trends from Lake
Michigan data due to multiple collection sites. These data display a general
decline in PCB concentrations in coho and chinook salmon fillets. No
OMOE samples were collected from Lake Michigan. Current EPA -
GLNPO concentrations fall into the one meal per month category.
Mixed
Improving
PCB levels in Lake Huron OMOE lake trout declined substantially between
1976 and 2004. In 1976 concentrations exceeded 4ppm, well above the "do
not eat" consumption limit of 1.22ppm for the general population. Current
PCB concentrations in 60cm lake trout slightly exceed 0.153 ppm,  allowing
for the safe consumption of a maximum of 4 meals per month.  Current
EPA - GLNPO concentrations range between the one meal per week and
the one meal per month categories.

Mercury levels in 45cm walleye from Lake Huron have ranged from 0.48 to
0.16 ppm between 1976 and 2004. With the exception of a maximum level
reached in 1984 (0.59 ppm), there has been a general decline over the last
few decades. During the last decade, levels  of mercury have remained
below the first level of consumption restriction (0.26ppm) for the sensitive
population.
Lake Erie
           Status:   Mixed
           Trend:   Improving
                        Draft for Discussion at SOLEC 2006

-------
   Primary Factors
      Determining
  Status and Trend
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Trend data are sparse for Lake Erie as lake trout are less abundant in this
lake. PCB levels in OMOE fish declined between 1984 and 2003.
Nevertheless, PCB concentrations in 60 cm lake trout currently restrict
consumption to 2 meals per month for the general population. The sensitive
population is advised not to consume these fish. Current EPA - GLNPO
concentrations range between the one meal per week and the one meal per
month categories.

Mercury levels in 45cm walleye have declined considerably over the period
of record, from 0.76 ppm in 1970 to 0.18 ppm in 2004. Over the past two
decades, levels of mercury have remained between 0.10 and 0.20 ppm, and
do not restrict consumption of 45cm walleye.
Mixed
Improving
Historically, the highest concentrations of PCBs have been found in Lake
Ontario. From the late 1970s to 1999, PCBs in 60 cm OMOE lake trout
from Lake Ontario were at or near the "do not eat" consumption limit.
Substantially lower concentrations have been found in the most recent
samples in 2002 and 2004, and the current levels would permit consumption
of 2 meals per month.  Current EPA - GLNPO concentrations fall into the
one meal per week category.

Mercury levels in 45cm walleye have fluctuated between 0.23 and 0.17 ppm
between 1975 and 2005. There has been no major decline in mercury
concentrations in walleye, however, maximum levels have only reached
0.32ppm. Over the past 3 years, mercury concentrations in 45cm walleye
have remained below the first level of consumption restriction for the
sensitive population.

High levels of mirex have been found in fish from Lake Ontario and it has
historically been a source offish consumption restrictions. Levels of mirex
in 60cm lake trout from Lake Ontario have  declined significantly from
0.302 to 0.036 ppm between 1978 and 2004, with a maximum of 0.387 ppm
reached in 1985. The current concentration of mirex no longer restricts
consumption of 60cm lake trout.  Photomirex is a breakdown product of
mirex, which also bioaccumulates in fish and has historically caused
consumption restrictions in some Lake Ontario species. Levels in 60cm
lake trout have declined from 0.044 to 0.015 ppm between 1994 and 2004.
Advice for the Protocol for a Uniform Great Lakes Sport Fish Consumption Advisory
was calculated for sensitive populations based on a weight of evidence of non-cancer
developmental effects. The general population is advised to follow the same advice
based on potential cancer risk.  Health Canada does not consider PCBs (especially
environmental levels) to be carcinogens. Therefore, non-cancer endpoints were used to
                        Draft for Discussion at SOLEC 2006

-------
                                                                                 	
calculate the Tolerable Daily Intakes (TDI) for PCBs. This TDI was applied more-or-less
equally to both sensitive and general populations. For mercury, Health Canada and US
states assign separate TDIs or RfDs for the general and sensitive populations.

Purpose
•To assess potential human exposure to persistent bioaccumulative toxic (PBT) contaminants
through consumption of popular sport species;
•To assess the levels of PBT contaminants in Great Lakes sport fish; and
•To identify trends over time of PBT contaminants in Great Lakes sport fish or in fish
consumption advisories.

In addition to an indicator of human health, contaminants in fish are an important indicator of
contaminant levels in an aquatic ecosystem because of the bioaccumulation of organochlorine
chemicals in their tissues. Contaminants that are often undetectable in water can be detected in
fish.

Ecosystem Objective
Great Lakes sport fish should be safe to eat and concentrations of toxic contaminants in sport fish
should not pose a risk to human health. Unlimited consumption of all Great Lakes sport fish
should be available to all citizens of the Great Lakes basin.

Annex 2 of the Great Lakes Water Quality Agreement (United States and Canada 1987) requires
Lakewide Management Plans (LaMPs) to define ".. .the threat to human health posed by critical
pollutants... including their contribution to the impairment of beneficial uses." Both the Protocol
for a Uniform Great Lakes Sport Fish  Consumption Advisory and the Guide to Eating Ontario
Sport Fish are used to assess the status of the ecosystem by comparing contaminant
concentrations to consumption advice.

State of the Ecosystem
Program History
Both the United States and Canada (Ontario)  collect and analyze sport fish to determine
contaminant concentrations, relate those concentrations to health protection values  and develop
consumption advice to protect human  health.  For U.S.-caught sport fish, the Protocol for a
Uniform Great Lakes Sport Fish Consumption Advisory for PCBs is used as a  standardized fish
advisory benchmark for this indicator, and it is applied to historical U.S. Environmental
Protection Agency (USEPA) Great Lakes National Program Office (GLNPO) data to track trends
in fish consumption advice. Individual Great Lakes  States and Tribes issue specific consumption
advice for how much fish and which fish are safe to eat for a wide variety of contaminants.
GLNPO salmon fillet data are used to  demonstrate this indicator. Due to gaps and variability in
GLNPO salmon fillet data, statistically significant trends are difficult to discern. For Canadian-
caught sport fish, Health Canada sets Tolerable Daily Intakes (TDI) for certain contaminants of
concern, including PCBs, mercury, dioxins (including dioxins, furans and dioxin-like PCBs),
mirex, photomirex, toxaphene and chlordane. TDIs are defined as the quantity of a  chemical that
can be consumed on a daily basis, for  a lifetime, with reasonable assurance that one's health will
not be threatened, and they are used in the calculation of sport fish consumption limits which are
listed in the Guide to Eating Ontario Sport Fish.
                        Draft for Discussion at SOLEC 2006

-------


The GLWQA, first signed in 1972 and renewed in 1978, expresses the commitment of Canada
and the United States to restore and maintain the chemical, physical and biological integrity of the
Great Lakes basin ecosystem.

Contaminants in Great Lakes Sport Fish
Since the 1970s, there have been declines in the levels of many PBT chemicals in the Great Lakes
basin due to bans on the use and/or production of harmful substances and restrictions on
emissions. However, because of their ability to  bioaccumulate and persist in the environment,
PBT chemicals continue to be a significant concern. Historically, PCBs have been the
contaminant that most frequently limited the consumption of Great Lakes sport fish. In some
areas, dioxins, toxaphene (Lake Superior) or mirex/photomirex (Lake Ontario) have been the
consumption-limiting contaminant. Recently Health Canada has revised downward its TDIs for
PCBs and dioxins, which has increased the frequency of consumption restrictions caused by
PCBs and dioxins and decreased the frequency  for toxaphene and mirex/photomirex.

Illustration note - Please note that differing species (coho salmon and lake trout) and units (ppm
and ppb) are presented in the accompanying graphs. Typically lake trout have higher contaminant
concentrations than coho salmon.

Pressures
Organochlorine contaminant levels in fish in the Great Lakes are generally decreasing. As these
contaminants  continue to decline, mercury will  become a more important contaminant of concern
in Great Lakes fish.

Concentrations of PBT contaminants such as PCBs have declined in lake trout throughout the
Great Lakes basin.  However, concentrations still exceed current consumption limits. Regular
monitoring must continue in the Great Lakes basin to maintain trend data.  In many areas of the
Great Lakes, dioxins (including dioxins, furans  and dioxin-like PCBs) are  now the consumption-
limiting contaminant and need to be monitored  more frequently. The focus should also turn to
PBT contaminants of emerging concern, such as brominated flame retardants, before their
concentrations in sport fish reach levels that may affect human health.

Consumption advisories and PCB concentrations in coho salmon (U.S. program)
State and tribal governments provide information to consumers regarding consumption of sport
caught fish. Neither the guidance nor advice of a state or tribal government is regulatory.
However, some states use the federal commercial fish guidelines for the acceptable level of
contaminants  when giving advice for eating sport-caught fish. Consumption advice offered by
most agencies is based on human health risk. This  approach involves interpretation of studies on
health effects  from exposure to contaminants. Each state or tribe is responsible for developing
fish consumption advisories for protecting the public from pollutants in fish and tailoring this
advice to meet the health needs of its citizens. As a result, the advice from different states and
tribal programs is sometimes somewhat different for the same lake and species within that lake.

Additional information about the toxicity  of a larger suite of chemicals is needed. The health
effects of multiple contaminants, including endocrine disrupters, also need to be addressed.
                         Draft for Discussion at SOLEC 2006

-------
Management Implications
Health risk communication is a crucial component to the protection and promotion of human
health in the Great Lakes. Enhanced partnerships between states and tribes involved in the issuing
offish consumption advice and USEPA headquarters will improve U.S. commercial and non-
commercial fish advisory coordination. In Canada, acceptable partnerships exist between the
federal and provincial agencies responsible for providing fish consumption advice to the public.

At present, PCBs and Chlordane are the only PBT chemicals that have uniform fish advisory
protocols across the U.S. Great Lakes basin, mercury is being drafted. There is a need to establish
additional uniform PBT advisories in order to limit confusion of the public that results from
issuing varying advisories for the same species of sport fish across the basin.

In order to best protect human health, increased monitoring and reduction of PBT chemicals need
to be made a priority. In particular, monitoring of contaminant levels  in environmental media and
biomonitoring  of human tissues need to be addressed, as well as assessments of frequency and
type offish consumed. This is of particular concern in  sensitive populations because contaminant
levels in some  fish are higher than in others. In addition, improved understanding of the potential
negative health effects  from exposure to PBT chemicals is needed.

In March, 2004, the U.S. Food and Drug Administration and the USEPA jointly released a
consumer advisory  on methylmercury in fish. The joint advisory advises women who may
become pregnant, pregnant women, nursing mothers, and young children to avoid eating some
types of fish and to  eat fish and shellfish that are lower in mercury. While this is a step forward
toward uniform advice regarding safe fish consumption, the national advisory is not consistent
with some Great Lakes State's advisories. Cooperation among National, State, and Tribal
governments to develop and distribute the same message regarding safe fish consumption needs
to continue. Health  Canada has had a similar advisory  since 1999.

Comments from the author(s)
Support is needed for the States from the Great Lakes National Program Office (GLNPO) and
U.S. Environmental Protection Agency (USEPA) headquarters to help facilitate a meeting to
review risk assessment protocols.

Evaluation of historical long term fish contaminant monitoring data sets, which were assembled
by several jurisdictions for different purposes, need to  be more effectively utilized. Relationships
need to be developed that allow for comparison and combined use of existing data from the
various sampling programs. These data could be used in expanding this indicator to other
contaminants and species and for supplementing the data used in this  illustration.

Coordination of future monitoring would greatly assist the comparison of fish contaminants data
among federal, provincial, state and tribal jurisdictions.

Agreement is needed on U.S.  fish advisory health benchmarks for the contaminants that cause
fish advisories  in the Great Lakes. Suggested starting points are: The  Great Lakes Protocol for
PCBs and Chlordane and USEPA's reference dose for  mercury. Ontario remains consistent with
Health Canada's TDIs throughout the province.
                         Draft for Discussion at SOLEC 2006

-------
Acknowledgments
Authors: Elizabeth Murphy, U.S. Environmental Protection Agency, Great Lakes National
Program Office;
Jackie Fisher, U.S. Environmental Protection Agency, Great Lakes National Program Office;
Emily Awad, Sport Fish Contaminant Monitoring Program, Ontario Ministry of Environment,
Etobicoke, ON;
Alan Hayton, Sport Fish Contaminant Monitoring Program, Ontario Ministry of Environment,
Etobicoke, ON; and

Data Sources
Elizabeth Murphy, U.S. Environmental Protection Agency, Great Lakes National Program Office,
murphy.elizabeth(Sjepa.gov.

De Vault, D.S. and Weishaar, J.A. 1984. Contaminant analysis of 1982 fall run coho salmon. U.S.
Environmental Protection Agency, Great Lakes National Program Office. EPA 905/3-85-004.

De Vault, D.S. and Weishaar, J.A. 1983. Contaminant analysis of 1981 fall run coho salmon. U.S.
Environmental Protection Agency, Great Lakes National Program Office. EPA 905/3-83-001.

De Vault, D.S., Weishaar, J.A., Clark, J.M., and Lavhis, G. 1988. Contaminants and trends in fall
run coho salmon. J. Great Lakes Res. 14:23-33.

Great Lakes Sport Fish Advisory Task Force. 1993. Protocol for a uniform Great Lakes sport fish
consumption advisory. http://fh.cfs.purdue.edu/anglingindiana/HealthRisks/TaskForce.pdf, last
accessed July 22, 2005.

Ontario Ministry of the Environment (OMOE). 2005. Guide to Eating Ontario Sport Fish 2005-
2006. http://www.ene.gov.on.ca/envision/guide/index.htm, last accessed July 19, 2006.

United States and Canada. 1987. Great Lakes Water Quality Agreement of 1978, as amended by
Protocol signed November  18, 1987. Ottawa and Washington, http://www.on.ec.gc.ca/glwqa/.

U.S. Environmental Protection Agency. 2004. Consumption Advice, Joint Federal Advisory for
Mercury in Fish, http://www.epa.gov/waterscience/fishadvice/advice.html, last accessed May 24,
2004.

Data
Great Lakes Fish Monitoring Program, Great Lakes National Program Office;
Sport Fish Contaminant Monitoring Program, Ontario Ministry of Environment;
Minnesota DNR salmon fillet data for Lake Superior.

List of Tables
Table 1. Contaminants on which the  fish advisories are based on by lake for  Canada and the
United States.
Source: Compiled by U.S. Environmental Protection Agency (USEPA) Great Lakes National
Program Office
                        Draft for Discussion at SOLEC 2006

-------

                                                                             to.
Table 2. Uniform Great Lakes Sport Fish Consumption Advisory.
Source: Great Lakes Sport Fish Advisory Task Force, 1993.

Table 3. Consumption limits used for the Guide to Eating Ontario Sport Fish (based on Health
Canada TDIs).
Source: Ontario Ministry of the Environment

Last updated
SOLEC 2006
Lake
Superior
Huron
Michigan
Erie
Ontario
Contaminants that Fish Advisories are
based on in Canada and the United
States
Dioxin, PCBs, toxaphene, mercury,
chlordane
Dioxin, PCBs, toxaphene, mercury,
chlordane
PCBs, mercury, dioxin, chlordane
PCBs, dioxin, mercury
PCBs, dioxin, mercury, mirex, toxaphene
Table 1. Contaminants on which the fish advisories are based on by lake for Canada and the
United States.
Source: Compiled by U.S. Environmental Protection Agency (USEPA) Great Lakes National
Program Office
                        Draft for Discussion at SOLEC 2006

-------
Consumption Advice
Groups
Sensitive* and General
Unrestricted
Consumption
2 meals/ week
1 meal/ week
1 meal/ month
6 meals/ year
Do not eat
Concentration
ofPCBs
(ppm)

0-0.05
NA
0.06-0.2
0.21-1.0
1.1-1.9
>1.9
Concentration
of Mercury
(ppm)**

0 <= 0.05
> 0.05 <= 0.11
>0.11 <=0.22
> .22 <= 0.95
NA
>0.95
Concentration
of Chlordane
(ppm)***

0-0.15
NA
0.16-0.65
0.66 - 2.82
2.83-5.62
>5.62
         * Women of childbearing age and children under 15
         **Draft Protocol for Mercury-based Fish Consumption Advice
         ***Discussion Paper for Chlordane HPV
Table 2. Uniform Great Lakes Sport Fish Consumption Advisory.
Source: Great Lakes Sport Fish Advisory Task Force, 1993
Advised meals per month
Sensitive*
8
4
Do not eat
Do not eat
Do not eat
General
8
4
2
1
Do not eat
Concentration of
PCBs (ppm)
<0.153
0.153-0.305
0.305-0.610
0.610-1.22
>1.22
                 * Women of childbearing age and children under 15

Table 3. Consumption limits used for the Guide to Eating Ontario Sport Fish (based on Health
Canada TDIs).
Source: Ontario Ministry of the Environment
                       Draft for Discussion at SOLEC 2006

-------

Air Quality
Indicator #4202

Overall Assessment
           Status:   Mixed
           Trend:   Improving

Purpose
•To monitor the air quality in the Great Lakes ecosystem; and
•To infer the potential impact of air quality on human health in the Great Lakes basin.

Ecosystem Objective
Air should be safe to breathe. Air quality in the Great Lakes ecosystem should be protected in
areas where it is relatively good, and improved in areas where  it is degraded. This is consistent
with ecosystem objectives being adopted by certain lakewide management plans, including Lake
Superior, in fulfillment of Annex 2 of the Great Lakes Water Quality Agreement (GLWQA). This
indicator also supports Annexes 1,13, and 15.

State of the Ecosystem
Overall, there has been significant progress in improving air quality in the Great Lakes basin. For
several substances of interest, both emissions  and ambient concentrations have decreased over the
last ten years or more. However, progress has not been uniform and differences in weather from
one year to the next complicate analysis of ambient trends. Ozone and fine particulate matter can
be particularly elevated during hot summers, and the trends are not consistent with those for
related pollutants. Drought conditions result in more fugitive dust emissions from roads and
fields, increasing the ambient levels of particulate matter.

In general, there has been significant progress with urban/local pollutants over the past decade or
more, though somewhat less in recent years, with a few remaining problem districts. Ground-
level ozone and fine particles remain a concern in the Great Lakes region, especially in the
Detroit-Windsor region and extending northward to Sault St. Marie  and eastward to Ottawa, the
Lake Michigan basin, and the Buffalo-Niagara area. These pollutants continue to exceed the
respective air quality criteria and standards at a number of monitoring locations in Southern
Ontario and in the lower Great Lakes region in the U.S.

For the purposes of this discussion, the pollutants can be divided into urban (or local)  and
regional pollutants. For regional pollutants, transport is a significant issue, from hundreds of
kilometers to the scale of the globe. Formation from other pollutants, both natural and man-made,
can also be important. Unless otherwise stated, references to the U.S. or Canada in this discussion
refer to nationwide averages.

Urban/Local Pollutants

Carbon Monoxide (CO)
Ambient Concentrations: In the U.S., CO levels for 2004 were the lowest recorded in the past 25
years. Ambient concentrations have decreased approximately  71% nationally from 1980 to 2004
and 42% nationally from 1993 to 2002. There are currently no nonattainment areas (areas where
                         Draft for Discussion at SOLEC 2006

-------
air quality standards are not met) in the U.S. for CO. In general, CO levels have decreased at the
same rate in the Great Lakes region as the nation as a whole.

In Ontario, the composite average of the one-hour maximum CO concentration decreased by 82
percent from 1971 to 2004, while the composite average of the eight-hour maximum
concentration decreased 87 percent. Since 1995, average CO concentrations have only decreased
16%. Ontario has not experienced an exceedence of the 1-hour and 8-hour criteria since 1991.

Emissions: In the U.S., nationwide emissions of CO have decreased 33% from 1990 to 2002, the
most recent year for which aggregate National Emissions Inventory (NEI) estimates are available.
The reductions in CO emissions are almost entirely due to decreased emissions from on-road
mobile sources, which have occurred despite yearly increases in vehicle miles traveled. In
general, CO emissions have decreased at the same rate in the Great Lakes region as the nation as
a whole.

In Canada, anthropogenic emissions (not including  open sources such as forest fires) have
decreased nationally by about 22% between 1990 and 2002, with a 29% decline in Ontario over
the same time period. These declines are mainly the result of more stringent transportation
emission standards.

Nitrogen Dioxide (NO2)
Ambient Concentrations: In Ontario, ambient average NO2 concentrations have decreased 31 %
from 1975 to 2004.  Over the last decade (1995 to 2004), average NO2 concentrations declined
13%. The Ontario 1-hour and 24-hour air quality criterion for NO2 were not exceeded at any of
Ontario's monitoring stations in 2004.

In the U.S., the annual mean concentrations decreased 37% from 1980 to 2004. NO2 levels in the
Great Lakes region decreased at a slightly higher pace during this time period. An analysis of
urban versus rural monitoring sites indicates that the declining trend seen nationwide and in the
Great Lakes region can mostly be attributable to decreasing concentrations of NO2 in urban areas
(similar results can be found in Ontario). There are  currently no NO2 nonattainment areas in the
U.S.

Emissions: In Canada, anthropogenic emissions (not including open sources such as forest fires)
have increased nationally by about 5% between 1990 and 2002; however, emissions have
decreased by about 11 % in Ontario over the same time period. These declines are mainly the
result of more stringent transportation emission standards.

In the U.S., emissions of NOx decreased by about 18% from 1990 to 2002. The downward trend
can be attributed to emissions reductions at electric  utilities and on-road mobile sources.
Although nationwide NOx emissions have decreased, emissions from some source categories
have increased including non-road engines.  In general, NOx emissions have decreased at a
slightly greater rate in the Great Lakes region as compared to the nation as a whole. (For more
information on oxides of nitrogen, please refer to the Great Lakes Indicator Report #9000 Acid
Rain.)
                         Draft for Discussion at SOLEC 2006

-------
Sulfur Dioxide (SO2)
Ambient Concentrations: In the U.S., annual mean concentrations of SO2 decreased 54% from
1983 to 2002. From 1993 to 2002, annual mean concentrations of SO2 in the U.S. decreased
39%. The Great Lakes region experienced reducing trends on par with the national averages.
Since the SOGL 2005 Report, the U.S. Environmental Protection Agency (USEPA)  approved the
redesignation of Lake County, Indiana, and Cuyahoga County, Ohio, to attainment areas. There
are currently no nonattainment areas for SO2 in the Great Lakes region.

In Ontario, the average ambient SO2 concentrations improved 86% from 1971 to 2004, with a
17% improvement since 1995. Ontario did not experience any violations of the one-hour SO2
criterion (250 ppb), 24-hour criterion (100 ppb), or the annual criterion (20 ppb) in 2004.

Emissions: In the U.S., national SO2 emissions were reduced 33% from 1990 to 2002 mostly in
response to regulations imposing cuts on coal-burning power plants. SO2 emissions in the Great
Lakes region have decreased at a much greater rate than the national trend over this time period.

Canadian emissions decreased 29% nationwide from 1990 to 2002, but have remained relatively
constant since 1995. Even with increasing economic activity, emissions remain about 29% below
the target national emission cap. From 1990 to 2002, the emissions of SO2  in Ontario decreased
47%. These reductions mostly were the result of the Canada Acid Rain Program which primarily
targeted major non-ferrous  smelters and fossil fuel-burning power plants in the seven eastern-
most provinces.

(For more information on sulfur dioxide, please refer to the Great Lakes Indicator Report #9000
Acid Rain.)

Lead
Ambient Concentrations: U.S. concentrations of lead decreased 97% from 1980 to 2004 with
most of the reductions occurring during the 1980s and early  1990s. Lead levels in the Great
Lakes region decreased at nearly the same rate as the national trend over this time. There are no
nonattainment areas for lead in the Great Lakes region.

Based on historical data, lead concentrations at urban monitoring stations in Ontario have
decreased over 95%.

Emissions: National lead emissions in the U.S. decreased 98% from 1980 to 1999 mostly as a
result of regulatory efforts to reduce the content of lead in gasoline.  The declines  since 1990 have
been from metals processing and waste management industries.

Similar improvements in Canada have followed with the usage of unleaded gasoline.

Total Reduced Sulfur (TRS)
Ambient Concentrations: This family of compounds is of concern in Canada due to odour
problems in some communities, normally near industrial or pulp  mill sources. Ontario did not
experience any violations of the one-hour TRS criterion (27 ppb) in 2004.
                         Draft for Discussion at SOLEC 2006

-------
                                                                                   	
Emissions: Hydrogen sulphide accounts for more than half of total reduced sulphur emissions.
There is no requirement to report TRS emissions in the NPRI; however, there has been a
requirement to report hydrogen sulphide emissions since 2000. Hydrogen sulphide emissions
have increased about 47 percent from 2000 to 2003.

PM10
Ambient Concentrations: PM10 is the fraction of particles in the atmosphere with a diameter of
10 microns or smaller. Annual average  PM10 concentrations in the U.S. have decreased 28%
from 1990 to  2004. Annual average concentrations in the  Great Lakes region have decreased at
nearly the same rate as the national trend over this time. The national 24-hour PM10
concentration was 31% lower than the 1990 level.  24-hour average concentrations in the Great
Lakes region  have decreased at nearly the same rate as the national trend over this time. There
are currently no nonattainment areas in the Great Lakes region. Since the SOGL 2003 report, the
USEPA approved the redesignation of 2 areas in Cook County, Illinois, to attainment areas.

Canada does not have an ambient target for PM10. However, Ontario has an interim standard of
50 (ig/m3 over a 24-hour sampling period to guide decision-making.

Emissions: In the U.S., national direct source man-made emissions decreased 29% from 1990 to
2002. The fuel combustion source category experienced the largest absolute decrease in
emissions (422,000 tons and 35%), while the on-road vehicle sector experienced the largest
relative decrease (183,000 tons  and 47%).  The Great Lakes region experienced reducing trends
on par with the national averages.

In Canada, anthropogenic emissions (not including open sources such as road dust) have
decreased nationally by about 15% between 1990 and 2002.  However, total PM10 emissions
including open sources such as road dust have actually increased by 34% in Canada over this time
period.  Ontario has experienced similar trends over this time period.

Air Toxics
This term captures a large number of pollutants that, based on the toxicity and likelihood for
exposure, have the potential to harm human health (e.g. cancer causing) or adverse environmental
and ecological effects. Some of these are of local importance, near to sources, while others may
be transported over long distances. Monitoring is difficult  and expensive, and usually limited in
scope as such toxics are usually present only at trace levels. Recent efforts in Canada and the U.S.
have focused  on better characterization of ambient levels and minimizing emissions. In the U.S.,
the Clean Air Act targets a 75% reduction in cancer "incidence" and a  "substantial" reduction in
non-cancer risks. The Maximum Available Control Technology (MACT) program sets emissions
standards on industrial sources to reduce emissions  of air toxics. Once  fully implemented, these
standards will cut emissions of toxic air pollutants by nearly 1.36 million metric tons per year
from 1990 levels.

In February 2006, EPA released the results of its national assessment of air toxics (NATA) using
1999 emissions. The purpose of the national-scale assessment is to identify and prioritize air
toxics, emission source types and locations which are of greatest potential concern in terms of
contributing to population risk.  From a national perspective, benzene is the most significant air
                         Draft for Discussion at SOLEC 2006

-------

toxic for which cancer risk could be estimated, contributing 25 percent of the average individual
cancer risk identified in this assessment. Based on EPA's national emissions inventory, the key
sources for benzene are onroad (49%) and nonroad mobile sources (19%), and open burning,
prescribed fires and wildfires (14%). EPA projects that onroad and nonroad mobile source
benzene emissions will decrease by about 60% between 1999 and 2020, as a result of motor
vehicle standards, fuel controls, standards for nonroad engines and equipment, and motor vehicle
inspection and maintenance programs.

Of the 40  air toxics showing the potential for respiratory effects, acrolein is the most significant,
contributing 91 percent of the nationwide average noncancer hazard identified in this assessment.
Note that the health information and exposure data for acrolein include much more uncertainty
than those for benzene. Based on the national emissions inventory, the key sources for acrolein
are open burning, prescribed fires and wildfires (61%), onroad (14%) and nonroad (11%) mobile
sources. The apparent dominance of acrolein as a noncancer "risk driver" in both the 1996 and
1999 national-scale assessment has led to efforts to develop an effective monitoring test method
for this pollutant. EPA projects that acrolein emissions from on-road sources will be reduced by
53% between 1996 and 2020 as a result of existing motor vehicle standards and fuel controls.
The assessment estimates that most people have a lifetime cancer risk between 1 and 25 in a
million from air toxics. This means that out of one million people, between 1 and 25 people have
increased  likelihood of contracting cancer as a result of breathing air toxics from outdoor sources,
if they were  exposed to 1999 levels over the course of their lifetime. The assessment estimates
that most urban locations have air toxics lifetime cancer risk greater than 25 in a million. Risk in
transportation corridors and some other locations are greater than 50 in a million. In contrast, one
out of every three Americans (330,000 in a million) will contract cancer during a lifetime, when
all causes  (including exposure to air toxics) are taken into account. Based on these results, the risk
of contracting cancer is increased less than 1 % due to inhalation of air toxics from outdoor
sources.

In Canada, key toxics such as benzene, mercury, dioxins, and furans are the subject of ratified and
proposed new standards, and voluntary reduction efforts.

Ambient Concentrations: A National Air Toxics Trend Site (NATTS) network was launched in
the U.S. in 2003 to detect trends in high-risk air toxics such as benzene, formaldehyde, 1,3-
butadiene, acrolein, and chromium. There are four NATTS monitoring sites in the  Great Lakes
region including Chicago, IL, Detroit, MI, Rochester, NY and Mayville, WI. Some ambient
trends have also been found from existing monitoring networks.  Average annual urban
concentrations of benzene have decreased 60% in the U.S.  from  1994 to 2004.

Manganese compounds are hazardous air pollutants of special concern in the Great Lakes region.
They  are emitted by iron and steel production plants, power plants, coke ovens, and many smaller
metal processing facilities. Exposures to elevated concentrations of manganese are harmful to
human health and have been associated with subtle neurological effects, such as slowed eye-hand
coordination. The  most recent NATA results identify manganese compounds  as the largest
contributor to neurological non-cancer health risk in the U.S. Modeled estimates of ambient
manganese compounds in all 3222 U.S. counties show that among the 50 counties  with the
highest concentrations nation-wide, 20 are located in Region 5. The median average annual
                         Draft for Discussion at SOLEC 2006

-------
                                                                                   	
manganese concentration at 21 trend sites showed a 14.7% decline between 2000 and 2004.
Additional years of data will be needed to confirm this apparent trend.

In Ontario, average annual urban concentrations of benzene, toluene, and xylene have decreased
about 45%, 42%, and 50% respectively from 1995 to 2004.

Emissions: The Great Lakes Toxics Inventory is an ongoing initiative of the regulatory agencies
in the eight Great Lakes States and the Province of Ontario. Emissions inventories have been
developed for 1996, 1997, 1998, 1999, 2001, and 2002 but different approaches were used to
develop these inventories making trend analysis difficult.

In Canada, emissions are also being tracked through the National Pollutant Release Inventory
(NPRI). The NPRI includes information on some of the substances listed by the Accelerated
Reduction/Elimination of Toxics (ARET) program. Significant voluntary reductions  in toxic
emissions have been reported through the ARET program.

In the U.S., emissions are also being tracked through the National Emissions Inventory (NEI) and
the Toxics Release Inventory (TRI). NEI data indicate that national U.S. air toxic emissions have
dropped approximately 42% between the 1990 and 2002, though emission estimates  are subject to
modification and the trends are different for different compounds.  The 1999 NEI also showed
that Region 5 had the highest manganese emissions of all EPA Regions, contributing 36.6% of all
manganese compounds emitted nation-wide.

The TRI, which began in 1988, contains information on releases of nearly 650 chemicals and
chemical categories from industries, including manufacturing, metal and coal mining, electric
utilities, and commercial hazardous waste treatment, among others. Although the TRI has
expanded and changed over the years, it is still possible to ascertain trends over time for core sets
of toxics. The total reported air emissions of the TRI 1988 Core Chemicals (299 chemicals) in the
eight Great Lakes states have decreased by about 78% from 1988 to 2004. According to the TRI
manganese emissions from point sources declined between 1988 and 2003 both nationally
(26.2%) and in Region 5 (36.7%). Year-to-year variability in manganese emissions is high,
however, and recent emissions data (1996-2003) suggest a weaker trend:  emissions dropped 7.6%
and 12.4% nation-wide and in Region 5, respectively.

Regional Pollutants

Ground-Level Ozone (O3)
Ozone is generally considered a secondary pollutant, which forms from reactions of precursors
(VOCs - volatile organic compounds and NOx - nitrogen oxides) in the presence of heat and
sunlight. Ozone is a problem pollutant over broad areas of the Great Lakes region, except for the
Lake Superior basin. Local onshore circulations around the Great Lakes can exacerbate the
problem, as pollutants can remain trapped for days below the maritime/marine inversion (this
forms when a layer of warm air moves to lie over colder marine air, thus trapping the colder air).
Consistently high levels are found in provincial parks near Lakes Huron and Erie, and western
Michigan is impacted by transport across Lake Michigan from Chicago.
                         Draft for Discussion at SOLEC 2006

-------

Ambient Concentrations'. In 2004, ozone levels in the U.S. showed continued improvement.
National assessments find some uneven improvement in peak levels, but with indications that
average levels may be increasing on a global scale. Ozone levels are still decreasing nationwide,
but the rate of decrease for 8-hour ozone levels has slowed since 1990. The Great Lakes region
has experienced smaller decreases than nationwide averages (Figure 1). Many of the
improvements in ozone concentrations during these times have been a result of local emission
reductions in urban areas.

To address the regional transport of ozone and ozone-forming pollutants in the eastern half of the
country, the U.S. EPA developed a program to reduce regional NOx emissions called the NOx
State Implementation Plan (SIP) Call in 2002. An analysis of 2002-2004 ozone data show that
the NOx SIP Call achieved an additional 4 percent reduction in seasonal 8-hour ozone
concentrations.  It is important to note that weather conditions in 2004 were not conducive to
ozone formation, and that ozone levels in 2005 and 2006 could be higher than in 2004 depending
on weather conditions. The NOx SIP Call also appears to have caused a gradual decline in 8-hour
daily maximum ozone concentrations (Figure 2).

Since the SOGL 2005 Indicator Report, the 1-hour ozone standard was revoked in the U.S. and
all 6 nonattainment areas in the Great Lakes basin were reclassified. Now there are 28 areas
covering 70 counties in the Great Lakes basin designated as nonattainment for the 8-hour ozone
standard (Chicago-Gary-Lake Co, IL-IN metropolitan area; South Bend/Elkhart, IN; LaPorte
County, IN; Fort Wayne, IN;  Detroit-Ann Arbor metro area, MI; Flint metro area, MI; Grand
Rapids metro area, MI; Muskegon County, MI; Allegan County, MI; Huron County, MI;
Kalamazoo-Battle Creek metro area, MI; Lansing-East Lansing metro area, MI; Benton Harbor
area, MI; Benzie County, MI; Cass County, MI; Mason County, MI; Jamestown, NY; Buffalo-
Niagara Falls metro area, NY; Rochester metro area, NY; Jefferson County, NY; Toledo metro
area, OH; Cleveland-Akron-Lorain metro area, OH; Erie, PA; Milwaukee-Racine metropolitan
area, WI; Sheboygan County, WI; Manitowoc County, WI; Kewaunee County, WI; and Door
County, WI).

In Ontario, ozone concentrations continued to exceed Ontario's Ambient Air Quality Criterion
(AAQC). In 2004, 28 of the 37 ambient Air Quality Index (AQI) monitoring stations in Ontario
recorded exceedences of the 1-hour ozone AAQC on  at least one occasion. Although the ozone
levels continue to exceed Ontario's AAQC, the 1-hour maximum ozone concentrations recorded
in Ontario have, on average, decreased by 13% from 1980 to 2004. Over the past 10 years (1995
to 2004), the annual composite means of one-hour ozone maximum concentrations have
decreased by about 4%.   In fact, the year 2004 recorded the lowest one-hour ozone maximum (84
ppb) over the last 25 years. This is partly related to the lack of weather conditions conducive to
formation of ground-level ozone in 2004; however, it also indicates that many of the efforts to
curb emissions and improve the air quality in Ontario are working.

However, Ontario has experienced an overall increasing trend in seasonal mean ozone
concentrations over the same  25-year period. The summer and winter seasonal ozone means have
increased by approximately 25% and 44%, respectively (Figure 3). The increase of the summer
mean is related to meteorological conditions and the transport of ozone and its precursors into
Ontario, whereas the increase of the winter mean indicates an increase in background
                         Draft for Discussion at SOLEC 2006

-------
                                                                                    	
concentrations of ozone throughout Ontario. Similar increases in the background concentrations
of ozone have been found in other parts of North America.

Although Ontario is not required to report on the new Canada-wide Standard (CWS) for ozone
until 2006, data from 2002-2004 indicate that all but one monitoring site (Thunder Bay) in
Ontario exceeded the ozone CWS of 65 ppb based on the 4th highest ozone eight-hour running
average over three consecutive years.

Emissions: In the U.S., VOC emissions from anthropogenic sources decreased 32% from 1990 to
2002. The rate of reduction in the Great Lakes basin was slightly less than the national average.
In 2002, VOC emissions from biogenic sources were estimated to determine the relative
contribution of natural versus anthropogenic sources. It was estimated that biogenic emissions
contributed approximately 71% of all VOC emissions in the country. NOx emissions in the U.S.
have also decreased 18% from 1990 to 2002.

In Ontario, man-made VOC emissions have decreased about 27 percent from 1990 to 2002. The
reductions are mostly attributable to the transportation and petroleum refining sectors. VOC
emissions in all of Canada have decreased 22 % over the same time period.  Canadian NOx
emissions have increased nationally by about 5% between 1990 and 2002; however, emissions
have decreased by about 1 1 % in Ontario over the same time period.
This fraction of particulate matter (diameter of 2.5 microns or less) is a health concern because it
can penetrate deeply into the lung, in contrast to larger particles. PM2.5 is primarily a secondary
pollutant produced from both natural and man-made precursors (SO2, NOX, and ammonia).

Ambient Concentrations: A CWS for PM2.5 of 30 (ig/m3 was established in June 2000.
Achievement of the standard is based on the 3 -year average of the annual 98th percentiles of the
daily, 24-hour (midnight to midnight) average concentrations. As PM2.5 monitoring has only
begun quite recently, there is not enough data to show any national long-term trends. Although
Ontario is not required to meet the CWS for fine particulate matter until 2010 and begin reporting
on progress towards meeting the new CWS until 2006, data from 2004 indicate that many areas in
Ontario have recorded 98th percentile daily averages of PM2.5 above 30 ug/m3 (Figure 4). In
Ontario, during summer episodes, PM2.5 mainly consists of sulphate particles.

In the U.S., annual average PM2 5 concentrations in 2004 were the lowest since nationwide
monitoring began in 1999. The trend is based on measurements collected at 707 monitoring
stations that have sufficient data to assess trends over that period. Concentrations in 2004
represent an 11% decrease since 1999.  The Great Lakes region has experienced a slightly greater
decline than the national average. In 2004, the average 24-hour PM2 5 concentration was also
1 1 % lower than the average 1 999 level. 24-hour PM2 5 concentrations in the Great Lakes region
decreased at nearly the same rate as the national trend over this  time. Despite some uncertainties,
the reductions in PM2.5 concentrations in the Great Lakes region appear to be largely a result of
emission reduction at sources that contribute to the formation of carbon-containing particles
(Figure  5). Direct emissions of carbon-containing particles include motor vehicles and fuel
combustion.
                         Draft for Discussion at SOLEC 2006

-------


There are three areas in the Great Lakes region that are designated nonattainment for the PM2.5
standard (Chicago-Gary-Lake Co, IL-IN metropolitan area; Detroit-Ann Arbor, MI metro area;
and the Cleveland-Akron-Lorain, OH metro area).

Emissions: In the U.S., direct emissions from anthropogenic sources decreased 27 percent
nationally between 1990 and 2002; however, this decreasing trend does not account for the
formation of secondary particles. The largest absolute reduction in PM2.5 emissions was seen in
the fuel combustion source category (347,000 tons and 38%); while, the largest relative reduction
in PM2.5 emissions was in the on-road vehicle category (175,000 tons and 54%).

In Canada, emissions (not including open sources such as road dust, construction operations, and
forest fires) have decreased nationally by about 14% between 1990 and 2002. However, total
PM2.5 emissions including open sources have increased by 6% in Canada over this time period.
Ontario has experienced similar trends over this time period.

Pressures
Continued economic growth, population growth, and associated urban sprawl are threatening to
offset emission reductions achieved by policies currently in place, through both increased energy
consumption and vehicles miles traveled. The changing climate may affect the frequency of
weather conditions conducive to high ambient concentrations of many pollutants. There is also
increasing evidence of changes to the atmosphere as a whole. Continuing health research is both
broadening the number of toxics, and producing evidence that existing standards should be
lowered.

Management Implications
Major pollution reduction efforts continue in both U.S. and Canada. In Canada, new ambient
standards for particulate matter and ozone have been endorsed, with a 2010 achievement date.
This will involve updates at the Federal level and at the provincial level (the Clean Air Action
Plan, and Ontario's Industry Emissions Reduction Plan). Toxics are also addressed at both levels.
The Canadian Environmental Protection Act (CEPA) was recently amended.

In the U.S., new, more protective ambient air standards have been promulgated for ozone and
particulate matter. MACT (Maximum Available Control Technology) standards continue to be
promulgated for sources of toxic air pollution. USEPA has also begun looking at the risk
remaining after emissions reductions for industrial sources take effect.

At the international level, Canada and the U.S. signed the Ozone Annex to the Air Quality
Agreement in December 2000. The Ozone Annex commits both countries to reduce emissions of
NOX and VOCs, the precursor pollutants to ground-level ozone, a major component of smog.
This will help both countries attain their ozone air quality goals to protect human health and the
environment.  Canada estimates that total NOX reduction in the Canadian transboundary region
will be between 35% and 39% of the 1990 levels by 2010. Under the Clean Air Action Plan,
Ontario is also committed to reducing  provincial emission of NOX and VOCs by 45%  of 1990
levels by 2015, with interim targets of 25% by 2005.
                         Draft for Discussion at SOLEC 2006

-------
The U.S. estimates that the total NOx reductions in the U.S. transboundary region will be 36%
year-round by 2010 and 43% during the ozone season. Canada and the U.S. have also undertaken
cooperative modeling, monitoring, and data analysis and developed a work plan to address
transboundary PM issues. PM2.5 networks will continue to develop in both countries, to
determine ambient levels, trends, and consequent reduction measures. Review of standards or
objectives will continue to consider new information. Efforts to reduce toxic pollutants will also
continue under North America Free Trade Agreement and through United Nations-Economic
Commission for Europe protocols. The U.S. is continuing its deployment of a national air toxics
monitoring network.

Comments from the author(s)
Updated 2005 emissions data from Canada's National Pollutant Release Inventory (NPRI) is
expected to become available in the fall of 2006. Environment Canada is also expected to release
a five-year comprehensive report on the progress towards the Canada-wide Standards (CWS) for
PM and ozone in the fall of 2006.

These new data will be incorporated into the indicator report before finalization.

Acknowledgments
Author: Todd Nettesheim, U.S. Environmental Protection Agency, Great Lakes National Program
Office, Chicago, IL.

Reviewers: Kate McKerlie, Diane Sullivan, Brenda Koekkoek, Ken Smith, John Ayres, Clarisse
Kayisire, Carmen Bigras, Philip Blagden, and Jay Barclay, Environment Canada, Gatineau,
Quebec.
Hong (Holly) Lin and Fred Conway, Environment Canada, Meteorological Service of Canada,
Downsview, Ontario.
Melynda Bitzos and Yvonne Hall, Ontario Ministry of the Environment, Ontario, Canada.

Data Sources
Canada-United States Air Quality Committee, Subcommittee on Scientific Cooperation, in
support of Canada-United States Air Quality Agreement. 2004. Canada-United States
Transboundary Particulate Matter Science Assessment. En56-203/2004E. ISBN: 0-662-38678-7.
http://www.ec.gc.ca/pdb/can us/canus links e.cfrn. last accessed September 5, 2006.

Environment Canada. 2006a. Criteria Air Contaminants (CAC) Emission Summaries, July 2006.
http://www.ec.gc.ca/pdb/cac/Emissionsl990-2015/emissionsl990-2015 e.cfrn. last accessed
September 6, 2006.

Environment Canada. 2006b. 2002 National Pollutant Release Inventory Data.
http://www.ec.gc.ca/pdb/npri/npri home e.cfrn. last accessed September 6, 2006.

Environment Canada. 2006c. National Air Pollution Surveillance Network, http://www.etc-
cte.ec.gc.ca/napsstations/main.aspx. last accessed September 6, 2006.
                         Draft for Discussion at SOLEC 2006

-------

Environment Canada. 2005. Border Air Quality Strategy: Great Lakes Basin Airshed
Management Framework Pilot Project. En4-48/2005E. ISBN: 0-662-40522-6.
http://www.ec.gc.ca/cleanair-airpur/caol/canus/great lakes/index e.cfm. last accessed September
5. 2006; and http://www.epa.gov/airmarkets/usca/pilotproiect.html. last accessed September 5,
2006.

Environment Canada. 2003a. Clean air in Canada: 2003 progress report on particulate matter and
ozone. ISBN 0-662-34514-2. http://www.ec.gc.ca/cleanair-
airpur/CAOL/air/PM resp  03/toc  e.html. last accessed September 6, 2006.

Environment Canada. 2003b. Cleaner air through cooperation: Canada - United States progress
under the Air Quality Agreement 2003. ISBN 0-662-34082-5.
http://www.epa.gov/airmarkets/usca/brochure/brochure.htm
, last accessed April 17, 2004.

Environment Canada. 2003c. Environmental signals: Canada's national environmental indicator
series 2003. http://www.ec. gc. ca/soer-ree/English/default.cfm. last accessed September 5, 2006.

Environment Canada. 2003d. Environment Canada Performance Report: for the period ending
March 31, 2003. David Anderson, Minister of the Environment.
http://www.ec.gc.ca/dpr/EC DPR  March 31 2003 EN-Oct6.pdf. last accessed June 17, 2004.

Great Lakes Commission. 2002 Inventory of Toxic Air Emissions: Point, Area and Mobile
Sources,  http://www.glc.org/air/inventory/2002/, last accessed September 5, 2006.

NARSTO. 2000. An assessment of tropospheric ozone: a North American perspective.
http://www.cgenv.com/Narsto/, last accessed June 30, 2004.

Ontario Ministry of the Environment (OMOE). 2006. Air Quality in Ontario  2004 Report.
Queen's Printer for Ontario. ISBN  1710-8128 or 0-7794-9921-2.
http://www.airqualitvontario.com/press/publications.cfm, last accessed September 6, 2006.

Ontario Ministry of the Environment (OMOE). 2006b. Rationale for the Development of Ontario
Air Standards for Lead and Lead Compounds, June 2006. Standards Development Branch.
http://www.ene.gov.on.ca/envision/AIR/airquality/standards.htmtfcontaminants. last accessed
September 6, 2006.

Ontario Ministry of the Environment (OMOE). 2006c. Rationale for the Development of Ontario
Air Standards for Total Reduced Sulphur, June 2006. Standards Development Branch.
http://www.ene.gov.on.ca/envision/AIR/airquality/standards.htmtfcontaminants. last accessed
September 6, 2006.

Ontario Ministry of the Environment (OMOE). 2005. Transboundary Air Pollution in Ontario.
Queen's Printer for Ontario, http://www.airqualityontario.com/press/publications.cfm. last
accessed September 6, 2006.
                         Draft for Discussion at SOLEC 2006

-------
Ontario Ministry of the Environment (OMOE). 2006. Air Quality in Ontario 2004 Report.
Queen's Printer for Ontario, http://www.airqualityontario. com/press/publications, cfm. last
accessed September 6, 2006.

U.S. Environmental Protection Agency (USEPA). 2006a. 2007 Report on the Environment
(ROE) Technical Document, http ://www.epa. gov/indicators/. last accessed September 5, 2006.

U.S. Environmental Protection Agency (USEPA). 2006b. Air Emission Trends - Continued
Progress Through 2005. http://www.epa.gov/airtrends/. last accessed September 6, 2006.

U.S. Environmental Protection Agency (USEPA). 2006c. National-Scale Air Toxics Assessment
for 1999: Estimated Emissions, Concentrations and Risk.
http://www.epa.gov/ttn/atw/natal999/index.html, last accessed September 5, 2006.

U.S. Environmental Protection Agency (USEPA). 2006d. Green book: non-attainment areas for
criteria pollutants. Office of Air Quality Planning and Standards.
http://www.epa.gov/air/oaqps/greenbk/. last accessed September 6, 2006.

U.S. Environmental Protection Agency (USEPA). 2006e. Toxics Release Inventory Program.
http://www.epa.gov/tri/, last accessed June 24, 2004.

U.S. Environmental Protection Agency (USEPA). 2006f. 2002 National Emissions Inventory
Data & Documentation, http ://www. epa. gov/ttn/chief/eiinformation.html, last accessed
September 6, 2006.

U.S. Environmental Protection Agency (USEPA). 2005a. Evaluating Ozone Control Programs in
the Eastern United States: Focus on the NOx Budget Trading Program, 2004. EPA454-K-05-001.
http://www.epa.gov/airtrends/2005/ozonenbp/. last accessed September 5, 2006.

U.S. Environmental Protection Agency (USEPA). 2005b. Border Air Quality Strategy: United
States-Canada Emissions Cap and Trading Feasibility Study. EPA 430-R-05-005.
http://www.epa.gov/airmarkets/usca/pilotproiect.html. last accessed September 5, 2006.

U.S. Environmental Protection Agency (USEPA). 2004a. The Particle Pollution Report: Current
Understanding of Air Quality and Emissions through 2003. EPA 454-R-04-002.
http://www.epa.gov/air/airtrends/aqtrnd04/pm.html. last accessed September 5, 2006.

U.S. Environmental Protection Agency (USEPA). 2004b. United States-Canada  Air Quality
Agreement: 2004 Progress Report. EPA 430-R-04-007.
http://www.epa.gov/airmarkets/usca/index.html. last accessed September 5, 2006; and
http://www.ec.gc.ca/pdb/can us/canus  links e.cfrn. last accessed September 5, 2006.

List of Figures
Figure 1. Trends in Fourth Highest Daily Maximum 8-hour ozone concentration (ppm) by EPA
Region 1980-2004.
                        Draft for Discussion at SOLEC 2006

-------
Source: Figure 004-4. Ambient ozone concentrations, 1980-2004, by EPA region; 2007 Report on
the Environment (ROE) Technical Document, http ://www. epa. gov/indicators/, last accessed
September 5, 2006.

Figure 2. Rural Seasonal Average 8-hour Maximum Ozone Concentrations by EPA Region,
1997-2004.
Source: Sidebar "Ozone Reduction in Rural Areas Shows Regional Improvements" on page 20 of
U.S. Environmental Protection Agency (USEPA). 2005a. Evaluating Ozone Control Programs in
the Eastern United States: Focus on the NOx Budget Trading Program, 2004. EPA454-K-05-001.
http://www.epa.gov/airtrends/2005/ozonenbp/. last accessed September 5, 2006.

Figure 3. Trend of Ozone Seasonal Means at Sites Across Ontario (1980-2004).
Source: Figure 2.5 of Ontario Ministry of the Environment. Air Quality in Ontario 2004 Report.
Queen's Printer for Ontario, 2006. . ISBN 1710-8128 or 0-7794-9921-2.
http://www.airqualitvontario.com/press/publications.cfm. last accessed September 6, 2006.

Figure 4. PM2.s Levels at Selected Sites Across Ontario, 98th Percentile PM2.5 Daily Average
(2004).
Source: Figure 3.4 of Ontario Ministry of the Environment. Air Quality in Ontario 2004 Report.
Queen's Printer for Ontario, 2006. . ISBN 1710-8128 or 0-7794-9921-2.
http://www.airqualitvontario.com/press/publications.cfm. last accessed September 6, 2006.

Figure 5. Trends of PM2.5 and its chemical constituents in the Industrial Midwest of the U.S.,
1999-2003.
Source: Figure 16 of U.S. Environmental Protection Agency (USEPA). 2004a. The Particle
Pollution Report: Current Understanding of Air Quality and Emissions through 2003.  EPA 454-
R-04-002. http://www.epa.gov/air/airtrends/aqtrnd04/pm.html. last accessed September 5, 2006.

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
                              State of the Great Lakes 2007 - Draft
     Figure 004-4 . Ambient o zone concent rations, 1980-2004,  by E PA region

      0.16.
   a  o. 14.
      5.12.
   I
   §
   "   0.10 -,
      0.08

  I
  E   0.06 -
      0.04 -
  I
      0.00 -
            NArt QS

                                                   Region 1
                                                 — Region 2
                                                   R egion 3
                                                   Region 4
                                                   R eglon 6
                                                   Region 7
                                                   R egion 8
                                                   R egion 9
                                                   Region 1 0
                                               ...... National
         80 93 Si 83 $4 85 86 8? 88 89 90 91 92 93 94 95 96 97 98 99 00 01 02 03 04
  Source: EPA's Air Quality System.
Figure 1. Trends in Fourth Highest Daily Maximum 8-hour ozone concentration (ppm) by EPA
Region 1980-2004.
Source: Figure 004-4. Ambient ozone concentrations, 1980-2004, by EPA region; 2007 Report on
the Environment (ROE) Technical Document, http ://www. epa. gov/indicators/. last accessed
September 5, 2006.
14
Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
    Rural Seasonal Average 8-hour Daily Maximum Ozone by Region,
                                  1997-2004
                                                     Northeast
                                                     Mid-Atlantic
                                                     Southeast
                                                     Midwest
  Source: EPA

  Note: Ozone concentrations are in parts per billion (ppb)
Figure 2. Rural Seasonal Average 8-hour Maximum Ozone Concentrations by EPA Region,
1997-2004.
Source: Sidebar "Ozone Reduction in Rural Areas Shows Regional Improvements" on page 20 of
U.S. Environmental Protection Agency (USEPA). 2005a. Evaluating Ozone Control Programs in
the Eastern United States: Focus on the NOx Budget Trading Program, 2004. EPA454-K-05-001.
http://www.epa.gov/airtrends/2005/ozonenbp/, last accessed September 5, 2006.
                       Draft for Discussion at SOLEC 2006
15

-------
                                    State of the Great Lakes 2007 - Draft
   40
   35
   30
   25
   20
o  15
O
    10
                          Trend of Ozone Seasonal Means at Sites Across Ontario
                                           (1980-2004)

Summer Mean

Winter Mean

                                                                                   •   •
      Note: Based on data from 22 ozone sites operated over 25 years.
          Seasonal definitions -Summer {May to September!: Winter (January to April, October to December).
    Figure 3. Trend of Ozone Seasonal Means at Sites Across Ontario (1980-2004).
    Source: Figure 2.5 of Ontario Ministry of the Environment. Air Quality in Ontario 2004 Report.
    Queen's Printer for Ontario, 2006. .  ISBN 1710-8128 or 0-7794-9921-2.
    http://www.airqualitvontario.com/press/publications.cfm. last accessed September 6, 2006.
    16
                Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
                           PM2i. Levels at Selected Sites Across Ontario
                                 SS9"' Percent! le PM:.E Daily Average
                                           (2004)
3& -


c
Q
'£• 20 -

Q
U

o -









































—















^















—















—















— 1















—















r—















— i















. —
































































—
































—





















 Note: Displayed sites are selected based on future requirements for Canada-wide Standard (CWS) reporting.
     Toronto reporting is based on Toronto Downtown, Toronto North. Toronto East and Toronto West sites.
Figure 4. PM2.5 Levels at Selected Sites Across Ontario, 98th Percentile PM2.5 Daily Average
(2004).
Source: Figure 3.4 of Ontario Ministry of the Environment. Air Quality in Ontario 2004 Report.
Queen's Printer for Ontario, 2006. . ISBN 1710-8128 or 0-7794-9921-2.
http://www.airqualityontario.com/press/publications.cfm, last accessed September 6, 2006.
                          Draft for Discussion at SOLEC 2006
17

-------
                           State of the Great Lakes 2007 - Draft
              Industrial Midwest
   16-


   14-


   12-

E
•& 10-
I  >
i
£  6>
      4-
    2-
      0-
               119 PM: 5 Monitor rig Sites
                                 -9%
         Sulfate
        PM2.E Remainder
         (mostly carbon)

        Nitrate
                                 -5%
                                -17%
                                 4%
       'Guslar
      1999  2000  2001   2002  2003

                    Year

Figure 5. Trends of PIVb.s and its chemical constituents in the Industrial Midwest of the
U.S., 1999-2003.
Source: Figure 16 of U.S. Environmental Protection Agency (USEPA). 2004a. The Particle
Pollution Report: Current Understanding of Air Quality and Emissions through 2003. EPA 454-
R-04-002. http://www.epa.gov/air/airtrends/aqtrnd04/pm.html, last accessed September 5, 2006.
18
                    Draft for Discussion at SOLEC 2006

-------
Coastal Wetland Invertebrate Community Health
Indicator #4501

Note: This indicator has not yet been put into practice. The fol-
lowing evaluation was constructed using input from investiga-
tors collecting invertebrate community composition data from
Great Lakes coastal wetlands over the last several years. Neither
experimental design nor statistical rigor has been used to specif-
ically address the status and trends of invertebrate communities
of coastal wetlands of the five Great Lakes.

Assessment: Not Assessed

Purpose
  To directly measure specific components of invertebrate com-
munity composition; and
  To infer the chemical, physical and biological integrity and
range of degradation of Great Lakes coastal wetlands.

State of the Ecosystem
Development of this indicator is still in progress. Thus, the state
of the ecosystem could not be determined using the wetland
invertebrate community health indicator during the last 2 years.

Teams of Canadian and American researchers from several
research groups (e.g. the Great Lakes Coastal Wetlands
Consortium, the Great Lakes Environmental Indicators project
investigators, the U.S. Environmental Protection Agency
(USEPA) Regional Environmental Monitoring and Assessment
Program (REMAP) group of researchers, and others) sampled
large numbers of Great Lakes wetlands during the last two years.
They have reported an array of invertebrate communities in
Great Lakes wetlands in presentations at international meetings,
reports, and peer-reviewed journals.

In 2002 the Great Lakes Coastal Wetlands Consortium conduct-
ed extensive surveys of wetland invertebrates of the 4 lower
Great Lakes. These data are not entirely analyzed to date.
However, the Consortium-adopted Index of Biotic Integrity (IBI,
Uzarski et al. 2004) was applied in wetlands of northern Lake
Ontario. The results can be obtained from Environment Canada
(Environment Canada and Central Lake Ontario Conservation
Authority 2004).

Uzarski et al. (2004) collected invertebrate  data from 22 wet-
lands in Lake Michigan and Lake Huron during 1997 through
2001. They determined that wetland invertebrate communities of
northern Lakes Michigan and Huron generally produced the
highest IBI scores. IBI  scores were primarily based on richness
and abundance of Odonata,  Crustacea plus Mollusca taxa rich-
ness, total genera richness, relative abundance Gastropoda, rela-
tive abundance Sphaeriidae, Ephemeroptera plus Trichoptera
taxa richness, relative abundance Crustacea plus Mollusca, rela-
tive abundance Isopoda, Evenness, Shannon Diversity Index,
and Simpson Index. Wetlands near Escanaba and Cedarville,
Michigan, scored lower than most in the area. A single wetland
near the mouth of the Pine River in Mackinac County, MI, con-
sistently scored low, also. In general, all wetlands of Saginaw
Bay scored lower than those of northern Lakes Michigan and
Huron. However, impacts are more diluted near the outer bay
and IBI scores reflect this. Wetlands near Quanicassee and
Almeda Beach, MI, consistently scored lower than other
Saginaw Bay sites.

Burton and Uzarski (unpublished) also studied drowned river
mouth wetlands of eastern Lake Michigan quite extensively
since  1998. Invertebrate communities of these systems show lin-
ear relationship with latitude. However, this relationship also
reflects anthropogenic disturbance. Based on the metrics used
(Odonata richness and abundance, Crustacea plus Mollusca rich-
ness, rotal genera richness, relative abundance Isopoda, Shannon
Index, Simpson Index, Evenness, and relative abundance
Ephemeroptera),  the sites studied were placed in increasing com-
munity health in  the order Kalamazoo, Pigeon, Muskegon,
White, Pentwater, Pere Marquette, Manistee, Lincoln, and
Betsie. The most impacted systems of eastern Lake Michigan are
located along southern edge and impacts decrease to the north.

Wilcox et al. (2002) attempted to develop wetland IBIs for the
upper Great Lakes using microinvertebrates. While they found
attributes that showed promise during a single year, they con-
cluded that natural  water level changes were likely to alter com-
munities and invalidate metrics. They found that Siskiwit Bay,
Bark Bay, and Port Wing had the greatest overall taxa richness
with large catches of cladocerans. They ranked microinvertebrate
communities of Fish Creek and Hog Island lower than the other
four western Lake Superior sites. Their work in eastern Lake
Michigan testing potential metrics placed the sites studied in
decreasing community health in the order Lincoln River, Betsie
River, Arcadia Lake/Little Manistee River, Pentwater River, and
Pere Marquette River. This order was primarily based on the
median number of taxa, the median Cladocera genera richness,
and also a macroinvertebrate metric (number of adult
Trichoptera species).

Pressures
Physical alteration  and eutrophication of wetland ecosystems
continue to be a threat to invertebrates of Great Lakes coastal
wetlands. Both can promote establishment of non-native vegeta-
tion, and physical alteration can destroy plant communities alto-
gether while changing the natural hydrology to the system.
Invertebrate community composition is directly related to vege-
tation type and densities; changing either of these components
will negatively impact the invertebrate communities.

                                                        177

-------

Acknowledgments
Authors: Donald G. Uzarski, Annis Water Resources Institute,
Grand Valley State University, Lake Michigan Center, 740 W.
Shoreline Dr., Muskegon, MI, 49441; and
Thomas M. Burton, Departments of Zoology and Fisheries and
Wildlife, Michigan State University, East Lansing, MI, 48824.

Sources
Environment Canada and Central Lake Ontario Conservation
Authority. 2004. Durham Region Coastal Wetland Monitoring
Project: year 2 technical report. Downsview, ON. ECB-OR.

Uzarski, D.G., Burton, T.M., and Genet,  J.A. 2004. Validation
and performance of an invertebrate index of biotic integrity for
Lakes Huron and Michigan fringing wetlands during a period of
lake level decline. Aquat. Ecosystem Health & Manage.
7(2):269-288.

Wilcox, D.A., Meeker, J.E., Hudson, P.L., Armitage, B.J., Black,
M.G., and Uzarski, D.G. 2002. Hydrologic variability and the
application of index of biotic integrity metrics to wetlands: a
Great Lakes  evaluation. Wetlands 22(3):588-615
Authors' Commentary
Progress on indicator development has been substantial, and
implementation of basin-wide sampling to indicate state of the
ecosystem should be possible before SOLEC 2006.

Last Updated
State of the Great Lakes 2005
                                                                                         2007
178

-------

Coastal Wetland Fish Community Health
Indicator ID: 4502

Overall Assessment: N/A

Note: This indicator has not yet been put into practice. The following evaluation was
constructed using input from investigators collecting fish community composition data from
Great Lakes coastal wetlands over the last several years. Neither experimental design nor
statistical rigor has been used to specifically address the status and trends offish
communities of coastal wetlands of the five Great Lakes.

Purpose
To assess the fish community composition and to infer suitability of habitat and water quality for
Great Lakes coastal wetland fish communities.

State of the Ecosystem
Development of this indicator is still in progress. Fish indices of biological integrity have been
proposed for selected parts of the ecosystem (e.g., Lake Erie river mouths (Thoma 1999)
Michigan and Ontario coastal wetlands (Uzarski et al. 2005), and coordinated basinwide
sampling has recently been completed by several groups. Thus, progress on indicator
development has been substantial, and assessment of data derived from sampling conducted
between 2002 and 2005 to indicate the state of the ecosystem should be possible before the next
SOLEC. Teams of Canadian and American researchers from several research groups (e.g., the
Great Lakes Coastal Wetlands Consortium of the Great Lakes Commission (GLCWC), the U.S.
EPA Star Grant funded Great Lakes Environmental Indicators group in Duluth, MN (GLEI), a
group of Great  Lakes Fishery Commission researchers led by Patricia Chow-Fraser of McMaster
University (GLFC), the U.S. EPA REMAP group of researchers led by Tom Simon, and others)
have sampled large numbers of Great Lakes wetlands during the last 5 years using comparable
methods. They have reported on an array of fish communities in Great Lakes wetlands in
presentations at international meetings and in reports. These data are now beginning to appear in
refereed journals as individual studies (Uzarski et al. 2005, Seilhamer and Chow Fraser 2006)
Work is also underway to integrate the datasets for true basinwide assessment (e.g., Brazner et al.
2006; Bhagat et al. in review). The composition offish communities is related to plant
community type within wetlands and, within plant community type, is related to amount of
certain types of anthropogenic disturbance (Uzarski et al. 2005; Wei et al. 2004, Seilhamer et al.
2006; Johnson  et al.  2006), especially water quality as affected by urban and agricultural
development (Seilhamer and Chow Fraser 2006; Bhagat et al. in review). Uzarski et al. (2005)
found no relationship between wetland fish composition and Great Lake suggesting that fish
communities of any single Great Lake were more impacted than any other. However, of the 61
wetlands sampled in 2002 from all five lakes, Lakes Erie and Ontario tended to have more
wetlands containing  cattail communities (a plant community type that correlates with nutrient
enrichment), and the fish communities found in cattails tended to have lower richness and
diversity than fish communities found in other  vegetation types. In contrast,  Thoma (1999) and
Johnson et al. (2006) were unable to find coastal wetlands on the US side of Lake Erie that
experienced minimal anthropogenic disturbances. Wetlands found in northern lakes Michigan and
Huron tended to have relatively high quality coastal wetland fish communities. The seven
wetlands sampled in Lake Superior contained relatively unique vegetation types so fish
                        Draft for Discussion at SOLEC 2006

-------

communities of these wetlands were not directly compared with those of wetlands of other lakes.
When the fish communities of reference wetlands are compared across the entire Great Lakes, the
most similar sites come from the same ecological province rather than from any single Great
Lake or specific wetland types. Data from several GLEI project studies indicate that the
characteristic groups of fish species in reference wetlands from each ecological province tend to
have similar water temperature and aquatic productivity preferences. When a wetland becomes
affected by human development, the fish community changes to the fish community typical of a
warmer, richer, more  southerly wetland. This finding may help us anticipate the likely effects of
regional climate change on the fish communities of Great Lakes coastal wetlands. Brazner et al.
looked at how 8 different candidate fish IBI components varied by lake, wetland type, ecological
province and anthropogenic stress at 80 wetlands across the entire US Great Lakes. Overall, each
of these 4 features explained approximately equal amounts of variation in those components.

John Brazner and co-workers from the U.S. EPA Laboratory in Duluth, MN sampled fishes of
Green Bay, Lake Michigan, wetlands in 1990, 1991, 1995, 2002, and in 2003. They sampled
three lower bay and one middle bay wetland in 2002 and 2003 and their data suggested that these
sites were improving in water clarity and plant cover, and supported a greater diversity of both
macrophyte and fish species, especially more centrarchid species, than they had in previous years.
They also noted that the 2002, and especially 2003, year classes of yellow perch were very large.
Brazner's observations suggest that the lower bay wetlands are improving slowly and the middle
bay site seems to be remaining relatively stable in moderately good condition (J. Brazner,
personal observation). The most turbid wetlands in the lower bay were characterized by mostly
warm-water, turbidity-tolerant species such as gizzard shad, Dorosomct cepedicmum; white bass,
Morone chrysops; freshwater drum, Aplodinotus grunniens; common shiners, Luxilus cornutus,
and common carp, Cyprinus carpio, while the least turbid wetlands in the upper bay were
characterized by several centrarchid species, golden shiner, Notemigonus chrysoleucas;  logperch,
Percina caprodes; smallmouth bass, Micropterus dolomieu, and northern pike, Esox lucius.
Green sunfish, Lepomis cyanellus, was the only important centrarchid in the lower bay in 1991,
while in 1995, bluegill and pumpkinseed sunfishes,i. macrochirus andi. gibbosus, had become
much more prevalent and a few largemouth bass, M. salmoides, were also present. There were
more banded killifish, Fundulus diaphanus, in 1995 and 2003 compared with 1991  and white
perch were very abundant in 1995, as this exotic species became dominant in the bay. The upper
bay wetlands were in  relatively good condition based on the fish and macrophyte communities
that were observed. Although mean fish species richness was significantly lower in developed
wetlands across the whole bay, differences between less developed and more developed wetlands
were most pronounced in the upper bay where the highest quality wetlands in Green Bay are
found (Brazner 1997).

Round gobies, Neogobius melanostomus, were introduced to the St. Clair River in 1990 (Jude et
al. 1992), and have since spread to all of the Great Lakes. Jude studied them in many tributaries
of the Lake Huron-St. Clair River-Lake Erie corridor and found that both species (round and
tubenose gobies Proterorhinus marmoratus} were very abundant at river mouths and colonized
far upstream. They were also found at the mouth of Old Woman Creek in Lake Erie, but not
within the wetland proper. Jude and Janssen's work in Green Bay wetlands showed that round
gobies had not invaded three of the five sites sampled, but few were found in lower Green Bay
along the sandy and rocky shoreline west of Little Tail Point.
                         Draft for Discussion at SOLEC 2006

-------
Uzarski and Burton (unpublished) consistently collected a few round gobies from a fringing
wetland near Escanaba, MI where cobbles were present. In the Muskegon River-Muskegon Lake
wetland complex on the eastern shoreline, round gobies are abundant in the heavily rip-rapped
harbor entrance to Lake Michigan, Muskegon Lake, and have just begun to enter the
river/wetland complex on the east side of Muskegon Lake (D. Jude, personal observations; Ruetz,
Uzarski, and Burton, personal observations). Based on intensive fish sampling prior to 2003 at
more than 60  sites spanning all of the Great Lakes, round gobies have not been sampled in large
numbers at any wetland or been a dominant member of any wetland fish community (Jude et al.
2005). Round gobies were collected at 11 of 80 wetlands sampled by the GLEI project (Johnson
et al. unpublished data). Lapointe (2005) assessed fish-habitat associations in the shallow (<3 m)
Canadian waters of the  Detroit River in 2004 and 2005 using boat-mounted electorfishing and
boat seining techniques. The round goby avoided complex macrophytes in all seasons at upper,
mid, and downstream segments of the Detroit River. However, in 2006 beach seining surveys at
shoreline sites in Canadian waters of Lake St. Clair, the Detroit River,  and western Lake Erie,
both tubenose and round gobies were collected in areas with aquatic vegetation (L.D. Corkum,
Univ. of Windsor, unpublished data). It seems likely that wetlands may be a refuge for native
fishes, at least with respect to the influence of round gobies (Jude et al. 2005).

There is little information on the habitat preferences of the tubenose goby within the Great Lakes
with the exception of studies on the Detroit River (Lapointe 2005), Lake St. Clair and the St.
Clair River (Jude and DeBoe 1996, Pronin et al. 1997; Leslie et al. 2002). Within the Great
Lakes, tubenose goby that were studied at a limited number of sites along the St. Clair River and
on the south shore of Lake St. Clair occurred in turbid water associated with rooted submersed
vegetation (Vallisneria  americana, Myriophyllum spicatum, Potamogeton richardsonii and
Chora sp.) (Leslie et al. 2002). Few specimens were found on sandy substrates devoid of
vegetation, supporting similar findings by Jude and DeBoe (1996). Leslie et al. (2002) collected
tubenose goby in water with no or slow flow on clay or alluvium substrates, where turbidity
varies and where rooted vegetation was sparse, patchy or abundant.  Lapointe (2005) found that
the association between tubenose goby and aquatic macrophytes differed seasonally in the Detroit
River. For example, tubenose goby was strongly negatively associated with complex macrophytes
in the spring and summer, but positively associated with complex macrophytes in the fall
(Lapointe 2005). Because tubenose goby shared habitats with fishes representing most
ecoethological guilds, Leslie et al. (2002) suggested that the tubenose goby would expand its
geographic range within the Great Lakes.

Ruffe have never been found in high densities in coastal wetlands anywhere in the Great Lakes.
In their investigation of the distribution and potential impact of ruffe on the fish  community of a
Lake Superior coastal wetland, Brazner et al. (1998) concluded that coastal wetlands in western
Lake Superior provide a refuge for native fishes from  competition with ruffe. The mudflat-
preferring ruffe actually avoids wetland habitats due to foraging inefficiency in dense vegetation
that characterizes healthy coastal wetland habitats. This suggests that further degradation  of
coastal wetlands or heavily vegetated littoral habitats could lead to increased dominance of ruffe
in shallow water habitats elsewhere in the Great Lakes.

There are a number of carp introductions (see Wetland Restoration and Rehabilitation or common
carp discussion) that have the potential for substantial impact on Great Lakes fish communities,
                         Draft for Discussion at SOLEC 2006

-------
including coastal wetlands. Goldfish, Carassius auratus, are common in some shallow habitats,
and occurred along with common carp young-of-the-year in many of the wetlands we sampled
along Green Bay. In addition, there are several other carp species, e.g., grass carp,
Ctenopharyngodon idella, bighead carp Hypophthalmichthys nobilis, and silver carp,
Hypophthalmichthys molitrix that escaped aquaculture operations and are now in the Illinois
River and migrating toward the Great Lakes through the Chicago Sanitary Canal. The black carp,
Mylopharygodon piceus, has also probably been released, but has not been recorded near the
Great Lakes yet. Most of these species attain large sizes; some are planktivorous, and also eat
phytoplankton, snails, and mussels, while the grass carp eats vegetation. These species represent
yet another substantial threat to food webs in wetlands and nearshore habitats with macrophytes
(USFWS 2002).

In 2003, Jude and Janssen (unpublished data) determined that bluntnose minnows, Pimephales
notatus, and johnny darters, Etheostoma nigrum, were almost absent from lower bay wetland
sites, but comprised 22% and 6% respectively, of upper bay catches. In addition, other species,
usually associated with plants and/or clearer water, such as rock bass, sand shiners Notropis
stramineus, and golden shiners Notemigonus crysoleucus, were also present in upper bay
samples, but not in  lower bay samples. In 2003, Jude and Janssen found that there were no
alewife Alosapseudoharengus or gizzard shad in upper Green Bay site catches when compared
with lower bay wetland sites, where they composed 2.1 and 34% respectively of the catches by
number.

Jude and Pappas (1992) found that fish assemblage structure in Cootes Paradise, a highly
degraded wetland area in Lake Ontario, was very different from other less degraded wetlands
analyzed. They used ordination analyses to detect fish-community changes associated with
degradation.

Acknowledgments
Authors: Donald G. Uzarski, Annis Water Resources Institute, Grand Valley State University,
Lake Michigan Center, 740 W.  Shoreline Dr., Muskegon, MI 49441;
Thomas M. Burton, Departments of Zoology and Fisheries and Wildlife, Michigan State
University, East Lansing, MI 48824;
John Brazner, US Environmental Protection Agency, Mid-Continent Ecology Division, 6201
Congdon Blvd., Duluth, MN 55804;
David Jude, School of Natural Resources and the Environment, 430 East University, University
of Michigan, Ann Arbor, MI 48109; and
Jan J.H. Ciborowski, Department of Biological Sciences, University of Windsor, Windsor, ON,
N9B 3P4

Data Sources
Brazner, J. C. 1997. Regional, habitat, and human development influences on coastal wetland and
beach fish assemblages in Green Bay, Lake Michigan. J. Great Lakes Res. 23 (1), 36-51.

Brazner, J. C., Tanner, D. K., Jensen, D. A., Lemke, A. 1998. Relative abundance and distribution
of ruffe (Gymnocephalus cernuus) in a Lake Superior coastal wetland fish assemblage. J. Great
Lakes Res.24 (2), 293-303.
                         Draft for Discussion at SOLEC 2006

-------
Jude, D.J., Albert, D., Uzarski, D.G., and Brazner, J. 2005. Lake Michigan's coastal wetlands:
Distribution, biological components with emphasis on fish and threats. In M. Munawar and T.
Edsall (Eds.). The State of Lake Michigan: Ecology, Health and Management. Ecovision World
Monograph Series, Aquatic Ecosystem Health and Management Society, p. 439-477

Jude, D. J., Pappas, J. 1992. Fish utilization of Great Lakes coastal wetlands . J. Great Lakes Res.
18(4), 651-672.

Jude, D. J., R. H. Reider, G. R. Smith. 1992. Establishment of Gobiidae in the Great Lakes basin.
Can. J.Fish. Aquat. Sci. 49, 416-421.

Uzarski, D.G., T.M. Burton, M.J. Cooper, J. Ingram, and S. Timmermans. 2005. Fish Habitat
Use Within and Across Wetland Classes in Coastal Wetlands of the Five Great Lakes:
Development of a Fish Based Index of Biotic Integrity. Journal of Great Lakes Research
31 (supplement 1): 171-187.

Bhagat, Y. 2005. Fish indicators of anthropogenic stress at Great Lakes coastal margins:
multimetric and multivariate approaches. M.Sc. Thesis, University of Windsor.  120 p.

Bhagat, Y., J.J.H. Ciborowski, L.B. Johnson, D.G. Uzarski, T.M. Burton, S.T.A.Timmermans,
and M.J. Cooper. In review. Testing a fish index of biotic integrity for Great Lakes coastal
wetlands: stratification by plant zones. Submitted to Wetlands (June 2006)

Brazner, J.C., N.  P. Danz, G. J. Niemi, R. R. Regal,  A. S. Trebitz, R. W. Howe, J. M. Hanowski,
L. B. Johnson, J.  J. H. Ciborowski, C. A. Johnston, E. D. Reavie, V. J. Brady, and G. V. Sgro.
2006. Evaluating geographic, geomorphic and human influences on Great Lakes wetland
indicators: multi-assemblage variance partitioning. Ecological Indicators In press.

Johnson, L.B., J.  Olker, J.J.H. Ciborowski, G.E. Host, D. Breneman, V.  Brady, J. Brazner, andN.
Danz. 2006. Identifying Response of Fish Communities in Great Lakes Coastal Regions to Land
Use and Local Scale Impacts. Bull. N. Am. Benthol. Soc. [also in prep for submission to J. Great
Lakes Research]

Lapointe, N.W.R. 2005. Fish-habitat associations in shallow Canadian waters of the Detroit
River. M.Sc. Thesis, University of Windsor, Windsor, Ontario.

Leslie, J.K., C.A. Timmins and R.G.  Bonnell. 2002. Postembryonic development of the tubenose
goby Proterorhinus marmoratus Pallas (Gobiidae) in the St. Clair River/Lake system, Ontario.
Arch. Hydrobiol. 154:341-352.

Thoma. R.F. 1999. Biological monitoring and an index of biotic integrity for Lake Erie's
nearshore waters. Pages 417-461 in T.P.  Simon (Editor). Assessing the sustainability and
biological integrity of water resources using fish communities. CRC Press, Boca Raton, FL.

Wei, A., Chow-Fraser, P. and Albert, D.  2004. Influence of shoreline features on fish
distribution in the Laurentian Great Lakes.  Can. J. Fish.  Aquat. Sci. 61:  111 3-1123.
                         Draft for Discussion at SOLEC 2006

-------
Seilheimer, T.S. and Chow-Fraser, P. 2006. Development and use of the Wetland Fish Index to
assess the quality of coastal wetlands in the Laurentian Great Lakes. Submitted to Can. J. Fish.
Aquat. Sci. 63:354-366.

Last Updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
         •••""     	
Wetland-Dependent Amphibian Diversity and Abundance
Indicator #4504
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Deteriorating
Species across the Great Lakes basin exhibited both positive and negative
population trend tendencies. Five species exhibited significantly negative
species population trends while only one species exhibited a significantly
positive species population trend.
Lake-by-Lake Assessment
Lake Superior
           Status:  Not Assessed
           Trend:  Undetermined
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Poor
Unchanging
Most species in this lake basin exhibited negative population trend
tendencies. However, of the only two significant species population trends,
one was positive and one was negative.
Mixed
Deteriorating
Species in this lake basin exhibited both positive and negative population
trend tendencies. However, four out of eight species exhibited significantly
negative population trends. There were no significantly positive species
population trends.
Mixed
Deteriorating
Species in this lake basin exhibited both positive and negative population
trend tendencies. Two focal species (Bullfrog and Northern Leopard Frog)
exhibited significant population trend declines. Only one species exhibited a
significantly positive population trend.
Mixed
Unchanging
Species in this lake basin exhibited both positive and negative population
trend tendencies. Two species exhibited significantly increasing population
trends, while only one species showed a significant declining species
population trend.
                         Draft for Discussion at SOLEC 2006

-------
Purpose
To directly measure species composition and relative occurrence of frogs and toads and to
indirectly measure the condition of coastal wetland habitat as it relates to factors that influence
the health of this ecologically important component of wetland biotic communities.

Ecosystem Objective
To restore and maintain diversity and self-sustaining populations of Great Lakes coastal wetland
amphibian communities.  Breeding populations of amphibian species across their historical range
should be sufficient to maintain populations of each species and overall species diversity
(Anonymous 1989).

State of the Ecosystem
Background
Numerous amphibian species occur in the Great Lakes basin and many of these are associated
with wetlands during part of their life cycle. Because  frogs and toads are relatively sedentary and
have semi-permeable skin, they are likely to be more sensitive to, and indicative of, local  sources
of wetland contamination and degradation than are most other vertebrates.  Assessing species
composition and relative abundance of calling frogs and toads in Great Lakes wetlands can
therefore help to infer wetland habitat quality.

Geographically extensive and long-term monitoring of calling amphibians is possible through the
enthusiasm, skill and coordination of volunteer participants trained in the application of
standardized monitoring protocols. Information about abundance, distribution and diversity of
amphibians provides data for calculating trends in population indices as well as investigating
habitat associations, which can contribute to effective  long-term conservation strategies.

Status of Amphibians
Since 1995, Marsh Monitoring Program (MMP) volunteers have collected amphibian data at 548
discrete routes across the Great Lakes basin. An annual summary of amphibian routes monitored
is provided in Table 1.

Thirteen amphibian species were recorded during the 1995 - 2005 period (Table 2).  Spring
Peeper was the most frequently detected species and was commonly recorded in full chorus (Call
Level Code 3) when it was encountered.  Green Frog was detected in more than half of the survey
stations and was most often recorded at Call Level Code 1 (calling individuals could be discretely
counted). Grey Treefrog, American Toad and Northern Leopard Frog were also common, being
recorded in approximately one-third or more of all survey stations. Grey Treefrog was recorded
with the second highest average calling code (1.8), indicating that  MMP observers usually heard
several individuals calling simultaneously at each survey station. Chorus Frog, Bullfrog and
Wood Frog were detected in approximately one-quarter of survey  stations, while the remaining
five species were detected in less than 3 percent of survey stations.

Trends in amphibian occurrence were assessed for eight species  commonly detected on MMP
routes (Figure 1). For each species, the annual proportion of stations where that species was
present within a route was calculated to derive annual  indices of occurrence. The overall
temporal trend in occurrence for each species was assessed by combining route-level trends in
                         Draft for Discussion at SOLEC 2006

-------

station occurrence. Statistically significant declining trends were detected for American Toad,
Bullfrog, Chorus Frog, Green Frog and Northern Leopard Frog. Spring Peeper exhibited a
statistically significant increasing population trend.

These data will serve as baseline data with which to compare future survey results. Anecdotal
and research evidence suggests that wide variations in occurrence of many amphibian species at a
given site is a natural and ongoing phenomenon. Additional years of data will help distinguish
whether the patterns observed (i.e., decline in American Toad, Bullfrog, Chorus Frog, Green Frog
and Northern Leopard Frog population indices) indicate significant long-term trends or simply
natural variation in population sizes inhabiting marsh habitats.  Bullfrog, for example, did not
experience a significant population index trend from 1995 to 2004 (Crewe et al. 2006; Archer et
al. 2006) but with the addition of 2005 data, its population index declined significantly.  Further
data are thus required to conclude whether Great Lakes wetlands are successfully sustaining these
amphibian populations. MMP amphibian data are being evaluated to determine how information
from their community composition can be used to gain a better understanding of Great Lakes
coastal wetland condition in response to various human induced stressors.

Future Pressures
Habitat loss and deterioration remain the predominant threat to Great Lakes amphibian
populations. Many coastal and inland Great Lakes wetlands are located along watersheds that
experience very intensive industrial, agricultural and residential development.  Therefore, these
wetlands are under continued stress as increased pollution from anthropogenic runoff is washed
down watersheds into these sensitive habitats.   Combined with other impacts such as water level
stabilization, sedimentation, contaminant and nutrient inputs, climate change and invasion of
exotic species, Great Lakes wetlands will likely continue to be  degraded and as such, should
continue to be monitored.

Future Activities
Because of the sensitivity of amphibians to their surrounding environment and the growing
international concern about amphibian population status, amphibians in the Great Lakes basin and
elsewhere will continue to be monitored. Wherever possible, efforts should be made to maintain
high quality wetland habitat as well as associated upland areas adjacent to coastal wetlands.
There is also a need to address other impacts that are detrimental to wetland health such as inputs
of toxic chemicals, nutrients and sediments. Restoration programs are underway for many
degraded wetland areas through the work of local citizens, organizations and governments.
Although significant progress has been made in this area, more work remains for many wetland
areas that have yet to receive restoration efforts.

Further Work Necessary
Effective monitoring of Great Lakes amphibians requires accumulation of many years of data,
using a standardized protocol, over a large geographic expanse. A reporting frequency  for
SOLEC of five years would be appropriate because amphibian  populations  naturally fluctuate
through time, and a five-year timeframe would be sufficient to indicate noteworthy changes in
population indices. More rigorous studies will relate trends in species occurrence or relative
abundance to environmental factors.  Reporting will be improved with establishment of a
network of survey routes that  accurately represent the full spectrum of marsh habitat in  the Great
Lakes basin.
                         Draft for Discussion at SOLEC 2006

-------
Most MMP amphibian survey routes have been georeferenced to the survey station level.
Volunteer recruitment has also improved significantly since the last status reporting period. Four
additional important tasks are in progress: 1) develop the SOLEC wetland amphibian indicator as
an index for evaluating coastal wetland health; 2) improve the program's capacity to monitor and
report on status of wetland specific Beneficial Use Impairments among Great Lakes Areas of
Concern; 3) develop and improve the program's capacity to train volunteer participants to
identify and survey amphibians following standard MMP protocols, and; 4) develop the capacity
to incorporate a regional MMP coordinator network component into the MMP to improve
regional and local delivery of the program throughout the Great Lakes basin. Also, further work
is required to determine the relationship between calling codes used to record amphibian
occurrence and survey count estimates.

Acknowledgments
Authors: Steve Timmermans and Ryan Archer, Bird Studies Canada.
The Marsh Monitoring Program is delivered by Bird Studies Canada in partnership with
Environment Canada and the U.S. Environmental Protection Agency's Great Lakes National
Program Office.  The contributions of all Marsh Monitoring Program volunteers are gratefully
acknowledged.

Sources
Anonymous.  1989.  Revised Great Lakes Water Quality Agreement of 1978.  Office of
Consolidation, International Joint Commission United States and Canada. Available online:
http://www.ijc.org/rel/agree/quality.html. Last accessed August 29, 2006.

Anonymous.  2003.  Marsh Monitoring Program training kit and instructions for
surveying marsh birds, amphibians, and their habitats. Revised in 2003 by Bird Studies Canada.
41pp.

Archer, R.W., T.L.  Crewe, and S.T.A.  Timmermans. 2006. The Marsh Monitoring Program
annual report, 1995-2004: annual indices and trends in bird abundance and amphibian occurrence
in the Great Lakes basin. Unpublished report by Bird Studies  Canada.  35pp.

Timmermans, S.T.A. 2002. Quality Assurance Project Plan for implementing the Marsh
Monitoring Program across the Great Lakes basin.  Prepared for United States Environmental
Protection Agency - Great Lakes National Program Office Assistance I.D. #GL2002-145.  31pp.

Timmermans, S.T.A., S.S. Badzinski, and K.E. Jones. 2004.  The Marsh Monitoring
Program annual report, 1995-2002: annual indices and trends in bird abundance and amphibian
occurrence in the Great Lakes basin. Unpublished report by Bird Studies Canada.  48pp.

Weeber, R.C., and M. Valliantos (editors). 2000. The Marsh Monitoring Program 1995-
1999: Monitoring Great Lakes wetlands and their amphibian and bird inhabitants. Published by
Bird Studies Canada in cooperation with Environment Canada and the U.S.  Environmental
Protection Agency. 47pp.
                         Draft for Discussion at SOLEC 2006

-------
List of Tables
Table 1. Number of routes surveyed for amphibians within the Great Lakes basin, from 1995 to
2005.
Source: Marsh Monitoring Program

Table 2. Frequency of occurrence (Percent Station-Years Present) and average Call Level Code
for amphibian species detected at MMP survey stations within the Great Lakes basin, from 1995
through 2005. Average calling codes are based on the three level call code standard for all MMP
amphibian surveys; Code 1 = little overlap among calls, numbers of individuals can be
determined, Code 2 = some overlap, numbers can be estimated, Code 3 = much overlap of calls,
too numerous to be estimated.
Source: Marsh Monitoring Program

List of Figures
Figure 1. Trends (percent annual change) in station occurrence (population index) of eight
amphibian species commonly detected at Marsh Monitoring Program routes, from 1995 to 2005.
Values in parentheses are upper and lower 95% confidence limits, respectively, for trend values
given.
Source: Marsh Monitoring Program

Last Updated
SOLEC 2006
                                 Year
Number of
  Routes
                                 1995
                                 1996
                                 1997
                                 1998
                                 1999
                                 2000
                                 2001
                                 2002
                                 2003
                                 2004
                                 2005
   115
   177
   208
   168
   163
   158
   166
   156
   156
   146
   177
Table 1. Number of routes surveyed for amphibians within the Great Lakes basin, from 1995 to
2005.
Source: Marsh Monitoring Program
                       Draft for Discussion at SOLEC 2006

-------
                                                  ,&^«£!^&*S5ra31:            *.*        ^
Species Percent Station-


Spring Peeper
Green Frog
Grey Treefrog
American Toad
Northern Leopard Frog
Chorus Frog
Bullfrog
Wood Frog
Fowler's Toad
Pickerel Frog
Cope's Grey Treefrog
Mink Frog
Blanchard's Cricket Frog
Years
Present 1
69.3
54.3
39.2
36.9
31.1
26.5
25.8
18.0
2.4
2.4
1.6
1.2
0.6
1 MMP survey stations monitored for multiple years
Average
Calling Code

2.5
1.3
1.8
1.5
1.3
1.7
1.3
1.6
1.4
1.1
1.4
1.2
1.5
considered as individual
Table 2. Frequency of occurrence (Percent Station-Years Present) and average Call Level Code
for amphibian species detected at MMP survey stations within the Great Lakes basin, from 1995
through 2005. Average calling codes are based on the three level call code standard for all MMP
amphibian surveys; Code 1 = little overlap among calls, numbers of individuals can be
determined, Code 2 = some overlap, numbers can be estimated, Code 3 = much overlap of calls,
too numerous to be estimated.
Source: Marsh Monitoring Program
                         Draft for Discussion at SOLEC 2006

-------


 X
 0)
c
o
 3
 Q.
 O
Q.
              American Toad
           -0.8 (-1.6,-0.1) P< 0.05
     60
     55-
     50
     45 •
     40 •
     35-
     30
    Green Frog
-1.2 (-2.0,-0.5) P< 0.01
   Spring Peeper
1.5(0.6, 2.4) P< 0.001
                           60
                           55-
                           50-
                           45-
                           40-
                           35-
                           30
                           70-
                           65-
                           60-
                           55-
                           50
                           45
                                      40

                                      35-

                                      30-

                                      25-

                                      20
                                       Bullfrog
                                 -1.5 (-2.4, 0.6) P< 0.01
   Grey Treefrog
0.5 (-0.5, 1.5)P = 0.30
    Wood Frog
0.1 (-0.8, 1.0)P = 0.92
                                    Chorus Frog
                                -1.2 (-2.2, -0.2) P< 0.05
                          65-
                          55-
                          45-
                          35-
                          25
                                                                                   1999  2001
Northern Leopard Frog
-1.3 (-2.2, -0.5) P< 0.01
                          65-
                          55-
                          45-
                          35-
                          25
                                                                        1999   2001
                                                   Year
  Figure 1. Trends (percent annual change) in station occurrence (population index) of eight
  amphibian species commonly detected at Marsh Monitoring Program routes, from 1995 to 2005.
  Values in parentheses are upper and lower 95% confidence limits, respectively, for trend values
  given.
  Source: Marsh Monitoring Program
                              Draft for Discussion at SOLEC 2006

-------
        •••""     	
Contaminants in Snapping Turtle Eggs
Indicator #4506
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Trend not assessed
Contaminants at AOCs exceeded concentrations at reference sites. Dioxin
equivalents and DDE concentrations in eggs exceeded the Canadian
Environmental Quality Guidelines, and sum PCBs exceeded partial
restriction guidelines for consumption from some sites.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
Not Assessed
Trend Not Assessed due to insufficient data
Lake Michigan
           Status:  Not Assessed
           Trend:  Trend Not Assessed due to insufficient data
Lake Huron
           Status:
           Trend:
Not Assessed
Trend Not Assessed due to insufficient data
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Trend Not Assessed
Contaminants at AOCs exceeded concentrations at reference sites. Dioxin
equivalents and DDE concentrations in eggs exceeded the Canadian
Environmental Quality Guidelines, and sum PCBs exceeded partial
restriction guidelines for consumption from some sites.
Mixed
Trend Not Assessed
Contaminants at AOCs exceeded concentrations at reference sites. Dioxin
equivalents and DDE concentrations in eggs exceeded the Canadian
Environmental Quality Guidelines, and sum PCBs exceeded partial
restriction guidelines for consumption from some sites.
Purpose
•To assess the accumulation of organochlorine chemicals and mercury in snapping turtle eggs;
•To assess contaminant trends and physiological and ecological endpoints in snapping turtles; and
•To obtain a better understanding of the impact of contaminants on the physiological and
ecological health of the individual turtles and wetland communities.
                        Draft for Discussion at SOLEC 2006

-------
Ecosystem Objective
Snapping turtle populations in Great Lakes coastal wetlands and at contaminated sites should not
exhibit significant differences in concentrations of organochlorine chemicals, mercury, and other
chemicals, compared to turtles at clean (inland) reference site(s). This indicator supports Annexes
1,2, 11 and 12 of the Great Lakes Water Quality Agreement.

State of the Ecosystem
Background
Snapping turtles inhabit (coastal) wetlands in the Great Lakes basin, particularly the lower Great
Lakes. While other Great Lakes wildlife species may be more sensitive to contaminants than
snapping turtles, there are few other species that are as long-lived, as common year-round, inhabit
such a wide variety of habitats, and yet are limited in their movement among wetlands. Snapping
turtles are also at the top in the aquatic food web and bioaccumulate contaminants. Plasma and
egg tissues  offer a nondestructive method to monitor recent exposure to chemicals as well as an
opportunity for long-term contaminant and health monitoring. Since they inhabit coastal wetlands
throughout  the lower Great Lakes basin, they allow for multi-site comparisons on a temporal and
spatial basis. Consequently, snapping turtles  are a very useful biological indicator species of local
wetland contaminant trends and the effects of these contaminants on wetland communities
throughout  the lower Great Lakes basin.

Status of Contaminants in Snapping Turtle Eggs
For more than 20 years, the Canadian Wildlife  Service (CWS) has periodically collected snapping
turtle eggs and examined the  species' reproductive success in relation to contaminant levels on a
research basis. More recently (2001-2005), CWS is examining the health of snapping turtles
relative to contaminant exposure in Canadian Areas of Concern (AOCs) of the lower Great Lakes
basin. The work by the CWS has shown that contaminants in snapping turtle eggs differ over time
and among  sites in the Great  Lakes basin, with significant differences observed between
contaminated and reference sites (Bishop et al.  1996, 1998). Snapping turtle eggs  collected at two
Lake Ontario sites (Cootes Paradise and Lynde Creek) had the greatest concentrations of
poly chlorinated dioxins and number of furans (Bishop et al. 1996, 1998). Eggs from Cranberry
Marsh (Lake Ontario) and two Lake Erie sites (Long Point and Rondeau Provincial Park) had
similar levels of polychlorinated biphenyls (PCBs) and organochlorines among the study sites
(Bishop et al. 1996, 1998). Eggs from Akwesasne (St. Lawrence River) contained the greatest
level of PCBs (Bishop et al. 1998). From 1984  to 1990/91, levels of PCBs and dichlorodiphenyl-
dichloroethene (DDE) increased significantly in eggs from Cootes Paradise and Lynde Creek, and
levels of dioxins and furans decreased significantly at Cootes Paradise (Struger et al.  1993;
Bishop et al. 1996). More recently, American researchers have also used snapping turtles as
indicators of contaminant exposure (Dabrowska et al. 2006).

Eggs with the greatest contaminant levels also showed the poorest developmental success (Bishop
et al. 1991,  1998). Rates of abnormal development of snapping turtle eggs from 1986-1991 were
highest at all four Lake Ontario sites compared to other sites studied (Bishop et al. 1998).

Lake Erie and connecting channels
                         Draft for Discussion at SOLEC 2006

-------


From 2001 to 2003, CWS collected snapping turtle eggs at or near three Canadian Lake Erie or
connecting channels AOCs: Detroit River, St. Clair River, and Wheatley Harbour AOCs, as well
as two reference sites. Mean sum PCBs ranged from 0.02 |o,g/g at Algonquin Park (reference site)
to 0.93 (ig/g at Detroit River. Sum PCB levels were highest at Turkey Creek (Detroit River),
followed by Wheatley Harbour, then St. Clair NWA (near St. Clair River AOC) and lastly,
Algonquin Provincial Park, an inland reference site (Figure 1). Dioxin equivalents of sum PCBs
in eggs from the Detroit River, Wheatley Harbour, and St. Clair River AOCs, and p,p'-DDE
levels in eggs from the Wheatley Harbour and the Detroit River AOCs, exceeded the Canadian
Environmental Quality Guidelines. Sum PCBs in eggs from the Detroit River and Wheatley
Harbour AOCs exceeded partial restriction guidelines for consumption (de Solla and Fernie
2004). An American study in 1997 funded by the Great Lakes Protection Fund found that sum
PCBs appeared to be higher in the American AOCs in Ohio, where concentrations ranged from
0.18 to 3.68 (ig/g; concentrations were highest from the Ottawa River AOC, followed by the
Maumee River AOC, Ashtabula River AOC, and the Black River within Maumee River AOC
(Dabrowska et al. 2006). The reference sites used near the American AOCs may have higher
contaminant exposure than the Canadian reference sites.

Lake Ontario  and connecting channels

From 2002 to 2003, CWS collected snapping turtle eggs at or near seven Lake Ontario and
connecting channel AOCs: Hamilton Harbour, Niagara River (Ontario), St. Lawrence River
(Ontario), and Toronto, as well as two  reference  sites. Mean sum PCBs varied ranged from 0.02
Hg/g at Algonquin Park (reference site) to 1.76 (ig/g at Hamilton Harbour (Grindstone Creek).
Sum PCB levels were highest at Hamilton Harbour (Grindstone Creek), followed by the second
site at Hamilton Harbour (Cootes Paradise), then Lyons Creek (Niagara River) (Figure 1). There
is evidence that PCB levels in snapping turtle eggs have been declining at the inland reference
site of Algonquin Park (1981-2003) and the heavily contaminated Hamilton Harbour AOC (1984-
2003). Long term trends at the St. Lawrence River AOC are difficult to determine, due to the high
degree of variability of contaminant sources in the area; PCBs have been reported as high as 738
(ig/g at Turtle Creek, Akwesasne (de Solla et al.  2001).

Flame retardants (polybrominated diphenyl ethers [PBDEs])  are one  of the chemicals  of emerging
concern because they are bioaccumulative and may potentially affect wildlife and human health.
Sum PBDE concentrations varied, but  they were an order of magnitude lower than sum PCBs in
snapping turtle eggs collected from the seven AOCs (2001-2003). Sum PBDE levels were lowest
at Algonquin Park (6.1 ng/g sum PDBE), where  airborne deposition is likely the main
contaminant source, and greatest at the Hamilton Harbour (Cootes Paradise; 67.6 ng/g) and
Toronto (Humber River;  107.0 ng/g) AOCs, indicative of urban areas likely being the main
source of PBDEs.

Pressures
Future pressures for this indictor include all sources of toxic contaminants that currently have
elevated concentrations (e.g. PCBs, dioxins), as well  as contaminants whose concentrations are
expected to increase in Great Lakes wetlands (e.g. PBDEs). Non-bioaccumulative compounds  in
which there are chronic exposures (e.g. PAHs) also pose a potential threat. Snapping turtle
populations face additional pressures from harvesting of adult turtles, road-side killings during
the nesting season in June, and habitat destruction.
                         Draft for Discussion at SOLEC 2006

-------

Management Implications
The contaminants measured by are persistent and bioaccumulative, with diet being the primary
source of exposure for snapping turtles, and thus indicate contamination that is available
throughout the aquatic food web. Although commercial collection of snapping turtles has ceased,
collection for private consumption persists. Therefore, consumption restrictions are required at
selected AOCs. Currently, only eggs are routinely sampled for contaminants, but body burdens of
females could be estimated using egg burdens, and thus used for determining if consumption
guidelines are needed. At some AOCs (i.e., Niagara River [Lyons Creek], Hamilton Harbour),
there are localized sediment sources of contaminants that may be rehabilitated through dredging
or capping. Mitigation of contaminant sources should eventually reduce contaminant burdens in
snapping turtles.

Comments from the author(s)
Contaminant status of snapping turtles should be monitored on a regular basis across the Great
Lakes basin where appropriate. Once the usefulness of the indicator is confirmed, a
complementary U.S. program is required to interpret basin-wide trends. This species offers an
excellent opportunity to monitor contaminant concentrations in coastal wetland populations.
Newly emerging contaminants also need to be examined in a long-term monitoring program. As
with all long-term monitoring programs, and for any indicator species used to monitor persistent
bioaccumulative contaminants, standardization of contaminant data is necessary for examining
temporal and spatial trends or combining data from different sources.

Acknowledgments
Authors: Shane de Solla, Canadian Wildlife Service, Environment Canada, Burlington, ON,
Shane.deSolla@ec.gc.ca, and Kim Fernie, Canadian Wildlife Service, Environment Canada,
Burlington, ON, kim.fernie@ec.gc.ca

Special thanks to Drs. Robert Letcher, Shugang Chu, and Ken Drouillard for chemical analyses,
particularly of the PBDEs. Thanks also go to other past and present CWS staff (Burlington,
Downsview, National Wildlife Research Centre), the wildlife biologists not associated with the
CWS, and private landowners.

Data Sources
Bishop,  C.A., Brooks, R.J., Carey, J.H., Ng, P., Norstrom, R.J., and Lean, D.R.S. 1991. The case
for a cause-effect linkage between environmental contamination and development in eggs of the
common Snapping Turtle (Chelydra s. serpentind) from Ontario, Canada. J. Toxic. Environ.
Health 33:521-547.

Bishop,  C.A., Ng, P., Norstrom, R.J., Brooks, R.J., and Pettit, K.E. 1996. Temporal and
geographic variation of organochlorine residues in eggs of the common Snapping Turtle
(Chelydra serpentina serpentind) (1981-1991) and comparisons to trends in the herring gull
(Larus argentatus) in the Great Lakes basin in Ontario, Canada. Arch. Environ. Contam. Toxicol.
31:512-524.
                         Draft for Discussion at SOLEC 2006

-------

Bishop, C.A., Ng, P., Pettit, K.E., Kennedy, S.W., Stegeman, J.J., Norstrom, R.J., and Brooks,
RJ. 1998. Environmental contamination and developmental abnormalities in eggs and hatchlings
of the common Snapping Turtle (Chelydra serpentina serpentind) from the Great Lakes-St.
Lawrence River basin (1989-1991). Environ. Pollut. 101:143-156.

Dabrowska, S., Fisher, W., Estenik, J., Kidekhel, R., and Stromberg, P. Polychlorinated biphenyl
concentrations, congener profiles, and ratios in the fat tissue, eggs, and plasma of snapping turtles
(Chelydra s. serpentina) from the Ohio basin of Lake Erie, USA. Arch Environ Contam Toxicol.
51:270-286.

de Solla,  S.R., Bishop, C.A., Lickers, H., and Jock, K. 2001. Organochlorine pesticide, PCB,
dibenzodioxin and furan concentrations in common snapping turtle eggs (Chelydra serpentina
serpentina) in Akwesasne, Mohawk Territory, Ontario, Canada. Arch Environ Contam Toxicol
40:410-417

de Solla,  S.R. and Fernie, KJ. 2004. Characterization of contaminants in snapping turtles
(Chelydra serpentina) from Canadian Lake Erie Areas of Concern: St. Clair, Detroit River, and
Wheatley Harbour. Environ Pollut.  132:101-112

Struger, J., Elliott, J.E., Bishop, C.A., Obbard, M.E., Norstrom, R.J., Weseloh, D.V.,  Simon, M.,
and Ng, P. 1993. Environmental contaminants in eggs of the common Snapping Turtles
(Chelydra serpentina serpentina) from the Great Lakes-St. Lawrence River Basin of Ontario,
Canada (1981, 1984).

List of Figures
Figure 1.  Sum PCB concentrations in snapping turtle eggs from various Canadian locations
throughout the lower Great Lakes basin, 2001 through 2003. Means ± standard errors are
presented.
Source: Canadian Wildlife Service

Last updated
SOLEC 2006
                         Draft for Discussion at SOLEC 2006

-------
S3- ? (1 ,
j- Z.U
O)
"0
§
0> 1.5-
5
3 LO-
CO
o
Q. 0.5'
E
c/>
o











_L rn m n
• •







•1-


i — |
1
	 1








11 s?i?Soofs
f"^:i"a = S'S^2
o> TO .. TO o> F o> -£ ° co S
< O £ tt- re
*J TO C *
W S TO §
I 1
CO o
^ TJ
C '
o is

c
o
5 c. 3 *r n  CO
CO I
_J
^_;
                                  O
                                        Location
Figure 1. Sum PCB concentrations in snapping turtle eggs from various Canadian locations
throughout the lower Great Lakes basin, 2001 through 2003. Means ± standard errors are
presented.
Source: Canadian Wildlife Service
                        Draft for Discussion at SOLEC 2006

-------
         •••""      	
Wetland-Dependent Bird Diversity and Abundance
Indicator #4507
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Deteriorating
Species across the Great Lakes basin exhibited both positive and negative
population trend tendencies. Significantly negative population trends
occurred for 14 species, while only six species exhibited significantly
positive population trends.
Lake-by-Lake Assessment
Lake Superior
           Status:   Not Assessed
           Trend:   Undetermined
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
Mixed
Deteriorating
Species in this lake basin exhibited both positive and negative population
trend tendencies. Despite an equal number of significantly positive and
negative trends among species, certain focal species did not occur at a level
sufficient for trend analysis, or were absent from monitoring stations.
Poor
Deteriorating
Most species in this lake basin exhibited a negative population trend. Eight
significantly negative species population trends occurred, while there were
  Status and Trend   no significantly positive species population trends.
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Deteriorating
Species in this lake basin exhibited both positive and negative population
trend tendencies. Significantly negative population trends occurred for
seven species, while only three species exhibited significantly positive
population trends.
Mixed
Deteriorating
Species in this lake basin exhibited both positive and negative population
trend tendencies. Significantly negative population trends occurred for six
species, while only two species exhibited significantly positive population
trends.
                         Draft for Discussion at SOLEC 2006

-------
                                                                                    	
Purpose
• To assess wetland bird species composition and relative abundance, and to infer condition of
  coastal wetland habitat as it relates to factors that influence the biological condition of this
  ecologically and culturally important component of wetland communities.

State of the Ecosystem
Background
Assessments of wetland-dependent bird diversity and abundance in the Great Lakes are used to
evaluate health and function of coastal and inland wetlands. Breeding birds are valuable
components of Great Lakes wetlands and rely on the physical, chemical and biological condition
of their habitats, particularly during breeding. Presence and abundance of breeding individuals
therefore provide a valuable source of information about wetland status and population trends.
Because several wetland-dependent birds are listed as species at risk due to the loss and
degradation of their habitats, the combination of long-term monitoring data and analysis of
habitat characteristics can help to assess how well Great Lakes coastal wetlands are able to
provide habitat for these sensitive species as well as other birds and wetland-dependent wildlife.

Geographically extensive and long-term monitoring of wetland-dependent birds is possible
through the enthusiasm, skill and coordination of volunteer participants trained in the application
of standardized monitoring protocols. Information about abundance, distribution and diversity of
marsh birds provides data for calculating trends in population indices as well as investigating
habitat associations which can contribute to effective, long-term conservation strategies.

Status of Wetland-Dependent Birds
Since 1995, Marsh Monitoring Program (MMP) volunteers have collected bird data at 508
discrete routes across the Great Lakes basin. An annual summary of bird routes monitored is
provided in Table 1.

From 1995 through 2005, MMP volunteers recorded 56 bird species that use marshes (wetlands
dominated by non-woody emergent plants) for feeding, nesting or both throughout the Great
Lakes basin. Red-winged Blackbird was the most commonly recorded non-aerial foraging bird
species observed by MMP participants, followed by Swamp Sparrow, Marsh Wren and Yellow
Warbler. Among birds that nest exclusively in marsh habitats, the most commonly recorded
species was Marsh Wren, followed by Virginia Rail, Common Moorhen,  Pied-billed Grebe,
American Coot and Sora. Among bird species that typically forage in the air above marshes, Tree
Swallow and Barn Swallow were the two most commonly recorded bird species.

With eleven years of data collected across the Great Lakes basin, the MMP is becoming an
established and recognized long-term marsh bird population monitoring program. Bird species
occurrence, abundance, activity and detectability vary naturally among years and within seasons.
Population indices and trends  (i.e., average annual percent change in population index) are
presented for several bird species recorded at Great Lakes MMP routes, from  1995 through 2005
(Figure 1). Species with significant basin-wide declines were American Coot (not shown), Black
Tern, Blue-winged Teal (not shown),  Common Grackle (not shown), Common Moorhen (not
                         Draft for Discussion at SOLEC 2006

-------
shown), Least Bittern, undifferentiated Common Moorhen/American Coot (calls of these two
species are difficult to distinguish from one another), Northern Harrier (not shown), Pied-billed
Grebe, Red-winged Blackbird, Sora, Tree Swallow and Virginia Rail (Figure 1). Statistically
significant basin-wide population increases were observed for Common Yellowthroat, Mallard,
Northern Rough-winged Swallow (not shown), Purple Martin (not shown), Trumpeter Swan (not
shown), Willow Flycatcher (not shown) and Yellow Warbler (not shown). American Bittern and
Marsh Wren populations did not show a significant trend in abundance indices from 1995 through
2005 (Figure 1). Declines in population indices of species that use wetlands almost exclusively
for breeding such as Least Bittern, Black Tern, Common Moorhen, American Coot, Sora, Pied-
billed Grebe and Virginia Rail, combined with an increase in some wetland edge and generalist
species (e.g., Common Yellowthroat, Willow Flycatcher and Mallard) suggest changes in wetland
habitat conditions may be occurring.  Difference in habitats, regional population densities, timing
of survey visits, annual weather variability and other factors likely interplay with water levels to
explain variation in wetland dependent bird populations. American Bittern, for example, showed
a significant declining population index from 1995 to 2004 (Crewe et al. 2006; Archer et al.
2006) but recently its population index has rebounded.  As such, further years of data will
hopefully help explain natural population variation from significant population trends.

Future Pressure
Future pressures on wetland-dependent birds will likely include continuing loss and degradation
of important breeding habitats through wetland loss, water level stabilization, sedimentation,
contaminant and nutrient inputs and invasion of exotic plants and animals.

Future Activities
Wherever possible, efforts should be made to maintain high quality wetland habitat and adjacent
upland areas. There is also a need to address other impacts that are detrimental to wetland health
such as water level stabilization, invasive species and inputs of toxic chemicals, nutrients and
sediments. Restoration programs are underway for many degraded wetland areas through the
work of local citizens, organizations and governments.  Although significant progress has been
made, considerably more  conservation and restoration work is needed to ensure maintenance of
healthy and functional wetland habitats throughout the Great Lakes basin.

Further Work Necessary
MMP wetland monitoring activities will continue across the Great Lakes basin. Continued
monitoring of at least 100 routes through 2006 is projected to provide good resolution for most of
the wetland-dependent birds recorded by MMP volunteers. Recruitment and retention of program
participants will therefore continue to be a high priority. Priority should also be placed on
establishing regional goals and acceptable thresholds for species-specific abundance indices  and
species community compositions. Assessments to determine relationships among survey indices,
bird population parameters and critical environmental parameters are also needed.

Previous studies have ascertained marsh bird habitat  associations using MMP bird and habitat
data. As more data is accumulated, these studies should be periodically updated in order to
provide a better understanding of the relationships between wetland bird species and habitat.
Most MMP bird survey routes have been georeferenced to the level of individual survey stations.
Volunteer recruitment has also improved significantly since the last status reporting period.   Five
additional important tasks are in progress: 1) develop the SOLEC wetland bird indicator as an
                         Draft for Discussion at SOLEC 2006

-------
index for evaluating coastal wetland health; 2) improve the program's capacity to monitor and
report on status of wetland specific Beneficial Use Impairments among Great Lakes Areas of
Concern; 3) improve and revise MMP bird survey protocols to coincide with continentally
accepted marsh bird monitoring survey standards; 4) develop and improve the program's capacity
to train volunteer participants to identify and survey marsh birds following standard MMP
protocols, and; 5) develop the capacity to incorporate a regional MMP coordinator network
component into the MMP to improve regional and local delivery of the program throughout the
Great Lakes basin.

Although more frequent updates are possible, reporting trends in marsh bird population indices
every five or six years is most appropriate for this indicator. A variety of efforts are underway to
enhance reporting breadth and efficiency.

Acknowledgments
Authors: Steve Timmermans and Ryan Archer, Bird Studies Canada
The Marsh Monitoring Program is delivered by Bird Studies Canada in partnership with
Environment Canada and the United States Environmental Protection  Agency - Great Lakes
National Program Office.  The contributions of all Marsh Monitoring Program volunteers are
gratefully acknowledged.

Sources
Anonymous.  1989.  Revised Great Lakes Water Quality Agreement of 1978. Office of
Consolidation, International Joint Commission United States and Canada.  Available online:
http://www.ijc.org/rel/agree/quality.html. Last accessed August 29, 2006.

Anonymous.  2003.  Marsh Monitoring Program training kit and instructions for
surveying marsh birds, amphibians, and their habitats. Revised in 2003 by Bird
Studies Canada. 41pp.

Archer, R.W., T.L. Crewe, and S.T.A. Timmermans. 2006. The Marsh Monitoring Program
annual report, 1995-2004:  annual indices and trends in bird abundance and amphibian occurrence
in the Great Lakes basin. Unpublished report by Bird Studies Canada. 35pp.

Timmermans, S.T.A. 2002. Quality Assurance Project Plan for implementing the Marsh
Monitoring Program across the Great Lakes basin. Prepared for United States
Environmental Protection Agency - Great Lakes National Program Office
Assistance I.D. #GL2002-145. 31pp.

Timmermans, S.T.A., S.S. Badzinski, and K.E. Jones. 2004. The Marsh Monitoring
Program annual report, 1995-2002:  annual indices and trends in bird abundance and amphibian
occurrence in the Great Lakes basin.  Unpublished report by Bird Studies Canada. 48pp.

Weeber, R.C., and M. Valliantos (editors). 2000. The Marsh Monitoring Program 1995-
1999: Monitoring Great Lakes wetlands and their amphibian and bird inhabitants. Published by
Bird Studies Canada in cooperation with Environment Canada and the U.S. Environmental
Protection Agency. 47pp.
                         Draft for Discussion at SOLEC 2006

-------

                                                   ^  ^ .,•_
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''  './ ''l3
                                                            s^
List of Tables
Table 1. Number of routes surveyed for marsh birds within the Great Lakes basin, from 1995 to
2005.

List of Figures
Figure 1. Trends (percent annual change) in relative abundance (population index) of marsh
nesting and aerial foraging bird species detected at Marsh Monitoring Program routes, from 1995
to 2005. Values in parentheses are upper and lower 95% confidence limits, respectively, for trend
values given.
Source: Marsh Monitoring Program

Last Updated
SOLEC 2006
Year Number of
Routes
1995
1996
1997
1998
1999
2000
2001
2002
2003
2004
2005
145
177
175
151
154
153
146
170
131
118
183
Table 1. Number of routes surveyed for marsh birds within the Great Lakes basin, from 1995 to
2005.
                        Draft for Discussion at SOLEC 2006

-------
                                                               ki4*f!vfiiiiFB; 	!•"                         %
                                                                                                           	
                American Bittern
             -5.0 (-10.6, 1.1) P = 0.10
                                         8.0-

                                         6.0-
                                                 Black Tern
                                         -12.4 (-16.1,-8.7) P< 0.0001
                                      Common Yellowthroat
                                       1.5(0.0, 3.0) P = 0.05
        1995   1997   1999  2001   2003   2005
                                            1995   1997   1999   2001   2003  2005
                                                                                 1995  1997   1999   2001   2003   2005
                  Least Bittern
           -10.7 (-15.1,-6.0) P< 0.0001
                                         6.0

                                         5.0

                                         4.0-

                                         3.0-

                                         2.0-

                                         1.0
                                                  Mallard
                                            5.4 (2.2, 8.8) P < 0.001
                                          Marsh Wren
                                      -1.5 (-3.1,-0.2) P = 0.07
        1995   1997   1999  2001   2003   2005
                                            1995   1997   1999   2001   2003  2005
                                                                                 1995  1997   1999   2001   2003   2005
 C
 o
 3
 Q.
 O
Q.
10.0
 9.0-
 8.0-
 7.0-
 6.0
 5.0-
 4.0-
 3.0-
 2.0
           Moorhen/Coot
       -4.8 (-7.2, -2.3) P < 0.001
   Pied-billed Grebe
-6.9 (-10.3, -3.4) P< 0.001
Red-winged Blackbird
-1.6 (-2.6, -0.6) P< 0.01
                             28.0
                             26.0
                             24.0-
                             22.0
                             20.0 •
                             18.0
                             16.0
         1995   1997  1999   2001   2003   2005
                                            1995   1997   1999   2001   2003  2005
                                                                                 1995  1997   1999  2001   2003  2005
                     Sora
             -4.7 (-8.3, -1.0) P = 0.01
                                                Tree Swallow
                                           -5.7 (-7.8, -3.7) P < 0.0001
                                         30.0
                                         25.0 •

                                         20.0 •

                                         15.0-

                                         10.0
                                                                        3.0

                                                                        2.5-

                                                                        2.0-

                                                                        1.5-

                                                                        1.0
                                          Virginia Rail
                                      -2.3 (-4.3, -0.3) P = 0.02
        1995   1997   1999  2001   2003   2005
                                             1995   1997   1999  2001   2003   2005
                                                                                 1995  1997   1999   2001   2003   2005
                                                        Year
  Figure 1. Trends (percent annual change) in relative abundance (population index) of marsh
  nesting and aerial foraging bird species detected at Marsh Monitoring Program routes, from 1995
  to 2005.  Values in parentheses are upper and lower 95% confidence limits, respectively, for trend
  values given.
  Source: Marsh Monitoring Program
                                 Draft for Discussion at SOLEC 2006

-------

Coastal Wetland Area by Type
Indicator #4510

Overall Assessment
           Status:  Mixed
           Trend:  Deteriorating

Lake-by-Lake Assessment
Lake Superior
           Status:  Not Assessed
           Trend:  Undetermined
Lake Michigan
           Status:
           Trend:
Not Assessed
Undetermined
Lake Huron
           Status:
           Trend:
Not Assessed
Undetermined
Lake Erie
           Status:
           Trend:
Not Assessed
Undetermined
Lake Ontario
           Status:
           Trend:
Not Assessed
Undetermined
Purpose
To assess the periodic changes in area (particularly losses) of coastal wetland types, taking into
account natural lake level variations.

Ecosystem Objective
Maintain total areal extent of Great Lakes coastal wetlands, ensuring adequate representation of
coastal wetland types across their historical range (Great Lakes Water Quality Agreement,
Annexes 2 and 13).

State of the Ecosystem
The status of this indicator has not been updated since the 2005 State of the Lakes report. Future
updates to the status of this indicator will require the repeated collection and analysis of remotely
sensed information. Currently, technologies and methods are being assessed for an ability to
estimate wetland extent. Next steps, including determination of funding and resource needs, as
well as pilot investigations must occur before an indicator status update can be made. The
timeline for this is not yet determined. However, once a methodology is established, it will be
applicable for long-term monitoring of this indicator, which is imperative for an improved
understanding of wetland functional responses and adaptive management. The 2005 assessment
of this indicator follows.
                         Draft for Discussion at SOLEC 2006

-------
                                                                                     	
Wetlands continue to be lost and degraded, yet the ability to track and determine the extent and
rate of this loss in a standardized way is not yet feasible.

In an effort to estimate the extent of coastal wetlands in the basin, the Great Lakes Coastal
Wetland Consortium (GLCWC) coordinated completion of a binational coastal wetland database.
The project involved building from existing Canadian and U.S. coastal wetland databases
(Environment Canada and Ontario Ministry of Natural Resources 2003, Herdendorf et al. 1981a-
f), and incorporating additional auxiliary Federal, Provincial and State data to create a more
complete, digital Geographic Information System (GIS) vector database. All coastal wetlands in
the database were classified using a Great Lakes hydrogeomorphic coastal wetland classification
system (Albert et al. 2005). The project was completed in 2004.The GIS database provides the
first spatially explicit seamless binational summary of coastal wetland distribution in the Great
Lakes system. Coastal wetlands totaling 216,743 ha have been identified within the Great Lakes
and connecting rivers up to Cornwall, Ontario (Figure 1). However, due to existing data
limitations, estimates of coastal wetland extent, particularly for the upper Great Lakes are
acknowledged to be incomplete.

Despite significant loss of coastal wetland habitat in some regions of the Great Lakes, the lakes
and connecting rivers still support a diversity of wetland types. Barrier protected coastal wetlands
are a prominent feature in the  upper Great Lakes, accounting for over 60,000 ha of the identified
coastal wetland area in Lake Superior, Lake Huron and Lake Michigan (Figure 2). Lake Erie
supports 22,057 ha of coastal wetland, with protected embayment wetlands accounting  for over
one third of the total area (Figure 2). In  Lake Ontario, barrier protected and drowned rivermouth
coastal wetlands account for 19,172 ha,  approximately three quarters of the total coastal wetland
area.

Connecting rivers within the Great Lakes system also support a diverse and significant  quantity of
wetlands (Figure 3). The St. Clair River delta occurs where the St. Clair River outlets into Lake
St. Clair, and it is the most prominent single wetland feature accounting for over 13,000 ha. The
Upper St. Lawrence River also supports a large area of wetland habitats that are typically
numerous small embayment and drowned rivermouth wetlands associated with the Thousand
Island region and St. Lawrence River shoreline.

Pressures
There are many stressors which have and continue to contribute to the loss and degradation of
coastal wetland area. These include: filling, dredging and draining for conversion to other uses
such as urban, agricultural, marina, and  cottage development; shoreline modification; water level
regulation; sediment and nutrient loading from watersheds; adjacent land use; invasive  species,
particularly non-native species; and climate variability and change. The natural dynamics of
wetlands must be considered in addressing coastal wetland stressors. Global climate variability
and change have the potential to amplify the dynamics by reducing water levels in the system in
addition to changing seasonal storm intensity and frequency, water level fluctuations and
temperature.
                         Draft for Discussion at SOLEC 2006

-------


Management Implications
Many of the pressures result from direct human actions, and thus, with proper consideration of
the impacts, can be reduced. Several organizations have designed and implemented programs to
help reduce the trend toward wetland loss and degradation.

Because of growing concerns around water quality and supply, which are key Great Lakes
conservation issues, and the role of wetlands in flood attenuation, nutrient cycling and sediment
trapping, wetland changes will continue to be monitored closely. Providing accurate useable
information to decision-makers from government to private landowners is critical to successful
stewardship of the wetland resource.

Comments from the author(s)
Development of improved, accessible, and affordable remote sensing technologies and
information, along with concurrent monitoring of other Great Lakes indicators will aid in
implementation and continued monitoring and reporting of this indicator.

The GLCWC database represents  an important step in establishing a baseline for monitoring and
reporting on Great Lakes coastal wetlands including extent and other indicators. Affordable and
accurate remote sensing methodologies are required to complete the baseline and begin
monitoring change in wetland area by type in the future. Other GLCWC-guided research efforts
are underway to assess the use of various remote sensing technologies in addressing this current
limitation. Preliminary results from these efforts indicate the potential of using radar imagery and
methods of hybrid change detection for monitoring changes in wetland type and conversion.

The difficult decisions on how to address human-induced stressors causing wetlands loss have
been considered for some time. Several organizations and programs continue to work to reverse
the trend, though much work remains. A better understanding of wetland functions, through
additional research and implementation of biological monitoring within coastal wetlands, will
help ensure that wetland quality is maintained in addition to areal extent. An educated public is
critical to ensuring that wise decisions about the stewardship of the Great Lakes basin ecosystem
are made.

Acknowledgments
Authors: Joel Ingram, Canadian Wildlife Service, Environment Canada;
Lesley Dunn, Canadian Wildlife Service, Environment Canada;
Krista Holmes, Canadian Wildlife Service, Environment Canada and
Dennis Albert, Michigan Natural Features Inventory, Michigan State University Extension.

Contributors: Greg Grabas and Nancy Patterson, Canadian Wildlife Service, Environment
Canada; Laura Simonson, Water Resources Discipline, U.S. Geological  Survey; Brian Potter,
Conservation and Planning Section-Lands and Waters Branch, Ontario Ministry of Natural
Resources; Tom Rayburn, Great Lakes  Commission, Laura Bourgeau-Chavez, General Dynamics
Advanced Information Systems.

Data Sources
Albert, D.A., Wilcox, D.A., Ingram, J.W., and Thompson, T.A. 2005. Hydrogeomorphic
classification for Great Lakes coastal wetlands.  J. Great Lakes Res.
                         Draft for Discussion at SOLEC 2006

-------

Environment Canada and Ontario Ministry of Natural Resources. 2003. The Ontario Great Lakes
Coastal Wetland Atlas: a summary of information (1983 - 1997). Canadian Wildlife Service
(CWS), Ontario Region, Environment Canada; Conservation and Planning Section-Lands and
Waters Branch, and Natural Heritage Information Center, Ontario Ministry of Natural Resources.

Herdendorf, C.E., Hartley, S.M., and Barnes, M.D. (eds.). 1981a. Fish and wildlife resources of
the Great Lakes coastal wetlands within the United States, Vol. 1: Overview. U.S. Fish and
Wildlife Service, Washington, DC. FWS/OBS-81/02-vl.
Herdendorf, C.E., Hartley, S.M., and Barnes, M.D. (eds.). 1981b. Fish and wildlife resources of
the Great Lakes coastal wetlands within the United States, Vol. 2: Lake Ontario. U.S. Fish and
Wildlife Service, Washington, DC. FWS/OBS-81/02-v2.

Herdendorf, C.E., Hartley, S.M., and Barnes, M.D. (eds.). 1981c. Fish and wildlife resources of
the Great Lakes coastal wetlands within the United States, Vol. 3: Lake Erie. U.S. Fish and
Wildlife Service, Washington, DC. FWS/OBS-81/02-v3.

Herdendorf, C.E., Hartley, S.M., and Barnes, M.D. (eds.). 1981d. Fish and wildlife resources of
the Great Lakes coastal wetlands within the United States, Vol. 4: Lake Huron. U.S. Fish and
Wildlife Service, Washington, DC. FWS/OBS-81/02-v4.

Herdendorf, C.E., Hartley, S.M., and Barnes, M.D. (eds.). 1981e. Fish and wildlife resources of
the Great Lakes coastal wetlands within the United States, Vol. 5: Lake Michigan. U.S. Fish and
Wildlife Service, Washington, DC. FWS/OBS-81/02-v5.

Herdendorf, C.E., Hartley, S.M., and Barnes, M.D. (eds.). 1981f. Fish and wildlife resources  of
the Great Lakes coastal wetlands within the United States, Vol. 6: Lake Superior. U.S. Fish and
Wildlife Service, Washington, DC. FWS/OBS-81/02-v6.

List of Figures
Figure 1. Great Lakes coastal wetland distribution and total area by lake and river.
Source: Great Lakes Coastal Wetlands Consortium

Figure 2. Coastal wetland area by geomorphic type within lakes of the Great Lakes system.
Source: Great Lakes Coastal Wetlands Consortium

Figure 3. Coastal wetland area by geomorphic type within connecting rivers of the Great Lakes
system.
Source: Great Lakes Coastal Wetlands Consortium

Last updated
SOLEC 2006
                         Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
                                    -

      f  '
                         _
Lake / River
Lake Superior
St. Marys River
Lake Huron
Lake Michigan
St. Clair Rvier
Lake St. Clair
Detroit River
Lake Erie
Niagara River
Lake Ontario
Upper St. Lawrence River
Total
Area (ha)
26,626
10,790
61,461
44,516
13,642
2,217
592
25,127
196
22,925
8,454
216,545
Figure 1. Great Lakes coastal wetland distribution and total area by lake and river.
Source: Great Lakes Coastal Wetlands Consortium
                      Draft for Discussion at SOLEC 2006

-------
                            State of the Great Lakes 2007 - Draft
                                              • Barrier Protected
                                              DOpen Embayment
                                              D Protected Embayment
                                              D Drowned River-Mouth
                                              D Delta
           Superior   Huron
Michigan   St. Clair
      LAKE
Erie
Ontario
Figure 2. Coastal wetland area by geomorphic type within lakes of the Great Lakes system.
Source: Great Lakes Coastal Wetlands Consortium
                      Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
   CD
   ro 3,500

  | 3,000

  ^ 2,500
  o:
  < 2,000
                              13,146
                                                 D Barrier Protected
                                                 • Open Embayment
                                                 D Protected Embayment
                                                 D Drowned River-Mouth
                                                 D Delta
            St. Marys     St. Clair      Detroit      Niagara

                               CONNECTING RIVER
Upper St.
Lawrence
Figure 3. Coastal wetland area by geomorphic type within connecting rivers of the Great Lakes
system.
Source: Great Lakes Coastal Wetlands Consortium
                      Draft for Discussion at SOLEC 2006

-------


Ice Duration on the Great Lakes
Indicator #4858

Overall Assessment
           Status:  Mixed
           Trend:  Deteriorating (with respect to climate change)

Purpose
•To assess the ice duration and thereby the temperature and accompanying physical changes to
each lake over time, in order to infer the potential impact of climate change.

Ecosystem Objective
This indicator is used as a potential assessment of climate change, particularly within the Great
Lakes basin. Changes in water and air temperatures will influence ice development on the Lakes
and, in turn, affect coastal wetlands, nearshore aquatic environments, and inland environments.

State of the Ecosystem
Background
Air temperatures over a lake are one of the few factors that control the formation of ice on that
surface. Colder winter temperatures increase the rate of heat released by the lake, thereby
increasing the freezing rate of the water. Milder winter temperatures have a similar controlling
effect, only the rate of heat released is slowed and the ice forms more slowly. Globally, some
inland lakes appear to be freezing up at later dates, and breaking-up earlier, than the historical
average, based on a study of 150 years of data (Magnuson et al. 2000). These trends add to the
evidence that the earth has been in a period of global warming for at least the last 150 years.

The freezing and thawing of lakes is a very important aspect to many aquatic and terrestrial
ecosystems. Many fish species rely on the ice to give their eggs protection against predators
during the late part of the ice season. Nearshore ice has the ability to change the shoreline as it
can encroach upon the land during winter freeze-up times. Even inland systems are affected by
the amount of ice that forms, especially within the Great Lakes basin. Less ice on the Great Lakes
allows for more water to evaporate and be spread across the basin in the form of snow. This can
have an affect on the foraging animals (like deer), that need to dig through snow during the winter
in order to obtain food.

Status of Ice Duration on the Great Lakes
Observations of the Great Lakes data showed no real conclusive trends with respect to the date of
freeze-up or break-up. A reason for this could be that due to the sheer size of the Lakes, it wasn't
possible to observe the whole lake during the winter season (at least before satellite imagery), and
therefore only regional observations were made (inner bays and ports). However, there was
enough data collected from ice charts to  make a statement concerning the overall ice cover during
the season. There appears to be a decrease in the maximum ice cover per season over the last
thirty years (Figure 1).

The trends on each of the five Lakes show that during this time  span the maximum amount of ice
forming each year has been decreasing, which,  in-fact, can be correlated to the average ice cover
per season observed for the same time duration (Table 1). Between the 1970s and 1990s there
                         Draft for Discussion at SOLEC 2006

-------
was at least a 10% decline in the maximum ice cover on each Lake, and almost as much as 18%
in some cases, with the greatest decline occurring during the 1990s. Since a complete freeze-up
did not occur on all the Great Lakes, a series of inland lakes (known to freeze every winter) in
Ontario were examined to see if there was any similarity to the results in the previous studies.
Data from Lake Nipissing and Lake Ramsey were plotted (Figure 2) based on the ice-on date
(complete freeze-over date) and the break-up date (ice-off date). As it turns out, the freeze-up
date for Lake Nipissing appears to have the same trend as the other global inland lakes: freezing
over later in the year. Lake Ramsey however, seems to be freezing over earlier in the season. The
ice-off date for both however, appear to be increasing, or occurring at later dates in the year.
These results contradict what is said to be occurring with other such lakes in the Northern
Hemisphere (see Magnuson et al. 2000).

The satellite data used in this analysis can be supplemented by on-the-ground citizen science
collected data. The IceWatch program of Environment Canada's Ecological Monitoring and
Assessment Network and Nature Canada have citizen scientists collecting ice-on and ice-off dates
of lakes throughout the Ontario portion of the Great Lakes basin. These volunteers use the same
criteria for ice-on  and ice-off as does the satellite data, although the volunteers only collect data
for the portion of the lake that is visible from a single vantage point on the shore. The IceWatch
program began in  2000 as a continuation of a program run by the Meteorological Service of
Canada. Data from this program date back to the 1850s. An analysis of data from this database
and the Canadian  Ice Database (Canadian Ice Services/Meteorological Service of Canada)
showed that ice break-up dates were occurring approximately one day earlier every seven years
between 1950 and 2004 for 341 lakes across Canada (Putter et al.  2006. In press). The data from
IceWatch is not as comprehensive as the satellite collected data, but does show some trends in the
Great Lakes basin. From two sites with almost 100  years of data, Lake Nipissing is shown to be
thawing later in the season (Figure 3). IceWatch data from near Lake Ramsay indicate that lakes
have been freezing later over the past thirty years.

Pressures
Based on the results of Figure 1 and Table 1, it seems that ice formation on the Great Lakes
should continue to decrease in total cover if the predictions on global atmospheric warming are
true. Milder winters  will have a drastic effect on how much of the lakes are covered in ice, which
in turn, will have an effect on many aquatic and terrestrial ecosystems that rely on lake ice for
protection and food  acquisition.

Management Implications
Only a small number of data  sets were collected and analyzed for this study, so this report is not
conclusive. To reach a level of significance that would be considered acceptable, more data on
lake ice formation would have to be gathered. While the data for the Great Lakes is easily
obtained from 1972-present, smaller inland lakes, which may be affected by climate change at a
faster rate, should be examined. As much historical information that is available should be
obtained. This data could come from IceWatch observers and the IceWatch database from
throughout the Great Lakes basin. The more data that are received will increase the statistical
significance of the results.
                         Draft for Discussion at SOLEC 2006

-------
Acknowledgments
Author: Gregg Ferris, Environment Canada Intern, Downsview, ON.
Updated by: Heather Andrachuk, Environment Canada, Ecological Monitoring and Assessment
Network (EMAN); Heather. Andrachuk@ec.gc.ca; (905)336-4411.
All data analyzed and charts created by the author.

Sources
Putter, M., B. Buckland, E. Kilvert, and H. Andrachuk. 2006. Earlier break-up dates of lake ice:
an indicator of climate change in Canada. In press.

Magnuson, J.J., Robertson, D.M., Benson, B.J., Wynne, R.H., Livingston, D.M., Arai, T., Assel,
R.A., Barry, R.G., Carad, V., Kuusisto, E., Granin, N.G., Prowse, T.D., Stewart, K.M., and
Vuglinski, V.S. 2000. Historical trends in lake and river ice covering the Northern Hemisphere.
Science 289(Sept. 8):1743-1746.

Ice charts obtained from the National Oceanic and Atmospheric Administration (NOAA) and the
Canadian Ice Service (CIS).

Data for Lake Nipissing and Lake Ramsey obtained from Walter Skinner, Climate and
Atmospheric Research, Environment  Canada-Ontario Region.

Comments from the author
Increased winter and summer air temperatures appear to be the greatest influence on ice
formation. Currently there are certain protocols, on a global scale, that are being introduced in
order to reduce the emission of greenhouse gases.

It would be convenient for the results to be reported every four to five years (at least for the Great
Lakes), and quite possibly a shorter time span for any new inland lake information. It may also be
feasible to subdivide the Great Lakes  into bays and inlets, etc., in order to get an understanding of
what is occurring in nearshore environments.

Last Updated
SOLEC 2006

List of Tables
Table 1. Mean ice coverage, in percent, during the corresponding decade.
Source: National Oceanic and Atmospheric Administration

List of Figures
Figure 1. Trends of maximum ice cover and the corresponding date on the Great Lakes, 1972-
2000. The red line represents the percentage of maximum ice cover and the blue line represents
the date of maximum ice cover.
Source: National Oceanic and Atmospheric Administration

Figure 2. Ice-on and ice-off dates for Lake Nipissing (red line) and Lake Ramsey (blue line). Data
were smoothed using a 5-year moving average.
Source: Climate and Atmospheric Research and Environment Canada
                         Draft for Discussion at SOLEC 2006

-------
Figure 3. Ice-off dates and trend line from 1900-2000 on Lake Nipising.
Source: Ecological and Monitoring Assessment Network (EMAN)
Lake
Erie
Huron
Michigan
Ontario
Superior
1970-1979
94.5
71.3
50.2
39.8
74.5
1980-1989
90.8
71.7
45.6
29.7
73.9
1990-1999
77.3
61.3
32.4
28.1
62.0
Change from
1970s to 1990s
-17.2
-10.0
-17.8
-11.7
-12.6
Table 1. Mean ice coverage, in percent, during the corresponding decade.
Source: National Oceanic and Atmospheric Administration
                       Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
                                   Lake Superior
              Lake Michigan
                                                         Lake Erie
               Lake Huron
i>
oS.
                                              inn
                                               90
                                               SO
                                               70
                                               60
                                               50
                                               ^0
                                               30
                                               K
                                               10
                                               0
                                                          Ice Season

                                                         Lake Ontario
200
180 3
160 Q
140 I
120 O
100 S
80 £
60 I
40 s
20 I
0
Figure 1. Trends of maximum ice cover and the corresponding date on the Great Lakes, 1972-
2000. The red line represents the percentage of maximum ice cover and the blue line represents
the date of maximum ice cover.
Source: National Oceanic and Atmospheric Administration
                       Draft for Discussion at SOLEC 2006

-------
                                 State of the Great Lakes 2007 - Draft
                   Ice-on Dates
                                                                  Ice-off Dates
   tK
   V«
   3SS
>,  350
0  345
JS  340-
'  335-

   --;y.
   ••;>:•
1945 1950 1955 1960 1965  1970  1975  1980 1985
                Ice Season
            Nipissing    -•- Ramsey
                                                   1900 1910 1920  1930 1940 1950  1960  1970 1980  1990 2000
                                                                   Ice Season
                                                           -f Nipissing      -»- Ramsey
                                                             Linear (Nipssing) ^^Linear (Ramsay)
Figure 2. Ice-on and ice-off dates for Lake Nipissing (red line) and Lake Ramsey (blue line).
Data were smoothed using a 5-year moving average.
Source: Climate and Atmospheric Research and Environment Canada
    145
    135
    125
    115
    105
            -Ice Off Date    Ice Off Trend Line
      1900
                     1920
                                    1940
                                                   1960
                                                                  1980
                                                                                 2000
                                             Year
Figure 3. Ice-off dates and trend line from 1900-2000 on Lake Nipisin^
Source: Ecological and Monitoring Assessment Network (EMAN)
                         Draft for Discussion at SOLEC 2006

-------
Effect of Water Level Fluctuations
Indicator #4861

Assessment: Mixed, Trend Not Assessed
Data are available for water level fluctuations for all Lakes. A
comparison of wetland vegetation along regulated Lake Ontario
to vegetation along unregulated Lakes Michigan and Huron pro-
vides insight into the impacts of water level regulation.

Purpose
  To  examine the historic water levels in all the Great
Lakes, and compare these levels and their effects on wet-
lands with post-regulated levels in Lakes Superior and
Ontario, where water levels have been regulated since
about 1914 and 1959, respectively; and
  To  examine water level fluctuation effects on wetland
vegetation communities over time as well as aiding in the
interpretation of estimates of coastal wetland area, especial-
ly in those Great Lakes for which water levels are not regu-
lated.

Ecosystem Objective
The ecosystem objective is to maintain the diverse array of
Great Lakes coastal wetlands by allowing, as closely as is
possible, the natural seasonal and long-term fluctuations of
Great Lakes water levels.

State of the Ecosystem
Background
Naturally fluctuating water levels are known to be essential
for maintaining the ecological health of Great Lakes shore-
line ecosystems, especially coastal wetlands. Thus, comparing
the hydrology of the Lakes  serves as an indicator of degradation
caused by the artificial alteration of the naturally fluctuating
hydrological cycle.

Great Lakes shoreline ecosystems are dependent upon natural
disturbance processes, such as water level fluctuations, if they
are to function as dynamic systems. Naturally fluctuating water
levels create ever-changing  conditions along the Great Lakes
shoreline, and the biological communities that populate these
coastal wetlands have responded to these dynamic changes with
rich and diverse assemblages of species.

Status of Great Lakes Water Level Fluctuations
Water levels in the Great Lakes have been measured since I860.
but 140 years is a relatively short period of time when assessing
the hydrological history of the Lakes. Sediment investigations
conducted by Baedke and Thompson (2000) on the Lake
Michigan-Huron system indicate quasi-periodic lake level  fluc-
tuations (Figure  1), both in period and amplitude, on an average
of about 160 years, but ranging from  120-200 years. Within this
160-year period, there also appear to be sub-fluctuations of
approximately 33 years. Therefore, to assess water level fluctua-
tions, it is necessary to consider long-term data.

Because Lake Superior is at the upper end of the watershed, the
fluctuations have less amplitude than the other lakes. Lake
Ontario (Figure 2), at the lower end of the watershed, more
clearly shows these quasi-periodic fluctuations and the almost
181.0
180.5
180.0
179.5
? 179.0
§ 178.5
B 178.0
5) 177.5
UJ
177.0
176.5
176.0
175.5
175.0















- Measured upper limit
Inferred lower limit









I
ijiyv/


; *
= ....!....





f.
N\
1 !\
M
V '






1
sM
i « /
J/\N
J





/
/

,1
f \
A


/

s
-3
\\\





1


178

ffi
\
\




It
i
~
^~
i
T
161
30






~~t

/
*v/
ri i
w


/
/
Al
l,V
/TIU
* u


f


\
i\
\ ^
\


year fluctuations







X
^r

s^




J
,

/\
\r\
>/
I




\
(V
\
-
_





-
_

592
590
588 m
586 1
584 |
3
582 "3
580
578
576
574
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Calendar year before 1950
I I I I AD BC I
1950 1500 1000 500 0 500
1000
1500 2000 2500 3000
Figure 1. Sediment investigations on the Lake Michigan-Huron system
indicates quasi-periodic lake level fluctuations.
Source: National Oceanic and Atmospheric Administration, 1992 (and
updates)
complete elimination of the high and low levels since the lake
level began to be regulated in 1959, and more rigorously since
1976. For example, the 1986 high level that was observed in the
other lakes was eliminated from Lake Ontario. The level in Lake
Ontario after 1959 contrasts with that of the Lake Michigan-
Huron system (Figure 3), which shows the more characteristic
high and low water levels.

The significance of seasonal and long-term water level fluctua-
tions on coastal wetlands is perhaps best explained in terms of
the vegetation, which, in addition to its own diverse composi-
tion, provides the  substrate, food, cover, and habitat for many
other species dependent on coastal wetlands.

Seasonal water level fluctuations result in higher summer water
levels and lower winter levels. Additionally, the often unstable
summer water levels ensure a varied hydrology for the diverse
plant species inhabiting coastal wetlands. Without the seasonal
variation, the wetland zone would be much narrower and less
diverse. Even very short-term fluctuations resulting from

                                                        197

-------

                                                                                             2007
                                           Year
Figure 2. Actual water levels for Lake Ontario. IGLD-International Great Lakes Datum.
Zero for IGLD is Rimouski, Quebec, at the mouth of the St. Lawrence River. Water level
elevations in the Great Lakes/St. Lawrence River system are measured above water level
at this site.
Source: National Oceanic and Atmospheric Administration, 1992 (and updates)
   177.5
   177.0
   176.5
   176.0
   175.5
                                           Year
Figure 3. Actual water levels for Lakes Huron and Michigan. IGLD-International Great
Lakes Datum. Zero for IGLD is Rimouski, Quebec, at the mouth of the St. Lawrence
River. Water level elevations in the Great Lakes/St. Lawrence River system are measured
above water level at this site.
Source: National Oceanic and Atmospheric Administration, 1992 (and updates)
                    inundation. At the same time, there is an
                    expansion of aquatic communities, notably
                    submergents, into the newly inundated
                    area. As the water levels recede, seeds
                    buried in the sediments germinate and
                    vegetate this newly exposed zone, while
                    the aquatic communities recede out-ward
                    back into the lake.  During periods of low
                    water, woody plants and emergents
                    expand again to reclaim their former area
                    as aquatic communities establish them-
                    selves further outward into the lake. The
                    long-term high-low fluctuation puts natu-
                    ral stress on coastal wetlands, but is vital
                    in maintaining wetland diversity. It is the
                    mid-zone of coastal wetlands that harbors
                    the greatest biodiversity. Under more sta-
                    ble water levels, coastal wetlands occupy
                    narrower zones along the lakes and are
                    considerably less diverse, as the more
                    dominant species, such as cattails, take
                    over to the detriment of those less able to
                    compete under a stable water regime. This
                    is characteristic of many of the coastal
                    wetlands of Lake Ontario, where water
                    levels are regulated.

                    Pressures
                    Future pressures on the ecosystem include
                    additional withdrawals or diversions of
                    water from the Lakes, or additional regu-
                    lation of the high and low water levels.
                    These potential future pressures will
                    require direct human intervention to
                    implement, and thus, with proper consid-
                    eration of the impacts, can be prevented.
                    The more insidious impact could be
                    caused by global climate change. The
                    quasi-periodic fluctuations of water levels
                    are the result of climatic effects, and glob-
                    al warming has the potential to greatly
                    alter the water levels in the Lakes.
  changes in wind direction and barometric pressure can substan-
  tially alter the area inundated, and thus, alter the coastal wetland
  community.

  Long-term water level fluctuations, of course, have an impact
  over a longer period of time. During periods of high water, there
  is a die-off of shrubs, cattails, and other woody or emergent
  species that cannot tolerate long periods of increased depth of

  198
Management Implications
The Lake Ontario-St. Lawrence River Study Board is undertak-
ing a comprehensive 5-year study (2000-2005) for the
International Joint Commission (IJC) to assess the current crite-
ria used for regulating water levels on Lake Ontario and in the
St. Lawrence River.
The overall goals of Environment/Wetlands Working Group of
the IJC study are (1) to ensure that all types of native habitats

-------
(floodplain, forested and shrubby swamps, wet meadows, shal-
low and deep marshes, submerged vegetation, mud flats, open
water, and fast flowing water) and shoreline features (barrier
beaches, sand bars/dunes, gravel/cobble shores, and islands) are
represented in an abundance that allows for the maintenance of
ecosystem resilience and integrity over all seasons, and (2) to
maintain hydraulic and spatial connectivity of habitats to ensure
that fauna have access, temporally and spatially, to a sufficient
surface of all the types of habitats they need to complete their
life cycles.

The environment/wetlands component of the IJC study provides
a major opportunity to improve the understanding of past water-
regulation impacts on coastal wetlands. The new knowledge will
be used to develop and recommend water level regulation crite-
ria with the specific objective of maintaining coastal wetland
diversity and health. Also, continued monitoring of water levels
in all of the Great Lakes is vital to understanding coastal wetland
dynamics and the ability to assess wetland health on a large
scale. Fluctuations in water levels are the driving force behind
coastal wetland biodiversity and overall wetland health. Their
effects on wetland ecosystems must be recognized and moni-
tored throughout the Great Lakes basin in both regulated and
unregulated lakes.

Acknowledgments
Author: Duane Heaton, U.S. Environmental Protection Agency,
Great Lakes National Programs Office, Chicago, IL.

Much of the information and discussion presented in this sum-
mary is based on work conducted by the following: Douglas A.
Wilcox, Ph.D. (U.S. Geological Survey / Biological Resources
Division); Todd A. Thompson, Ph.D. (Indiana Geological
Survey); Steve J. Baedke, Ph.D. (James Madison University).

Sources
Baedke, S. J., and Thompson, T.A. 2000. A 4,700-year record of
lake level and isostasy for Lake Michigan. /. Great Lakes Res.
26(4):416-426.

International Joint Commission. Great Lakes Regional Office,
Windsor, ON and Detroit, MI.

International Lake Ontario-St. Lawrence River Study Board,
Technical Working Group on Environment/Wetlands.
http: //www. ijc.org.

Maynard, L., and Wilcox, D. 1997. Coastal wetlands  of the
Great Lakes. State of the Lakes Ecosystem Conference 1996
Background Paper. Environment Canada and U.S.
Environmental Protection Agency.
National Oceanic and Atmospheric Administration (NOAA).
1992 (and updates). Great Lakes water levels, 1860-1990.
National Ocean Service, Rockville, MD.
Authors' Commentary
Human-induced global climate change could be a major cause of
lowered water levels in the Lakes in future years. Further study
is needed on the impacts of water level fluctuations on other
nearshore terrestrial communities. Also, an educated public is
critical to ensuring wise decisions about the stewardship of the
Great Lakes basin ecosystem are made, and better platforms to
getting understandable information to the public are needed.

Last Updated
State of the Great Lakes 2003
                                                                                                                     199

-------

Coastal Wetland Plant Community Health
Indicator #4862

Overall Assessment
           Status:  Mixed
           Trend:  Undetermined
Lake-by-Lake Assessment
Lake Superior
           Status:   Good
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Unchanging
Degradation around major urban areas
Mixed
Unchanging
High quality wetlands in north part of lake
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Deteriorating
Plowing, raking, and mowing on Saginaw Bay wetlands during low water
causing degradation. Northern wetlands high quality
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Unchanging
Generally poor on US shore with some restoration at Metzger marsh -
Presque Isle, PA and Long Pt, Ontario high quality wetlands
Poor
Unchanging
Degraded by nutrient loading and water level control.  Some scattered
Canadian wetlands of higher quality.
Purpose
•To assess the level of native vegetative diversity and cover for use as a surrogate measure of
quality of coastal wetlands which are impacted by coastal manipulation or input of sediments.
                        Draft for Discussion at SOLEC 2006

-------

Ecosystem Objective
Coastal wetlands throughout the Great Lakes basin should be dominated by native vegetation,
with low numbers of invasive plant species that have low levels of coverage. (Great Lakes Water
Quality Agreement, United States and Canada 1987).

State of the Ecosystem
Background
To understand the condition of the plant community in coastal wetlands it is necessary to
understand the natural differences that occur in the plant community across the Great Lakes
basin. The characteristic size and plant diversity of coastal wetlands vary by wetland type, lake,
and latitude, due to differences in geomorphic and climatic conditions. Major factors will be
described below.

Lake: The water chemistry and shoreline characteristics of each Great Lake differ, with Lake
Superior being the most distinct due to its low alkalinity and prevalence of bedrock shoreline.
Nutrient levels also increase in the lake basins further to the east, that is, in Lake Erie, Lake
Ontario, and in the upper St. Lawrence River.

Geomorphic wetland type: There are several different types of wetland based on the
geomorphology of the shoreline where the wetland forms. Each landform has its characteristic
sediment, bottom profile, accumulation of organic material, and exposure to wave activity. These
differences result in differences in plant zonation and breadth, as well as species composition. All
coastal wetlands contain different zones (swamp, meadow, emergent, submergent), some  of
which may be typically absent in certain geomorphic wetland types. All Great Lakes wetlands
have recently been classified and mapped (Albert et al. In Press).
http://glc.org/wetlands/inventory.html

Latitude: Latitudinal differences in temperature result in floristic differences between the
southern and northern Great Lakes. Probably more important is the increased agricultural activity
along the shoreline of the southern Great Lakes, resulting in increased sedimentation and non-
native species introductions.

There are characteristics of coastal wetlands that make usage of plants as indicators difficult in
certain conditions. Among these are:

Water level fluctuations: Great Lakes water levels fluctuate greatly from year to year. Either an
increase or decrease in water level can result in changes in numbers of species or overall species
composition in the entire wetland or in specific zones. Such a change makes it difficult to monitor
change over time. Changes are great in two zones, the wet meadow where grasses and sedges
may disappear in high water or new annuals may appear in low water, and in shallow emergent or
submergent zones, where submergent and floating plants may disappear when water levels drop
rapidly.

Lake-wide alterations: For the southern lakes, most wetlands have been dramatically altered by
both intensive agriculture and urban development of the  shoreline. For Lake Ontario,  water level
                         Draft for Discussion at SOLEC 2006

-------

control has resulted in major changes to the flora. For both of these cases, it is difficult to identify
base-line high quality wetlands for comparison to degraded wetlands.

There are several hundred species of plant that occur within coastal wetlands. To evaluate the
status of a wetland using plants as indicators, several different plant metrics have been suggested.
Several of these are discussed briefly here.

Native plant diversity: The number of native plant species in a wetland is considered by many as
a useful indicator of wetland health.  The overall diversity of a site tends to decrease from south to
north. Different hydrogeomorphic wetland  types support vastly different levels of native plant
diversity, complicating the use of this metric.

Non-native species: Non-native species are considered signs of wetland degradation, typically
responding to increased sediment, nutrients, physical disturbance, and seed source. The amount of
non-native species coverage appears to be a more effective measure of degradation than number
of non-native species, except in the most heavily degraded sites.

Submergent species:  Submergent plants respond to high levels of sediment, nutrient enrichment,
and turbidity, and plant species have been identified that respond to each of these changes.
Floating species, such as Lemna spp., are similarly responsive to nutrient enrichment.  While
submergents are valuable indicators whose response to changing environmental conditions is well
documented, they also respond dramatically to natural fluctuations in the water level, making
them less dependable as indicators in the Great Lakes than in other wetland settings.

Nutrient responsive species: Several species from all plant zones are known to respond to nutrient
enrichment. Cattails (Typha spp.) are the best known responders.

Salt tolerance: Many species are not tolerant to salt, which is introduced along major coastal
highways. Narrow-leaved cattails are known to be very tolerant to high salt levels.

Floristic Quality Index (FQI): Many of the  states and provinces along the Great Lakes have
developed indices based on the "conservatism" of all plants growing there. A species is
considered conservative if it only grows in  a specific, high quality environment. FQI has proved
effective for comparing similar wetland sites. However, FQI of a given wetland can change
dramatically in response to a water level change, limiting its usefulness in monitoring  the
condition of a given wetland from year to year without development of careful sampling
protocols. Another problem associated with FQIs is that the conservatism values for a given plant
vary between states and provinces.

Status of Wetland Plant Community Health
The state of the wetland plant community is quite variable, ranging from good to poor across the
Great Lakes basin. The wetlands  in individual lake basins are often similar in their characteristics
because of water level controls and lake-wide near-shore management practices. There is
evidence that the plant component in some  wetlands is deteriorating in response to extremely low
water levels in some  of the Great Lakes, but this deterioration is not seen in all wetlands within
these lakes. In general, there is slow deterioration in many wetlands as shoreline alterations
introduce non-native species. However, the turbidity of the southern Great Lakes has reduced
                         Draft for Discussion at SOLEC 2006

-------
                                                                                    	
with expansion of zebra mussels, resulting in improved submergent plant diversity in many
wetlands.

Trends in wetland health based on plants have not been well established. In the southern Great
Lakes (Lake Erie, Lake Ontario, and the Upper St. Lawrence River), almost all wetlands are
degraded by either water level control, nutrient enrichment, sedimentation, or a combination of
these factors. Probably the strongest demonstration of this is the prevalence of broad zones of cat-
tails, reduced submergent diversity and coverage, and prevalence of non-native plants, including
reed (Phragmites australis), reed canary grass (Phalaris  arundinacea), purple loosestrife (Lythrum
salicaria), curly pondweed (Potamogeton crispus), Eurasian milfoil (Myriophyllum spicatum),
and frog bit (Hydrocharis morsus-ranae). In the remaining Great Lakes (Lake St. Clair, Lake
Huron, Lake Michigan, Georgian Bay, Lake Superior, and their connecting rivers), intact, diverse
wetlands can be found for most geomorphic wetland types. However, low water conditions have
resulted in the  almost explosive expansion of reed in many wetlands, especially in Lake St. Clair
and southern Lake Huron, including Saginaw Bay. As water levels rise, the response of reed
should be monitored.

One of the disturbing trends is the expansion of frog bit, a floating plant that forms dense mats
capable of eliminating submergent plants, from the St. Lawrence River and Lake Ontario
westward into  Lake Erie. This expansion will probably  continue into all or many of the remaining
Great Lakes.

Studies in the northern Great Lakes have demonstrated that non-native species like reed, reed
canary grass, and purple loosestrife have established throughout the Great Lakes, but that the
abundance of these species is low, often restricted to only local disturbances such as docks and
boat channels.  It appears that undisturbed marshes are not easily colonized by these species.
However, as these species become locally established, seeds or fragments of plant may be able to
establish when water level changes create appropriate sediment conditions.

Pressures
There are several pressures that lead to degradation of coastal wetlands.

Agriculture: Agriculture degrades wetlands in several ways, including nutrient enrichment from
fertilizers, increased sediments from erosion, increased  rapid runoff from drainage ditches,
introduction of agricultural non-native species (reed canary grass), destruction of inland wet
meadow zone by plowing and diking, and addition of herbicides. In the southern lakes, Saginaw
Bay, and Green Bay, agricultural sediments have resulted in highly turbid waters which support
few or no submergent plants.

Urban development: Urban development degrades wetlands by hardening shoreline, filling
wetland, adding a broad diversity of chemical pollutants, increasing stream runoff, adding
sediments, and increased nutrient loading from sewage treatment plants. In most urban settings
almost complete wetland loss has occurred along the shoreline.

Residential shoreline development: Along many coastal wetlands, residential development has
altered wetlands by nutrient enrichment from fertilizers and septic systems, shoreline alterations
                         Draft for Discussion at SOLEC 2006

-------

for docks and boat slips, filling, and shoreline hardening. While less intensive than either
agriculture or urban development, local physical alteration often results in introduction of non-
native species. Shoreline hardening can completely eliminate wetland vegetation.

Mechanical alteration of shoreline: Mechanical alteration takes a diversity of forms, including
diking, ditching, dredging, filling, and shoreline hardening. With all of these alterations non-
native species are introduced by construction equipment or in introduced sediments. Changes in
shoreline gradients and sediment  conditions are often adequate to allow non-native species to
become established.

Introduction of non-native species: Non-native species are introduced in many ways. Some were
purposefully introduced as agricultural crops or ornamentals, later colonizing in native
landscapes. Others came in as weeds in agricultural seed. Increased sediment and nutrient
enrichment allows many of our worst aquatic weeds to out-compete native species. Most of our
worst non-native species are either prolific seed producers or reproduce from fragments of root or
rhizome. Non-native animals have also been responsible for increased degradation of coastal
wetlands. One of the worst invasive species has been Asian carp, who's mating and feeding result
in loss of submergent vegetation in shallow marsh waters.

Management Implications
While plants are currently being evaluated as indicators of specific types of degradation, there are
limited examples of the effects of changing management on plant composition. Restoration
efforts at Coots Paradise, Oshawa Second, and Metzgers marsh have  recently evaluated a number
of restoration approaches to restore submergent and emergent marsh  vegetation, including carp
elimination, hydrologic restoration, sediment control, and plant introduction. The effect of
agriculture and urban sediments may be reduced by incorporating buffer strips along streams and
drains. Nutrient enrichment could be reduced by more effective fertilizer application, reducing
algal blooms. However, even slight levels of nutrient enrichment cause dramatic increases in
submergent plant coverage. For most urban areas it may prove impossible to reduce nutrient loads
adequately to restore native aquatic vegetation. Mechanical disturbance of coastal sediments
appears to be one  of the primary vectors for introduction of non-native species. Thorough
cleaning of equipment to eliminate seed source and monitoring following disturbances might
reduce new introductions of non-native plants.

Acknowledgments
Authors: Dennis Albert, Michigan Natural Features Inventory, Michigan State University
Extension.
Contributors: Great Lakes Coastal Wetlands Consortium

Data Sources
Albert, D.A., and Mine, L.D. 2001. Abiotic and floristic characterization of Laurentian Great
Lakes'  coastal wetlands. Stuttgart, Germany. Verh. Internal. Verein. Limnol. 27:3413-3419.

Albert, D.A., Wilcox, D.A., Ingram, J.W., and Thompson, T.A. 2006. Hydrogeomorphic
Classification for Great Lakes Coastal Wetlands. J. Great Lakes Res.
                         Draft for Discussion at SOLEC 2006

-------

Environment Canada and Central Lake Ontario Conservation Authority. 2004. Durham Region
Coastal Wetland Monitoring Project: Year 2 Technical Report. Environment Canada,
Downsview, ON: ECB-OR.

Herdendorf, C.E. 1988. Classification of geological features in Great Lakes nearshore and coastal
areas. Protecting Great Lakes Nearshore and Coastal Diversity Project. International Joint
Commission and The Nature Conservancy, Windsor, ON.

Herdendorf, C.E., Hakanson, L., Jude, D.J., and Sly, P.O. 1992. A review of the physical and
chemical components of the Great Lakes: a basis for classification and inventory of aquatic
habitats. In The development of an aquatic habitat classification system for lakes., eds. W.-D. N.
Busch and P. G. Sly, pp. 109-160. Ann Arbor, MI: CRC Press.

Herdendorf, C.E., Hartley,  S.M., and Barnes, M.D. (eds.). 1981a. Fish and wildlife resources of
the Great Lakes coastal wetlands within the United States, Vol. 1: Overview. U.S. Fish and
Wildlife Service, Washington, DC. FWS/OBS- 81/02-vl.

Jaworski, E., Raphael, C.N., Mansfield, P.J., and Williamson, B.B. 1979. Impact of Great Lakes
water level fluctuations on  coastal wetlands. U.S. Department of Interior, Office of Water
Resources and Technology, Contract Report 14-0001-7163, from Institute of Water Research,
Michigan State University, East Lansing, MI, 351pp.

Keough J.R., Thompson, T.A., Guntenspergen, G.R., and Wilcox, D.A. 1999. Hydrogeomorphic
factors and ecosystem responses in coastal wetlands of the Great Lakes. Wetlands 19:821-834.

Mine, L.D. 1997. Great Lakes coastal wetlands: An overview of abiotic factors affecting their
distribution, form, and species composition. Michigan Natural Features Inventory, Lansing, MI.

Mine, L.D., and Albert, D.A.  1998. Great Lakes coastal wetlands: abiotic and floristic
characterization. Michigan Natural Features Inventory, Lansing, MI.

United States and Canada.  1987. Great Lakes Water Quality Agreement of 1978, as amended by
Protocol signed November 18, 1987. Ottawa and Washington.
http://www.ijc.org/rel/agree/quality.html, last accessed March 15, 2005.

Wilcox, D.A., and Whillans, T.H. 1999. Techniques for restoration of disturbed coastal wetlands
of the Great Lakes. Wetlands 19:835-857.

Last updated
SOLEC 2006
                         Draft for Discussion at SOLEC 2006

-------

Land Cover Adjacent to Coastal Wetlands
Indicator # 4863
Overall Assessment
           Status:  Not Fully Assessed
           Trend:
   Primary Factors
      Determining
  Status and Trend
Undetermined
The status and trends are currently under investigation and proposed for
additional investigation for the full basin (see Data Sources). Although
other results exist for Canada (see Data Sources), "Land Cover Adjacent to
Coastal Wetlands" results are currently unavailable for Canada.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Fully Assessed
Undetermined
The status and trends are currently under investigation and proposed for
additional investigation in the Lake Superior Basin (see Data Sources)
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Fully Assessed
Undetermined
The status and trends are currently under investigation and proposed for
additional investigation in the Lake Michigan Basin (see Data Sources)
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Fully Assessed
Undetermined
The status and trends are currently under investigation and proposed for
additional investigation in the Lake Huron Basin (see Data Sources)
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Fully Assessed
Undetermined
The status and trends are currently under investigation and proposed for
additional investigation in the Lake Erie Basin (see Data Sources)
Not Fully Assessed
Undetermined
The status and trends are currently under investigation and proposed for
additional investigation in the Lake Ontario Basin (see Data Sources)
                         Draft for Discussion at SOLEC 2006

-------
Purpose
Assess the basin-wide presence, location, and/or spatial extent of land cover in close proximity to
coastal wetlands. Infer the condition of coastal wetlands as a function of adjacent land cover.
Relevant coastal areas in the Great Lakes Basin have been mapped to assess the presence and
proximity of general land cover in the vicinity of wetlands using satellite remote-sensing data and
geographic information systems (GIS), providing a broad scale measure of land cover in the
context of habitat suitability and habitat vulnerability for a variety of plant and animal species.
For example, upland grassland and/or upland forest areas adjacent to wetlands may be important
areas for forage, cover, or reproduction for organisms. Depending upon the particular
physiological and sociobiological requirements of the different organisms, the wetland-adjacent
land cover extent (e.g., the width or total area of the upland area around the wetland) may be used
to describe the potential for suitable habitat, or the vulnerability of these areas of habitat to loss or
degradation. Although other SOLEC Indicators are described for Canada (see Data Sources) at a
broad scale, basin-wide "Land Cover Adjacent to Coastal Wetlands" results are currently
unavailable for Canada.

Ecosystem Objective
Restore and maintain the ecological (i.e., hydrologic and biogeochemical) functions of Great
Lakes coastal wetlands. Presence, wetland-proximity, and/or spatial extent of land cover should
be such that the hydrologic and biogeochemical functions of wetlands continue.

State of the Ecosystem
The state of the Great Lakes Ecosystem (i.e., the sum of ecological functions for the full Great
Lakes Basin) is currently under investigation and proposed for additional investigation (see Data
Sources).  Differences in the regional status of "Habitat Adjacent to Coastal Wetlands" can be
determined using the existing data (see Pressures), but the results are preliminary and
observations are not conclusive. Nor can the regional trends be extrapolated to determine the
state of the ecosystem as a whole.

Percent forest adjacent to wetlands
The amount of forest land cover on the periphery of wetlands may indicate the amount of upland
wooded habitat for organisms that may travel relatively short distances to and from nearby
forested areas and wetland areas for breeding, water, forage, or shelter. Also, the affects of runoff
on wetlands from nearby areas (e.g., nearby agricultural land) may be ameliorated by
biogeochemical processes that occur in the forests on the periphery of the wetland. For example,
forest vegetation may contribute to the uptake, accumulation, and transformation of chemical
constituents in runoff. Broad-scale approaches to assessing percentage of forest directly adjacent
to wetlands may be calculated by summing the total area of forest land cover directly adjacent to
wetland regions in a reporting unit (e.g., an Ecoregion, a watershed, or a state) and dividing by
wetland total area in the reporting unit. This calculation ignores those upland areas of forest
outside of the adjacent "buffer zone"  for wetlands within each reporting unit. Other buffer
distances may be appropriate for other habitat analyses, depending on the type of organism; for
runoff analyses the chemical constituent(s), flow dynamics, soil conditions, position of wetland in
the landscape, and other landscape characteristics should be carefully considered. Coastal wetland
areas may be generally assessed by calculating forest wetland-adjacency in specifically targeted
                         Draft for Discussion at SOLEC 2006

-------


coastal wetlands of interest, by targeting narrow coastal areas such as areas within 1 km of the
lake shoreline (Figure 1), or by targeting all wetlands in a specific inland and coastal region of the
historical lake plain (Figure 2).

Percent grassland adjacent to wetlands
The amount of grassland on the periphery of wetlands may indicate the amount of upland
herbaceous plant habitat for organisms that might travel relatively short distances to and from
nearby upland grassland and wetland areas for breeding, water, forage, or shelter. As with
forested areas, the affect of runoff on wetlands from areas nearby (e.g., agricultural) land may be
ameliorated by biogeochemical processes that occur in herbaceous areas that are on the periphery
of the wetland.  For example, herbaceous vegetation stabilizes soils and may reduce erosional soil
loss to nearby wetlands and other surface water bodies. As with forest calculations, broad-scale
approaches to assessing percentage of grassland directly adjacent to wetlands may be calculated
by summing the total area of grassland directly adjacent to wetland regions in a reporting unit.
Other buffer distances may be more appropriate for habitat analyses, depending on the type of
organism; for runoff analyses the chemical constituent(s), flow dynamics, soil conditions,
position of wetland in the landscape, and other landscape characteristics should be carefully
considered. Coastal wetland areas may be generally assessed by calculating grassland wetland-
adjacency in specifically targeted coastal wetlands of interest; by targeting narrow coastal areas
such as areas within 1 km of the lake shoreline (Figure 3), or by targeting all wetlands in a
specific inland and coastal region of the historical lake plain (Figure 4).

Standard Deviation
Classes describe the distribution of percentage of forest or percentage of grassland adjacent to
wetlands (among reporting units) relative to the mean value for the metric distribution. Class
breaks are generated by successively described by standard deviations from the mean value for
the  metric. A two-color ramp (red to blue) emphasizes values (above to below) the mean value
for  a metric, and is a useful method for visualizing spatial variability of a metric.

Pressures
Although several causal relationships have been postulated for changes in "Land Cover Adjacent
to Coastal Wetlands" for the Great Lakes Basin (see Data Sources), it is undetermined as to the
relative contribution of the various factors. However, some preliminary regional trends exist.
For example, in the 1 km coastal region of southern Lake Superior there is a relatively high
percent of forest adjacent to coastal wetlands, and in the 1 km coastal region of western Lake
Michigan there is a relatively low percent of forest adjacent to coastal wetlands. Differences in
percent forest between these two coastal zones generally track with respect to percent of
agricultural land cover or urban land cover, as measured with similar techniques (see Data
Sources). These results are preliminary and observations are not conclusive. Similar phenomena
are  currently under investigation and proposed for additional regional and full-basin investigation.

Management Implications
Because  critical forest and grassland habitat areas on the periphery of coastal wetlands may
influence the presence and fitness of localized and migratory organisms in the Great Lakes,
natural resource managers may use these data to determine the ranking of their areas of interest,
such as areas where they are responsible for coastal wetland resources, among other areas in the
Great Lakes. It is important for managers to understand that results for their areas of interest are
                          Draft for Discussion at SOLEC 2006

-------
reported among a distribution for the entire Great Lakes Basin (USA) and that caution should be
used when interpreting the results at finer scales.

Comments from the author(s)
To conduct such measures at a broad scale, the relationships between wetland-adjacent land cover
and the functions of coastal wetlands need to be verified. This measure will need to be validated
fully with thorough field sampling data and sufficient a priori knowledge of such endpoints and
the mechanisms of impact. The development of indicators (e.g., a regression model using
adjacent vegetation characteristics and wetland hydroperiod) is an important goal, and requires
uniform measurement of field parameters across a vast geographic region to determine accurate
information to calibrate such models.

Acknowledgments
Authors: Ricardo D. Lopez, U.S. Environmental Protection Agency, National Exposure Research
Laboratory, Environmental Sciences Division, Landscape Ecology Branch, Las Vegas, Nevada,
USA

Data Sources
Lopez, R.D., D.T. Heggem, J.P. Schneider, R. Van Remortel, E. Evanson, L.A. Bice, D.W. Ebert,
J.G. Lyon, and R.W. Maichle. 2005. The Great Lakes Basin Landscape Ecology Metric Browser
(v2.0). EPA/600/C-05/011. The United States Environmental Protection Agency, Washington,
D.C. Compact Disk and Online at http://www.epa.gov/nerlesdl/land-
sci/glb browser/GLB  Landscape Ecology Metric Browser.htm.

Citation/Source
Lopez, R.D., D.T. Heggem, J.P. Schneider, R. Van Remortel, E. Evanson, L.A.  Bice,
D.W. Ebert, J.G. Lyon, and R.W. Maichle. 2005. The Great Lakes Basin Landscape
Ecology Metric  Browser (v2.0). EPA/600/C-05/011. The United States Environmental
Protection Agency, Washington, D.C. Compact Disk and Online at
http ://www.epa. gov/nerlesd 1 /land-
sci/glb  browser/GLB Landscape Ecology  Metric Browser.htm

List of Figures
Figure 1. Percent forest adjacent to wetlands, among 8-digit USGS Hydrologic Unit Codes
(HUCs), measured within 1 km of shoreline; data are reported as standard deviations from the
mean.
Source: Lopez et  al, 2006

Figure 2. Percent forest adjacent to wetlands, among 8-digit USGS Hydrologic Unit Codes
(HUCs), measured within 5 km of shoreline; data are reported as standard deviations from the
mean.
Source: Lopez et  al., 2006

Figure 3. Percent grassland adjacent to wetlands, among 8-digit USGS Hydrologic Unit Codes
(HUCs), measured within 1 km of shoreline; data are reported as standard deviations from the
mean.
                        Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
Source: Lopez et al, 2006
Last updated
SOLEC 2006
|-2 - -1 Std. Dev.
1-1-0 Std. Dev.
 Mean
JO - 1 Std. Dev,
|l - 2 Std, Dev,
ll - 3 Std, Dev.
|> 3 Std. Dev.
 Not Available
                              GLB Landscape Metrics
                                1 km of Shoreline
                               Standard Deviation
                                    Percent forest
                                  adjacent to wetlands
3   100 233


Kilometers
Figure 1. Percent forest adjacent to wetlands, among 8-digit USGS Hydrologic Unit Codes
(HUCs), measured within 1 km of shoreline; data are reported as standard deviations from the
mean.
Source: Lopez et al., 2006
                       Draft for Discussion at SOLEC 2006

-------
                              State of the Great Lakes 2007 - Draft
               I-2--1 Std. Dev.
               1-1-0 Std. Dev.
                Mean
               ] 0 - 1 Std. Dev.
                1-2 Std. Dav.
               j 2 - 3 Std. Dev.
              J > 3 Std. Dev.
              _J Not Available
GLB Landscape Metrics
 5km of Shoreline
 Standard Deviation
      Percent forest
    adjacent to wetlands
        200
0  100 200

 Kilometers
Figure 2. Percent forest adjacent to wetlands, among 8-digit USGS Hydrologic Unit Codes
(HUCs), measured within 5 km of shoreline; data are reported as standard deviations from the
mean.
Source: Lopez et al, 2006
                       Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
I-2 - -1 Std. Dev.
  -1-0 Std- Dev,
  Mean
  D - 1 Std. Dev.
  il - 2 Std. Dev.
  2 - 3 Std. Dev.
  :-• 3 Std. Dev.
  Nat Available
                               GLB Landscape Metrics
                                1 km of Shoreline
                               Standard Deviation
                                   Percent grassland
                                  adjacent to wetlands
0  tOO 2OT


 Kilcrre:ers
Figure 3. Percent grassland adjacent to wetlands, among 8-digit USGS Hydrologic Unit Codes
(HUCs), measured within 1 km of shoreline; data are reported as standard deviations from the
mean
Source: Lopez et al, 2006
                       Draft for Discussion at SOLEC 2006

-------
                              State of the Great Lakes 2007 - Draft
               I-2--1 Std. Dev.
                -1-0 Std. Dev.
                Mean
               JO- 1 Std. Dev.
                1-2 Std. Dav.
                12 - 3 Std. Dev.
                > 3 Std. Dev.
                Not Available
GLB Landscape Metrics
  5 km of Shoreline
 Standard Deviation
    Percent grassland
    adjacent to wetlands
0  100 ZOO

 Kilometers
Figure 4. Percent grassland adjacent to wetlands, among 8-digit USGS Hydrologic Unit Codes
(HUCs), measured within 5 km of shoreline; data are reported as standard deviations from the
mean (Lopez et al, 2006).
                       Draft for Discussion at SOLEC 2006

-------
Urban Density
Indicator #7000

Overall Assessment
           Status:  Mixed/ Trend Not Assessed
           Trend:  Improving, Unchanging, Deteriorating or Undetermined
   Primary Factors
      Determining
  Status and Trend

Lake by Lake Assessment
Trends on a lake-to-lake basis are unavailable due to insufficient data.

Purpose
To assess the urban human population density in the Great Lakes basin, and to infer the degree of
land use efficiency for urban communities in the Great Lakes ecosystem.

Ecosystem Objective
Socio-economic viability and sustainable development are the generally acceptable goals for
urban growth in the Great Lakes basin. Socio-economic viability indicates that development
should be sufficiently profitable and social benefits are maintained over the long term.
Sustainable development requires that we plan our cities to grow in a way so that they will be
environmentally sensitive, and not compromise the environment for future generations. Thus, by
increasing the densities in urban areas while maintaining low densities in rural and fringe areas,
the amount of land consumed by urban sprawl will be reduced.

State of the Ecosystem
Background
Urban density is defined as the number of people per square kilometer of land for urban use in a
municipal or township boundary. Low urban density indicates urban sprawl that is low-density
development beyond the edge of service and employment, which separates residential areas from
commercial, educational, and recreational areas - thus requiring automobiles for transportation
(TCRP, 1998; TCRP, 2003; Neill et al. 2003). Urban sprawl has many detrimental effects on the
environment. This process consumes large quantities of land, multiplies the required
infrastructure, and increases the use of personal vehicles as the feasibility  of alternate
transportation declines. When there is an increased dependency on personal vehicles,
consequentially, there is an increased demand for roads and highways, which in turn, produce
segregated land uses, large parking lots, and urban sprawl. These implications result in the
increased consumption of many non-renewable resources, the creation of impervious surfaces and
damaged natural habitats, and the production of many harmful emissions.  Segregated land use
also lowers the quality of life as the average time spent traveling increases and the sense of
community diminishes. For this assessment, the population data used was derived from 1990-
2000 U.S. census and 1996 - 2001 Canadian census.

This indicator offers information on the presence, location, and predominance of human-built
land cover and implies the intensity of human activity in the urban area. It may provide
                         Draft for Discussion at SOLEC 2006

-------
information about how such land cover types affect the ecological characteristics and functions of
ecosystems, as demonstrated by the use of remote-sensing data and field observations.

Status of Urban Density
Within the Great Lakes basin there are 10 Census Metropolitan Areas (CMAs) in Ontario and 24
Metropolitan Statistical Areas (MSAs) in the United States. In Canada, a CMA is defined as an
area consisting of one or more adjacent municipalities situated around a major urban core with a
population of at least 100,000. In the United States, an MSA must have at least one urbanized
area of 50,000 or more inhabitants and at least one  urban cluster of at least a population of 10,000
but less than 50,000. The urban population growth  in the Great Lakes basin show consistent
patterns in both the United States and Canada.  The  population in both countries has been
increasing over the past five to ten years.  According to the 2001 Statistics Canada report, between
1996 and 2001, the population of the Great Lakes basin CMAs grew from 7,041,985 to
7,597,260, an increase of 555,275 or 7.9% in five years. The 2000 U.S. census reports that from
1990 to 2000 the population contained in the MSAs of the Great Lakes basin grew from
26,069,654 to 28,048,813, an increase of 1,979,159 or 7.6% in  10 years.

In the Great Lakes basin, as there has been an increase in population, there has also been an
increase in the average population densities of the CMAs  and MSAs. However, using the CMA
or MSA as urban delineation has two major limitations. First, CMA and MSA contain substantial
land areas that is rural and by themselves result in over-estimation of the land area occupied by a
city or town. Second, these area delineations are based on a population density threshold and
hence provide information on residential  distribution and not necessarily on other urban land
categories such as commercial land, recreational land. If within the CMAs and MSAs the amount
of land being developed is escalating  at a greater rate than the population growth rate, the average
amount of developed land per person is increasing. For example, "In the GTA (Greater Toronto
Area) during the 1960s, the average amount  of developed land per person was a modest 0.019
hectares. By 2001 that amount tripled to 0.058 hectares per person" (Gilbert et al. 2001).

Population densities illustrate the development patterns of an area. If an urban area has a low
population density this indicates that the city has  taken on a pattern of urban sprawl and
segregated land uses. This conclusion can be made  as there is a greater amount of land per
person; however, it is important to not only look at the overall urban density of an area, but also
the  urban dispersion. For example, a CMA or MSA with a relatively low density could have
different dispersion characteristics than another CMA or MSA with the same density. One CMA
or MSA could have the distribution of people centred around an urban core, while another could
have a generally consistent sparse dispersion across the entire area and both would have the same
average density. Therefore, to properly evaluate the growth pattern of an area, it is necessary to
examine not only at the urban density but also  at the urban dispersion.

While density is a readily understandable measure, it is challenging to quantify because of the
difficulty in estimating true urban extent in a consistent and unbiased way. The geographic
extents of MSAs and CMAs give approximate indications of relative city size, however, they tend
to contain substantial areas of rural land use. Recently satellite remote sensing data has been used
to map landuse of Canadian cities as part of a program to  develop an integrated urban database,
the  Canadian Urban Land Use Survey (CUrLUS). In  southern Ontario a total of 11 cities have
                         Draft for Discussion at SOLEC 2006

-------

been mapped using Landsat data acquired in the 1999-2002 timeframe and densities estimated
using population statistics from the 2001 Canadian census (Figure 1). Population density is
related with the city size. Bigger cities with higher population pressure have higher population
density and more efficient land use. Comparing the population densities of 11 cities (or CMAs) in
southern Ontario, derived from remote sensing mapping and 2001 census (Zhang and Guindon,
2005),  the Great Toronto Area (GTA) has a higher population density (2848 km")  than other
smaller cities.

The growth characteristics of 5 large Canadian cities have also been studied for the period 1986-
2000. Preliminary analyses (Figure 2) indicate that the areal extents of these communities have
grown  at a faster rate than their populations and thus that sprawl continues to be a major problem.
A comparison of the ten CMAs and MSAs with the highest densities to the ten CMAs and MSAs
with the lowest densities in the Great Lakes basin shows there is a large range between the higher
densities and lower densities. Three of the ten lowest density areas have experienced a population
decline while the others have experienced very little population growth over the time period
examined. The areas with population declines and areas of little growth are generally occurring in
northern parts of Ontario and eastern New York State. Both of these areas have had relatively
high unemployment rates (between 8% and 12%) which could be linked to the slow growth and
decreasing populations.

Overall, the growing urban areas in the Great Lakes basin seem to be increasing their
geographical area at a faster rate than their population.  This trend has many detrimental effects as
outlined previously, namely urban sprawl and its implications. Such trends may continue to
threaten the Great Lakes basin ecosystem unless this pattern is reversed. However, there is a need
for a solid definitive information about relying on relatively fine-scale urban delineation data as it
pertains to broad-scale trends for the Great Lakes region.
Pressures
Under the pressure of rapid population growth in the Great Lakes region, mostly in the
metropolitan cities, the urban development has been undergoing unprecedented growth. For
instance, the urban built-up area of the Greater Toronto Area (GTA) has been doubled since
1960s.  Sprawl is increasingly becoming a problem in rural  and urban fringe areas of the Great
Lakes basin, placing a strain on infrastructure and consuming habitat in areas that tend to have
healthier environments than those that remain in urban areas. This trend is expected to continue,
which will exacerbate other problems, such as increased consumption of fossil fuels, longer
commute times from residential to work areas, and  fragmentation of habitat. For example, at
current rates in Ontario, residential building projects will consume some 1,000 square kilometres
of the province's countryside, an area double the size of Metro Toronto, by 2031. Also, gridlock
could add 45% to commuting times, and air quality could suffer due to a 40% increase in vehicle
emissions (Loten 2004). The pressure urban sprawl exerts on the ecosystem has not yet been fully
understood. It may be years before all of the implications have been realized.

Management Implications
                         Draft for Discussion at SOLEC 2006

-------
Urban density impacts can be more thoroughly explored and explained if they are linked to the
functions of ecosystems (e.g., as it relates to surface water quality). For this reason, interpretation
of this indicator is correlated with many other Great Lakes indicators and their patterns across the
Great Lakes. Urban density impacts on ecosystem functions should be linked to the ecological
endpoint of interest, and this  interpretation may vary as a result of the specificity of land cover
type and the contemporaneous nature of the data. Thus, more detailed land cover specificity is
required.

To conduct such measures at a broad scale, the relationships between land cover and ecosystem
functions need to be verified. This measure will need to be validated fully with thorough field-
sampling data and sufficient a priori knowledge of such endpoints and the mechanisms of impact
(if applicable). The development of indicators (e.g., a regression model) is an important goal, and
requires uniform measurement of field parameters across a vast geographic region to determine
accurate information to calibrate such models.

The governments of the United States and Canada have both been making efforts to ease the
strain caused by pressures of urban sprawl by proposing policies and creating strategies. Although
this is the starting point in implementing a feasible plan to deal with the environmental and social
pressures of urban sprawl, it does not suffice. Policies are not effective until they are put into
practice and in the meantime our cities continue to grow at unsustainable rates. In order to
mitigate the pressures of urban sprawl, a complete set of policies, zoning bylaws and
redevelopment incentives must be developed, reviewed and implemented. As noted in the Urban
Density indicator report from 2000, policies that encourage infill and brownfields redevelopment
within urbanized areas will reduce sprawl. Compact development could save 20% in
infrastructure costs (Loten 2004). Comprehensive land use planning that incorporates "green"
features, such as cluster development and greenway areas, will help to alleviate the pressure from
development.

For urban sustainable development, we should understand fully the potential negative impacts of
urban high density development. High urban density indicates intensified human activity in the
urban area, which would be potential threads to the urban environment quality. Therefore, the
urbanization strategies should be based on the concept of sustainable development on the balance
the costs and benefits.

Comments from the author(s)
A thorough field-sampling protocol, properly validated geographic information, and other
remote-sensing-based data could lead to successful development of urban density as an indicator
of ecosystem  function and ecological vulnerability in the Great Lakes basin. This indicator could
be applied to select sites, but would be most effective if used at a regional or basin-wide scale.
Displaying U.S. and Canadian census population density on a GIS map will allow increasing
sprawl to be documented  over time in the Great Lakes basin on a variety of scales. For example,
the maps included with the 2003 Urban Density report show the entire Lake Superior basin and a
closer view of the southwestern part of the basin.

To best quantify the indicator for the whole Great Lakes watershed, a watershed-wide consistent
urban built-up database is needed.
                         Draft for Discussion at SOLEC 2006

-------
Acknowledgments
Authors:
Bert Guindon, Natural Resources Canada, Ottawa, ON;
Ric Lopez, U.S. Environmental Protection Agency, Las Vegas, NV
Lindsay Silk, Environment Canada Intern, Downsview, ON; and
Ying Zhang, Natural Resources Canada, Ottawa, ON.

Data Sources
Bradof, K. GEM Center for Science and Environmental Outreach, Michigan Technological
University, MI, and James G. Cantrill, Communication and Performance Studies, Northern
Michigan University, MI.

GEM Center for Science and Environmental Outreach. 2000. Baseline Sustainability Data for the
Lake Superior Basin: Final Report to the Developing Sustainability Committee, Lake Superior
Binational Program, November 2000. Michigan Technological University, Houghton, MI.
http://emmap.mtu.edu/gem/community/planning/lsb.html.

Gilbert, R., Bourne, L.S., and Gertler, M.S. 2001. The State of GTA in 2000. A report for the
Greater Toronto Services Board. Metropole Consultants, Toronto, ON.

Loten, A. 2004. Sprawl plan our 'last chance:' Caplan. Toronto Star, July 29, 2004.

Neill, K.E., Bonser, S.P., and Pelley, J. 2003. Sprawl Hurts Us All! A guide to the costs of sprawl
development and how to create livable communities in Ontario. Sierra Club of Canada, Toronto,
ON.

Statistics Canada. 2001. Community Profiles and 1996 census subdivision area profiles.
http://wwwl2.statcan.ca/english/profil01/PlaceSearchForml.cfm.

TCRP, 1989: The cost of Sprawl-Revisited, Transportation Research Board,  TCRP report 39,
p40.

TCRP, 2002: Cost of Sprawl-2000. Transportation Research Board, TCRP report 74, p84.

U.S. Census  Bureau. American Fact Finder, Census 2000 Summary File 1  (SF 1) 100-Percent
Data, Detailed Tables.
http://factFinder.census.gov/servlet/DTGeoSearchByRelationshipServlet? _ts=l 09848346281.

Y. Zhang and B. Guindon, 2005: Using satellite remote sensing to survey transportation-related
urban Sustainability. Part I: Methodology for indicator quantification. Accepted by Applied Earth
Observation  and Geoinformation.

List of Figures
Figure  1. Population densities of cities with population more than 100,000 in southern Ontario of
the Great Lakes watershed for 2001.
                        Draft for Discussion at SOLEC 2006

-------
                                                      "if1* f^*?/5^v^"-*"*" * ^iSlJi Jiiiiii ilttjyi  i^iiiiihiif'^^^
Figure 2. Growth characterization of 5 urban areas in the period of 1986-2001.
Last updated
SOLEC 2006


C\l
'p
^
>,
•t-f
'(/)
C
CD
Q
g
:ra
13
Q.
O
Q.
1
I)




°nnn
2800 -


2600 -
2400 -

2200 -


2000 -

1800 -

1600 -

1400 -

/•
b
...--
..•*
y ^...«e
i 	
..L.i 	

h e

f


1 '
105 106
a
b

d

e
f
g
h
i
j
1

Toronto (GTA)
Hamilton
London
Kitchener











St. Catharines-Niagara
Windsor
Oshawa
Barrie
Kingston
Guelph
Peterborough








                                   Urban Population

Figure 1. Population densities of cities with population more than 100,000 in southern Ontario of
the Great Lakes watershed for 2001. Source: 'Y. Zhang and B. Guindon, private communication'
                         Draft for Discussion at SOLEC 2006

-------

    150
    140 -
 o  130
 CD
 Q.
 m  12°
 c
 CD
 J2
    110 -
    100
             St. Catharines
                        Hamilton^"*
                          London
                            Urban Growth 1986-2000
        100       110      120       130      140

                  Urban Population Growth (%)
150
Figure 2. Growth characterization of 5 urban areas in the period of 1986-2001. Source: 'Y. Zhang
and B. Guindon, private communication'
                       Draft for Discussion at SOLEC 2006

-------
        •••""      	
Land Cover/Land Conversion
Indicator #7002
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
Low-intensity development increased 33.5%, road area increased
7.5%, and forest decreased 2.3% from 1992 and 2001. Agriculture
lost 210,000 ha of land to development. Approximately 50% of
forest losses were due to management and 50% to development.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Good
Undetermined
Lowest conversion rate of non-developed land to development and highest
conversion rate of non-forest to forest. Of the 4.2 million ha watershed area
on the U.S. side, 1,676 ha of wetland, 2,641 ha of agricultural land, and
14,300 ha of forest land were developed between 1992 and 2001.
Mixed
Undetermined
Intermediate to high rate of land conversions to development.  Of the 1.2
million ha watershed, 9,724 ha of wetland, 78,537 ha of agricultural land,
and 57,529 ha of forest land were developed between 1992 and 2001.
Fair
Undetermined
Second lowest rate of conversion of land to development. Of the 4.1
million ha watershed area on the U.S. side, 4,314 ha of wetland, 17,881 ha
of agricultural land, and 17,730 ha of forest land were developed between
1992 and 2001.
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Poor
Undetermined
Highest conversion rate of non-developed to development LULC. Of the
5.0 million ha watershed area on the U.S. side, 3,352 ha of wetland, 52,502
ha of agricultural land, and 27,869 ha of forest land were developed
between 1992 and 2001.
Lake Ontario
           Status:   Mixed
           Trend:   Undetermined
                        Draft for Discussion at SOLEC 2006

-------

   Primary Factors  Intermediate to high conversion rate of non-developed to development
      Determining  LULC coupled with the lowest rates of wetland development. Of the 3.4
  Status and Trend  million ha watershed area on the U.S. side, 458 ha of wetland, 24,883 ha of
                   agricultural land, and 20,670 ha of forest land were developed between
                   1992 and 2001.

Purpose
•To document the proportion of land in the Great Lakes basin under major land use classes, and
assess the changes in land use over time; and
•To infer the potential impact of existing land cover and land conversion patterns on basin
ecosystem health.

Ecosystem Objective
Sustainable development is a generally accepted land use goal. This indicator supports Annex 13
of the Great Lakes Water Quality Agreement.

State of the Ecosystem
Binational land use data from the early 1990s was developed by Guindon (Natural Resources
Canada).  Imagery data from the North American Landscape Characterization and the Canada
Centre for Remote Sensing archive were combined and processed into land cover using
Composite Land Processing System software. This data set divides the basin into four major land
use classes - water, forest, urban, and agriculture and grasses.

Later, finer-resolution satellite imagery allowed analysis to be conducted in greater detail, with a
larger number of land use categories. For instance, the Ontario Ministry of Natural Resources  has
compiled Landsat TM (Thematic Mapper) data, classifying the Canadian Great Lakes basin into
28 land use classes.

On the U.S. side of the basin, the Natural Resources Research Institute (NRRI)  of the University
of Minnesota - Duluth has developed a 25-category classification scheme (Table 1) based on
1992 National Land Cover Data (NLCD) from the U.S.  Geological Survey supplemented by 1992
WISCLAND, 1992 GAP, 1996 C-CAP and raw Landsat TM data to increase resolution in
wetland classes (Wolter et al. 2006). The 1992 Topologically Integrated Geographic Encoding
and Reference  (TIGER) data were also used to add roads on to the map. Within the U.S. basin,
the NRRI found the following:

Between two nominal time periods (1992 and 2001), the U.S. portion of the  Great Lakes
watershed has undergone substantial change in many key LULC categories (Fig. 1). Of the total
change that occurred (798,755 ha, 2.5 % of watershed area), salient transition categories included
a 33.5 % increase in area of low-intensity development, a 7.5% increase in road area, and a
decrease of forest area by over 2.3 % - the largest LULC category and area of change within the
watershed.  More than half of the forest losses involved transitions into early successional
vegetation (ESV), and hence, will likely remain in forest production of some sort. However,
nearly as  much forest area was, for all practical purposes, permanently converted to developed
land. Likewise, agriculture lost over 50,000 more hectares of land to development than
forestland, much of which involved transitions into urban/suburban sprawl (See: Fig. 2).
                         Draft for Discussion at SOLEC 2006

-------

Approximately 210,068 ha (81 %) of agricultural lands were converted to development, and 16.3
% of that occurred within 10 km of the Great Lakes shoreline.

Land use/land cover transitions between 1992 and 2001 within near-shore zones of the Great
Lakes (0-1, 1-5, 5-10 km) largely parallel those of the overall watershed. While the same
transition categories dominated, their proportions varied by buffered distance from the lakes.
Within the 0-1 km zone from the Great Lakes shoreline, conversions of forest to both ESV (9,087
ha, 5.0 % of total category change (TCC)) and developed land (8,657 ha, 5.6 % of TCC) were
the largest transitions, followed by conversion of 3,935 ha (1.9 % of TCC) of agricultural land to
developed.  For the 1-5 km zone inland from the shore, forest to developed conversion was the
largest of the three transitions (17,049 ha, 11.0 % of TCC), followed by agricultural to developed
(14,279 ha, 6.8 % of TCC) and forest to ESV (13,116 ha, 7.3 % of TCC). Within the 5-10 km
zone from shoreline, transition category dominance was most similar to the trend for the whole
watershed, with 16,113 ha (7.7 % of TCC) of agriculture converted to developed, 14,516 ha (8.0
% of TCC) of forest converted to ESV, and 14,390 ha (9.3 % of TCC) of forestland being
developed by 2001.  When all buffers form shoreline out to 10 km are combined, the forest to
developed transition category was the largest (40,099 ha, 25.9 % of TCC), followed by forest to
ESV (36,726 ha, 20.3 % of TCC), and agricultural to developed (34,328 ha, 16.3 % of TCC).

Contrary to previous decadal estimates showing an increasing forest area trend from the early
1980s to the early 1990s, due to agricultural abandonment and transitions of forest land away
from active management, we observed an overall decrease (~2.3 %) in forest area between 1992
and 2001.  Explanation of this trend is largely unclear; however, both increased forest harvesting
practices in parts of the region coupled with forest clearing for new developments may be
overshadowing gains from the agricultural sources observed in previous decades.

When analyzed on a lake-by-lake basis (Fig. 3, Table 2), Michigan's watershed naturally has
experienced the greatest area of change from 1992 to 2001 (286587 ha, ~2.5 %), as its watershed
is entirely within the U.S., and hence, the largest analyzed. Michigan's watershed leads in all
LULC transition categories but two:  1) misc. veg. to flooded and 2) ESV to forest (Fig. 3).
When normalized by area, however, Michigan's proportion of LULC change  is intermediate
when compared to the other Great Lakes watersheds on the U.S. side of the boarder. Although
not a Great Lake, and largely metropolitan (See: Fig 2), Lake St. Clair's watershed shows the
highest rates of change into development from wetland, ESV, agriculture, and forest sources (Fig.
4).

Of the Great Lakes, Erie's watershed shows the greatest proportion of land conversion to
development (87,077 ha, 1.74 %), while Superior's watershed had the lowest proportion (20,351,
0.48 %) (Table 2). For example, Erie had the highest proportion of agricultural land conversion
to development.  However, Ontario's watershed showed the greatest proportion of forest
conversion to development (Fig. 4).  Superior's watershed reflects a high proportion of lands
under forest management in that it has both the highest proportion of forest conversion to ESV
and visa-versa. Lastly, Huron's watershed had the highest proportion of wetlands being
converted to development, followed closely by Michigan and Erie (Fig.  4).
                         Draft for Discussion at SOLEC 2006

-------

Management Implications
As the volume of data on land use and land conversion grows, stakeholder discussions will assist
in identifying the associated pressures and management implications.

Comments from the author(s)
Land classification data must be standardized. The resolution should be fine enough to be useful
at lake watershed and sub-watershed levels.  LULC classification updates need to be completed in
a timely manner to facilitate effective remedial action if necessary.

Acknowledgments
Author: Peter Wolter, Department of Forest Ecology and Management
University of Wisconsin-Madison

Data Sources
Data courtesy of: Bert Guindon (Natural Resources Canada), Lawrence Watkins (Ontario
Ministry of Natural Resources) and Peter Wolter (Natural Resources Research Institute at the
University of Minnesota - Duluth). Forest Inventory and Analysis statewide data sets
downloaded from USDA Forest Service website and processed by the author to extract data
relevant to Great Lakes basin.

List of Tables
Table 1. Classification scheme used to analyze LULC change in the U.S. portion of the Great
Lakes basin. Original 25 classes are listed in the left column, while aggregated LULC categories
are listed in the right column. Numbers in parentheses indicate aggregated class membership.
Miscellaneous vegetation class was generated (code 6) to represent land that was vegetated, but
not mature forest or annual row crop.
Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ. 2006. Land use land cover change in the
U.S. Great Lakes basin 1992 to 2001. J. Great Lakes Res. 32: 607-628.

Table 2. Total area (ha) and proportion of watershed converted from non-developed to developed
LULC from 1992 to 2001 for each of the Great Lakes and Lake St. Clair.
Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ. 2006. Land use land cover change in the
U.S. Great Lakes basin 1992 to 2001. J. Great Lakes Res. 32: 607-628.

List of Figures
Figure 1. LULC type changes for the U.S. Great Lake basin by area and percent change
since 1992 (numbers above and below bars).
Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ. 2006. Land use land cover change in the
U.S. Great Lakes basin 1992 to 2001. J. Great Lakes Res. 32: 607-628.

Figure 2. LULC change in the lower Green Bay basin of Lake Michigan (A) and the area
surrounding Detroit, MI (B).
Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ. 2006. Land use land cover change in the
U.S. Great Lakes basin 1992 to 2001. J. Great Lakes Res. 32: 607-628.

Figure 3. Lake-by-lake LULC transitions for the U.S. portion of the Great Lakes basin.
                        Draft for Discussion at SOLEC 2006

-------

Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ.  2006. Land use land cover change in the
U.S. Great Lakes basin 1992 to 2001. J. Great Lakes Res. 32: 607-628.

Figure 4. Lake-by-lake LULC transitions for the U.S. portion of the Great Lakes basin as a
percent of respective watershed area.
Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ.  2006. Land use land cover change in the
U.S. Great Lakes basin 1992 to 2001. J. Great Lakes Res. 32: 607-628.

Last updated
SOLEC 2006
                                              1 Developed
                                              2 Agriculture
                                              3 Early Successional Vegetation
                                              4 Forest
                                              5 Wetland
                                              6 Miscellaneous Vegetation
 (1)     Low Intensity Residential
 (1)     High Intensity Residential
 (1)     Commercial/Industrial
 (1)     Roads (Tiger 1992)
 (3)     Bare Rock/Sand/Clay
 (1)     Quarries/Strip Mines/Gravel Pits
 (6)     Urban/Recreational Grasses
 (2)     Pasture/Hay
 (2)     Row Crops
 (2)     Small Grains
 (3,6)   Grasslands/Herbaceous
 (2,6)   Orchards/Vineyards/Other
 (4)     Deciduous Forest
 (4)     Evergreen Forest
 (4)     Mixed Forest
 (3,6)   Transitional
 (3,6)   Shrubland
 (5)     Open Water
 (5)     Unconsolidated Shore
 (5)     Emergent Herbaceous Wetlands
 (5)     Lowland Grasses
 (5)     Lowland Scrub/Shrub
 (5)     Lowland Conifers
 (5)     Lowland Mixed Forest
 (5)     Lowland Hardwoods

Table 1. Classification scheme used to analyze LULC change in the U.S. portion of the Great Lakes basin.
Original 25 classes are listed in the left column, while aggregated LULC categories are listed in the right
column. Numbers in parentheses indicate aggregated class membership. Miscellaneous vegetation class
was generated (code 6) to represent land that was vegetated, but not mature forest or annual row crop.
Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ. 2006
                         Draft for Discussion at SOLEC 2006

-------
                             State of the Great Lakes 2007 - Draft

Total




watershed area
Non-dev. to
developed
% of watershed
Erie
4994413

87077
1.74
Huron
4114697

42857
1.04
Michigan
11702442

155936
1.33
Ontario
3428229

46507
1.36
Superior
4226924

20351
0.48
Stdair
564825

16112
2.85
Erie/Stdair
5559238

103189
1.86
Table 2. Total area (ha) and proportion of watershed converted from non-developed to developed
LULC from 1992 to 2001 for each of the Great Lakes and Lake St. Clair.
Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ. 2006



150 -


1
U
* 0
0
tn
•o
o c
"/I :
esnoiil

-200 -








n













s. >
•^ W
a
_i














p









W WPfer
i^pp.1 L^*1 ***"*j '^-***WKl
F^SCV H#i H i nJ"j" 0 i'j ^cp r P
^ ^X«cy
	 T 	 <- 	 ^^^^^ 	 1
Mrfff) n»-a«»i *iy»« P*5T Pip^ij^ilpj ^(5 L*riP"4 (VHWY
CON CaHhs^FHH HOW R«*Cn>n IB LMkMSm*
kHX I*«FWII S3 SntfCnn ICON KntardCeMfVHl 1
en SmxuH lAi LMnfKBMUH tkHI Ldn LMWdlUFVHI


D

> Q UJ >-
UJ «^ [£ tJL
Q O 
-------
        State of the Great Lakes 2007 - Draft
                                 Detroit, Ml    /
                                        LULC Transitions 1992-2001
                                           Agriculture to Developed
                                           Forest to Developed
                                          'ESVto Developed
                                       \ —"Forest to Agriculture
                                       -j ~— No Change (Developed)
                                        • ^ No Change (Undeveloped)
Figure 2. LULC change in the lower Green Bay basin of Lake Michigan (A) and the area
surrounding Detroit, MI (B).
Source: Wolter, P.T., Johnston, C.A., andNeimi, GJ.  2006
                       Draft for Discussion at SOLEC 2006

-------
                                  State of the Great Lakes 2007 - Draft
                              Great Lakes LULC Transitions 1992-2001




of hectare;
3 §
i
(0
tfl
.c
H












r
D Superior
• Michigan
D Huron
D St. Clair
• Erie
D Ontario
D U.S. Total





1-1 	 1-1 1 r-l K









-m









A,









i-



















[








.
ill









_^_









j,J








~i
I








PI
t^
       Misc. veg.to Wetland to   ESVto  Agriculture  Forest to   Developed   ESVto   Agriculture  Forest to  Forest to
        Flooded  Developed  Developed    to    Developed  to Misc.   Forest   to Forest    ESV    Agriculture
                              Developed           Veg.
                                    LULC Transition Category


Figure 3. Lake-by-lake LULC transitions for the U.S. portion of the Great Lakes basin.

Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ. 2006
                          Draft for Discussion at SOLEC 2006

-------
          State of the Great Lakes 2007 - Draft
     1.6
     1.4
   g  1.0
   s
   > 0.8
  £ 0.6
  o
    0.2
    0.0
                         LULC Transitions as Percent of Respective Watershed Area
D Superior
• Michigan
D Huron
D St. Clair
• Erie
D Ontario
        Misc. veg.to Wetland to  Early Succ.  Agriculture   Forest to  Developed Early Succ. Agriculture  Forest to   Forest to
         Flooded  Developed   Veg.to     to    Developed to Misc. Veg.  Veg.to    to Forest   Early Succ.  Agriculture
                         Developed  Developed                   Forest              Veg.
                                       LULC Transition Category

Figure 4. Lake-by-lake LULC transitions for the U.S. portion of the Great Lakes basin as a
percent of respective watershed area.
Source: Wolter, P.T., Johnston, C.A., and Neimi, GJ.  2006
                            Draft for Discussion at SOLEC 2006

-------
Brownfields Redevelopment
Indicator #7006
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving
Data from multiple sources are not consistent. Inventories of existing
brownfields are not available in Ontario so it is difficult to determine a trend
for the redevelopment of brownfields. Since more sites are being
redeveloped and/or are being planned, there is some trend of an
improvement in the Great Lakes basin, but it is not based on a quantitative
assessment. Funding and liability issues are obstacles for brownfields
redevelopment and can hinder progress.
Purpose
    •  To assess the area of redeveloped brownfields; and
    •  To evaluate over time the rate at which society remediates and reuses former developed
       sites that have been degraded or abandoned.

Ecosystem Objective
The goal of brownfields redevelopment is to remove threats of contamination associated with
these properties and to bring them back into productive use. Remediation and redevelopment of
brownfields results in two types of ecosystem improvements:

1. reduction or elimination of environmental risks from contamination associated with these
properties; and

2. reduction in pressure for open space conversion as previously developed properties are reused.

State of the Ecosystem
Brownfields  are abandoned, idled, or under-used industrial and commercial facilities where
expansion, redevelopment or reuse is complicated by real or perceived environmental
contamination. In 1999, 21,178 brownfields sites were identified in the United States which was
equivalent to approximately 33,010 hectares (81,568 acres) of land (The United States
Conference of Mayors). Although similar research does not exist for Canada and no inventory
exists for either contaminated or brownfields sites in Ontario, it is estimated that approximately
50,000 to  100,000 brownfields sites may exist in Canada (Globe 2006).

All eight Great Lakes states, Ontario and Quebec have programs to promote remediation or clean-
up and redevelopment of brownfields sites. Several of the brownfields clean-up programs have
been in place since the mid-to-late 1980s, but establishment of more comprehensive brownfields
programs that focus on remediation and redevelopment has occurred during the 1990s. Today,
each of the Great Lakes states has a voluntary clean-up or environmental response program and
there are over 5,000 municipalities with some type of brownfields program in the U.S. (Globe
2006).  These clean-up programs offer a range of risk-based, site-specific background and health
clean-up standards that are applied based on the specifics of the contaminated property and its
intended reuse.
                         Draft for Discussion at SOLEC 2006

-------
                                                -**    ''*''  ~        ''"'"''"*JiitJiJ(
In Quebec, the Revi-Sols program was established in 1998 and is aimed at assessing and cleaning
urban contaminated sites for the purpose of reuse. Through this program, it was possible to
collect some data on the number of contaminated sites in Quebec as it was compulsory for the
land owner to report this information to complete the application for financing. Based on this
program, more than 7,000 sites are included in this inventory.

To encourage redevelopment, Ontario's environmental legislation provides general protection
from environmental orders for historic contamination to  municipalities, creditors and others.
Ontario Regulation  153/04, which came into effect on October 1, 2004, details the requirements
that property owners must meet in order to file a record of site condition. Two technical
documents are referenced by this regulation, one providing applicable site condition standards,
the other providing laboratory analytical protocols for the analysis of soil, sediment and ground
water. A Brownfields Environmental Site Registry offers property owners the opportunity to
complete an online record of site condition with this information then being publicly accessible.
This registry is currently voluntary. As of October 2005, property owners are required to file a
record of site condition before a property or commercial use to a more sensitive area, such as
residential. A record of site condition ensures that a property meets regulated site-assessment and
clean-up standards that are appropriate for the new use (Ontario Legislation Promotes Stronger
Healthier Community).

The 2003 enactment of the New York State Brownfield Law has resulted in increased interest by
private developers and municipalities in the redevelopment of contaminated properties.

Efforts to track brownfields redevelopment are uneven among Great Lakes states and provinces.
Not all jurisdictions track brownfields activities and methods vary where tracking does take place.
States, provinces and municipalities track the amount of funding assistance provided as well as
the number of sites that have been redeveloped. They also  track the number of applications that
have been received for brownfields redevelopment funding. These are indicators of the level of
brownfields redevelopment activity in general, but they do not necessarily reflect land renewal
efforts (i.e., area of land redeveloped), the desired measure for this indicator.  Compiling  state and
provincial data to report a brownfields figure that represents the collective eight states and two
provinces is challenging. Several issues are prominent. First, state and provincial clean-up data
reflect different types of clean-ups, not all of which are "brownfields" (e.g. some include leaking
underground storage tanks and others do not). Second, some jurisdictions have more than one
program, and not necessarily all relevant programs engage in such tracking. Third, program
figures do not include clean-ups that have not been part of a state or provincial clean-up program
(e.g. local or private clean-ups). That said, several states and provinces do track acres of
brownfields remediated, although no Great Lakes state or province tracks acres of brownfields
redeveloped.

Information on area of brownfields remediated from Illinois, Minnesota, New York, Ohio,
Pennsylvania, Quebec and Ontario indicate that, as of August, 2002, a total of 13,413 hectares
(33,143 acres) have been remediated. Available data from  eight Great Lakes states, Quebec and
Ontario indicate that almost 27,000 brownfields sites have participated in brownfields clean-up
programs since the mid-1990s, although the degree of remediation varies considerably. In
                         Draft for Discussion at SOLEC 2006

-------
Ontario, brownfields redevelopment is planned for 108 hectares (267 acres) of land between 2006
and 2008 for the municipalities that participated in this assessment.

Remediation is a necessary precursor to redevelopment. Remediation is often used
interchangeably with "clean-up," though brownfields remediation does not always involve
removing or treating contaminants. Many remediation strategies utilize either engineering or
institutional controls (also known as exposure controls) or adaptive reuse techniques that are
designed to limit the spread of, or human exposure to, contaminants left in place. In many cases,
the cost of treatment or removal of contaminants would prohibit reuse of land. All Great Lakes
states and provinces allow some  contaminants to remain on site as long as the risks of being
exposed to those contaminants are eliminated or reduced to acceptable levels. Capping a site with
clean soil or restricting the use of groundwater are examples of these "exposure controls" and
their use has been a major factor in advancing brownfields redevelopment. Several jurisdictions
keep track of the number and location of sites with exposure controls, but monitoring the
effectiveness of such controls occurs in only three out of the ten jurisdictions.

Redevelopment is a criterion for eligibility under many state brownfields clean-up programs.
Though there is inconsistent and inadequate data on area of brownfields remediated and/or
redeveloped, available data indicate that both brownfields clean-up and redevelopment efforts
have risen dramatically in the mid-1990s and steadily since 2000. The increase is due to risk-
based clean-up standards and the widespread use of state liability relief mechanisms that allow
private parties to redevelop, buy or sell properties without being liable for contamination they did
not cause. Canadian law does not provide liability exemptions for new owners such as those in
the U.S. Small Business Liability Relief and Brownfields Revitalization Act (Globe 2006).
Environmental liability is a major barrier to successful brownfields redevelopment in Canada.
Current owners do not want to sell brownfields sites for fear of liability issues in the future,
purchasers of land do not want to buy sites without some level of protection and municipalities
assume liability when they become site owners (Brownfields Redevelopment versus Greenfield
Development). The Ontario Ministry of Finance has proposed changes under Bill 130 (Municipal
Statute Law Amendment Act, 2006) which would allow brownfields to be advertised as "free" of
any provincial crown liens if a municipality assumes ownership of a property with a failed tax
sale. Also, under certain circumstances, this new policy will allow for the removal of crown liens
on brownfields properties at tax sale. If passed, this change in legislation would reduce some of
the issues related to civil and regulatory liabilities. One recommendation is that once a property
owner has met regulatory standards in the cleanup phase that they are not forced to meet stricter
standards in the future.

In 2005, the Government of Canada allocated $150 million for brownfields remediation. Other
initiatives include the Sustainable Technologies Canada Funding, and the Federal Contaminated
Sites Action Plan. Also, more financial tools for brownfields redevelopment are available though
a Community Improvement Plan (CIP), which allows municipalities to encourage brownfields
redevelopment by offering financial incentives. Other grants and loans can be provided to
supplement the CIP including an exemption or a reduction in the cost of fees associated with
permits, parkland  dedications and zoning amendments. Tax incentives can also be provided by
municipalities to encourage the cleanup of contaminated sites (Financial Tools for Brownfields
Redevelopment).
                         Draft for Discussion at SOLEC 2006

-------
Data also indicate that the majority of clean-ups in the Great Lakes states and provinces are
occurring in older urbanized areas, many of which are located on the shoreline of the Great Lakes
and in the basin. Based on the available information, the state of brownfields redevelopment is
mixed and improving.

Pressures
Laws and policies that encourage new development to occur on undeveloped land instead of on
urban brownfields, are significant and on-going pressures that can be expected to continue.
Programs to monitor, verify and enforce effectiveness of exposure controls are in their infancy,
and the potential for human exposure to contaminants may inhibit the redevelopment of
brownfields. Several Great Lakes states allow brownfields redevelopment to proceed without
cleaning up contaminated groundwater as long as no one is going to use or come into contact with
that water. However, where migrating groundwater plumes ultimately interface with surface
waters, some surface water quality may continue to be at risk from brownfields contamination
even where brownfields have been remediated.

Management Implications
Programs to monitor and enforce exposure controls need to be fully developed and implemented.
More research is needed to determine the relationship between groundwater supplies and
Great Lakes surface waters and their tributaries. Because brownfields redevelopment results in
both reduction or elimination of environmental risks from past contamination and reduction in
pressure for open space land conversion, data should be collected that will enable an evaluation of
each of these activities. For every hectare (2.5 acres) developed in a brownfields project, it can
save an estimated minimum of 4.5 hectares (11 acres) of land from being developed in an
outlying area (Cleaning Up the Past, Building the Future).

Ontario is expected to add 3.7 million more people to its population in the next 25 years with
most of the growth occurring in the  Greater Golden Horseshoe (western end of Lake Ontario)
(Places to Grow: Better Choices, Brighter Future). Brownfields redevelopment needs to be a part
of the planning and development reform in order to address the issue of urban sprawl.

Comments from the author(s)
Great Lakes states and provinces have begun to track brownfields remediation and or
redevelopment, but the data is generally inconsistent or not available in ways that are helpful to
assess progress toward meeting the terms of the Great Lakes Water Quality Agreement. Though
some jurisdictions have begun to implement web-based searchable applications for users to query
the status of brownfields sites, the data gathered are not necessary consistent, which presents
challenges for assessing progress in the entire basin.  States and provinces  should develop
common tracking methods and work with local jurisdictions  incorporating local data to online
databases that can be searched by: 1) area remediated; 2) mass of contamination removed or
treated (i.e., not requiring an exposure control); 3) type of treatment; 4) geographic location; 5)
level of urbanization; and 6) type of reuse (i.e., commercial, residential, open, none, etc). A recent
development in the province of Ontario is the designation of a Provincial Brownfields
Coordinator who will coordinate provincial brownfields activities and provide a single point of
access on brownfields in Ontario.
                         Draft for Discussion at SOLEC 2006

-------

Acknowledgments
Authors: Victoria Pebbles, Senior Project Manager, Transportation and Sustainable Development,
Great Lakes Commission, Ann Arbor, MI, vpebbles@glc.org, www.glc.org.
Updated by: Stacey Cherwaty-Pergentile, A/Science Liaison Officer, Environment Canada,
Burlington, ON, Stacey.Cherwaty@ec.gc.ca, and Elizabeth Hinchey Malloy, Great Lakes
Ecosystem Extension Specialist, Illinois-Indiana Sea Grant,  Chicago, IL,
Hinchev.Elizabeth@epa.gov. www.iisgcp.org.

Contributors
Personal communication with Great Lakes State Brownfields/Voluntary Cleanup Program
Managers:
David E. Hess, Director, Land Recycling Program, Pennsylvania Department of Environmental
Protection
Andrew Savagian, Outreach Specialist ,Remediation and Redevelopment (RR) Program
Wisconsin Department of Natural Resources
Ron Smedley, Brownfield Redevelopment Coordinator, Michigan DEQ Remediation and
Redevelopment
Gerald Stahnke, Project Leader, Voluntary Investigation and Cleanup Unit, Minnesota Pollution
Control Agency
Susan Tynes Harrington, Indiana Brownfields Program, Indiana Finance Authority
Amy Yersavich, Manager, Voluntary Action Program, Ohio EPA

Personal communication with Provincial as well as Canadian municipalities within the Great
Lakes basin including:
City of Barrie, Nancy Farrer, Policy Planner
City of Cornwall, Ken Bedford, Senior Planner
City of Hamilton, Carolynn Reid, Brownfields Coordinator
City of Mississauga, Jeff Smylie, Environmental Engineer
City of Kingston, Joseph Davis, Manager, Brownfields and Initiatives
City of Kitchener, Terry Boutilier, Brownfields Coordinator
City of London, Terry Grawey, Planning Division
City of Thunder Bay, Katherine Dugmore, Manager of Planning Division
City of Toronto, Glenn Walker, Economic Development Officer
City of Toronto Economic Development Corporation (TEDCO)
Province of Quebec, Michel Beaulieu

Data Sources
Selected Annual Reports of state cleanup programs.

Association of Municipalities  of Ontario Report on Brownfields Redevelopment - What has been
Achieved, What Remains to be done, May 2006.
http://www.amo. on.ca/AM/Template.cfm?Section=Eventsl&Template=/CM/HTMLDisplay.cfm
&ContentID=65396. last accessed October 11, 2006.

Brownfields Redevelopment versus Greenfield Development, City of Hamilton Planning and
Development Department, http://www.vision2020.hamilton.ca/downloads/POINTS-TO-
PONDER-Brownfields-vs-Greenfield-
                        Draft for Discussion at SOLEC 2006

-------
Development.pdf#search=%22Brownfields%20Redevelopment%20versus%20Greenfield%20De
velopment%2C%20Citv%20oP/o20Hamilton%20Planning%20and%20Development%20Departm
ent%22. last accessed October 11, 2006.

Brownfields Redevelopment in Small Urban and Rural Municipalities, Summer 2006. Ministry of
Municipal Affairs and Housing. Government of Ontario, ISBN 1-4249-1635-6.
www.brownfields.ontario.ca.

Brownfields Ontario website www.mah.gov.on.ca/userfiles/HTML/nts 1 3305_l.html, last
accessed October 11, 2006.

Cleaning Up the Past, Building the Future. A National Brownfields Redevelopment Strategy for
Canada. National Round Table on the Environment and the Economy 2003, ISBN 1-894737-05-9,
http://www.nrtee-trnee.ca/Publications/HTML/SOD Brownfields-StrategyE.htm. last accessed
October 11,2006.

Delcan, Golder Associates Ltd., and McCarthy - Tetrault. Urban Brownfields: Case Studies for
Sustainable Economic Development. The Canadian Example. Canada Mortgage and Housing,
p. 1.

Financial Tools for Brownfields Redevelopment, Summer 2006. Ministry of Municipal Affairs
and Housing. Government of Ontario, ISBN 104249-1956-8. www.brownfields.Ontario.ca.

Globe 2006, Vol. 27, No. 7, pp 254 - 259, ISSN 0149-8738, Bureau of National Affairs, Inc.,
Washington, D.C., 2006.

Ministry of Municipal Affairs and Housing, Remarks from Honourable John Gerretsen,
Association of Municipalities of Ontario  Annual Conference, August 15, 2006.
www.mah.gov.on.ca/userfiles/HTML/nts 1  276111 .html, last accessed October 11, 2006.

Ontario's Brownfields Legislation Promotes Stronger, Healthier Communities - June 2006, News
Release, Ontario Ministry of the Environment,
www.ene.gov.on.ca/envision/news/2005/062201.htm, last accessed October 11, 2006.

Places to Grow: Better Choices, Bright Futures - A Proposed Growth Plan for the Greater Golden
Horseshoe, November 2005, Ministry of Public Infrastructure and Renewal, ISBN 0-7794-9089-4.

Stakeholders Urge Government to Limit  Brownfields Liability,
http://www.willmsshier.com/newsletters.asp?id=30, last accessed October 11, 2006.

The United States Conference  of Mayors. A National Report in Brownfields Redevelopment -
Volume 3. Feb. 2000, p. 12.

List of Tables
Table 1. Summary of acres remediated and number of sites remediated in the Great Lakes basin,
1990-2006.
                        Draft for Discussion at SOLEC 2006

-------


Source: Various state, municipal and provincial brownfields coordinators and city planners

List of Figures
Figure 1. Redeveloped brownfields site, Spencer Creek, Hamilton, Ontario.
Source: City of Hamilton

Last updated
SOLEC 2006
State/Province
WI
PA
OH
MI
IN
MN
IL
NY
ON
QC
Total
Acres remediated
1,220
13,229
4,204
not tracked
not tracked
7,047
6,412
55
92
741
33,143
Hectares remediated
494
5354
1701
not tracked
not tracked
2852
2595
22
37
300
13,413
Time frame
2004-2006
2000- 2006
1994-2006


1998-2002
1990-2001
2000-2002
2002-2005
1998-2002

Sites remediated
18,000
1,097
156
5,539f
382
462
899
16
13
309
26,873
Time frame
1994-2005
1996-2002
1996-2002
1995-2002
1997-2002
1998-2002
1990-2001
2000-2002
2002-2005
1998-2005

Table 1. Summary of acres remediated and number of sites remediated in the Great Lakes basin,
1990-2006.
Source: Various state, municipal and provincial brownfields coordinators and city planners
                        Draft for Discussion at SOLEC 2006

-------
                                                         •  jEz/f '"                 A    I
                                                   ^yjstfjtA'»ifM^waSi-l'i W>"' **"—^'T^"-"~""^mML^
                                                   "^^^^^M^^^f'^^^M^	
Figure 1. Redeveloped brownfields site, Spencer Creek, Hamilton, Ontario.
Source: City of Hamilton
                         Draft for Discussion at SOLEC 2006

-------
Sustainable Agriculture Practices
Indicator #7028

Assessment: Not Assessed

Purpose
  To assess the number of environmental and conservation farm
plans and environmentally friendly practices in place such as:
integrated pest management to reduce the potential adverse
impacts of pesticides; conservation tillage and  other soil preser-
vation practices to reduce energy consumption and sustain natu-
ral resources and to prevent ground and surface water contami-
nation.

Ecosystem Objective
The goal is to create a healthy and productive land base that sus-
tains food and fiber, maintains functioning watersheds and natu-
ral systems, enhances the environment and improves the rural
landscape. The sound use and management of soil, water,  air,
plant, and animal resources is needed to prevent degradation of
agricultural resources. The process integrates natural resource,
economic, and social considerations to meet private and public
needs. This indicator supports Annex 2, 3, 12 and 13 of the
Great Lakes Water Quality Agreement.
State of the Ecosystem
Background
Agriculture accounts for approxi-
mately 35% of the land area of
the Great Lakes basin and domi-
nates the southern portion of the
basin. In years past, excessive
tillage and intensive crop rota-
tions led to soil erosion and the
resulting sedimentation of major
tributaries. Inadequate land man-
agement practices contributed to
approximately 57 metric tons of
soil eroded annually by the
1980s.  Ontario estimated its costs
of soil erosion and nutrient/pesti-
cide  losses at $68 million (CA)
annually. In the United States,
agriculture is a major user of pes-
ticides, with an annual use of
24,000 metric tons. These prac-
tices lead to a decline of soil
organic matter. Since the late
1980s,  there has been increasing
participation by Great Lakes
basin farmers in various soil and
water quality management pro-
                             grams. Today's conservation systems have reduced the rates of
                             U.S. soil erosion by 38% in the last few decades. The adoption
                             of more environmentally responsible practices has helped to
                             replenish carbon in the soils back to 60% of turn-of-the-century
                             levels.

                             Both the Ontario Ministry of Agriculture and Food (OMAF) and
                             the U.S. Department of Agriculture (USDA), Natural Resources
                             Conservation Service (NRCS) provide conservation planning
                             advice, technical assistance and incentives to farm clients and
                             rural landowners. Clients develop and implement conservation
                             plans to protect, conserve, and enhance natural resources that
                             harmonize productivity, business objectives and the environ-
                             ment. Successful implementation of conservation planning
                             depends largely upon the voluntary participation of clients.
                             Figure 1 shows the number of acres of cropland in the U.S. por-
                             tion of the Great Lakes basin that are covered under a conserva-
                             tion plan.

                             The Ontario Environmental Farm Plan (EFP) encourages farm-
                             ers to develop action plans and adopt environmentally responsi-
                             ble management practices and technologies. Since 1993, the
                             Ontario  Farm Environmental Coalition (OFEC), OMAF, and the
                             Ontario  Soil and Crop Improvement Association (OSCIA) have
                             cooperated to deliver EFP workshops. The Canadian federal
                             government, through various programs over the years, has pro-
                                                            Total Atres Planned

                                                            CHO-5,OCO Acres
                                                                b.UIJC  Ib-X'jAtri-a
                                                            I    I 15,000 -25.000 Acres
                                                            •I 25,000 - 50.000 Acres
Figure 1. Acres of cropland in U.S portion of the basin covered under a conservation plan, 2003.
Source: Natural Resource Conservation Service, U.S. Department of Agriculture
                                                                                                                     211

-------
                                             OF   THE   GREAT
                             2007
vided funding for EFP. As can be seen from Figure 2 the number
of EFP incentive claims rose dramatically from 1997 through
2004, particularly for the categories of soil management, water
wells, and storage of agricultural wastes. As part of Ontario's
Clean Water Strategy, the Nutrient Management Act (June 2002)
is setting province-wide standards to address the effects of agri-
cultural practices on the environment, particularly as they relate
to land-applied materials containing nutrients.
3000
2500
2000
in
| 1500
O
1000
500
0


_.-»
^••*" 2763
- .^*
.^^^^^2488
^^^2338
4^2097 m
.' *f~ 2021
/ ^»'''~
y _^^^*68°
y ^^^^ 1506 1191
f° VH252
X"" S' 	 	 "
*' 	 "L ' 1029 802
^^..'•^•^^''-^

-------
Ministry of Agriculture and Food (OMAF), Guelph, Ontario
Canada, peter.roberts@omaf.gov.on.ca;
Ruth Shaffer, United States Department of Agriculture (USDA),
Natural Resource Conservation Service (NRCS),
ruth.shaffer@mi.usda.gov; and
Roger Nanney, United States Department of Agriculture
(USDA), Natural Resources Conservation Service (NRCS),
roger.nanney@in.usda.gov.

Sources
Ontario Soil and Crop Improvement Association. 2004.
Environmental Farm Plan Database.

Last Updated
State of the Great Lakes 2005
                                                                                                                 213

-------
   Economic Prosperity
   Indicator #7043

   Assessment: Mixed (for Lake Superior Basin), Trend Not
   Assessed
   Data are not system-wide.

   Purpose
     To assess the unemployment rates within the Great Lakes
   basin; and
     To infer the capacity for society in the Great Lakes region to
   make decisions that will benefit the Great Lakes ecosystem
   (when used in association with other Great Lakes indicators).

   Ecosystem Objective
   Human economic prosperity is a goal of all governments. Full
   employment (i.e. unemployment below 5% in western societies)
   is a goal for all economies.

   State of the Ecosystem
   This information is presented to supplement the report on
   Economic Prosperity in SOLEC 2000 Implementing Indicators
   (Draft for Review, November 2000). In 1975, 1980, 1985, 1990,
   1995 and 2000 the civilian unemployment rate in the 16 U.S.
   Lake Superior basin counties averaged about 2.0 points above
   the U.S. average, and above the averages for their respective
   states, except occasionally Michigan (Figure 1). For example.
   the unemployment rate in the four Lake Superior basin counties
         1975
                  1980
                           1985      1990
                                Year
                                             1995
                                                      2000
           • United States             DMichigan
           • Minnesota               nWisconsin
           ]U.S. Lake Superior Counties   nOntario L. Superior Basin 1996
Figure 1. Unemployment rate in the U.S. (national), Michigan.
Wisconsin, and the U.S. portion and Ontario portion of the Lake
Superior basin, 1975-2000.
Source: U.S. Census Bureau and Statistics Canada
in Minnesota was consistently higher than for Minnesota overall.
2.7 points on average but nearly double the Minnesota rate of
6.0% in 1985. Unemployment rates in individual counties
ranged considerably, from 8.6% to 26.8% in 1985, for example.

In the 29 Ontario census subdivisions mostly within the Lake
Superior watershed, the 1996 unemployment rate for the popula-
tion  15 years and over was 11.5%. For the population 25 years
and older, the unemployment rate was 9.1%. By location the
rates ranged from 0% to 100%; the extremes, which occur in
adjacent First Nations communities, appear to be the result of
small populations and the 20% census sample. The most popu-
lated areas, Sault Ste. Marie and Thunder Bay, had unemploy-
ment rates for persons 25 years and older of 9.4% and 8.6%.
respectively. Of areas with population greater than 200 in the
labour force, the range was from 2.3% in Terrace Bay Township
to 31.0% in Beardmore Township. Clearly, the goal of full
employment (less than 5% unemployment) was not met in either
the Canadian or the U.S. portions of the Lake Superior basin
during the years examined.
Acknowledgments
Authors: Kristine Bradof, GEM Center for Science and
       16.0
                                ft
                                                                             Individuals  Individuals Individuals  Families
                                                                               1979      1989      1999      1996
                                                                                              Year
                                                                             • USA      • Minnesota    D U.S. L. Superior Basin
                                                                             D Michigan   • Wisconsin    • Ontario L. Superior Basin
Figure 2. Individuals below poverty level in the U.S. (national).
Michigan, Wisconsin, and the U.S. Great Lakes basin counties.
1979-1999, and families below poverty level in Ontario Great
Lakes basin subdivisions, 1996.
Source: U.S. Census Bureau and Statistics Canada
Environmental Outreach, Michigan Technological University.
MI; and
James G. Cantrill, Communication and Performance Studies.
Northern Michigan University, MI.
   214

-------
                                                 1999
              • USA
              D Michigan
I Minnesota
I Wisconsin
D Lake Superior Basin
Figure 3. Children under age 18 below the poverty level, 1979-
1999, U.S. (national), Michigan, Minnesota, Wisconsin and U.S.
portion of the Lake Superior basin.
Source: U.S. Census Bureau
 Sources
 GEM Center for Science and Environmental Outreach. 2000.
 Baseline Sustainability Data for the Lake Superior Basin: Final
 Report to the Developing Sustainability Committee, Lake
 Superior Binational Program, November 2000. Unpublished
 report,  Michigan Technological University, Houghton, MI.
 htto://emmap.mtu.edu/gem/communitv/Dlanning/lsb.html.
for the overall population, children under age 18, families, and
persons age 65 and older. Two examples of trends in those meas-
ures are shown in Figures 2 and 3. For persons of all ages within
the U.S. Lake Superior basin for whom poverty status was estab-
lished, 10.4% were below the poverty level in 1979. That figure
had risen to 14.5% in 1989, a rate of increase higher than the
states of Michigan, Minnesota,  and Wisconsin and the U.S. over-
all over the same period. Poverty rates for individuals and chil-
dren in the U.S. Lake Superior basin in 1979, 1989, and 1999
ranged from 10.4% to  17.1%, while 12.8% of families in the
Ontario Lake Superior basin had incomes below the poverty
level in 1996. Poverty rates in all areas were lower in 1999, but
the U.S. Lake Superior basin (and Ontario portion of the basin in
1996) was higher than any of the three states. The 1979 poverty
rate for counties within the Lake Superior basin ranged from a
low of 4.4% in Lake County, Minnesota, to a high of 17.0% in
Houghton County, Michigan. In 1989  and 1999, those same
counties again were the extremes. Similarly, among children
under age 18, poverty rates in the Great Lakes basin portions of
the three  states in 1979, 1989, and 1999 exceeded the rates of
Minnesota and Wisconsin as a whole,  though they remained
below the U.S. rate. In a region where one-tenth to one-sixth of
the population lives in poverty,  environmental Sustainability is
likely to be perceived by many  as less important than economic
development.

Last Updated
State of the Great Lakes 2003
 Statistics Canada. 1996. Beyond 20/20 Census Subdivision Area
 Profiles for the Ontario Lake Superior Basin.

 U.S. Census Bureau. 2002. Population by poverty status in 1999
 for counties: 2000.
 http://www.census.gov/hhes/poverty/2000census/poppvstatOO.ht
 ml.

 U.S. Census Bureau. State & County Quick Facts 2000. Table
 DP-3. Profile of Selected Economic Characteristics.
 http://censtats.census.gov/data/MI/ O4026.pdf#page=3.

 U.S. Census Bureau. USA Counties 1998 CD-ROM (includes
 unemployment data from Bureau of Labor Statistics).
 Authors' Commentary
 As noted in the State of the Great Lakes 2001 report for this
 indicator, unemployment may not be sufficient as a sole meas-
 ure. Other information that is readily available from the U.S.
 Census Bureau and Statistics Canada includes poverty statistics
                                                                                                                     215

-------
Water Withdrawals
Indicator #7056

Assessment: Mixed, Unchanging

Purpose
  To use the rate of water withdrawal to help evaluate the sus-
tainability of human activity in the Great Lakes basin.

Ecosystem Objective
The first objective is to protect the basin's water resources from
long-term depletion. Although the volume of the Great Lakes is
vast, less than one percent of their waters are renewed annually
through precipitation, run-off and infiltration. Most water with-
drawn is returned to the watershed, but water can be lost due to
evapotranspiration, incorporation into manufactured goods, or
diversion to other drainage basins. In this sense, the waters of
the Great Lakes can be considered a non-renewable resource.

The second objective is to minimize the ecological impacts
stemming from water withdrawals. The act of withdrawing water
can shift the flow regime, which in turn can affect the health of
aquatic ecosystems. Water that is returned to the basin after
human use can also introduce contaminants,  thermal pollution or
invasive species into the watershed. The process of withdrawing,
treating and transporting water also requires  energy.
State of the Ecosystem
Water was withdrawn from the Great Lakes basin at a rate of
46,046 million gallons per day (MGD) in 2000 (or 174 billion
litres per day), with almost two-thirds withdrawn in the U.S. side
(30,977 MGD) and the remaining one-third in Canada (15,070
MGD). Self-supplying thermoelectric  and industrial users with-
drew over 80% of the total. Public water systems, which are the
municipal systems that supply households, commercial users and
other facilities, comprised 13% of withdrawals. The rural sector,
which includes both domestic and agricultural users, withdrew
2%, with the remaining 3% used for environmental, recreation,
navigation and quality control purposes. Hydroelectric use,
which is considered "in-stream use" because water is not actual-
ly removed from its source, accounted for additional with-
drawals at a rate of 799,987 MGD (Figure 1) (GLC 2004).
                                                              Withdrawal rates in the late 1990s were below their historical
                                                              peaks and do not appear to be increasing at present. On the U.S.
                                                              side, withdrawals have dropped by more than 20% since 1980,
                                                              following rapid increases from the 1950s onwards (USGS  1950-
                                                              2000)1. Canadian withdrawals continued rising until the mid-
                                                              1990s, but have decreased by roughly 30% since then (Harris
                                                              and Tate 1999)2. In both countries, the recent declines have been
                                                              caused by the shutdown of nuclear power facilities, advances in
                                                              water efficiency in the industrial sector, and growing public
                                                              awareness on resource conservation. Part of the decrease, how-
                                                              ever, may be attributed to improvements in data collection meth-
                                                              ods over time (USGS 1985). Refer to Figures 2,3  and 4.

                                                              The majority of waters withdrawn are returned to the basin
                                                              through run-off and discharge. Approximately 5% is made
                                                              unavailable, however, through evapotranspiration or

-------
"ro
ra
40000 -i

38000 -

36000 -

34000 -

32000 -

30000 -

28000 -

26000 -

24000 -
         1950
              1960
1970     1980
   Year
1990
2000
                       -USGS
                                          -GLC
Figure 3. U.S. basin water withdrawals, 1950-2000.
Source: U.S. Geological Survey, 1950-2000. Great Lakes
Commission (GLC).
     30000 -i
     25000 -
     20000 -
     15000 -
^   10000 -
      5000 -
         0
         1970    1975
                        1980
                               1985    1990
                                 Year
                                              1995   2000
                           -Gaia
                                       -GLC
Figure 4. Canadian basin water withdrawals, 1972-2000.
Source: Gaia Economic Research Associates, 1999 (based on data
from Environment Canada and Statistics Canada). Great Lakes
Commission (GLC).
  incorporation into manufactured products. This quantity, referred
  to as "consumptive use," represents the volume of water that is
depleted due to human activity. It is argued that consumptive
use, rather than total water withdrawals, provides a more suitable
indicator on the sustainability of human water use in the region.
Basin-wide consumptive use was estimated at 3,166 MGD in
2000. Although there is no consensus on an optimal rate of con-
sumptive use, a loss of this magnitude does not appear to be
placing significant pressure on water resources. The long-term
Net Basin Supply of water (sum of precipitation and run-off.
minus natural evapotranspiration), which represents the maxi-
mum volume that can be consumed without permanently reduc-
ing the availability of water, and equals the volume of water dis-
charged from Lake Ontario into the St. Lawrence River, is esti-
mated to be 132,277 MGD (estimate is for 1990-1999 period
Environment Canada 2004). It should be noted, however, that
focusing on these basin-wide figures can obscure pressures at
the local watershed level.

Moreover, calculating consumptive use is a major challenge
because of the  difficulty in tracking the movement of water
through the hydrologic cycle. Consumptive use is currently
inferred by multiplying withdrawals against various coefficients.
depending on use type. For instance, it is assumed that thermo-
electric users consume as little as 1% of withdrawals, compared
to a loss rate of 70-90% for irrigation (GLC 2003). There  are
inconsistencies in the coefficients used by the various states and
provinces. Estimating techniques were even more rudimentary in
the past, making it problematic to discuss historical consumptive
use trends. Due to these data quality concerns,  it may not yet be
appropriate to consider consumptive use as a water use  indicator.

Water removals from diversions, by contrast, are monitored
more closely, a result of the political attention that prompted the
region's governors and premiers to sign the Great Lakes Charter
in 1985. The Charter and its Annexes require basin-wide notifi-
cation and consultation for water exports, while advocating that
new  diversions be offset by a commensurate return of water to
the basin. The two outbound diversions approved since  1985
have accommodated this goal by diverting water in from exter-
nal basins. The outbound diversions already in operation by
1985, most notably the Chicago diversion, were not directly
affected by the Charter, but these losses are more than offset by
inbound diversions located in northwestern Ontario. Thus, there
is currently no net loss of water due to diversions.

There is growing concern over the depletion of groundwater
resources, which cannot be replenished following withdrawal
with the same ease as surface water bodies. Groundwater was
withdrawn at a rate of 1,541 MGD in 2000, making up  3% of
total water withdrawals (GLC 2004). This rate  may not have a
major effect on the basin as a whole, but high-volume with-
drawals have outstripped natural recharge rates in some loca-
tions. Rapid groundwater withdrawals in the Chicago-

                                                       219

-------
Milwaukee region during the late 1970s produced cones of
depression in that local aquifer (Visocky 1997). However, the
difficulty in mapping the boundaries of groundwater supplies
makes unclear whether the current groundwater withdrawal rate
is sustainable.

Pressures
The Great Lakes Charter, and its domestic legal corollaries in the
U.S. and Canada, was instituted in response to concerns over
large-scale water exports to markets such as the arid southwest-
ern U.S. There does not appear to be significant momentum  for
such long  distance shipments due to legal and regulatory barri-
ers, as well as technical difficulties and prohibitive costs. In  the
immediate future, the greatest pressure will come from commu-
nities bordering the basin, where existing water supplies are
scarce or of poor quality. These localities might look to the Great
Lakes as a source of water. Two border-basin diversions have
been approved under the Charter and have not resulted in net
losses of water to the basin. This outcome, however, was
achieved through negotiation and was not proscribed by treaty or
law.

As for withdrawals within the basin, there is no clear trend in
forecasting regional water use.  Reducing withdrawals, or at least
mitigating further increases, will be the key to lessening con-
sumptive use. Public water systems currently account for the
bulk of consumptive use, comprising one-third of the total, and
withdrawals in this category have been increasing in recent years
despite the decline in total withdrawals. Higher water prices
have been widely advocated in order to reduce water demand.
Observers have noted that European per-capita water use is only
half the North American level, while prices in the  former are
twice as high. However, economists have found that both resi-
dential and industrial water demand in the U.S. and Canada are
relatively insensitive to price changes (Renzetti 1999, Burke et
al. 2001)3. The over-consumption of water in North America
may be more  a product of lifestyle and lax attitudes. Higher
prices may still be crucial for providing public water systems
with capital for repairs; this can prevent water losses by fixing
system leaks, for example.  But reducing the underlying demand
may require other strategies in addition to price increases, such
as public education on resource conservation and promotion of
water-saving technologies.

Assessing the availability of water in the basin will be compli-
cated by factors outside local or human control. Variations in cli-
mate and precipitation have produced long-term fluctuations in
surface water levels in the past. Global climate change could
cause similar impacts; research suggests that water levels may be
permanently lower in the future as a result. Differential move-
ment of the Earth's crust, a phenomenon known as isostatic
rebound, may exacerbate these effects at a local level. The crust

220
                                                                                   •i' s:  2007
is rising at a faster rate in the northern and eastern portions of
the basin, shifting water to the south and west. These crustal
movements will not change the total volume of water in the
basin, but may affect the availability of water in certain areas.

Acknowledgments
Author: Mervyn Han, Environmental Careers Organization, on
appointment to U.S. Environmental Protection Agency, Great
Lakes National Program Office.

Rebecca Lameka (Great Lakes Commission), Thomas Crane
(Great Lakes Commission), Wendy Leger (Environment
Canada), and Fabien Lengelle (International Joint Commission)
assisted in obtaining data for this report. Steven  Renzetti (Brock
University) and Michel Villeneuve (Environment Canada) assist-
ed in explaining water consumption economics.

Site-specific water withdrawal data courtesy of James Casey
(Illinois Department of Natural Resources), Sean Hunt
(Minnesota Department of Natural Resources), Paul Spahr (Ohio
Department of Natural Resources) and Ralph Spaeth (Indiana
Department of Natural Resources). Ontario water permit map
courtesy of Danielle Dumoulin (Ontario Ministry of Natural
Resources).

Sources
Burke, D., Leigh,  L., and Sexton, V. 2001. Municipal water pric-
ing, 1991-1999. Environment Canada, Environmental
Economics Branch.

Environment Canada. 2004. Great Lakes-St. Lawrence
Regulation Office.

Gaia Economic Research Associates. 1999. Water demands in
the Canadian section of the Great Lakes basin 1972-2021.

Great Lakes Commission (GLC). 2004. Great Lakes regional
water use database.
http://www.glc.org/wateruse/database/search.html.

Great Lakes Commission (GLC). 2003. Toward  a water
resources management decision support system for the Great
Lakes-St. Lawrence River basin: status of data and information
on water resources,  water use, and related ecological impacts.
Chapter.3, pp.58-62.
http://www.glc.org/wateruse/wrmdss/finalreport.html.

Harris, J., and Tate, D. 1999. Water demands in  the Canadian
section of the Great Lakes  basin, 1972-2021. Gaia Economic
Research Associates (GERA) Report, Ottawa, ON.

Mills, E.L., Leach, J.H., Carlton, J.T., and Secor, C.L. 1993.

-------
Exotic species in the Great Lakes: a history of biotic crises and
anthropogenic introductions./. Great Lakes Res.  19(1): 1-54.
Renzetti, S. 1999. Municipal water supply and sewage treatment:
costs, prices and distortions. The Canadian Journal of
Economics. 32(3):688-704.

U.S. Geological Survey (USGS). 1950-2000. Estimated Water
Use in the United States: circulars published at 5-year intervals
since 1950. http://water.usgs.gov/watuse/.

U.S. Geological Survey (USGS). 1985. Estimated use of water
in the United States in 1985. 68pp.

Visocky, A.P. 1997. Water-level trends and pumpage in the deep
bedrock aquifers in the Chicago region, 1991-1995. Illinois State
Water Survey Circular 182. Cited in International Joint
Commission. 2000. Protection of the waters of the Great Lakes:
final report to the governments of Canada and the United States.
Chapter.6, pp 20-26. http://www.ijc.org/php/publications/
html/finalreport.html.

Endnotes
1 USGS estimates show water withdrawals in the U.S. Great
Lakes watershed increasing from 25,279 MGD in 1955 to a peak
in the 36-39,000 MGD range during the 1970-80 period, but
dropping to the 31-32,000 MGD range for 1985-1995. GLC
reported U.S. water withdrawals in the 32-34,000 range for
1989-1993, and around 30,000 MGD since 1998, with 30,977
MGD in 2000.

2 Historical Canadian data from Gaia Economic Research
Associates (GERA) report, and are based on data from Statistics
Canada and Environment Canada. GERA reported that Canadian
water withdrawals increased from 8,136 MGD in 1972 to  21,316
MGD in 1996. GLC reported Canadian withdrawals of 21-
24,000 MGD in 1989-1993, around 17,000 MGD for 1998 and
1999, and  15,070 MGD in 2000.

3 Econometric studies of both residential and industrial water
demand consistently display relatively small price elasticities.
Literature review on water pricing economics can be found in
Renzetti (1999). However, the relationship between water
demand and price structure is complex. The introduction of vol-
umetric pricing (metering), as opposed to flat block pricing
(unlimited use), is indeed associated with lower water use, per-
haps because households become more aware of their water
withdrawal rate (Burke et al. 2001).
Authors' Commentary
Water withdrawal data is already being compiled on a systemic
basis. However, improvements can be made in collecting more
 accurate numbers. Reporting agencies in many jurisdictions do
 not have, or do not exercise, the statutory authority to collect
 data directly from water users, relying instead on voluntary
 reporting, estimates, and models. Progress is also necessary in
 establishing uniform and defensible measures of consumptive
 use, which is the component of water withdrawals that most
 clearly signals the sustainability of current water demand.

 Mapping the point sources of water withdrawals could help
 identify local watersheds that may be facing significant pres-
 sures. In many jurisdictions, water permit or registration pro-
 grams can provide suitable geographic data. However, only in a
 few states (Minnesota, Illinois, Indiana and Ohio) are withdraw-
 al data available per registered facility. Permit or registration
 data, moreover, has limited utility in locating users that are not
 required to register or obtain permits, such as the rural sector, or
 facilities with a withdrawal capacity below the statutory thresh-
 old (100,000 gallons per day in most jurisdictions.) Refer to
 Figures 5 and 6.

 Further research into the ecological impact of water withdrawals
 should also be a priority. There is evidence that discharge from
 industrial and thermoelectric plants, while returning water to the
 basin, alters the thermal and chemical integrity of the lakes. The
 release of water at a higher than normal temperature has been
 cited as facilitating the establishment of non-native species
 (Mills et al. 1993). The changes to the flow regime of water,
 through hydroelectric dams, internal diversions  and canals, and



     •••,.:   '
   • Withdrawal Capacities exceeding 100 Million Litres per Day
   • Water Withdrawal locations
Figure 5. Permitted water withdrawal capacities in the Ontario
portion of the Great Lakes basin.
Source: Ontario Ministry of Natural Resources
                                                                                                                      221

-------
                                            Wat.r Withdrawal! ptr Rtgiiurtd Facility
                                               Million of «.illoiit |>ei D.
-------
Energy Consumption
Indicator #7057

Assessment: Mixed, Trend Not Assessed

Purpose
  To assesses the energy consumed in the Great Lakes basin
per capita; and
  To infer the demand for resource use, the creation of waste
and pollution, and stress on the ecosystem.

Ecosystem Objective
Sustainable development is a generally accepted goal in the
Great Lakes basin. Resource conservation minimizing the
unnecessary use of resources is an endpoint for ecosystem
integrity and sustainable development. This indicator supports
Annex 15 of the Great Lakes Water Quality Agreement.

State of the Ecosystem
Energy use per capita and total consumption by the commercial.
residential, transportaion, industrial, and electricity sectors  in
the Great Lakes basin can be calculated using data extracted
from the Comprehensive Energy Use Database (Natural
Resources Canada),  and the State Energy Data 2000
Consumption tables  (U.S. EIA2000).  Table 1 lists populations
and total consumption in the Ontario and U.S. basins, with the
U.S. basin broken down by states. For this report, the U.S.  side
of the basin is defined as the portions  of the eight Great Lakes
states within the basin boundary (which totals 214 counties
either completely or partially within the basin boundary). The
Ontario basin is defined by eight sub-basin watersheds. The
most recent data available are from 2002 for Ontario and 2000
for the U.S. The largest change between 2000 and 2002 energy
consumption by sector in Ontario was a 4.4% increase in the
commercial sector (all other sectors changed by less than 2% in
either direction).

In Ontario, the per capita energy consumption increased by 2%
between 1999 and 2000. In the U.S. basin, per capita consump-
tion decreased by an average of 0.875% from 1999 to 2000.
Five states showed decreases in per capita energy consumption.
while three states  had increases (Figure 1). Electrical energy
consumption per capita was fairly similar on both sides of the
basin in 2000 (Figure 2). Over the last four decades, consump-
tion trends in the U.S. basin have been fairly steady, although
per capita consumption increased in each state from 1990 to
2000 (Figure 3). Interestingly, New York and Ohio consumed
less per capita in 2000 than in 1970. Looking at the trends  in
Ontario from 1970 to 2000, the per capita energy consumption
has stayed relatively consistent, with the exception of an
increase seen in 1980. The per capita energy consumption fig-
ures for Ontario do not include the electricity generation sector
  due to an absence of data for this sector up until 1978. It is
  important to note that the quality of data processing and valida-
  tion has improved over the last four decades and therefore the
  data quality may be questionable for the 1970s.

  Total secondary energy consumption by the five sectors on the
160


140

120


100
               1
                            State/Province
 Figure 1. Total energy consumption per capita 1999-2000. 1 MWh :
 lOOOkWh.
 Source: Energy Information Administration (2000) and Natural
 Resources Canada (2000)
                             State/Province
Figure 2. Electric energy consumption per capita 2000. 1 MWh =
lOOOkWh.
Source:  Energy Information Administration (2000) and Natural
Resources Canada (2000)
                                                                                                                     223

-------
   160 -,


   140 -


   120 -
 § 100 -
 £
                           rfl
                             State/Province
                         11970 n1980 • 1990 • 2000
Figure 3. Total per capita energy consumption 1970-2000.1 MWh =
1000 kWh. Other energy sources include geothermal, wind, photo-
voltaic and solar energy. The Ontario data do not include the elec-
tricity generation sector due to an absence of data for this sector
until 1978.
Source: Energy Information Administration (2000) and Natural
Resources Canada (2000)
                                                                     E
                                                                     •£  200
                                                                                            State/Province
                   • Residential   n Industrial     • Electricity
                   • Commercial  • Transportation   Generation
Figure 4. Secondary energy consumption within the Great Lakes
basin by sector. Note: all data are from 2000, although 2002 data
from Ontario are discussed in the report.
Source: Energy Information Administration (2000) and Natural
Resources Canada (2000)
    Canadian side of the basin in 2002 was 930,400,000 Megawatts-
    hours (MWh) (Table 1). Secondary energy is the energy used by
    the final consumer. It includes energy used to heat and cool
    homes and workplaces, and to operate appliances, vehicles and
State/Province
Ontario (2002 data)
U.S. Basin Total (2000 data)
Illinois (IL)
Indiana (IN)
Michigan (Ml)
Minnesota (MN)
New York (NY)
Ohio (OH)
Pennsylvania (PA)
Wisconsin (Wl)
Total energy consumption by
State/Province within the Great
Lakes basin (MWh)
930,400,000
3,364,000,000
669,400,000
304,900,000
998,500,000
36,600,000
309,600,000
614,000,000
43,700,000
387,300,000
Population within the
Great Lakes basin*
9,912,707
31,912,867
6,025,752
1,845,344
9,955,795
334,444
4,506,223
5,325,696
389,210
3,530,403
* The U.S. side of the basin is defined as the portions of the 8 Great Lakes states within the basin boundary
(which totals 214 counties either completely or partially within the basin boundary).
Table 1 : Energy consumption and population within the Great Lakes basin, by state
for the year 2000 (U.S.) and 2002 (Ontario). The U.S. basin population was calcu-
lated from population estimates by counties (either completely or partially within
the basin) from the 2000 U.S. Census (U.S. Census Bureau 2000). Ontario basin
populations were determined using sub-basin populations provided by Statistics
Canada.
Source: U.S. Energy Information Administration and Natural Resources Canada
factories. It does not include intermediate uses of energy for
transporting energy to market or transforming one energy form
to another, this is primary energy. Accounting for 33% of the
total secondary energy consumed in the Canadian basin, electric-
ity generation was the largest end user of all the sectors. The
other four sectors account for the remaining energy consumption
              as follows: industrial,  22%; transportation 20%;
              residential, 15%; and commercial, 12% (Table 2).
              Note that due to rounding, these figures do not
              add up to 100. There was a 0.5% increase in total
              energy consumption by all sectors in Ontario
              between  2000 and 2002.
                                                                                Total secondary energy consumption by the five
                                                                                sectors on the U.S. side of the basin in 2000 was
                                                                                3,364,000,000 MWh (Table 1). As in the
                                                                                Canadian basin, electricity generation was the
                                                                                largest consuming sector in the U.S. basin, using
                                                                                28% of the total secondary energy in the U.S.
                                                                                side of basin. The U.S. industrial sector con-
                                                                                sumed only slightly less energy, 27% of the total.
                                                                                The remaining three U.S. sectors account for
                                                                                44% of the total, as follows: transportation, 21%;
                                                                                residential, 14%; and commercial, 9% (Table 2).
                                                                                Note that due to rounding, these percentages do
                                                                                not add up to 100. Figure 4 shows the total ener-
                                                                                gy consumption by sector for both the Ontario
                                                                                and U.S. sides of the Great Lakes basin in 2000.
   224

-------
Sector
Residential
Commercial
Industrial
Transportation
Electricity Generation
U.S. Basin Total Energy
Consumption - 2000*
478,200,000
314,300,000
903,900,000
714,000,000
953,600,000
Canadian Basin Total Energy
Consumption - 2002
127,410,000
107,800,000
206,410,000
184,950,000
303,830,000
* Note: 2000 is the most recent data available on a consistent basis for the U.S. More recent data is
available for some energy sources from the EIA, but survey and data compilation methods may
vary.
Table 2: Total Secondary Energy Consumption in the Great Lakes basin, in
Megawatts -hours (MWh).
Source: U.S. Energy Information Administration and Natural Resources
Canada
                            State/Province
                   n Electricity  • Petroleum
Figure 5. Commercial sector energy consumption by source, 2000.
Wood and coal were minor sources in this sector.
Source: Energy Information Administration (2000) and Natural
Resources Canada (2000)
   The commercial sector includes all activities related to trade.
   finance, real estate services, public administration, education.
   commercial services (including tourism), government and insti-
   tutional living and is the smallest energy consumer of all the sec-
   tors in both Canada and the U.S. (Table 2). Of the total second-
   ary energy use by this sector in the Ontario basin, 57% of the
   energy consumed was supplied by fossil fuel (natural gas, 50%;
   and petroleum, 7%) and 43% was supplied by electricity. In
   Ontario, this sector had the largest increase in total energy con-
   sumption, 4.4%, between 2000 and 2002. By source, on the U.S.
                                                                       side of the basin, 61% was supplied by fossil fuel (natural
                                                                       gas, 53%; and petroleum, 8%) and 39% was supplied by
                                                                       electricity. On both sides of the basin, the commercial
                                                                       sector had the highest proportion of electricity use of any
                                                                       sector. Figure 5 shows energy consumption by source for
                                                                       the commercial sector for the Canadian and the U.S.
                                                                       basins in 2000.

                                                                       The residential sector includes four major types of
                                                                       dwellings: single detached homes, single attached homes.
                                                                       apartments and mobile homes, and excludes all institu-
                                                                       tional living facilities. Fossil fuels (natural gas, petroleum.
                                                                       and coal)  are the dominant energy source for residential
                                                                       energy requirements in the Great Lakes basin. Of the total
                                                                 secondary energy use by the residential sector in the Ontario
                                                                 basin in 2002 (Table 2), the source for 67%  of the energy con-
                                                                 sumed was supplied by fossil fuel (natural gas, 61%; and petro-
                                                                 leum, 6%), 30% by electricity and 3% by wood (Figure 6).
                                                                                            State/Province
                                                                                    • Wood D Electricity • Petroleum D Natural Gas
 Figure 6. Residential sector energy consumption by source,
 2000. Coal, geothermal, and solar energy were minor sources in
 this sector.
 Source: Energy Information Administration (2000) and Natural
 Resources Canada (2000)
There was a 0.3% increase in total energy consumption by the
Ontario residential sector between 2000 and 2002. On the U.S.
side of the basin, fossil fuels are the leading source of energy
accounting for 75% of the total residential sector consumption.
Natural gas and petroleum are both consumed by this sector, but
it is important to note that this  sector has the highest natural gas
consumption of all five sectors. The remaining energy sources
were electricity, 22% and wood, 3% (Figure 6).
                                                                                                                         225

-------
                                              OF   THE   GREAT
                             2007


millions)
:D
Consumption (MWh in
D O O
N


n n
M
p!__|iQiy
' X X / / ' / / X

-------
                            State/Province
         D Hydroelectric Power  n Nuclear Power • Coal  • Petroleum  • Natural Gas
Figure 9. Electricity generation sector energy consumption by
source, 2000. Wood and wood waste were very minor energy
sources in this sector.
Source: Energy Information Administration (2000) and Natural
Resources Canada (2000)
petroleum), and 7% was supplied by hydroelectric energy. There
was an increase in total energy use of 1.9% between 2000 and
2002 in Ontario. It is important to note that the Great Lakes
basin contains the majority of Canada's nuclear capacity. Of the
total secondary energy use by this sector in the U.S. basin (Table
2), 70% was supplied by the following types of fossil fuel: coal
(66%), natural gas (2%), and petroleum (2%). The other two
major sources, nuclear and hydroelectric energy, provided 27%
and 3% respectively. This sector consumed 75% of the coal used
in the entire U.S.  basin. Figure 9 shows energy consumption by
source for the electricity generation sector for the Canadian and
U.S. sides of the basin in 2000.

The overall trends in energy consumption by sector were quite
similar on both sides of the basin. Ranked from highest to lowest
energy consumption, the pattern  for the sectors was the same for
the U.S. and Canadian basins (Table 2). Analyses of the sources
of energy within each sector and trends in resources consump-
tion also indicate  very similar trends.

Pressures
In 2001, Canada was ranked as the fifth largest energy  producer
and the eighth largest energy consuming nation in the world.
  Comparatively, the United States is ranked as "the world's
  largest energy producer, consumer, and net importer" (U.S.
  EIA 2004). The factors responsible  for the high energy con-
  sumption rates in Canada and the U.S. can also be attributed
  to the Great Lakes basin. These include a high standard of liv-
  ing, a cold climate, long travel  distances, and a large industrial
  sector. The combustion of fossil fuels, the dominant source of
  energy for most sectors in the basin, releases greenhouse gases
  such as carbon dioxide and nitrous oxide into the air contribut-
  ing to smog, climate change, and acid rain.

  Canada's Energy Outlook 1996-2020
  fhttp://nrnl.nrcan.gc.ca:80/es/ceo/toc-96E.html) notes that "a
  significant amount of excess generating capacity exists in all
  regions of Canada" because demand has not reached the level
  predicted when new power plants were built in the 1970s and
  1980s. Demand is projected to  grow at an average annual rate
  of 1.3 percent in Ontario and 1.0 percent in Canada overall
  between 1995 and 2020. From  2010-2020, Ontario will add
  3,650 megawatts of new gas-fired and 3,300 megawatts of
  clean coal-fired capacity.  Several hydroelectric plants will be
  redeveloped.  Renewable resources are projected to quadruple
  between 1995 and 2020, but will contribute only 3  percent of
  total power generation.
  The pressures the U.S. currently faces will continue into the
  future, as the U.S. works to renew its aging energy infrastruc-
  ture and develop renewable energy sources. Over the next two
  decades, U.S. oil consumption is estimated to grow by 33%.
  and natural gas consumption will increase by more than 50%.
Electricity demand is forecast to increase by 45% nationwide
(National Energy Policy 2001). Natural gas demand currently
outstrips domestic production in the U.S. with imports (largely
from Canada) filling the gap. 40% of the total U.S. nuclear out-
put is generated within five states, including three within the
Great Lakes basin (Illinois, Pennsylvania, and New York) (U.S.
EIA 2004). Innovation and creative problem solving will be
needed to work towards balancing economic growth and energy
consumption in the Great Lakes basin in the future.
Management Implications
Natural Resources Canada, Office of Energy Efficiency has
implemented several programs that focus on energy efficiency
and conservation within the residential, commercial, industrial.
and transportation sectors. Many of these programs work to pro-
vide consumers and businesses with useful and practical infor-
mation regarding  energy saving methods for buildings, automo-
biles, and homes.  The U.S. Department of Energy Office of
Energy Efficiency and Renewable Energy recently launched an
educational website (http://www. eere.energy, gov/consumerinfo/l.
which provides homes and businesses with ways to improve effi-
ciency, tap into renewable and green energy supplies, and reduce

                                                       227

-------
                                                                                         2007
energy costs. In July 2004, Illinois, Minnesota, Pennsylvania,
and Wisconsin were awarded $46.99 million to weatherize low-
income homes, which is expected to save energy and cost
(EERE 2004). The U.S. Environmental Protection Agency
Energy Star program, a government/industry partnership initiat-
ed in 1992, also promotes energy efficiency through product cer-
tification. In 2002, Americans saved more than $7 billion in
energy costs through Energy Star, while consuming less power
and preventing greenhouse gas emissions (USEPA 2003).

In addition to these programs, the Climate Change Plan for
Canada challenges all Canadians to reduce their greenhouse gas
emissions by one tonne, approximately 20% of the per capita
production on average each year. The One-Tonne Challenge
offers a number of ways to reduce the greenhouse gas emissions
that contribute to climate change and in doing so will also
reduce total energy consumption.

Renewable energy sources such as solar and wind power  are
available in Canada, but constitute only a fraction of the total
energy consumed. Research continues to develop these as alter-
nate sources of energy, as well as developing more efficient
ways of burning energy. In the United States, according to the
U.S. Energy Information Administration,  6% of the  total 2002
energy consumption came from renewable energy sources (bio-
mass, 47%; hydroelectric, 45%; geothermal, 5%; wind, 2%; and
solar, 1%).  The U.S. has invested almost a billion dollars, over
three years, for renewable energy technologies (Garman 2004).
Wind energy, cited as one of the fastest growing renewable
sources worldwide, is a promising source for the Great Lakes
region. The U.S. Department of Energy, its laboratories, and
state programs are working to advance research and develop-
ment of renewable energy technologies.

Acknowledgments
Authors:  Susan Arndt, Environment Canada, Ontario Region,
Burlington, ON;
Christine McConaghy, Oak Ridge Institute for Science and
Education,  on appointment to U.S. Environmental Protection
Agency, Great Lakes National Program Office, Chicago, IL; and
Leena Gawri, Oak Ridge Institute for Science and Education, on
appointment to U.S. Environmental Protection Agency, Great
Lakes National Program Office, Chicago, IL.

Sources
Canada and U.S. Country Analysis Briefs. 2005. Energy
Information Administration.
http://www. eia.doe. gov/emeu/cabs/canada.html, last accessed
October 4,  2005.

Energy Efficiency and Renewable Energy (EERE) Network
News.  2004. DOE Awards $94.8 Million to  Weatherize Homes in

228
20 States. U.S. Department of Energy.
http://www.eere. energy. gov/news/news_detail.cfm/news_id=743
S, last accessed October 4, 2005.

Environment Canada. 2003. Environmental Signals, Canada's
National Environmental Indicator Series 2003, Energy
Consumption, pp 56-59. http://www.ec.gc.ca/soer-ree.

Garman, D.K. 2004. Administration s views on the role that
renewable energy technologies can play in sustainable electricity
generation. United States Senate, Testimony before the
Committee on Energy and Natural Resources.
http://www.eere.energy. gov/office_eere/congressional_test_0427
04.html.

National Energy Policy Development Group (NEPDG). 2001.
Report of the National Energy Policy Development Group.
http://energy.gov/engine/content.do?BT_CODE=AD_AP.

Natural Resources Canada. 2002. Energy Efficiency Trends in
Canada 1990-2000. http://oee.nrcan.gc.ca/neud/dpa/home.cfm.

Natural Resources Canada. Comprehensive Energy Use
Database.
http ://oee.mean. gc.ca/neud/dpa/comprehensive_tables/.

Statistics Canada. 2000. Human Activity and the Environment
2000. [CDRom].

U.S. Census Bureau and Texas State Data Center. 2000. U.S.
2000 decennial census data. Department of Rural Sociology,
Texas A&M University, http ://www.census. gov/dmd/www/resap-
port/states/indiana.pdf and
http://www.txsdc.tamu.edu/txdata/apport/hist_a.php.

U.S. Energy Information Administration (EIA). 2004. State ener-
gy data 2000 consumption tables, http://www.eia.doe.gov.

U.S. Environmental Protection Agency (USEPA). 2003. ENER-
GY STAR - The power to protect the environment through energy
efficiency.
http ://www. energystar. gov/ia/partners/downloads/energy_star_re
port aug 2003.pdf.
Authors' Commentary
Ontario data are available through Natural Resources Canada,
Office of Energy Efficiency. Databases include the total energy
consumption for the residential, commercial, industrial, trans-
portation, agriculture and electricity generation sectors by energy
source and end use. Population numbers for the Great Lakes
basin, provided by Statistics Canada, were used to calculate the

-------
energy consumption numbers within the Ontario side of the        Last Updated
basin. This approach for the residential sector should provide a     gtate Oft^e Qreat Lakes 2005
reasonable measure of household consumption. For the commer-
cial, transportation and especially industrial sectors, it may be a
variable estimation of the total consumption in the basin. The
data are provided on nation-wide, or province-wide basis.
Therefore it provides a great challenge to disaggregate it by any
other methods to provide a more precise representation of the
Great Lakes basin total energy consumption.

Energy consumption, price, and expenditure data are available
for the United States (1960-2000) through the Energy
Information Administration (EIA). The EIAis updating the State
Energy Data 2000 series to 2001 by August 2004. There may be
minor discrepancies in how the sectors were defined in the U.S.
and Canada, which may need further investigation (such as
tourism in the U.S. commercial sector, and upstream oil and  gas
in the U.S. industrial sector). Actual differences in consumption
rates may be difficult to distinguish from minor differences
between the U.S. and Canada in how data were collected and
aggregated. Hydroelectric energy was not included in the indus-
trial sector analysis, but might be considered in future analyses.
In New York  State, almost as much energy came from hydro-
electric energy as from wood. Wisconsin and Pennsylvania also
had small amounts of hydropower consumption.

In the U.S. the current analysis of the total basin consumption is
based on statewide per capita energy consumption, multiplied by
the basin population. The ideal estimate of this indicator would
be to calculate the per capita consumption within the basin, and
would require energy consumption data at the county level or by
local utility reporting areas. Such data may be quite difficult to
obtain, especially when electricity consumption per person is
reported by utility service area. The statewide per capita con-
sumption may be different than the actual per capita consump-
tion within the basin, especially for the states with only small
areas within the basin (Minnesota and Pennsylvania). The pro-
portion of urban to rural/agricultural land in the basin is likely to
influence per capita consumption within the basin. Census data
are available at the county and even the block level, and may in
the future be combined with the U.S. basin boundary using GIS
to refine the basin population estimate.

Additionally,  the per capita consumption data for the U.S.  in
Figures 1, 2, and 3 are based on slightly different energy con-
sumption totals than the data in Tables 1 and 2. The next update
of this indicator should examine whether it is worthwhile to
include the minor sources in the sector analysis on both sides of
the basin or to exclude them from the per capita figures.
                                                                                                                      229

-------
Solid Waste Disposal
Indicator #7060

Overall Assessment
           Status:  Trend Not Assessed
           Trend:  Undetermined
   Primary Factors
      Determining
  Status and Trend
This year the indicator report focuses only on disposal data in the U.S.
instead of generation or recycling data.  Disposal data was the most
consistently collected by the counties/states in the U.S. Generation and
recycling data were available for Ontario, Canada. Over time, a
change in disposal tonnages can be used as an indicator for solid waste
in the Great Lakes, however more consistent and comparable data
would improve this indicator.
Lake-by-Lake Assessment
Due to insufficient data, a lake-by-lake assessment is not available for this indicator.

Purpose
•To assess the amount of solid waste disposed in the Great Lakes basin; and
•To infer inefficiencies in human economic activity (i.e. wasted resources) and the potential
adverse impacts to human and ecosystem health.

Ecosystem Objective
Solid waste provides a measure of the inefficiency of human land based activities and the degree
to which resources are wasted. In order to promote sustainable development, the amount of solid
waste disposed of in the basin needs to be assessed and ultimately reduced. Because a portion of
the waste disposed of in the basin is generated outside of basin counties, efforts to reduce waste
generation or increase recycling need to occur regionally.  Reducing volumes of solid waste via
source reduction or recycling is indicative of a more efficient industrial ecology and a more
conserving society. This indicator supports Annex 12 of the Great Lakes Water Quality
Agreement (United States and Canada 1987).

State of the Ecosystem
Canada and the United States are working towards improvements in waste management by
developing strategies to prevent waste  generation and reuse and recycle more of the generated
waste.  The data available to support this indicator are limited in some areas of the basin and not
consistent from area to area. For example, while most of the U.S. states in the basin track amount
of waste disposed in a landfill or incinerator located in a county, they may define the wastes
differently. Some track all non-hazardous waste disposed and some only track municipal solid
waste.  Because the wastes disposed of in each county in the basin were not necessarily generated
by the  county residents, per capita estimates are not meaningful. Not all of the U.S. counties
provide generation and recycling rates information. Canada provides estimates of waste
generation rate for each of its Provinces for residential, industrial/commercial, and construction
and demolition sources. The summary statistics report also provided disposal data, however the
disposal data included wastes that were disposed of outside the Province, some of which is
captured in the U.S. county disposal data within the basin. For this reason, generation and
                         Draft for Discussion at SOLEC 2006

-------
                                              '. V > & ftC     i	i     Mlllllllllllllllllllllllllllllllll^lllllllllllllllllnlnnnnnnnnnnnnnnnnnnnnnnnnnnnnnnnilMJMl^Mill^
diversion estimates were used only for Ontario, Canada; disposal data were used for the U.S.
counties. Types of waste included in the disposal data are identified below.

Statistics for the generation of waste in Ontario were gathered from the Annual Statistics 2005
report. More than 11 million tonnes of wastes were generated in Ontario in 2000 and slightly
more than 12 million tonnes were generated in 2002.  These figures include residential wastes,
commercial/industrial wastes, and construction and demolition wastes. Diversion information
was also provided in the report and can be seen in Figure 1.  In 2000, 20.8% of the residential
waste generated was diverted to recycling  and in 2002 that figure increased to 21.6%.  The
industrial/commercial recycling rate was 22.7% in 2000 and 20.2% in 2002.  Finally, the C&D
recycling rate was 11.6% in 2000 and 12.5% in 2002.  Ontario has a goal to divert 60% of its
waste by 2008.

Minnesota Great Lakes basin counties provided data on the amounts of waste disposed of in the
county as well as an estimate of the amount of waste buried by residents (on their own property).
Data are provided in Figure 2.  In 2003, 124,931  tons of waste were disposed of or buried in the 7
basin counties in MN. In 2004, there was  a 5% increase to 132,128 tons disposed or buried.
Each county showed an increase in waste disposed. These figures  only include municipal solid
waste (not construction and demolition debris or other industrial wastes).

The Indiana Department of Environmental Management's data regarding amounts disposed of at
permitted facilities were used to determine the total amount disposed in each Indiana Great Lakes
Basin county. The data are provided in Figure 3. The disposal in 2004 was approximately 9%
greater than in 2003.  The 15 basin counties disposed of 2,468,913 tons of waste in 2004 and
2,224,581 tons in 2005. About 15% was generated outside of the counties in 2004. The data
include municipal solid waste, construction and demolition wastes, and some industrial byproduct
waste.

The Illinois Environmental Protection Agency, Bureau of Land,  reported the amounts disposed of
in permitted landfills in the 2 Great Lakes  basin counties.  Data were compiled for 2004 and 2003
and are shown in Figure 4. There was less than a 2% change in total materials.  In 2004
1,814,529 tons were disposed and in 2003  slightly less waste (1,784,452 tons) was disposed.
The data include municipal solid waste, construction and demolition waste, and some industrial
waste.

The Michigan Department of Environmental Quality reports on total waste disposed in Michigan
landfills in cubic yards. General conversion factors (to translate cubic yards to tons) could not be
used because the waste totals include a variety of waste sources (municipal solid waste,
construction and demolition debris, and some industrial byproducts). Data for the 83 Great Lakes
basin counties were compiled and are presented in Figure 5.  There was less than a 1% difference
between the total cubic yards disposed in 2004 and 2005 in these counties. The total for 2005
was slightly smaller.  For both years, approximately 64 million cubic yards were disposed of in
the 83 counties  in the Great Lakes Basin.

The New York Department of Environmental Conservation provided municipal solid waste
disposal data for facilities located in the 32 Great Lakes basin counties for the years 2004 and
                         Draft for Discussion at SOLEC 2006

-------
2002. The data are presented in Figure 6. There was an approximate 5% increase in waste
disposed.  The total waste disposed was 7,853,087 tons in 2004 and 7,333,685 tons in 2002.  This
data includes municipal solid waste only. More than 65% of the states waste is managed in the
basin counties.

The Pennsylvania Department of Environmental Protection provided disposal data for the three
Great Lakes basin counties. Municipal solid waste and construction and demolition debris are
combined in these annual totals which are presented in Figure 7. For 2004, 282,004 tons were
disposed in the three basin counties. There was a 25% decrease in waste disposed in the counties
in 2005 to 209,229 tons.

The Wisconsin Department of Natural Resources collects data on the amount disposed of in each
facility located in the Great Lakes basin counties.  Data were compiled for the 26 basin counties
and are presented in Figure 8.  In 2005, 7,663,187 tons of wastes were disposed,  within 1% of the
total disposed in 2004.  Totals include a wide variety of wastes such as municipal solid waste,
sludges, and foundry  sand.

The Ohio Environmental Protection Agency collects data for waste disposed of in landfills and
incinerators. The data for the 36 Great Lakes basin counties was compiled for 2003 and 2004 and
are presented in Figure 9. There was an approximate 5% increase in waste disposed. More than
60% of these waste disposed in the counties came from outside the  counties. The data includes
municipal solid waste, some industrial wastes, and tires. Construction and demolition debris is
not included. In 2004, the 36 basin counties disposed of 8,791,802 tons and in 2003 8,334,865
tons were disposed.

Pressures
The generation and management of solid waste raise important environmental, economic and
social issues for North Americans. Waste disposal costs billions of dollars and the entire waste
management process  uses energy and contributes to land, water, and air pollution. The U.S. EPA
has developed tools and information linking waste management practices to  climate change
impacts. Waste prevention and recycling reduce greenhouse gases associated with these activities
by reducing methane  emissions, saving energy, and increasing forest carbon sequestration. Waste
prevention and recycling save energy when compared to disposal of materials.

The state of the economy has a strong impact on consumption and waste generation. Municipal
solid waste generation in the U.S. continued to increase through the 1990s and has remained
steady since 2000 (USEPA 2003). Generation of other wastes, such as construction and
demolition debris and industrial wastes is also strongly linked to the economy. The U.S. EPA is
developing a methodology to better estimate the generation, disposal, and recycling of
construction and demolition debris in the U.S.

Because waste disposed of in the Great Lakes Basin may be generated outside of the Basin or
moved around within the Basin, efforts to reduce waste generation and increase recycling need to
focus on a broad area, not just the Basin. Continued collaboration of state, local,  and federal
efforts is important for long term success.

Management Implications
                         Draft for Discussion at SOLEC 2006

-------

The U.S. EPA supports a bi-annual study that characterizes the municipal solid waste stream and
estimates the national recycling rate.  The latest study (2003) estimates a 30.6% national
recycling rate. The U.S. EPA has established a goal of reaching a 35% recycling rate by 2008.
The 2003 study indicated that paper, yard and food waste, and packaging represent large portions
of the waste stream. The U.S. EPA's is concentrating its efforts on these materials; working with
stakeholders to determine activities that may support increased recovery of those materials.   The
federal government is also working to promote  strategies that support recycling programs in
general, including Pay-As-You-Throw (generators pay per unit of waste rather than a flat fee);
innovative contracting mechanisms such as resource management (includes incentives for
increased recycling), and supporting demonstration projects and research on various end markets
and collection strategies for waste materials.  The States are also working to increase recycling
rates and provide support for local jurisdictions. Each state with counties in the Great Lakes
basin provides financial and technical support for local recycling programs.  Many provide
significant market development support as well.

Canada and the U.S. both support integrated  solutions to the waste issue and look for innovative
approaches that involve the public and private sectors. Extended Producer Responsibility (EPR),
also known as Product Stewardship is one  approach that involves manufacturers of products.
EPR efforts have focused on many products including electronics, carpets, paints, thermostats,
etc.

Ontario's Waste Diversion Act was passed in 2002 and created Waste Diversion Ontario, a
permanent, non-government corporation.  The Act gave WDO the mandate to develop,
implement and operate waste diversion programs-to reduce, reuse or recycle waste.

The City of Toronto has set ambitious waste  diversion goals and reported a 40% diversion rate in
2005. The development of a green bin system (allowing residents to separate out the organic
fraction of the waste stream from traditional recyclables) is credited for the high diversion rate
achieved.

Improved and consistent data collection would help to better inform decisionmakers regarding
effectiveness of programs as well as determining where to target efforts.

Comments from the author(s)

During the process of collecting data for this indicator, it was found that U.S. states and Ontario
compile and report on solid waste information in different formats. Future work to organize a
standardized method of collecting, reporting  and accessing data for both the  Canadian and U.S.
portions of the Great Lakes basin will aid in the future reporting of this indicator and in the
interpretation of the data and trends.  More consistent data may also support strategic planning.

Acknowledgments
Authors: Susan Mooney, Julie Gevrenov, and Christopher Newman U.S. Environmental
Protection Agency, Waste, Pesticides, and Toxics Division, Region 5, Chicago, IL.

Data Sources
                         Draft for Discussion at SOLEC 2006

-------
The United States data regarding national recycling rate and municipal solid waste characteristics
was collected from Municipal solid waste in the United States: 2003 facts and figures; available
on the U.S. EPA's web site at http://www.epa.gov/epaoswer/non-hw/muncpl/msw99.htm.

Solid waste data for Ontario was collected from Human Activity and the Environment. Annual
Statistics 2005, Featured Article: Solid Waste in Canada, Catalogue number 16-201XIE, Statistics
Canada.

Illinois waste disposal data for the 2 basin counties was compiled from the Illinois Environmental
Protection Agency, Bureau of Land's 2004 Landfill Capacity report found on their web site at:
http://www.epa.state.il.us/land/landfill-capacity/2004/index.html. The 2 Great Lakes Basin
counties are located in Illinois EPA's Region 2.

Indiana waste disposal data for the basin counties were compiled from the Indiana Department of
Environmental Management's permitted solid waste facility reports found at
http://www.in.gov/idem/programs/land/sw/index.html.

Michigan waste disposal data for the basin counties were compiled from the Michigan
Department of Environmental Quality's Annual Report on Solid Waste Landfills. Data from the
2005  and 2004 studies were compiled.  The author accessed the data via the Border Center's
WasteWatcher web site (http://www.bordercenter.org/wastewatcher/mi-waste.cfm ) to more
easily search for the appropriate county - level data.

Minnesota municipal solid waste disposal  data for the basin counties was compiled from the 2004
and 2003 SCORE data available on the Minnesota Pollution Control Agency's web site at:
http://www.moea.state.mn.us/lc/score04.cfm   The SCORE report is a report to the Legislature,
the main components of this report are to identify and target source reduction, recycling, waste
management and waste generation collected from all 87 counties in Minnesota.

New York municipal solid waste disposal data for the basin counties were compiled from New
York State Department of Environmental Conservation's capacity data for landfills and waste to
energy facilities available on their website at:
http://www.dec.state.ny.us/website/dshm/sldwaste/newsw2.htm.

Ohio waste disposal data for the basin counties were compiled from Ohio Environmental
Protection Agency's  2003 and 2004 facility data reports which are available on their web site at
http://www.epa.state.oh.us/dsiwm/pages/general.html.

Pennsylvania waste disposal data  for the basin counties were compiled from the Pennsylvania
Department of Environmental Protection, Bureau of Land Recycling and Waste Management's
disposal data located on their web site at:
http://www.depweb.state.pa.us/landrecwaste/cwp/view.asp?a=1238&Q=464453&landrecwasteNa
Wisconsin municipal solid waste disposal data for the basin counties were compiled from the
Wisconsin Department of Natural Resources, Bureau of Waste Management's Landfill Tonnage
Report found on their website at:, http://www.dnr.state.wi.us.
                         Draft for Discussion at SOLEC 2006

-------
                                                tMi*! ' '*'  '••	—'===	JifmmMlmmmmmmmmimgmm
United States and Canada. 1987. Great Lakes Water Quality Agreement of 1978, as amended by
Protocol signed November 18, 1987. Ottawa and Washington.

List of Figures
Figure 1.  Ontario Waste Diversion Rates.
Source: Statistics Canada, Catalogue number 16-201XIE, Human Activity and the Environment.
Annual Statistics 2005, Featured Article: Solid Waste in Canada.

Figure 2.  Minnesota Basin County Disposal.
Source: Minnesota Pollution Control Agency, Score Report, 2003 and 2004.

Figure 3.  Indiana Basin County Disposal.
Source: Indiana Department of Environmental Management, Permitted Solid Waste Facility
Report.

Figure 4.  Illinois  Basin County Disposal.
Source: Illinois Environmental Protection Agency, 2004 Landfill Capacity Report.

Figure 5.  Michigan Basin County Disposal.
Source: Michigan Department of Environmental Quality, 2005 and 2004 Annual Report on Solid
Waste Landfills.

Figure 6.  New York Basin County Disposal.
Source: New York State Department of Conservation Capacity data for Landfills and Waste to
Energy Facilities.

Figure 7.  Pennsylvania Basin County Disposal.
Source: Pennsylvania Department of Environmental Protection Landfill Disposal Data.

Figure 8 Wisconsin Basin County Disposal
Source: Wisconsin Department of Natural Resources, Landfill Tonnage Report.

Figure 9.  Ohio  Basin County Disposal.
Source: Ohio Environmental Protection Agency, 2003 and 2004 Facility Data Reports.

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
       State of the Great Lakes 2007 - Draft
           Figure 1: Ontario Waste Diversion Rates
         25

         20 -H
      4M
      §  15 -H
      9
      Q-  10 4-

          5 I-

          0
             year
             2002
year
2000
                 Residential
year
2002
year
2000
           industrial/commercial
year
2002
year
2000
                     C&D
Figure 1. Ontario Waste Diversion Rates.
Source: Statistics Canada, Catalogue number 16-201XIE, Human Activity and the Environment.
Annual Statistics 2005, Featured Article: Solid Waste in Canada.
           Figure 2 : Minnesota  Basin County
                           Disposal
      150000
   M 100000
   e
   I-  50000
            0
                          Basin County
                                 D 2004 totals
                                            tf3
Figure 2. Minnesota Basin County Disposal.
Source: Minnesota Pollution Control Agency, Score Report, 2003 and 2004.
                    Draft for Discussion at SOLEC 2006

-------
Figure 3: Indiana Basin County Disposal
'inn nnn
ftnn nnn -
"O yrin r\r\f\ .
O (UU.UUU
o fiOQ 000 -
O. pnn npn .
Q~ 4nn onn -
« onn nnn
2 -3UU,UUU -
° ?nn nnn -
i nn nnn
IUU,UUU
n











^
	 D2003
D23D4



ffi C
1 5
County
Figure 3. Indiana Basin County Disposal.
Source: Indiana Department of Environmental Management, Permitted Solid Waste Facility
Report.
                       Draft for Discussion at SOLEC 2006

-------
       State of the Great Lakes 2007 - Draft
                Figure 4: Illinois Basin County
                           Disposal
        2000000
         1500000
      g  1000000
         500000
                   Cook    La
-------

- z: <- n -> ijfVj .
16 000 00" -
- • Q o 0 00 C -
* '2 000 003 •
93
o 3 000 00n •

4 000 00" -
t f\f\n An-

Figure 5: Michigan Basin County Disposal






D2005cata
B2CO- cats








rvrfL^m


ftl ™rn ralfl™


n fin.
n. rfl m m



B-fcj
Cj O -L*^" **" ^* v^
Basin County

(l 1 ill
I_IL ^fn^^llihrJI







—

x
-------
        State of the Great Lakes 2007 - Draft
Figure 6: NY Basin County Disposal
Q nnn nnn -
7 nnn nnn
^ c nnn nnn
to
o
A nnn nnn -
c
o
i— q nnn nnn
o nnn nnn .
1 nnn nnn -








j (
fc tfl Jl ML fflEfl




• 2002
D2004

O* 
-------
        Figure 7: Pennsylvania Basin
                County Disposal
      300000
      250000
   w  200000
   §  150000
   *~"  100000
       50000
          0
Figure 7. Pennsylvania Basin County Disposal.
Source: Pennsylvania Department of Environmental Protection Landfill Disposal Data.
12
Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
FlgureS: Wsconsin Basin County
9 nnn mri -
1 ^nn mn
gj I JUU.UJU
|Bj
i nnn nrifi .
cnn nnn _
v'UU.UUU •
0.
Disposal




l 1

1 1 J
. Um_ tx».





nil

• 2005 tons
• 2004 tons

courtly
Figure 8 Wisconsin Basin County Disposal
Source: Wisconsin Department of Natural Resources, Landfill Tonnage Report.
                     Draft for Discussion at SOLEC 2006
13

-------
Figure 9: Ohio Basin County Disposal



TJ
d>
O
a














fl n Fl™

1 1 J



„ ~~[l™


02003
• 2004

County
Figure 9. Ohio Basin County Disposal.
Source: Ohio Environmental Protection Agency, 2003 and 2004 Facility Data Reports.
14
Draft for Discussion at SOLEC 2006

-------
Nutrient Management Plans
Indicator #7061

Assessment: Not Assessed

Purpose
  To determine the number of Nutrient Management Plans; and
  To infer environmentally friendly practices that help to pre-
vent ground and surface water contamination.

Ecosystem Objective
This indicator supports Annexes 2, 3, 11,  12 and 13 of the Great
Lakes Water Quality Agreement. The objective is sound use and
management of soil, water, air, plants and animal resources to
prevent degradation of the environment. Nutrient Management
Planning guides the amount, form, placement and timing of
applications of nutrients for uptake by crops as part of an envi-
ronmental farm plan.

State of the Ecosystem
Background
Given the key role of agriculture in the Great Lakes ecosystem,
it is important to track changes in agricultural practices that can
lead to protection of water quality, the sustainable future of agri-
culture and rural development, and better ecological integrity in
the basin. The indicator identifies the degree  to which agricul-
ture is becoming more sustainable and has less potential to
adversely impact the Great Lakes ecosystem.
As more  farmers embrace environmental plan-
ning over time, agriculture will become more
sustainable through nonpolluting, energy effi-
cient technology and best management prac-
tices for efficient and high quality food pro-
duction.
Status of Nutrient Management Plans
The Ontario Environmental Farm Plans (EFP)
identify the need for best nutrient management
practices. Over the past 5 years farmers,
municipalities and governments and their
agencies have made significant progress.
Ontario Nutrient Management Planning soft-
ware (NMAN) is available to farmers and con-
sultants wishing to develop or assist with the
development of nutrient management plans.

In 2002 Ontario passed the Nutrient
Management Act (NM Act) to establish
province-wide  standards to ensure that all
land-applied materials will be managed in a
sustainable manner resulting in environmental
and water quality protection. The NM Act
                 requires standardization, reporting and updating of nutrient man-
                 agement plans through a nutrient management plan registry. To
                 promote a greater degree of consistency in by-law development,
                 Ontario developed a model nutrient management by-law for
                 municipalities. Prior to the NM Act, municipalities enforced
                 each nutrient management by-law by inspections performed by
                 employees of the municipality or others under authority of the
                 municipality.

                 In the United States, the two types of plans dealing with agricul-
                 ture nutrient management are the Comprehensive Nutrient
                 Management Plans (CNMPs) and the proposed Permit Nutrient
                 Plans (PNP) under the U.S. Environmental Protection Agency's
                 (USEPA) National Pollution Discharge Elimination System
                 (NPDES) permit requirements. Individual States also have addi-
                 tional nutrient management programs. An agreement between
                 USEPA and U.S. Department of Agriculture (USDA) under the
                 Clean Water Action plan called for a Unified National Strategy
                 for Animal Feeding Operations. Under this strategy, USDA-
                 Natural Resources Conservation Service has leadership for the
                 development of technical standards  for CNMPs. Funds from the
                 Environmental Quality Incentives Program can be used to devel-
                 op CNMPs.

                 The total number of nutrient management plans developed annu-
                 ally for the U.S. portion of the basin is shown in Figure  1. This
                 includes nutrient management plans for both livestock and non-
                 livestock producing farms. The CNMPs are tracked on an annual
                                                 Nutrient Management Applied
                                                    1 0- 1,500 Acres
                                                    J 1,500-5,000 Acres
                                                    . 5,000-10,000 Acres
                                                   I 10,000-25,000 Acres
Figure 1. Annual U.S. Nutrient Management Systems total number of nutrient manage-
ment plans developed annually for the U.S. portion of the basin, 2003.
Source: U.S. Department of Agriculture, Natural Resources Conservation Service
(NRCS), Performance and Results Measurement System

                                                                        235

-------
 basis due to the rapid changes in farming opera-
 tions. This does not allow for an estimate of the
 total number of CNMPs. USEPA will be tracking
 PNP as part of the Status's NPDES program.

 Figure 2 shows the number of Nutrient
 Management Plans by Ontario County for the years
 1998-2002, and Figure 3 shows cumulative acreage
 of Nutrient Management Plans for the Ontario por-
 tion of the basin. The Ontario Nutrient
 Management Act is moving farmers toward the
 legal requirement of having a nutrient management
 plan in place. Prior to 2002 the need for a plan was
 voluntary and  governed by municipal by-laws. The
 introduction of the Act presently requires new.
 expanding, and existing large farms to have a nutri-
 ent management plan. This has brought the expec-
 tation, which is reflected in Figure 2,  that there will
 be on-going needs to have nutrient management
 plans in place.

 Having completed a NMP provides assurance farm-
 ers are considering the  environmental implications
 of their management decisions. The more plans in place the bet-
 ter. In the future there may be a way to grade plans by impacts
 on the ecosystem. The first year in which this information is col-
 lected will serve as the base line year
                                                      Counties
                                                   n Bruce
                                                   • Elgin
                                                   • Huron
                                                   n Lambton
                                                   • Middlesex
                                                   n Oxford
                                                   • Perth
                                                   n Dundas
                                                   • Lennox & Addington
                                                   • Niagara
                                                   n Northumberland
                                                   n Peterborough
                                                   • Prescott
Figure 3. Cumulative acreage of Nutrient Management Plans for selected
Ontario Counties in the basin. Over 75% NMP acreages found in Huron, Perth.
Oxford and Middlesex Counties.
Source: Ontario Ministry of Agriculture and Food
-Cur
nulative Acrea
e by Year
Figure 2. Nutrient Management Plans by Ontario County, 1998-
2002.
Source:  Ontario Ministry of Agriculture and Food
 Pressures
 As livestock operations consolidate in number and increase in
 size in the basin, planning efforts will need to keep pace with
 236
              changes in water and air quality standards and technology.
              Consultations regarding the provincial and U.S. standards and
              regulations will continue into the near future.

              Acknowledgments
              Authors: Peter Roberts, Water Management Specialist, Ontario
              Ministry of Agriculture and Food, Guelph, Ontario Canada.
              peter.roberts@omaf. gov. on. ca;
              Ruth Shaffer, U.S. Department of Agriculture, Natural Resource
              Conservation Service, ruth.shaffer@mi.usda.gov; and
              Roger Nanney, Resource Conservationist, U.S. Department of
              Agriculture, Natural Resource Conservation Service.
              Authors' Commentary
              The new Nutrient Management Act authorizes the establishment
              and phasing in of province-wide standards for the management
              of materials containing nutrients and sets out requirements and
              responsibilities for farmers, municipalities and others in the busi-
              ness of managing nutrients. It is anticipated that the regulations
              under this act will establish a computerized NMP registry; a tool
              that will track nutrient management plans put into place. This
              tool could form a part of the future "evaluation tool box" for
              nutrient management plans in place in Ontario. The phasing in
              requirements of province-wide standards for nutrient manage-
              ment planning in Ontario and the eventual adoption over time of
              more sustainable farm practices should allow  for ecosystem
              recovery with time.
              The USDA's Natural Resources Conservation Service has

-------
formed a team to revise its Nutrient Management Policy. The
final policy was issued in the Federal Register in 1999. In
December 2000, USDA published its Comprehensive Nutrient
Management Planning Technical Guidance (CNMP Guidance) to
identify management activities and conservation practices that
will minimize the adverse impacts of animal feeding operations
on water  quality.  The CNMP Guidance is a technical guidance
document and does not establish regulatory requirements for
local, tribal, State, or Federal programs. PNPs are complementa-
ry to and leverage the technical expertise of USDA with its
CNMP Guidance. USEPA is proposing that Concentrated Animal
Feeding Operations, covered by the effluent guideline, develop
and implement a PNP. There is  an increased availability of tech-
nical assistance for U.S. farmers via Technical Service Providers,
who can provide  assistance directly to producers and receive
payment  from them with funds  from the  Environmental Quality
Incentives Program.

Last Updated
State of the Great Lakes 2005
                                                                                                                   237

-------

                                                                                           2007
Integrated Pest Management
Indicator # 7062

Assessment: Not Assessed

Purpose
  To assess the adoption of Integrated Pest Management (IPM)
practices and the effects IPM has had toward preventing surface
and groundwater contamination in the Great Lakes basin by
measuring the acres of agricultural pest management applied to
agricultural crops to reduce adverse impacts on plant growth,
crop production and environmental resources.

Ecosystem Objective
A goal for agriculture is to become more sustainable through the
adoption of more non-polluting, energy efficient technologies
and best management practices for efficient and high quality
food production. The sound use and management of soil, water,
air, plant, and animal resources is needed to prevent degradation
of agricultural resources. The process integrates natural resource,
economic, and social considerations to meet private and public
needs. This indicator supports Article VI (e) - Pollution from
Agriculture, as well as Annex 1, 2, 3, 11, 12 and 13 of the Great
Lakes Water Quality Agreement.

State of the Ecosystem
Background
Pest Management is controlling organisms that cause damage or
annoyance. Integrated pest management is utilizing environmen-
tally sensitive prevention, avoidance, monitoring and suppres-
sion strategies to manage weeds, insects, diseases, animals and
other organisms (including  invasive and non-invasive species)
that directly or indirectly cause damage or annoyance.
Environmental risks of pest management must be evaluated for
all resource concerns identified in the conservation planning
process, including the negative impacts of pesticides in ground
and surface water, on humans, and non-target plants and ani-
mals. The pest management component of an environmental
conservation farm plan must be designed to minimize negative
impacts of pest control on all identified resource concerns.

Agriculture accounts for approximately 35% of the  land area of
the Great Lakes basin and dominates the southern portion of the
basin. Although field crops  such as corn and soybeans comprise
the most crop acreage, the basin also supports a wide diversity
of specialty crops. The mild climate created by the Great Lakes
allows for production of a variety of vegetable and fruit crops.
These include tomatoes  (for both the fresh and canning markets),
cucumbers, onions and pumpkins. Orchard and tender fruit crops
such as cherries, peaches and apples are economically important
commodities in the region,  along with grape production for juice
or wine. The farmers growing these agricultural commodities are
major users of pesticides.

Research has found that reliance on pesticides in agriculture is
significant and that it would be impossible to abandon their use
in the short term. Most consumers want to be able to purchase
inexpensive yet wholesome food. Currently, other than organic
production, there is no replacement system readily available at a
reasonable price for consumers, and at a lesser cost to farmers,
that can be brought to market without pesticides. Other research
has shown that pesticide use continues to decline  as measured by
total active ingredient, with broad-spectrum pest control prod-
ucts being replaced by more target specific technology, and with
lowered amounts of active ingredient used per acre. Reasons for
these declines are cited as changing acreages of crops, adoption
of integrated pest management (IPM) and alternative pest con-
trol strategies such as border sprays for migratory pests, mating
disruption, alternative row spraying and pest monitoring.

With continued application of pesticides in the Great Lakes
basin, non-point source pollution of nearshore wetlands and the
effects on fish and wildlife still remains a concern. Unlike point
sources of contamination, such as at the outlet of  an effluent
pipe, nonpoint sources are more difficult to define. An estimated
21 million kg of pesticides are used annually on agricultural
crops in the Canadian and American Great Lakes  watershed
(GAO  1993). Herbicides account for about 75% of this usage.
These pesticides are frequently transported via sediment, ground
or surface water flow from agricultural land into the aquatic
ecosystem. With mounting concerns and evidence of the effects
of certain pesticides on wildlife and human health, it is crucial
that we determine the occurrence and fate of agricultural pesti-
cides in sediments, and in aquatic and terrestrial life found in the
Great Lakes basin. Atrazine and metolachlor were measured in
precipitation at nine sites in the Canadian  Great Lakes basin in
1995 (OMOE 1995). Both were detected regularly at all nine
sites monitored. The detection of some pesticides  at sites where
they were not used provides evidence of atmospheric transport
of pesticides.

Cultural controls (such as crop rotation and sanitation of infested
crop residues), biological controls, and plant selection and
breeding for resistant crop cultivars have always been an integral
part of agricultural IPM. Such practices were very important and
widely used prior to the advent of synthetic organic pesticides.
Indeed, many of these practices are still used today as compo-
nents of pest management programs. However, the great success
of modern pesticides has resulted in their use as the dominant
pest control practice for the past several decades,  especially
since the 1950s. Newer pesticides are generally more water solu-
ble, less strongly adsorbed to particulate matter, and less persist-
238

-------
ent in both the terrestrial and aquatic environments than the
older contaminants, but they have still been found in precipita-
tion at many sites.

Status of Integrated Pest Management
The Ontario Pesticides Education Program (OPEP) provides
farmers with training and certification through a pesticide safety
course. Figure 1 shows  survey results for 5800 farmers who took
pesticide certification courses over a three-year period (2001-
2004). Three sustainable practices (alter spray practices/manage
drift from spray, mix/load equipment in order to protect surface
and/or groundwater, and follow label precautions) and the farm-
ers' responses are shown. Results suggest that in 2004 more
farmers "do or plan to do now" these three practices after being
educated about their respective benefits. These practices have
significant value for reducing the likelihood of impairing rural
surface and groundwater quality. Figure 2 shows the acres of
pest management practice applied to cropland in the U.S. Great
Lakes basin for 2003.

Pressures
Pest management practices may be compromised by changing
land use and development pressures (including higher taxes);
flooding or seasonal drought; and lack of long-term financial
incentives for adoption  of environmentally friendly practices. In
order for integrated pest management to be successful, pest man-
agers must shift from practices focusing on purchased inputs
(using commercial sources of soil nutrients (i.e.  fertilizers) rather
than manure) and broad-spectrum pesticides to those using tar-
geted pesticides and knowledge about ecological processes.
Future pest management will be more knowledge intensive and
focus on more than the  use of pesticides. Federal, provincial and
state agencies, university Cooperative Extension programs, and
grower organizations are important sources for pest management
information and dissemination. Although governmental agencies
are more likely to conduct the underlying research, there is sig-
nificant need for private independent pest management consult-
ants to provide technical assistance to the farmer.

Management Implications
All phases of agricultural pest management, from research to
field implementation, are evolving from their current product-
based orientation to one that is based on ecological principles
and processes. Such pest management practices will rely more
on an understanding of  the biological interactions that occur
within every crop environment and the knowledge of how to
manage the cropping systems to the detriment of pests. The opti-
mum results would include fewer purchased inputs (and there-
fore a more sustainable  agriculture),  as well as fewer of the
human and environmental hazards posed by the broad-spectrum
pesticides so widely used today. Although pesticides will contin-
ue to be a component of pest management, the following are sig-
                       Follow Label Precaution/Safety
                          Percentage of participants
            D   10    20   30   40   50   60    70   80    90
I do this now/would do
    anyway
 I plan to do this now
   Don't plan to do
  this/No comment
                        Alter Spray Practices Manage Drift
                            Percentage of participants
            0    10    20    30    40    50    60    70    80
I do this now/would do
    anyway
 I plan to do this now
   Don't plan to do
  this/No comment
                  Mix/Load Equipment Protect Surface/Ground Water
                          Percentage of participants
            0   10   20   30   40   50    60   70   80
I do this now/would do
    anyway
  Don't plan to do
  this/No comment
Figure 1. Ontario selected grower pesticide safety training
course evaluation results from 2001-2004.
Source: Ontario Ministry of Agriculture and Food, Ontario
Ministry of the Environment (OMOE) and the University of
Guelph
                                                                                                                         239

-------
                                              OF   THE   GREAT
                             2007
nificant obstacles to the continued use of broad-spectrum pesti-
cides: pest resistance to pesticides; fewer new pesticides; pesti-
cide-induced pest problems; lack of effective pesticides; and
human and environmental health concerns.

Based upon these issues facing pesticide use, it is necessary to
start planning now in order to be less reliant on broad-spectrum
pesticides in the future. Society is requiring that agriculture
become more environmentally responsible through such things
as the adoption of Integrated Pest Management.  This will require
effective evaluations of existing policies and implementing pro-
grams for areas such as Integrated Pest Management. To reflect
these demands there is a need to further develop this indicator.
The following types of future activities could assist with this
process:
      Indicate and track future adoption trends of IPM best
    management practices;
      Analyze rural water quality data for levels of pesticide
    residues;
      Evaluate the success of the Ontario Pesticide Training
     Course, such as adding and evaluating survey questions
    regarding IPM principles and practices  to course evaluation
    materials; and
      Evaluate the number of farmers and vendors who attend-
     ed, were certified, or who failed the Ontario Pesticides
    Education Program.
Note: Grower pesticide certification is mandatory in Ontario and
in all Great Lakes States, and it applies to individual farmers as
well as custom applicators.

Acknowledgments
Authors: Peter Roberts, Water Quality Management Specialist,
Resources Management, Ontario Ministry of Agriculture and
Food, Guelph, peter.roberts@omaf.gov.on.ca;
Ruth Shaffer United States Department of Agriculture, Natural
Resources Conservation Service, ruth.shaffer@mi.usda.gov; and
Roger Nanney, Resource Conservationist, United States
Department of Agriculture, Natural Resources Conservation
Service.

Sources
U.S. General Accounting Office (GAO). 1993. Pesticides -
Issues concerning pesticides used in the Great Lakes watershed.
GAO/RCED-93-128. Washington, DC. 44pp.

Ontario Ministry of the Environment (OMOE). 1995. Water
monitoring 1995. Environmental Monitoring and Reporting
Branch.

Last Updated
State of the Great Lakes 2005
                                                   Pesticide Management Applied

                                                   ID 0-1,500 Acres
                                                   ^J 1,500-5,000 Acres
                                                      5,000 - 10,000 Acres
                                                   • 10,000-17,500 Acres
  Figure 2. Annual U.S. Pesticide Management Systems Planned for 2003.
  Source: U.S. Department of Agriculture, Natural Resources Conservation Service
  (NRCS), Performance and Results Measurement System
240

-------
Vehicle Use
Indicator # 7064

Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Poor
Deteriorating
Population growth and urban sprawl in the Great Lakes Basin has led
to an increase in the number of vehicles on roads, fuel consumption,
and kilometers spent on the road by residents. Vehicle use is a driver of
fossil fuel consumption, deteriorating road safety, and ecological
impacts such as climate change and pollution.
Purpose
To assess the amount and trends in vehicle use in the Great Lakes Basin (GLB) and to infer the
societal response to the ecosystem stressed caused by vehicle use.

Ecosystem Objective
This indicator supports Annex 15 of the Great Lakes Water Quality Agreement. An alternative
objective is to reduce stress on the environmental integrity of the Great Lakes region caused by
vehicle use.

State of the Ecosystem
A suite of indicators monitoring vehicle use, the number of registered vehicles licensed, and fuel
consumption is measured by governments in Canada and the United States to capture trends
linked to fossil fuel consumption, deteriorating road safety, and ecological impacts such as
climate change and pollution. Figure 1 shows the estimated total distance travelled by vehicles on
roads in Ontario during 1993-2003 and the number of licensed vehicles registered in Ontario for
the same period. The number of registered vehicles in Ontario rose 21% from over 6.3 million in
1990 to 7.6 million in 2004. More significant, however, is the  estimated 122 million vehicle
kilometers travelled (VKT) in Ontario, up 62% from 75 million in 1993. The greatest increase in
VKT occurred between 1999 and 2000 (an increase of 39%). From this data, it is evident that
Canadians in the Great Lakes Basin are increasingly spending more time on the road.

Looking to the U.S., Figure 2 shows the estimated trends in registered vehicles, licensed drivers,
and vehicle kilometers travelled in the Great Lakes States from 1994 to 2004. The number of
registered vehicles increased approximately 11 % during this time period, while the number of
licensed drivers only increased 8%.  These increasing trends are somewhat lower than national
averages in the U.S., showing increases of 20% and 13%, respectively. Just as in Ontario, VKT
increased at a greater rate than the number of registered vehicles or licensed drivers. VKT
increased in the Great Lakes States approximately 20% from 1994 to 2004, as compared to a 24%
national U.S. increase. In 2004, U.S. residents in the Great Lakes States gained 7% more
kilometres per vehicle than were driven in 1994.

A snapshot of the total registered vehicles in Ontario points abundantly to a societal dependence
on private vehicles. Of the total registered vehicles in Ontario, passenger vehicles  continually
dominate road traffic, accounting for 74% of the total registered vehicles in 2004.  As anyone who
has  driven on basin highways might guess, commercial freight traffic was the runner-up,
                         Draft for Discussion at SOLEC 2006

-------

accounting for 14% of road traffic in the same year. Notably, trucking flows of inter-provincial
trade through Quebec and Ontario (both directions) also accounted for $41 billion worth of
commodities or 30 per cent of total inter-provincial trade in Canada.

The movement of people is undoubtedly a driving force behind the economic profitability of the
GLB. However, the tradeoffs of unsustainable modes of transport are evident. In Canada, road
transportation, including private vehicles, represented 77% of total transportation in terms of
energy use in 2004. As a result, energy-related GHGs rose by 25%, from 135.0 megatonnes to
168.8 megatonnes. In that same time period, the number of vehicles rose 8% faster than the
number of people (Canada, 2005). In Ontario, sale of motor gasoline increased by 22% between
1989 and 2004 (Figure 3), on par with the national average. Gasoline sales rose from more than
12 billion litres to more than 15 billion litres between 1990 and 2003, with diesel fuel sales in
Ontario alone doubling during the same period, from more than 12 million to almost 15 million
litres. In the Great Lakes States, fuel (gasoline and gasohol) consumption for vehicles increased
by 17% from 1994 to 2004, as compared to a 24% increase nationally in the U.S. It is noteworthy
to point out that use of ethanol blended fuels (gasohol) in the Great Lakes States increased 160%
over this time period. Gasohol now comprised approximately 39% of fuel consumption in the
Great Lakes States. The increased demand for fuel in both countries is driven by a rise in number
of vehicles on highways, increased power of automobile engines, and the growing popularity of
sports utility vehicles and large-engine cars (Menard, 2006)

Over the last decade, consumers have also shown a strong preference for high-performance
vehicles.  Since 1999, the production of Sport Utility Vehicles (SUVs) has dominated the
automotive industry, surpassing the output of both minivans and pickup trucks nation wide. For
the period of January to September 2004, SUVs accounted for 18% of total light-duty vehicle
manufacturing, which assembles passenger cars, vans, minivans, pickup trucks and SUVs in
Canada (Magnusson, 2005). In the Great Lakes States, the registrations of private and
commercially owned trucks, which include personal passenger vans, passenger minivans, and
sport-utility vehicles, have increased approximately 50% from 1994 to 2004.  Private and
commercially owned trucks now comprise about 37% of all registered vehicles in the Great Lakes
States. Although the fuel economy of the average new car has improved more than 76% since
1975, the automotive industry has traded off fuel consumption improvements in new vehicles for
more powerful engines. This improved performance reduced the fuel economy that otherwise
could have been achieved, meaning, cars collectively get worse gas mileage today than they did
in the mid-1980's (NRC, 1992)—the effects of which are experienced with diminished air quality
locally.

Pressures
Suburban development has become the predominant form of growth in the Great Lakes Basin.
The mixed assessment in the GLB urban air quality can be directly linked to the increase in traffic
congestion. Presently, transportation GHG emissions are increasing at a slower rate than activity
because of the more efficient travel of people and goods. However, all modes  of transport are still
greatly dependent on GHG-intensive hydrocarbons to provide them with energy. As a major
driver of ecological stress, vehicles are the single largest domestic source of the smog-causing
greenhouse gas (GHG) emissions. These emissions include nitrogen oxides (NOx) and volatile
organic compounds (VOCs) as well as carbon monoxide (CO), all which contribute contaminants
                         Draft for Discussion at SOLEC 2006

-------

to air and water systems (MOE, 2005). Such pollutants have been connected with respiratory
problems and premature death. There is strong evidence that atmospheric deposition is a source
of pollutants in storm water runoff, and that this runoff reaches streams, rivers and other aquatic
resources (UC, 2004). Congestion caused by automobiles and vehicle-related development also
degrades the liveability of urban environments by contributing noise, pollution, and fatalities.
Positive trends in road use may also lead to further fragmentation of natural areas in the basin.

Management Implications
There is a need to reduce the volume and congestion of traffic in the GLB. While progress has
been made through less polluting fuels and emission reduction technologies, and economic tools
such as the tax incentives  that encourage the purchase of fuel-efficient vehicles (e.g. Tax  for Fuel
Conservation), issues of urban sprawl must also be managed. Recent studies by the U.S. EPA
found that infill development and re-development of older suburbs could reduce VKT per capita
by 39% to 52% (depending on the metropolitan area studied) (Chiotti, 2004). The success of
current strategies will assist managers and municipalities protect natural areas, conserve valuable
resources (such as agriculture  and fossil fuels), ensure the stability of ecosystem services, and
prevent pollution. Under the Kyoto Protocol, Canada is committed to reducing its GHG emissions
by 6% below 1990 levels  by the year 2010, even though the government may consider new
targets.

Over the next 30 years,  the number of people living in Ontario is expected to grow by
approximately four million, the majority of which are expected to reside in the GLB. In the
Golden Horseshoe Area alone, 2031 forecasts predict that the population of this area is to grow
by an additional 3.7 million (from 2001) to 11.5 million.  The McGuinty government has invested
in the several initiatives (including, Bill 26, the Strong Communities Act, 2004) in order to
manage regional growth and development, and municipalities and regions within the GLB are
developing their own plans within the common mandate.

Improving public transit is the first investment priority, however there is an acknowledgment that
improving population growth forecasts, intensifying land use, revitalizing urban spaces,
diversifying employment  opportunities, curbing sprawl, protecting rural areas, and improving
infrastructure are all part of the solution. Urban development strategies must be supported by
positive policy and financial frameworks that allow municipalities to remain profitable, while
creating affordable housing and encouraging higher density growth in the right locations. Further
research, investment and action are needed to explore multi-modal corridors and modes for
transporting goods in the basin.

Comments from the author(s)
It should be noted that Canadian Vehicle Kilometres Travelled (VKM) data is based on a
voluntary vehicle-based survey conducted by Transport Canada. The measure of vehicle-
kilometres travelled does not take into account occupancy rates, which affect the sustainability of
travel.

It also should be noted that U.S. motor fuel data come from the records of State agencies  that
administer the State taxes on motor fuel are the underlying  source for most of the data presented
in these tables.  Over the last several years, there have been numerous changes in  State fuel tax
laws and procedures that have resulted in improved fuel tax compliance, especially for diesel fuel.


                         Draft for Discussion at SOLEC 2006                        3

-------

The improved compliance has resulted in increased fuel volumes being reported by the States to
FHWA. The trends shown in the tables reflect both improvements in tax compliance and changes
in consumption.

U.S. VKT data - These data are derived from the Highway Performance Monitoring System
(HPMS). The HPMS is a combination of sample data on the condition, use, performance and
physical characteristics of facilities functionally classified as arterials and collectors (except rural
minor collectors) and system level data for all public roads within each State.

Although data about VKT, registered vehicles, and fuel consumption was only available up to
2004, the authors feel this indicator should be updated in future to examine potential shifts in
vehicle-use behaviours based on the recent rise in gasoline prices, which began climbing in late
2002. A 2005 report by Transport Canada, based on partial data, suggest that gas prices post-
Hurricane Katrina had an impact on fuel consumption nationally.

Acknowledgments
Authors: Katherine Balpataky, Environment Canada, Burlington; and Todd Nettesheim, U.S.
EPA, Great Lakes National Program Office, Chicago, IL.

Data Sources
Ontario data for Vehicle Kilometres Travelled was obtained from the Ministry of Transportation,
Ontario Road Safety Annual Reports. Original source of VKT data Statistics Canada, Canadian
Vehicle Survey, Statistics Canada Catalogue No.  53-223-XIE, 2000 to 2003.

Chiotti, Q. Toronto's Environment: A Discussion on Urban Sprawl and Atmospheric Impacts.
Pollution Probe. 2004. www.pollutionprobe.org.  Last viewed August 25, 2006.

Committee on Fuel Economy of Automobiles and Light Trucks, National Research Council.
Automotive Fuel Economy: How Far Can We Go? 1992. The National Academic Press.

Davis,  William B., Levine, Mark D. and Train, Kenneth. 1993. Feebates: Estimated Impacts on
Vehicle Fuel Economy, Fuel Consumption, CO2 Emissions,  and Consumer Surplus, Lawrence
Berkeley Laboratory, Berkeley, California.

Government of Canada (Environment Canada, Statistics Canada, Health Canada). Canadian
Environmental Sustainability Indicators. 2005.

International Joint Commission. IJC Air Quality Report. 2004.
http://www.iic.org/php/publications/pdf/ID 1544.pdf, last viewed 28 August 2006.

Magnusson, E. Sport Utility Vehicles: Driving Change. Statistics Canada. Manufacturing,
Construction and Energy Division. No. 11-621-MIF2005020. 2005.
http://www.statcan.ca/english/research/ll-621-MIE/ll-621-MIE2005020.htm. last accessed 25
August 2006.
                         Draft for Discussion at SOLEC 2006

-------

Menard, M. Canada, a Big Energy Consumer: A Regional Perspective, Manufacturing,
Construction and Energy Division. Statistics Canada. Manufacturing, Construction and Energy
Division. http://www.statcan.ca/english/research/l 1-621-MIE/l l-621-MIE2005023.htm, last
viewed 28 August 2006.

Ministry of the Environment. Drive Clean Reduced Harmful Emissions. 2005.
http://www.ene.gov.on.ca/envision/news/2005/111801fs.htm. last viewed 28 August 2006.

Ministry of Public Infrastructure Renewal.  Ontario. Growth Plan for the Greater Horseshoe
Area. 2006. http://www.pir.gov.on.ca. last viewed 28 August 2006.

Natural Resources Canada. Energy Efficiency Trends in Canada, 1990 to 2003. June 2005.
http://oee.nrcan.gc.ca/corporate/statistics/neud/dpa/data e/trends05/index.cfm?attr=0, last viewed
28 August 2006.

Statistics Canada, Canadian Vehicle Survey, Statistics Canada Catalogue No. 53-223-XIE, 2000
to 2003.

Statistics Canada, Road motor vehicles, fuel sales, CANSIM Table 405-0002.
http://www40.statcan.ca/101 /cstO 1 /trade37a.htm, last viewed 28 August 2006.

Statistics Canada. Statistics Canada's Energy Statistics Handbook.  2006.
http://www.statcan.ca/english/freepub/57-601-XIE/57-601-XIE2006001.pdf

Transport Canada. Integration Technologies for Sustainable Urban Goods Movement. 2004.
http://www.tc.gc.ca/pol/en/Report/UrbanGoods/Report.htm. last viewed 28 August 2006.

Transport Canada. The Full Cost Investigation of Transportation in Canada. 2005.
http://www.transport-canada.org/pol/en/aca/fci/menu.htm. Last viewed 5 October 2006.

U.S.  Department of Transportation. Federal Highway Administration. Office of Highway Policy
Information. Highway Statistics Publications.
http://www.fhwa.dot.gov/policy/ohpi/hss/hsspubs.htm

List  of Tables
Table 1: Primary energy consumption of Motor Gasoline  and Diesel Fuel, Canada, 1990 and
2003.
Source: Report on energy supply-demand in Canada, Statistics Canada Catalogue No. 57-003-
XIB, 1990 and 2003; population estimates, CANSIM Table 051-0001; Real GDP, CANSIM
Table 384-0013.

List  of Figures
Figure 1. Number of Licensed Vehicles and Vehicle Kilometres Travelled in Ontario.
Data Source: Statistics Canada Canadian Vehicle Survey.

Figure 2. Number of Registered Vehicles, Licensed Drivers and Vehicle Kilometres Travelled in
Great Lakes States.
                         Draft for Discussion at SOLEC 2006

-------

Data Source: U.S. Department of Transportation. Federal Highway Administration. Office of
Highway Policy Information. Highway Statistics Publications.

Figure 3. Petroleum Consumption in Ontario.
Data source: Statistics Canada's Energy Statistics Handbook. 2006

Last updated
SOLEC 2006
Variable
Level
1990
2003
Variation from
1990 to 2003
value
%
% share of
energy
consumed
1990
2003
%
contribution
to change
Primary energy consumption in terajoules
Motor gasoline
Diesel fuel
432,446
169,466
539,230
248,437
106,784
78,971
25
47
15
6
16
8
22
16
Table 1. Primary energy consumption of Motor Gasoline and Diesel Fuel, Canada, 1990 and
2003.
Source: Report on energy supply-demand in Canada, Statistics Canada Catalogue No. 57-003-
XIB, 1990 and 2003; population estimates, CANSIM Table 051-0001; Real GDP, CANSIM
Table 384-0013.
                        Draft for Discussion at SOLEC 2006

-------
           State of the Great Lakes 2007 - Draft
  9,000,000 T
  8,000,000 --
  7,000,000 --
o

$
  6,000,000 --
•= 5,000,000 --

•o
£

« 4,000,000
5)
at
o:
•s
j; 3,000,000
Si

3
z
  2,000,000 --
  1,000,000 --
T 140,000
                                                                                            -- 120,000
                                                                                            -- 100,000
        _c

        •o
        _2


-- 80,000  2

        1
        "3


-- 60,000  i

        o
-- 40,000  o
        "fi


        UJ

-- 20,000
            1993    1994    1995    1996    1997    1998   1999   2000   2001    2002    2003    2004

                                                  Year
                   -Number of registered vehicles in Ontario  *  Estimated Vehicle Kilometres Travelled (in millions)
 Figure 1. Number of Licensed Vehicles and Vehicle Kilometres Travelled in Ontario.
 Data Source: Statistics Canada Canadian Vehicle Survey
                            Draft for Discussion at SOLEC 2006

-------
~ 40,000,000
•5, 30,000,000
                               State of the Great Lakes 2007 - Draft
                                         1999
                                         Year
                             Registered Vehicles + Licensed drivers —•—VMT
                                                                              1,000,000 °
Figure 2. Number of Registered Vehicles, Licensed Drivers and Vehicle Kilometres Travelled in
Great Lakes States.
Source: U.S. Department of Transportation. Federal Highway Administration. Office of Highway
Policy Information. Highway Statistics Publications.
                        Draft for Discussion at SOLEC 2006

-------


  18,000
_ 16,000
Cfl
0

•s



." 14,000
•o 12,000
- 10,000
   8,000
o  6,000
   4,000
   2,000
      1988
               1990
                         1992
                                  1994
                                           1996       1998


                                                Year
                                                              2000
                                                                       2002
                                                                                 2004
                                                                                          2006
  Figure 3. Petroleum Consumption in Ontario.

  Data source: Statistics Canada's Energy Statistics Handbook. 2006
                            Draft for Discussion at SOLEC 2006

-------
Wastewater Treatment and Pollution
Indicator # 7065

Note: This is a progress report towards implementation of this indicator.

Overall Assessment
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  Data to support this indicator have not been summarized according to
      Determining  quality control standards.  Compilation of a comprehensive report on
  Status and Trend  wastewater treatment and pollution in the Great Lakes will require a
                   substantial amount of additional time and effort.

Lake-by-Lake Assessment
A lake-by-lake assessment is not available at this time as data summarization is incomplete and
not available for analysis on a lake-by-lake basis.

Purpose (proposed)
    •  To measure the proportion of the population served by municipal sewage treatment
       facilities
    •  To evaluate the level of municipal treatment provided
    •  To measure the percent of collected wastewater that is treated; and
    •  To assess the loadings of metals, phosphorus, Biochemical Oxygen Demand (BOD), and
       organic chemicals released by wastewater treatment plants into the water courses of the
       Great Lakes basin.

Ecosystem Objective
The purpose of this indicator is to assess (1) the reduction of pressures induced on the ecosystem
by insufficient wastewater treatment networks and procedures and (2) to further the progression
of wastewater treatment towards  sustainable development.

This indicator is also intended to  (3) assess the scope of municipal sewage treatment and the
commitment to protecting freshwater quality in the Great Lakes basin. The quality of wastewater
treatment determines the potential adverse impacts to human and ecosystem health as a result of
the loadings of pollutants discharged into the Great Lakes basin.

State of the Ecosystem
Background Information
Wastewater refers to the contents of sewage systems drawing liquid wastes from a variety of
sources, including municipal, institutional and industrial, and stormwater discharges. After
treatment, wastewater is released into the environment from a treatment plant as effluent into
receiving waters such as lakes, ponds, rivers, streams and estuaries.

Wastewater contains a large number of potentially harmful pollutants, including some that are the
result of biological activity as well as others that are part of the over 200 identified chemicals
from industries, institutions, households, and other sources. Wastewater systems are designed to
                         Draft for Discussion at SOLEC 2006

-------
collect and treat these wastes using various levels of treatment to remove pollutants prior to
discharge, ranging from no treatment to very sophisticated and thorough treatments. Despite
treatment, effluents released from wastewater systems can still contain pollutants of concern,
since even advanced treatment systems do not necessarily remove all pollutants and chemicals.

The following constituents, mostly associated with human waste, are present in all sewage
effluent to some degree:
    •   biodegradable oxygen-consuming organic matter (measured as biochemical oxygen
       demand or BOD);
    •   suspended solids (measured as total suspended solids or TSS);
    •   nutrients, such as phosphorus (usually measured as total phosphorus) and nitrogen-based
       compounds (nitrate, nitrite, ammonia, and ammonium, which are measured either
       separately or in combination as total nitrogen);
    •   microorganisms (which are usually measured in terms of the quantity of representative
       groups of bacteria, such as fecal coliforms or fecal streptococci, found  in human wastes);
    •   sulphides;
    •   assorted heavy metals; and
    •   trace amounts of other toxins and emerging chemicals of concern that have yet to be
       consistently monitored for in wastewater effluents.

Municipal Wastewater Effluent (MWWE) is one of the largest sources of pollution, by volume,
discharged to surface water bodies in Canada (CCME, 2006). Reducing the discharge of pollution
through MWWE requires a number of interventions ranging from source control to end of pipe
measures.

Levels of Treatment in the U.S. and Canada
The concentration and type of effluent released into the receiving body of water depends heavily
on the type of sewage treatment used. As a result, information regarding the level of treatment
that was used on wastewater is integral in assessments of potential impacts on water quality.  In
both the United States and Canada, the main levels of wastewater treatment used  include primary,
secondary, and advanced or tertiary.

In primary wastewater treatment, solids are removed from raw sewage primarily through
processes involving sedimentation. This process typically removes  roughly 25-35% of solids and
related organic matter (U.S. EPA 2000).

In the U.S., pretreatment may also occur preliminary to primary treatment, in which contaminants
are reduced and large debris is removed from industrial wastewater  before it is  discharged to
municipal treatment systems to undergo regular treatment. U.S. federal regulations require that
Publicly Owned Treatment Works (POTW) Pretreatment Programs  include the development of
local pretreatment limits in situations where industrial pollutants could potentially interfere with
municipal treatment facility operations or contaminate sewage sludge. The U.S. EPA can
authorize the states to implement their own Pretreatment Programs as well.  Of the eight states
that are part  of the Great Lakes basin, Michigan, Minnesota, Ohio and Wisconsin currently hold
an approved State Pretreatment Program, (U.S.  EPA, NPDES 2006).
                         Draft for Discussion at SOLEC 2006

-------

Secondary wastewater treatment includes an additional biological component in which oxygen-
demanding organic materials are removed through bacterial synthesis enhanced with oxygen
injections.  About 85% of organic matter in sewage is removed through this process, after which
the excess bacteria are removed, (U.S. EPA 1998).  Effluent can then be disinfected with chlorine
prior to discharge in an effort to kill potentially harmful bacteria.  Subsequent dechlorination is
also often required to remove excess chlorine that may be harmful to aquatic life.

Secondary treatment effluent standards are established by the EPA and have technology-based
requirements for all direct discharging facilities.  These standards are expressed as a minimum
level of effluent quality in terms of biochemical oxygen demand measurements over a five-day
interval (BOD5), total suspended solids (TSS)  and pH.  Secondary treatment of municipal
wastewater is the minimum acceptable level of treatment according to U.S. federal law unless
special considerations dictate otherwise (U.S. EPA 2000).

Advanced, or tertiary, levels of treatment often occur as well and are capable of producing high-
quality water.  Tertiary treatment can include the removal of nutrients such as phosphorus and
nitrogen and essentially all suspended and organic matter from wastewater through combinations
of physical and chemical processes.  Additional pollutants can also be removed when processes
are tailored for those purposes.

Data on the level of treatment utilized in the United States are  available from the Clean Water
Needs Survey (CWNS).  This cooperative effort between the U.S. EPA and the states resulted in
the creation and maintenance of a database with technical and  cost information on the 16,000
publicly owned wastewater treatment facilities in the nation. According to the  results of the 2000
CWNS, the total population served by POTWs in the U.S. portion of the Great  Lakes basin was
17,400,897 in 2000. Of this number, 0.7% received treatment from facilities that do  not
discharge directly into Great Lakes waterways and dispose of wastes by other means, 14.1%
received secondary treatment, and 85.3% received treatment that was greater than secondary,
making advanced treatment the type  used most extensively. Please see Figure 1 for the complete
distribution of population served according to level of treatment by major lake  and river basins
within the U.S. Great Lakes watershed. These values do not include  a possible additional 12,730
people who were reportedly served by facilities in New York for which watershed locations are
unknown within the CWNS database. Although the facilities are in counties at least partially
within the U.S. portion of the Great Lakes region, their location within Great Lakes watersheds
can not be easily verified.

Wastewater Treatment Plants (WTPs) in Ontario also use primary, secondary, and tertiary
treatment types.  The processes are very similar, if not the same to those used in the U.S.
(described above), but Canadian regulatory emphasis is placed on individual effluent quality
guidelines as opposed to mandating that a specific treatment type be utilized across the province.

A complete distribution of population served according to level of treatment is  not available in
the Great Lakes basin for Ontario at this time. However, for a general understanding, a
distribution of the population served  by each treatment type for all of Canada is available in
Figure 2.
                         Draft for Discussion at SOLEC 2006

-------
Tertiary or advanced treatment is the most common type of sewage treatment across the basin,
which can be inferred from the combined trends demonstrated in both Figures 1 and 2. This
indicates the potential for high effluent water quality, which can only be verified through analysis
of regulatory and monitoring programs.

Condition of Wastewater Effluent in Canada and the United States:
Regulation. Monitoring,  and Reporting
Canada does not regulate effluent conditions through treatment level requirements, but instead
sets specific limits for each individual WTP, no matter which type of treatment is used.  In the
U.S., effluent limits are standardized by the  Federal Government, but the states have the power to
make alterations as long  as minimum guidelines are met.

Each federally managed wastewater treatment plant (WTP) in Canada must also follow guidelines
given by the Federal Government.  Effluent guidelines for wastewater from Federal facilities are
to be equal to or more stringent than the established standards or requirements of any Federal or
Provincial regulatory agency (Environment  Canada, 2004). The guidelines indicate the degree of
treatment and the effluent quality applicable to the wastewater discharged from the specific WTP.
Use of the Federal guidelines is intended to  promote a consistent wastewater approach towards
the cleanup and prevention of water pollution and ensure that the best practicable control
technologies are used (Environment Canada, 2004).

Table 1 lists the pollutant effluent limits specified for all federally approved WTPs in Ontario.
The effluents discharged to the receiving water should receive treatment such that an effluent of
minimum quality is achieved. In general, compliance with the numerical limits should be based
on 24 hour composite samples (Environment Canada, 2004).

In Ontario, wastewater treatment and effluents are monitored through a Municipal Water Use
Database (MUD) through Environment Canada. This database uses a survey for all municipalities
to report on wastewater treatment techniques. Unfortunately, the last complete survey is from
1999 and this data are not sufficient to use for this report. The most up to date municipal water
use survey will be released within the next few months and would useful to examine the treatment
results  within Canada. Unfortunately, the survey is not yet available, and other methods have
been chosen to examine wastewater treatment in Ontario, which are explained in the Attempted
Eperimental Protocols section of this progress report.

The U.S. regulates and monitors wastewater treatment systems and effluents through a variety of
national programs. The U.S. EPA's Office of Wastewater Management promotes  compliance
with the Clean Water Act through the National Pollutant Discharge Elimination System (NPDES)
Permit Program. These permits regulate wastewater discharges from POTWs by setting effluent
limits, monitoring and reporting requirements, and they can lead to enforcement actions when
excessive violations  occur. The U.S. EPA can authorize the states to implement all or part of the
NPDES program, and all US states in the Great Lakes region are currently approved to do so
provided they meet minimum federal requirements, (U.S. EPA, NPDES 2006). This distribution
of implementation power can create difficulties when specific assessments are attempted across
regions spanning several states.
                         Draft for Discussion at SOLEC 2006

-------
Large scale national assessments of wastewater treatment have been completed in the past by
using BOD and dissolved oxygen (DO) levels as indicators of water quality.  Since DO levels
have been proved to be related to BOD output from wastewater discharges (increased BOD
loadings lead to greater depletion of oxygen and lower DO levels in the water), historical DO
records can be a useful indicator of water quality responses to wastewater loadings. According to
a national assessment of wastewater treatment completed in 2000, the U.S. Great Lakes basin had
a statistically significant improvement in worst-case DO levels after the Clean Water act (U.S.
EPA 2000).  The study's design estimates also showed that the national discharge of BOD5 in
POTW effluent decreased by about 45%, despite a significant increase of 35% in the population
served and the influent loadings. This improving general trend supported assumptions made in the
1996 CWNS Report to Congress that the efficiency of BOD removal would increase due to the
growing proportion of POTWs using advanced treatment processes across the nation.

Although specific case studies do exist, unfortunately comprehensive studies such as the
examples listed above have not been conducted for pollutants other than BODs, and have not
been completed to an in-depth level for the  Great Lakes region.

An extensive investigation of the Permit Compliance System (PCS) database is one way such a
goal can be accomplished. This national information management system tracks NPDES data
including permit issuance, limits, self-monitoring, and compliance. The PCS database can
provide the information necessary to calculate the loadings of specific chemicals present in
wastewater effluent from certain POTWs in the U.S. portion of the Great Lakes basin, providing
the relevant permits exist.

Attempted Experimental Protocol for Calculating Pollutant Loadings from Wastewater Treatment
Plants to the Great Lakes
This calculation was attempted for the U.S. and Canadian portion of the G.L. basin during the
compilation of this report, and although an extensive amount of data  are  available and have been
retrieved, their summarization to an appropriate level of quality control is substantially difficult
and is not complete at this time. The protocol followed thus far is outlined below.

The initial procedure for mining the U.S. data from the PCS database began with the compilation
of a list of all the municipal wastewater treatment facilities located within the Great Lakes basin.
The determination of which pollutants were most consistently permitted for across the basin
followed, and the effluent loadings data for all facilities that monitored for those parameters were
obtained for 2000 and 2005. These pollutant parameters were often referred to by various
common names in the database, which additionally complicated extraction of concise data. The
resulting mass of data was extremely large and could not be feasibly  summarized due to internal
inconsistencies such as difference in units of measurement, monitoring time frames, extreme
outliers, and apparent data entry mistakes.

In an effort to decrease the amount of U.S. data requiring analysis at  a more precise level, (as a
result of the problems mentioned above,) several specific facilities throughout the basin were
chosen to hopefully serve as representative  case studies off which total loadings estimates could
be calculated.  These facilities were chosen by two sets of criteria.  The first was according to
location within the basin, to ensure that all states and each Great Lake were represented. The
second criteria was the greatest average level of effluent flow, as the  selected facilities could
                         Draft for Discussion at SOLEC 2006

-------
                                           f^^^M^ii^^MiSi^^^	
potentially have the greatest impact due to sheer volume of effluent, and these values could also
be used to calculate loadings in cases where pollutant measurements were gathered as a
concentration as opposed to by quantity (as was often the case).  Fifteen facilities were eventually
selected for further analysis, and corresponding effluent measurements for basic pollutants were
extracted from the PCS database. Calculation of percent change in pollutant loadings and the
number of violations from 2000 to 2005 was attempted for these data, but results are not available
yet due to the data quality issues described earlier.

With total effluent loadings being so difficult to calculate independently from database records,
government generated historical records of effluent limit violations can provide some insight into
the performance of U.S. Great Lakes wastewater treatment facilities.  The Enforcement and
Compliance History Online (ECHO) is a publicly accessible data system funded by the EPA that
was used to obtain violation information by quarter over a three year time span for the group of
15 U.S. facilities previously selected for loadings calculations.

The resulting compliance data are presented in Figure 3 according to each pollutant for which
violations of permitted effluent levels occurred during the 12 possible quarters under
investigation from 2003-2006. This information is further separated out into quarters that
demonstrated basic violations of effluent limits and those that had a significant level of non-
compliance with permitted  effluent limits.  Chloride, fecal coliform, and solids violations were
the most common, with copper, cyanide, and mercury having higher numbers of violations as
well.  Chloride, copper, mercury, and solids violations showed the most significant non-
compliance with permitted  levels.

In Ontario, wastewater treatment plants must report on the operation of the system and the quality
of the wastewater treatment procedures on an annual basis to satisfy the requirements of the
Ontario Ministry of Environment and the Certificate of Approval. Each report fulfills the
reporting requirements established in section 10(6) of the Certificate of Approval made under the
Ontario Water Resources Act (R.S.O. 1990, c. O.40). As a result of these requirements, effluent
limit violations for BOD, phosphorus, and suspended solids should be available for future
analysis.  Data are too extensive to summarize at this time to a sufficient level of quality control.

Since results from the Municipal Water Use Database were not available at this time, the data
used for the Canadian component of this report were provided by 10 municipalities in the Great
Lakes basin. Municipalities were randomly chosen based on their proximity to the great lakes
and their population of over 10,000. Most of the chosen municipalities had about one to three
wastewater treatment plants, which compiled to 24 treatment plants being examined in total for
this indicator report. Data from 2005 annual reports for each wastewater treatment plant were
used to analyze wastewater treatment procedures and associated effluent quality for this indicator,
with special focus on four specific pollutant parameters. These include BOD, phosphorus and
suspended solids, all of which are indicators of potential health hazards.

These parameters are regulated by most wastewater treatment plants, which when exceeded, have
the potential to have serious effects on human health. Current targets exist to minimize
environmental and health impacts. For example, Ontario WTPs  have a target of 50% for the
removal of BOD and limits must not exceed 20mg/L in a 5 day span. The target for the removal
                         Draft for Discussion at SOLEC 2006

-------
of suspended solids is 70%, with a limit of 25 mg/L in a 24 hour sample period. And although
some nutrients are essential for plant production in all aquatic ecosystems, an oversupply of
nutrients, particularly from municipal wastewater effluents, can lead to the growth of large algal
blooms and extensive weed beds (Environment Canada 2001).  Resulting wastewater effluent
limits for phosphorus in Ontario have been set at l.Omg/L accordingly. Completed results
corresponding to the exceedences of these limits are not available for Ontario at this time.

Pressures
There are numerous challenges to providing adequate levels of wastewater treatment in the Great
Lakes basin. These include: facility aging, disrepair and outdatedness; population growth that
stresses the capabilities of existing plants and requires the need for more facilities; new and
emerging contaminants that are more complex and prolific than in the past; and new development
that is located away from urban areas and served by decentralized systems (such as septic
systems) that are much harder to  regulate and monitor.  The escalating costs associated with
addressing these challenges continue to be a problem for both U.S. and Canadian municipalities,
(U.S. EPA, 2004 and Government of Canada, 2002).

Management Implications
Despite demonstrated significant progress with wastewater treatment across the basin,  substantial
problems remain with regards to  nutrient enrichment, sediment contamination,  heavy metals, and
toxic organic chemicals that still  pose threats to the environment and human health. It is
therefore important to continually invest in wastewater treatment infrastructure improvements, so
any current achievements in water pollution control are not overwhelmed by the demands of
future urban population growth and so other remaining concerns can be addressed such as
polluted urban runoff and untreated municipal stormwater. These sources have emerged as prime
contributors to local water quality problems throughout the basin (Environment Canada, 2004).
WTPs are having difficulties keeping up with demands created by urban development which
cause an increasing amount of bypass into the Great Lakes. The governments of Canada and
Ontario and municipal authorities, working under the auspices of the Canada-Ontario Agreement
Respecting the Great Lakes Basin Ecosystem (COA), have been developing and evaluating new
stormwater control technologies and sewage treatment techniques to resolve water quality
problems (Environment Canada,  2004). Under the new COA, Canada  and Ontario will continue
to build on this work, implementing efficient and cost effective projects to reduce the
environmental damage of a rapidly expanding urban population (Environment Canada, 2004).

Municipal wastewater effluent (MWWE) is currently managed through a variety of policies, by-
laws and legislation at the federal, provincial/territorial and municipal levels  (CCME, 2006). This
current variety of policies unfortunately creates confusion and complex situations for regulators,
system owners and operators.  As a result, the Canadian Council of Ministers of the Environment
(CCME) has established a Development Committee to develop a Canada-wide  Strategy for the
management of MWWE by November 2006. An integral part of the strategy's development will
be to consult with a wide variety of stakeholders to ensure that management strategies  for
MWWE incorporate their interests, expertise and vision. The strategy will address a number of
governance and technical issues,  resulting in a harmonized management approach (CCME, 2006).

The presence of emerging chemicals of concern in wastewater effluent is another developing
issue that requires attention. Current U.S. State and municipality permit requirements are based
                         Draft for Discussion at SOLEC 2006

-------
on state water quality laws that are developed according to pollutants anticipated to exist in the
community. This is also true for the WTP in Ontario. This means the existence of new
potentially toxic substances can be overlooked.  So even in areas with a high degree of municipal
wastewater treatment, pollutants such as endocrine-disrupting substances can inadvertently pass
through wastewater treatment systems and into the environment. These substances are known to
disrupt or mimic naturally occurring hormones and may have an impact on the growth,
reproduction, and development of many species of wildlife. Additional monitoring for these
pollutants and corresponding protection and regulation measures need to be investigated and
implemented.

The methodologies used in the U.S. national assessments of wastewater treatment could
potentially be reproduced and used to detect loadings trends and performance measures for
additional pollutants in the Great Lakes. The QA/QC safeguards included in such methods could
lead to very useful analyses of watershed-based point source  controls. Sufficient resources in
terms of time and funding need to be allocated in order to accomplish this task.

Comments from  the author(s)
The actual proportion of the entire population receiving treatment can not be calculated until a
definite population for the Great Lakes basin can be obtained for the same time period. Several
different population estimates exist for the region, but they were compiled according to county in
the U.S., and therefore represent a skewed total for the population that actually resides within the
boundaries of the  Great Lakes watershed.  GIS analysis of census data needs to be completed in
order to obtain a more accurate value for the Great Lakes population.

In Canada, only one year was assessed due to lack of available data.  In future years, data from
the Environment Canada MWWS survey would be useful to use, but the survey is currently only
updated to 1999, which unfotunately would not be useful for this report. The newest survey will
be out within the next year and it should be examined in future assessments for this indicator.

Several problems  exist in the calculation of effluent loadings. For example, actual flow through
effluent is not consistently monitored for in the U.S. Although influent levels are obtainable for
every facility, there is no way to ensure that the  effluent is comparable, since a substantial volume
may be removed during treatment processes.  Since effluent flow is sometimes necessary to
calculate loadings from concentration values of pollutants, precise estimates of total loadings to
Great Lakes waters may be next to impossible to obtain on a  large scale.

Another future effort towards the implementation of this indicator would be to use a consistent
guideline when analyzing wastewater treatment in both the U.S. and Canada. In the U.S. portion
of the basin, data were compiled from several different databases, with population information
derived from a separate source than effluent monitoring reports.  For Ontario, data from randomly
chosen municipalities serving a population of 10,000 or greater were available  for analysis.
Focusing on this criterion for wastewater treatment can  only provide a fragmented view of the
treatment patterns in the Canadian Great Lakes basin; however, by using a consistent wastewater
treatment analysis guideline, bias results would be avoided.
                         Draft for Discussion at SOLEC 2006

-------


Furthermore, a more organized monitoring program must be implemented in order to successfully
correlate wastewater treatment quality with the status of the Great Lakes basin. Although the
wastewater treatment plants provide useful monitoring information regarding the quality of
wastewater, they only state the quality of that specific municipality, rather than the overall quality
of the great lakes. Implementation of a more standardized, updated approach to monitoring
contaminants in effluent and reporting data for wastewater treatment is needed to address this
issue. Additionally, the difference in monitoring requirements between Canada and the U.S.
make it difficult to assess the quality of wastewater treatment on a basin-wide scale. A
standardized reporting format and inclusive database, accessible to all municipalities, researchers,
and the general public, should be established for binational use. This would make trend analysis
easier, and thus provide a more effective assessment of the potential health hazards associated
with wastewater treatment for the Great Lakes as a whole.

Considering all the difficulties encountered while attempting to adequately summarize the  vast
amount of U.S. effluent monitoring data contained in the PCS database, the logical solution
would be to request an application that could automate accurate calculations. Interestingly, such
an application previously existed that was capable of producing effluent data mass loadings
reports from the PCS database, but it was discontinued due to the modernization of the PCS
system that is currently underway. While the PCS system is being updated, adequate resources
have not been available to extend this overhaul to the previously mentioned application as  of yet,
and the lack of substantial use of the application in the past raised concern over its cost-
effectiveness. Additionally, incorporating this component into the current modernization could
take years due to various logistical problems, including the inherent quality assurance controls
needed for PCS metadata before potential loadings results could be accepted as reliable, high
quality data (personal communication with James Coleman, 2006).Despite these problems, the
reinstatement of such a tool would solve the data summarization needs presented in this indicator
report and could lead to an effective, comprehensive, and time-efficient analysis of pollutant
loadings to the Great Lakes from wastewater treatment plants.

Acknowledgments
Authors: Chiara Zuccarino-Crowe, Environmental Careers Organization Associate, on
appointment to U.S. Environmental Protection Agency, Great Lakes National Program Office,
Chicago, IL;
Tracie Greenberg, Environment Canada Intern, Burlington, ON

Contributors
James Coleman, U.S. EPA, Region 5 Water Division, Water Enforcement and Compliance
Assurance Branch
Paul Bertram, U.S. EPA, Great Lakes National Program Office
Sreedevi Yedavalli, U.S. EPA, Region 5 Water Division, NPDES Support and Technical
Assistance Branch

Data Sources
2000 Clean Watershed Needs Survey data was supplied in 2006 by William Tansey, U.S. EPA,
and was compiled for the Great Lakes basin by Tetra Tech, Inc.
                         Draft for Discussion at SOLEC 2006

-------
Canadian Council of Ministers of the Environment. 2006. Municipal Wastewater Effluent. Last
accessed September 7, 2006 from: http://www.ccme.ca/initiatives/water.html?categorv  id=81

City of Hamilton. 2006. Woodward Wastewater Treatment Plant Report 2005 Annual Report.
Woodward Wastewater Treatment Plant, Hamilton, Ontario.

City of Toronto. 2006. Ashbridges Bay Treatment Plant 2005 Summary. Toronto, Ontario.

City of Toronto. 2006. Highland Creek Wastewater Treatment Plant 2005 Summary. Toronto,
Ontario.

City of Toronto. 2006. Number Wastewater Treatment Plant 2005 Summary. Toronto, Ontario.

City of Sault Ste Marie. 2006. East End Water Pollution Control Plant 2005 Annual Report.
Sault Ste Marie, Ontario.

City of Sault Ste Marie. 2006. West End Water Pollution Control Plant 2005 Annual Report.
Sault Ste Marie, Ontario.

City of Windsor. 2006. Little River  Water Pollution Control Plant 2005 Annual Report. Windsor,
Ontario.

City of Windsor. 2006. Lou Romano Water Reclamation Plant 2005 Annual Report. Windsor,
Ontario.

County of Prince Edward. 2006. Picton Water Pollution Control Plant - Monitoring and
Compliance Report 2005. The corporation of the country of Prince Edward, Belleville, Ontario.

County of Prince Edward. 2006. Wellington Water Pollution Control Plant - Monitoring and
Compliance Report 2005. The corporation of the country of Prince Edward, Belleville, Ontario.

Environment Canada.  2004. Guidelines for Effluent Quality and Wastewater Treatment at
Federal Establishments. Last accessed September 5, 2006 from:
http://www.ec.gc. ca/etad/default.asp?lang=En&n=023194F5-l#general

Environment Canada.  2001. The State of Municipal Wastewater Effluents in Canada. Last
Accessed August 31, 2006 from: http://www.ec.gc.ca/soer-ree/English/soer/MWWE.pdf

Government of Canada. 2002. Municipal Water Issues in Canada. Last accessed September 14,
2006 from: http://dsp-psd.pwgsc.gc.ca/Collection-R/LoPBdP/BP/bp333-e.htmtfTREATING

Halton Region. 2006. Acton WWTP Performance Report, 2005. Regional Municipality of Halton,
Halton, Ontario.

Halton Region. 2006. Skyway WWTP Performance Report, 2005. Regional Municipality of
Halton, Halton Ontario.
                        Draft for Discussion at SOLEC 2006

-------
Halton Region. 2006. Georgetown WWTP Performance Report, 2005. Regional Municipality of
Halton, Halton Ontario.

Halton Region. 2006. Milton WWTP Performance Report, 2005. Regional Municipality of
Halton, Halton Ontario.

Halton Region. 2006. Mid-Halton WWTP Performance Report, 2005. Regional Municipality of
Halton, Halton Ontario.

Halton Region. 2006. Oakvitte South East WWTP Performance Report, 2005. Regional
Municipality of Halton, Halton Ontario.

Halton Region. 2006. Oakville South West WWTP Performance Report, 2005. Regional
Municipality of Halton, Halton Ontario.

PCS data supplied by James Coleman, Information Management Specialist, U.S. EPA, Region 5
Water Division, Water Enforcement and Compliance Assurance Branch.

Peel Region.  2006. Clarhson Compliance Report 2005. Mississauga, Ontario.

Peel Region.  2006. Lakeview Compliance Report 2005. Mississauga, Ontario.

Region of Durham. 2006. Corbett Creek Wastewater Treatment Plant Operational Data 2005.
Town of Whitby, Ontario.

Region of Durham. 2006. Duffin Creek Wastewater Treatment Plant Operational Data 2005.
Town of Whitby, Ontario.

Region of Durham. 2006. Newcastle Creek Wastewater Treatment Plant Operational Data 2005.
Town of Whitby, Ontario.

Region of Durham. 2006. Port Darlington Wastewater Treatment Plant Operational Data 2005.
Town of Whitby, Ontario.

Region of Durham. 2006. Harmony Creek Wastewater Treatment Plant Operational Data 2005.
Town of Whitby, Ontario.

U.S. EPA.  1998.  Wastewater Primer. U.S. EPA, Office of Water.
http://www.eap. gov/owm/

U.S. EPA.  2000.  Progress in Water Quality: An Evaluation of the National Investment in
Municipal Wastewater Treatment. U.S. Environmental Protection Agency, Washington, DC.
EPA-832-R-00-008.
                       Draft for Discussion at SOLEC 2006                      11

-------
U.S. EPA. 2004. Primer for Municipal Wastewater Treatment Systems. U.S. EPA, Office of
Water and Office of Wastewater Management, Washington, DC.  EPA 832-R-04-001.

U.S. EPA. "Compliance and Enforcement Water Data Systems." Data, Planning and Results.
July 03, 2006.  U.S. EPA, Office of Enforcement and Compliance Assurance.
http://www.epa.gov/compliance/data/svstems/index.html (Accessed September 27, 2006).

U.S. EPA. "Enforcement & Compliance History Online (ECHO)." Compliance and
Enforcement. September 2006. U.S. EPA, Office of Enforcement and Compliance Assurance.
http://www.epa.gov/echo/index.html (Accessed September27, 2006).

U.S. EPA. "NPDES Permit Program Basics." National Pollutant Discharge Elimination System
(NPDES). July 05, 2006. U.S. EPA, Office of Wastewater Management.
http://cfpub.epa.gov/npdes/index.cfm (Accessed July 25, 2006).

List of Tables
Table 1. Canadian Pollutant Effluent Limits
Source:  Environment Canada,  2004 http://www.ec.gc.ca/etad/default.asp?lang=En&n=023194F5-
l#general

List of Figures
Figure 1. Population served by Publicly Owned Treatment Works (POTWs) by treatment level in
the U.S. Great Lakes basin
Caption: (a)= "No discharge" facilities do not discharge treated wastewater to the Nation's
waterways. These facilities dispose of wastewater via methods such as industrial re-use,
irrigation, or evaporation.
* Lake St. Clair and Detroit River watersheds are also considered part of the Lake Erie basin
** MI Unknown refers to the population served by facilities in the state of Michigan for which
exact watershed locations are unknown, so the data could not be grouped with a specific lake
basin. Population could potentially be distributed between the Lakes Michigan, Huron, or Erie.
Source:  2000 Clean Watershed Needs Survey

Figure 2. Percent of Population Served in Canada by Each Treatment Type in 1999.
Source:  Municipal Water Use Database Web site:
(http://www.ec.gc.ca/water/en/manage/use/e_data.htm)

Figure 3. Total number of quarters with reported effluent limit violations by pollutant for
selected U.S. facilities
Caption: Data was compiled from 15 different facilities according to the total number of quarters
that were in non-compliance of at least one pollutant effluent limit permit during 2003-2006.
* = combination of violations for 5-day BOD listed as total % removal and total
** = combination of violations for fecal coliform listed as general and broth totals
*** = combination of violations for cyanide listed as A and CN totals
**** _ combination of violations for total nitrogen listed as N and as NH3
***** = combination of violations for solids as listed as total settleable, total dissolved, total
suspended, and suspended % removal
                         Draft for Discussion at SOLEC 2006

-------

Source:  U.S. EPA. "Enforcement & Compliance History Online (ECHO)."  Compliance
and Enforcement.  September 2006. U.S. EPA, Office of Enforcement and Compliance
Assurance,  http://www.epa.gov/echo/index.html (Accessed September27, 2006).

Last updated
SOLEC 2006
Pollutant Effluent
5 day Biochem Biochemical Oxygen Demand
Suspended Solids
Fecal Coliforms
Chlorine Residual
PH
Phenols
Oils & Greases
Phosphorus (Total P)
Temperature
Limit
20 mg/L
25 mg/L
400 per 1 00 mL (after disinfection)
0. 50 mg/L minimum after 30 minutes contact
time
6 to 9
20 micrograms/L
15 mg/L
1 .0 mg/L
Not to alter the ambient water temperature
by more than one degree Centigrade (1°C).
Table 1. Canadian Pollutant Effluent Limits
Source: Environment Canada, 2004 http://www.ec.gc.ca/etad/default.asp?lang=En&n=023194F5-
l#general
                      Draft for Discussion at SOLEC 2006

-------
                                   State of the Great Lakes 2007 - Draft
               Population served by POTWs by treatment level in the U.S. Great Lakes basin
                DNo Discharge (a)
               I Secondary
D Greater than Secondary
   6,000,000
   5,000,000
SJ 4,000,000

(/}
O 3,000,000
3
§• 2,000,000
0.
   1,000,000
           Lake Superior  Lake Michigan   Lake Huron
               Lake St.
              Clair/Detroit
                River*
                                                    Lake Erie
                                                              Lake Ontario   St. Lawrence  Ml Unknown*'
                                          Lake/River Basin
  Figure 1. Population served by Publicly Owned Treatment Works (POTWs) by treatment level
  in the U.S. Great Lakes basin
  Caption: (a)= "No discharge" facilities do not discharge treated wastewater to the Nation's
  waterways.  These facilities dispose of wastewater via methods such as industrial re-use,
  irrigation, or evaporation.
  * Lake St. Clair and Detroit River watersheds are also considered part of the Lake Erie basin
  ** MI Unknown refers to the population served by facilities in the state of Michigan for which
  exact watershed locations are unknown, so the data could not be grouped with a specific lake
  basin.  Population could potentially be distributed between the Lakes Michigan, Huron, or Erie.
  Source:  2000 Clean Watershed Needs Survey
  14
Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
             Percent of Population Served by Each Treatment Type (in
                                       1999)
       D primary
       • stabilizing ponds
       • secondary
       • tertiary
                         71.31%
                                                                           21.26%
Figure 2. Percent of Population Served in Canada by Each Treatment Type in 1999.
Source: Municipal Water Use Database Web site:
(http://www.ec.gc. ca/water/en/manage/use/e_data.htm)
        D significant non-compliance with effluent limits
[general limit violations
             o
                                         Pollutant

Figure 3. Total number of quarters with reported effluent limit violations by pollutant for
selected U.S. facilities
Caption: Data was compiled from 15 different facilities according to the total number of quarters
that were in non-compliance of at least one pollutant effluent limit permit during 2003-2006.
                        Draft for Discussion at SOLEC 2006
                        15

-------
* = combination of violations for 5-day BOD listed as total % removal and total
** = combination of violations for fecal coliform listed as general and broth totals
*** = combination of violations for cyanide listed as A and CN totals
**** _ combination of violations for total nitrogen listed as N and as NH3
***** = combination of violations for solids as listed as total settleable, total dissolved, total
suspended, and suspended % removal
Source: U.S. EPA.  "Enforcement & Compliance History Online (ECHO)."  Compliance
and Enforcement. September 2006. U.S. EPA, Office of Enforcement and Compliance
Assurance, http://www.epa.gov/echo/index.html (Accessed September27, 2006).
                        Draft for Discussion at SOLEC 2006

-------
Natural Groundwater Quality and Human-
Induced Changes
Indicator #7100

Assessment: Not Assessed
Note: This indicator report uses data from the Grand River
watershed only and may not be representative of groundwater
conditions throughout the Great Lakes basin.

Purpose
  To measure  groundwater quality as determined by the natural
chemistry of the bedrock and overburden deposits, as well as
any changes in quality due to anthropogenic activities; and
  To address groundwater quality impairments, whether they
are natural or human induced in order to ensure a safe and
clean supply of groundwater for human consumption and
ecosystem functioning.

Ecosystem Objective
The ecosystem objective for this indicator is to ensure that
groundwater quality remains at or approaches natural condi-
tions.

State of the Ecosystem
Background
Natural groundwater quality issues and human induced changes
in groundwater quality both have the potential to affect our
ability to use groundwater safely. Some constituents found nat-
urally in groundwater renders some groundwater reserves inap-
propriate for certain uses. Growing urban populations, along
with historical and present industrial and agricultural activity,
have caused significant harm to groundwater quality, thereby
obstructing the use of the resource and damaging the environ-
ment. Understanding natural groundwater quality provides a
baseline from which to compare, while monitoring anthro-
pogenic changes can allow identification of temporal trends and
assess any improvements or further degradation in quality.

Natural Groundwater Quality
The Grand River watershed can generally be divided into three
distinct geological areas; the northern till plain, the central
region of moraines with complex sequences of glacial,
glaciofluvial and glaciolacustrine deposits, and the southern
clay plain. These surficial overburden deposits are underlain by
fractured carbonate rock (predominantly dolostone). The
groundwater resources of the watershed include regional-scale
unconfined and confined overburden and bedrock aquifers as
well as discontinuous local-scale deposits  which contain suffi-
cient groundwater to meet smaller users needs. In some areas of
the watershed  (e.g. Whitemans Creek basin) the presence of high
permeability sands at ground surface and or a high water table
leads to unconfined aquifers which are highly susceptible to
  degradation from surface contaminant sources.

  The natural quality of groundwater in the watershed for the most
  part is very good. The groundwater chemistry in both the over-
  burden and bedrock aquifers is generally high in dissolved inor-
  ganic constituents (predominantly calcium, magnesium, sodium,
  chloride and sulphate). Measurements of total dissolved solids
  (TDS) suggest relatively "hard" water throughout the watershed.
  For example, City of Guelph production wells yield water with
  hardness measured from 249 mg/1 to 579 mg/1, which far
  exceeds the aesthetic Ontario Drinking Water Objective of 80
  mg/1 to 100 mg/1. Elevated concentrations of trace metals (iron
  and manganese) have also been identified as ambient quality
  issues with the groundwater resource.
                                          Ambienl Wa»r Quality IBUSC
                                          • Salt,
                                          A Sulphm
                                          • Mural
                                          * Gas
                                            DUNDEE
                                            DNONDAOA - AMHERSTBURG
                                          •1 BOIS BLANC
                                            ORISKANY
                                            BUSS ISLANDS - BERTIE
                                            SAUNA
                                            EUELPH
                                            LOCKPORT - AMABEL
B                                            CLINTON • CATARACT GROUP
                                            MANITOULIN
                                            OUEENSTON
                   Kilometres
Figure 1. Bedrock wells with natural quality issues in the Grand
River watershed.
Source: Grand River Conservation Authority
  Figures 1 and 2 illustrate water quality problems observed in
  bedrock and overburden wells, respectively. These figures are
                                                                                                                     241

-------
  based on a qualitative assessment of well water at the time of
  drilling as noted on the Ontario Ministry of Environment's water
  well record form. The majority of these wells were installed for
  domestic or livestock uses. Overall, between 1940 and 2000, less
  than 1% (approximately 1131 wells) of all the wells drilled in
  the watershed reported having a water quality problem. Of the
  wells  exhibiting a natural groundwater problem about 90% were
  bedrock wells while the other 10% were completed in the over-
  burden. The most frequently noted quality problem associated
  with bedrock wells was high sulphur content (76% of bedrock
  wells  with quality problems). This is not surprising, as sulphur is
  easy to detect due to  its distinctive and objectionable odour.
  Generally, three bedrock formations commonly intersected with-
  in the watershed contain most of the sulphur wells: the Guelph
  Formation, the Salina Formation, and the Onondaga-
  Amherstburg Formation. The Salina Formation forms the shal-
  low bedrock under the west side of the watershed while the
                                         Ambient Water Quality Issues
                                          •  Salty
                                          A  Sulphur
                                          •  Mineral
                                          •  Gas
                                         Generalized Surficial Geology
                                         m Bedrock
                                             Clay
                                             Gravel
                                             Organic
                                             Sand
                                             Sandy Till
                                         M Silty Till
                                           • Water
                  Kilometres
Figure 2. Overburden wells with natural quality issues in the Grand
River watershed.
Source: Grand River Conservation Authority
Guelph underlies the east side of the watershed.

Additional quality concerns noted in the water well records
include high mineral content and salt. About 20% of the reported
quality concerns in bedrock wells were high mineral content
while 4% reported salty water. Similar concerns were noted in
overburden wells where reported problems were sulphur (42%),
mineral (34%), and salt (23%).

Human Induced Changes to Groundwater Quality
Changes to the quality of groundwater from anthropogenic activ-
ities associated with urban sprawl, agriculture and industrial
operations have been noted throughout the watershed. Urban
areas within the Grand River watershed have been experiencing
considerable growth over the past few decades. The groundwater
quality issues associated with human activity in the watershed
include: chloride, industrial chemicals (e.g. trichloroethylene
(TCE)), and agricultural impacts (nitrate, bacteria, and pesti-
cides). These contaminants vary in their extent from very local
impact (e.g. bacteria) to widespread impact (e.g. chloride).
Industrial contaminants tend to be point sources, which general-
ly require very little concentration to impact significant ground-
water resources.

Chloride
Increasing chloride concentrations in groundwater have been
observed in most municipal wells in the urban portions of the
watershed. This increase has been attributed to winter deicing of
roads with sodium chloride (salt). Detailed studies carried out by
the Regional Municipality of Waterloo have illustrated the
impact of road salting associated with increased urban develop-
ment to groundwater captured by two municipal well  fields.
Figure 3 shows the temporal changes in chloride concentration
for the two well fields investigated in this study. Wells A, B, and
C, are from the first well field while wells D and E are from the
second well field. In 1967 land use within the capture zone of
the first field was 51% rural and 49% urban, while in the second
well field capture zone the land use was 94% rural and 6%
urban. By 1998, the area within the first well field capture zone
had been completely converted to urban land while in the  second
well field capture zone 60% of the land remained rural.

Although wells from both well fields show increased chloride
levels, wells A, B, and C  in the heavily urbanized capture  zone
show a greater increase in chloride concentrations than do wells
D and E in the predominantly rural capture zone. For  example,
well B showed a change in chloride concentration from 16.8
mg/1 in 1960, to 260 mg/1 in 1996, where as well D showed a
change from 3 mg/1 in 1966, to  60 mg/1 in  1996. This indicates
that chloride levels in groundwater can be linked to urban
growth and its associated land uses (i.e. denser road network).
The Ontario Drinking Water Objective for chloride had been
  242

-------
                                                      Wells
                                                      -•-A
                                                      -•-B
                                                      -*-C
                                                      -*-D
                                                      -•-E
        1960
                  1970
                            1980
                            Year
                                      1990
                                               2000
Figure 3. Chloride levels in selected groundwater wells in the
Regional Municipality of Waterloo. Red indicates wells from one
area/well field. Green indicates wells from a different area/well
field.
Source: Stanley Consulting, 1998
   established at 250 mg/1, although this guideline is predominantly
   for aesthetic reasons, the issue of increasing chloride levels
   should be addressed.

   Industrial Contaminants
   Groundwater resources in both the overburden and bedrock
   deposits within the Grand River watershed have been impacted
   by contamination of aqueous and non-aqueous contaminants
   which have entered the groundwater as a result of industrial
   spills or discharges, landfill leachates, leaky storage containers,
   and poor disposal practices. A significant number of these chem-
   icals are volatile organic compounds (VOCs). Contamination by
VOCs such as TCE, have impacted municipal groundwater sup-
plies in several communities in the watershed. For example, by
the year 1998, five of the City of Guelph's 24 wells were taken
out of service due to low-level VOC contamination. These wells
have a combined capacity of 10,000 to 12,000 mVday and repre-
sent about  15% of the City's permitted water-taking capacity. As
a second example, contamination of both a shallow aquifer and a
deeper municipal aquifer with a variety of industrial chemicals
(including toluene, chlorobenzene, 2,4-D, 2,4,5-T) emanating
from a chemical plant in the Region of Waterloo led to the
removal of municipal wells from the water system in  the town of
Elmira.

Agricultural and Rural Impacts
Groundwater quality in agricultural areas is affected by activities
such as pesticide application, fertilizer and manure applications
on fields, storage and disposal of animal wastes and the improp-
er disposal and spills of chemicals. The groundwater contami-
nants from these activities can be divided into three main
groups: nitrate, bacteria and pesticides. For example, the applica-
tion of excessive quantities of nutrients to agricultural land may
impact the  quality of the groundwater. Excess nitrogen applied
to the soil to sustain crop production is converted to nitrate with
infiltrating water and hence transported to the water table.
Seventy-six percent of the total land area in the Grand River
watershed is used for agricultural purposes and thus potential
and historical contamination of the groundwater due to these
activities is a concern.

Land use and nitrate levels measured in surface water from two
sub-watersheds, the Eramosa River and Whitemans Creek, are
                                                                      Eramosa River
                                                                       Sub-Basin
                             Whitemans Creek

                                  Other
                                        Urban and
                                      /-Developed
                                                                                               Eramosa River
          Figure 4. Land cover on moraine systems and areas that facilitate high to very high groundwater recharge of the
          Whitemans Creek and Eramosa River sub-watersheds: (a) Spatial distribution and (b) Percent distribution of classi-
          fied land use.
          Source: Grand River Conservation Authority
                                                                                                                         243

-------
used to illustrate the effects of agricultural activities on ground-
water quality and the quality of surface water.

In the Whitemans Creek sub-watershed, approximately 78% of
the land classified as groundwater recharge area is covered with
agricultural uses, and only 20% is forested. In the Eramosa sub-
watershed about 60% of the significant recharge land is used for
agricultural purposes with approximately 34% of the land being
covered with forest (Figure 4). Both of these tributary streams
are considered predominantly groundwater-fed streams, meaning
that the majority of flow within them is received directly from
groundwater discharge.

Average annual concentrations of nitrate measured in the
Eramosa River and Whitemans Creek from 1997 to 2003 are
shown in Figure 5. Average annual concentration of nitrate
measured in Whitemans Creek between 1997 and 2003 were 2.5
to 8 times higher than those measured in the Eramosa River. The
higher nitrate levels measured in  Whitemans Creek illustrate the
linkage between increased agricultural activity and groundwater
contamination and its impact on surface water quality. In addi-
tion to the agricultural practices in the Whitemans Creek sub-
watershed, the observed nitrate concentrations may also be
linked to rural communities with a high density of septic sys-
tems that leach nutrients to the subsurface.
Manure spreading on fields, runoff from waste disposal sites.

"5> 19
I
c m
.1
'•&
S. ft -
*J
§ 6
E
O
O 4
0)
*j
2 2
*j £-
z
0










T
J
M\
1997


J
19

dn
98


jl
-



|
-


.p
i--i r
T
P

1 rTi
I
f

! 1999 2000 2001 2002 2003
Year
ED Eramosa



River D Whitemans Cree


k


Figure 5. Average annual concentrations of nitrate measured in the
Eramosa River and Whitemans Creek from 1997 to 2003. (Also shown
on the bar graphs is the standard error of measurement)
Source: Ontario Provincial Water Quality Monitoring Network, 2003.
and septic systems may all provide a source of bacteria to
groundwater. Bacterial contamination in wells in agricultural
areas is common, however, this is often due to poor well con-
struction allowing surface water to enter the well and not indica-
tive of widespread aquifer contamination. Shallow wells are par-
ticularly vulnerable to bacterial contamination.

Pressures
The population within the Grand River watershed is expected to
increase by over 300,000 people in the next 20 years. The urban
sprawl and industrial development associated with this popula-
tion growth, if not managed appropriately, will increase the
chance for contamination of groundwater resources.
Intensification of agriculture will lead to increased potential for
pollution caused by nutrients, pathogens and pesticides to enter
the groundwater supply and eventually surface water resources.
While largely unknown at this time, the effects of climate
change may lead to decreased groundwater resources, which
may concentrate existing contaminant sources.

Management Implications
Protecting groundwater resources generally requires multi-
faceted strategies including regulation, land use planning, water
resources management, voluntary adoption of best management
practices and public education. Programs to reduce the amount
of road salt used for deicing will lead to reductions in chloride
    contamination in groundwater. For example, the Regional of
    Waterloo (the largest urban community in the watershed) in
    cooperation with road maintenance departments has been
    able to decrease the amount of road salt applied to Regional
    roads by 27% in just one winter season.

    Acknowledgments
    Authors: Alan Sawyer, Grand River Conservation Authority.
    Cambridge, ON;
    Sandra Cooke, Grand River Conservation Authority.
    Cambridge, ON;
    Jeff Pitcher, Grand River Conservation Authority.
    Cambridge, ON;  and
    Pat Lapcevic, Grand River Conservation Authority.
    Cambridge, ON.

    Alan Sawyer's position was partially funded through a grant
    from Environment Canada's Science Horizons internship
    program. The assistance of Samuel Bellamy of the Grand
    River Conservation Authority,  as well as Harvey Shear.
    Nancy Stadler-Salt and Andrew Piggott of Environment
    Canada is gratefully acknowledged.
                                                                  Sources
                                                                  Braun Consulting Engineers, Gartner Lee Limited, and
244

-------
Jagger Hims Limited Consulting Engineers. 1999. City of
Guelph Water System Study Resource Evaluation Summary.
Report prepared for the City of Guelph.

Crowe, A.S., Schaefer, K.A., Kohut, A.,  Shikaze, S.G., and
Ptacek, CJ. 2003. Groundwater quality.  Canadian Council of
Ministers of the Environment, Winnipeg, Manitoba. Canadian
Council of Ministers of the Environment (CCME), Linking
Water Science to Policy Workshop Series. Report No. 2, 52pp.

Holysh, S., Pitcher, J., and Boyd, D. 2001. Grand River regional
groundwater study. Grand River Conservation Authority,
Cambridge, ON, 78pp+ appendices.

Ontario Provincial Water Quality Monitoring Network. 2003.
Grand  River Conservation Authority Water Quality Stations.

Region of Waterloo. Official Municipal Website.
http://www.region.waterloo.on.ca.

Stanley Consulting. 1998. Chloride Impact Assessment Parkway
and Strasburg Creek Well Fields Final Report. Prepared for the
Regional Municipality of Waterloo.

Whiffin, R.B., and Rush, RJ.  1989. Development and demon-
stration of an integrated approach to aquifer remediation at an
organic chemical plant. In Proceedings of the FOCUS
Conference on Eastern Regional Ground Water Issues, October
17-19,  1989, Kitchener, ON, Canada, pp. 273-288.
Authors' Commentary
While there is a large quantity of groundwater quality data avail-
able for the various aquifers in the watershed, this data has not
been consolidated and evaluated in a comprehensive or system-
atic way. Work is needed to bring together this data and incorpo-
rate ongoing groundwater monitoring programs. An assessment
of the groundwater quality across Ontario is currently being
undertaken through sampling and analysis of groundwater from
the provincial groundwater-monitoring network  (PGMN) wells
(includes monitoring stations in the Grand River watershed).
Numerous watershed municipalities also have had ongoing mon-
itoring programs, which examine the quality of groundwater as a
source of drinking water in place for a number of years.
Integrating this data along with data contained in various site
investigations will allow for a more comprehensive picture of
groundwater quality in the watershed.

Last Updated
State of the Great Lakes 2005
                                                                                                                    245

-------
                                              OF   THE   GREAT
                                 2007
Groundwater and Land: Use and Intensity
Indicator #7101

Assessment:  Not Assessed
Note: This indicator report uses data from the Grand River
watershed only and may not be representative of groundwater
conditions throughout the Great Lakes basin.

Purpose
  To measure water use and intensity and land use and intensity;
  To infer the potential impact of land and water use on the
quantity and quality of groundwater resources as well as evalu-
ate groundwater  supply and demand; and
  To track the main influences on groundwater quantity and
quality such as land and water use to ensure sustainable high
quality groundwater supplies.

Ecosystem Objective
The  ecosystem objective for this indicator is to ensure that land
and water use does not negatively impact groundwater
supplies/resources.

State of the Ecosystem
Background
Land use and intensity has the potential to affect both groundwa-
    ter quality and quantity. Similarly, water use and intensity (i.e.
    demand) can impact the sustainability of groundwater supplies.
    In addition, groundwater use and intensity can impact streams
    and creeks, which depend on groundwater for base flows to sus-
    tain aquatic plant and animal communities.

    Land use and intensity
    The Grand River watershed can generally be divided into three
    distinct geological areas; the northern till plain, central moraines
    with complex sequences of glacial, glaciofluvial and glaciolacus-
    trine deposits, and the southern clay plain. These surficial over-
    burden deposits are underlain by fractured carbonate rock (pre-
    dominantly dolostone). The groundwater resources of the water-
    shed include regional-scale unconfined and confined overburden
    and bedrock aquifers as well as discontinuous local-scale
    deposits which contain sufficient groundwater to meet smaller
    users' needs. In some areas of the watershed (e.g. Whiteman's
    Creek basin) the presence of high permeability sands at ground
    surface and/or a high water table leads to unconfined aquifers
    which are highly susceptible to contamination from surface con-
    taminant sources.

    Agricultural and rural land uses predominate in the Grand River
    watershed. Approximately 76% of the watershed land area is
    used for agriculture (Figure 1).  Urban development covers about
   A
B
                                                                    Urban and
                                                                    Developed
                                                                       5%
                                                             Other
                                                        (e.g. golf courses)
                                                              Open Water and
                                                                 Wetland
                                                                   2%
                                                                       Forested
                                                                         17%
                                                                                                                Agricultural
                                                                                                                   76%
                                                        Figure 1. Land cover in the Grand River watershed: (a) Spatial distri-
                                                        bution and (b) Percent distribution of classified land use.
                                                        Source: Grand River Conservation Authority
246

-------
5% of the watershed area while forests cover about 17%. The
largest urban centres, including Kitchener, Waterloo, Cambridge
and Guelph, are located in the central portion of the watershed
and are situated on or in close proximity to many of the complex
moraine systems that stretch across the watershed (Figure 1).
The moraines and associated glacial outwash area in the water-
shed form a complex system of sand and gravel layers separated
by less permeable till layers. Together with the sand plain in the
southwest portion of the watershed these units provide signifi-
cant groundwater resources.  The majority of the groundwater
recharge in the watershed is concentrated in a land area that cov-
ers approximately 38% of the watershed. Figure 2 illustrates the
land cover associated with those areas that have high recharge
potential.

Land use on these moraines and significant recharge areas can
have major influence on both groundwater quantity and quality
(Figure 2). Intensive cropping practices with repeated manure
and fertilizer applications have the potential to impact ground-
water quality while urban development can interrupt groundwa-
ter recharge and impact groundwater quantity. About 67% of the
significant recharge areas are in agricultural production while
23% and 8% of the recharge areas are covered with forests and
urban development respectively. Since the moraine systems and
recharge areas in the Grand River watershed provide important
       ecological, sociological and economical services to the water-
       shed, they are important watershed features that must be main-
       tained to ensure sustainable groundwater supplies.

       Land use  directly influences the ability of precipitation to
       recharge shallow aquifers. Urban development such as the
       paving of roads and building of structures intercepts precipita-
       tion and facilitates the movement of water off the land in surface
       runoff, which subsequently reduces groundwater recharge of
       shallow aquifers. A significant portion (62%) of the urban area
       in the Grand River watershed tends to be concentrated in the
       highly sensitive groundwater recharge areas (Figure 3).
       Development is continuing in these sensitive areas. For example,
       of the total kilometres of new roads built between 2000 and
       2004 in the Region of Waterloo, about half of them were situated
       in the more sensitive areas.

       Land uses that protect groundwater recharge such as some agri-
       cultural land use and forested areas need to be protected to
       ensure groundwater recharge. About 34% and 51% of the water-
       shed's agricultural and forested land cover is located in the sig-
       nificant recharge areas.  Strategic development is needed to pro-
       tect these recharge areas to protect groundwater recharging func-
       tion in the watershed.
                Kilometres
                                                         B
                                                                   Urban and
                                                                   Developed-,
                                                                     8%    \
                                                           Other
                                                        e.g. golf courses^
                                                            1%
                                                          Open Water and  /
                                                             Wetland
                                                               2%
                                                                                                               Agricultural
                                                                                                                 67%
                                                                           Forested
                                                                             23%
Figure 2. Land cover on moraine systems and areas that facilitate high
or very high groundwater recharge of the Grand River watershed:
(a) Spatial distribution and (b) Percent distribution of classified land use.
Source: Grand River Conservation Authority

                                                               247

-------
                                              OF   THE   GREAT
                             2007

90



-------
Figure 6. Changes in amount of irrigated land in the Grand
River watershed (percentage of total watershed area irrigated).
Source: Statistics Canada data for 1986, 1991, and 1996
impact the quantity of groundwater supplies for watershed resi-
dents. Therefore, it is essential that municipalities and watershed
residents protect the moraine systems and significant recharge
areas to ensure future groundwater supplies.

Population growth with continued urban development and agri-
cultural intensification are the biggest threats to groundwater
supplies in the Grand River watershed. It is estimated that the
population of the watershed will increase by approximately
300,000 people in the next 20 years (Figure 8). The biggest sin-
gle users of groundwater are municipalities for municipal drink-
ing water supplies, although industrial users, including aggregate
and dewatering operations, use a significant amount of ground-
water. Municipalities, watershed residents and industries will
need to increase their efforts in water conservation as well as
continue to seek out new or alternate supplies.


-o 180° 1
a) -|600
1 1400 -

80 -s«
CO 9)
i|
~ 5i
ra
v>

-•— Percent Average Annual River Flow
Figure 7. Number of new wells drilled each year (bars). Annual
stream flow
(as a percentage on long term
watershed illustratin
(green line)

g below average, and


average
average) in the Grand River
average climatic conditions


Source: Ontario Ministry of the Environment Water Well Database,
2003





              1971
                       1981      1991
                             Year
                                        2001
                                                 2021
   Figure 8. Estimated population in the Grand River water-
   shed including future projections (burgundy bar).
   Source: Dorfman, 1997 and Grand River Conservation
   Authority, 2003
   Climate influence on groundwater resources in the Grand
   River watershed cannot be underestimated. It is evident that
   during times with below average precipitation, there is
   increased demand for groundwater resources for both the nat-
   ural environment and human uses. In addition, climate
   change will likely redistribute precipitation patterns through-
   out the year, which will likely impact groundwater resources
   in the watershed.

   Management Implications
   Land use and development has a direct effect on groundwater
   quantity and quality. Therefore, land use planning must con-
   sider watershed functions such as groundwater recharge when
   directing future growth. Municipal growth strategies should
   direct growth and development away from sensitive water-
   shed landscapes such as those areas that facilitate groundwa-
   ter recharge. Efforts in recent years have focussed on delin-
   eating wellhead protection zones, assessing the threats and
   understanding the regional hydrogeology. Through the plan-
   ning process, municipalities such as the Region of Waterloo.
   City of Guelph and the County of Wellington have recog-
nized the importance of protecting recharge to maintain ground-
water resources and have been taking steps to protect this water-
shed function. These initiatives include limiting the amount of
impervious cover in sensitive areas and capturing precipitation
with rooftop collection systems. By permitting development that
facilitates groundwater recharge or redirecting development to
landscapes that are not as sensitive, important watershed func-
tions can be protected to ensure future groundwater supplies.

Water conservation measures should be actively promoted and
adopted in all sectors of society. Urban communities must
actively reduce consumption while rural communities require
management plans to strategically irrigate using high efficiency
methods and appropriate timing.
                                                        249

-------
Acknowledgments
Authors: Alan Sawyer, Grand River Conservation Authority,
Cambridge, ON;
Sandra Cooke, Grand River Conservation Authority, Cambridge,
ON;
Jeff Pitcher Grand River Conservation Authority, Cambridge,
ON; and
Pat Lapcevic, Grand River Conservation Authority, Cambridge,
ON.

Alan Sawyer's position was partially funded through a grant
from Environment Canada's Science Horizons internship pro-
gram. The assistance of Samuel Bellamy of the Grand River
Conservation Authority, as well as Harvey Shear, Nancy Stadler-
Salt and Andrew Piggott of Environment Canada is gratefully
acknowledged.

Sources
Bellamy, S., and Boyd, D. 2004. Water use in the Grand River
watershed. Grand River Conservation Authority, Cambridge,
ON.

Dorfman, M.L., and Planner Inc. 1997. Grand River Watershed
Profile.  Prepared for the Grand River Conservation Authority.

Grand River Conservation Authority (GRCA). 2003. Watershed
Report.  Grand River Conservation Authority, Cambridge, ON.

Holysh, S., Pitcher, J., and Boyd, D. 2001. Grand River
Regional Groundwater Study. Grand River Conservation
Authority, Cambridge, ON.

Ontario Ministry of the Environment. 2003. Water Well
Information System Database. Ministry of Environment,
Toronto, ON.

Region  of Waterloo. Official Municipal Website.
http://www.region.waterloo.on.ca

Statistics Canada. Census of Agriculture. 1986, 1991, 1996.
Statistics Canada, Ottawa, ON.
Consistent and improved monitoring and data collection are
required to accurately estimate groundwater demand as well as
determine long-term trends in land use. For example, linking
groundwater permits to actual well log identification numbers
will assist with understanding the spatial distribution of ground-
water takings. Furthermore, groundwater permit holders should
be required to report actual water use as opposed to permitted
use. This will help estimate actual water use and therefore the
true impact on the groundwater system.

Last Updated
State of the Great Lakes 2005
Authors' Commentary
Understanding the impact of water use on the groundwater
resources in the watershed will require understanding the avail-
ability of water to allow sustainable human use while still main-
taining healthy ecosystems. Assessing groundwater availability
and use at appropriate scales is an important aspect of water bal-
ance calculations in the watershed. In other words, assessing
water and land use at the larger watershed scale masks more
local issues such as the impact of extensive irrigation.

250

-------

Base Flow Due to Groundwater Discharge
Indicator #7102
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Deteriorating
It is estimated that human activities have detrimentally impacted
groundwater discharge on at least a local scale in some areas of the
Great Lakes basin and that discharge is not significantly impaired in
other areas.
Lake-by-Lake Assessment
Lake Superior
           Status:  Not Assessed
           Trend:  Undetermined
Lake Michigan
           Status:
           Trend:
Not Assessed
Undetermined
Lake Huron
           Status:
           Trend:
Not Assessed
Undetermined
Lake Erie
           Status:
           Trend:
Not Assessed
Undetermined
Lake Ontario
           Status:
           Trend:
Not Assessed
Undetermined
Purpose
•To measure the contribution of base flow due to groundwater discharge to total stream flow; and
•To detect the impacts of anthropogenic factors on the quantity of the groundwater resource.

Ecosystem Objective
Base flow due to the discharge of groundwater to the rivers and inland lakes and wetlands of the
Great Lakes basin is a significant and often major component of stream flow, particularly during
low flow periods. Base flow frequently satisfies flow, level, and temperature requirements for
aquatic species and habitat. Water supplies and the capacity of surface water to assimilate
wastewater discharge are also dependent on base flow. Base flow due to groundwater discharge is
therefore critical to the maintenance of water quantity and quality and the integrity of aquatic
species and habitat.
                        Draft for Discussion at SOLEC 2006

-------
                                           f^^^M^ii^^MiSi^^^	
State of the Ecosystem
Background
A significant portion of precipitation over the inland portion of the Great Lakes basin returns to
the atmosphere by evapo-transpiration. Water that does not return to the atmosphere either flows
across the ground surface or infiltrates into the subsurface and recharges groundwater. Some of
this water is subsequently removed by consumptive uses such as irrigation and water bottling.
Water that flows across the ground surface discharges into surface water features (rivers, lakes,
and wetlands) and then flows toward and eventually into the Great Lakes. The component of
stream flow due to runoff from the ground surface is rapidly varying and transient, and results in
the peak discharges of a stream.

Water that infiltrates into the subsurface and recharges groundwater also results in flow toward
the Great Lakes. Most recharged groundwater flows at relatively shallow depths at local scales
and discharges into adjacent surface water features. However, groundwater also flows at greater
depths at regional scales and discharges either directly into the Great Lakes or into distant surface
water features. The quantities of groundwater flowing at these greater depths can be significant
locally but are generally believed to be modest relative to the quantities flowing at shallower
depths. Groundwater discharge to surface water features in response to precipitation is greatly
delayed relative to surface runoff. The stream flow resulting from groundwater discharge  is,
therefore, more uniform.

Base flow is the less variable and more persistent component of total stream flow. In the Great
Lakes region, groundwater discharge is often the dominant component of base flow; however,
various human and natural factors also contribute to base flow. Flow regulation, the storage and
delayed release of water using dams and reservoirs, creates a stream flow signature that is similar
to that of groundwater discharge. Lakes and wetlands also moderate stream flow, transforming
rapidly varying surface runoff into more slowly varying flow that approximates the dynamics of
groundwater discharge. It is important to note that these varying sources of base flow affect
surface water quality, particularly with regard to temperature. All groundwater discharge
contributes to base flow but not all base flow is the result of groundwater discharge.

Status of Base Flow
Base flow is frequently determined using a mathematical process known as hydrograph
separation.  This process uses stream flow monitoring information as input  and partitions the
observed flow into rapidly and slowly varying components, surface runoff and base flow,
respectively. The stream flow data that are used in these analyses are collected across the  Great
Lakes basin using networks of stream flow gauges that are operated by the United States
Geological Survey (USGS) and Environment Canada. Neff et al. (2005) summarize the
calculation and interpretation of base flow for 3,936 gauges in Ontario and the Great Lakes states
using six methods of hydrograph separation and length-of-record stream flow monitoring
information for the periods ending on December 31, 2000 and September 30, 2001, respectively.
The results reported by Neff et al. (2005) are the basis for the majority of this report. Results
corresponding to the UKIH method of hydrograph separation (Piggott et al. 2005) are referenced
throughout this report in order to maintain consistency with the previous report for this indicator;
however, results calculated using the five other methods are considered to be equally probable
outcomes. Figure 1 illustrates the daily stream flow monitoring information and the results of
                         Draft for Discussion at SOLEC 2006

-------
hydrograph separation for the Nith River at New Hamburg, Ontario for January 1 to December
31, 1993. The rapidly varying response of stream flow to precipitation and snow melt are in
contrast to the more slowly varying base flow.

Application of hydrograph separation to daily stream flow monitoring information results in
lengthy time series of output. Various measures are used to summarize this output; for example,
base flow index is a simple, physical measure of the contribution of base flow to stream flow that
is appropriate for use in regional scale studies. Base flow index is defined as the average rate of
base flow relative to the average rate of total stream flow, is unitless, and varies from zero to one
where increasing values indicate an increasing contribution of base flow to stream flow. The
value of base flow index for the data shown in Figure 1 is 0.28, which implies that 28% of the
observed flow is estimated to be base flow. Neff et al. (2005) used a selection of 960 gauges in
Ontario and the Great Lakes states to interpret base flow. Figure 2 indicates the distribution of the
values of base flow index calculated for the selection of gauges relative to the gauged and
ungauged portions of the Great Lakes basin. The variability of base flow within the basin is
apparent; however, further processing of the information is required to differentiate the
component of base flow that is  due to groundwater discharge and the component that is due to
delayed flow through lakes and wetlands upstream of the gauges. An approach to the
differentiation of base flow calculated using hydrograph separation into  these two components is
summarized in the following paragraphs of this report. Variations in the density of the stream
flow gauges and discontinuities in the coverage of monitoring are also apparent in Figure 2 and
may have significant implications relative to the  interpretation of base flow.

The values of base flow index calculated for the selection of gauges using hydrograph separation
are plotted relative to the extents of surface water upstream of each of the gauges in Figure 3
where the extents of surface water are defined as the area of lakes and wetlands upstream of the
gauges relative to the total area upstream of the gauges. While there is considerable scatter among
the values, the expected tendency for larger values of base flow index to be associated with larger
extents of surface water is confirmed. Neff et al.  (2005) modeled base flow index as a function of
surficial geology and the spatial extent of surface water. Surficial geology is assumed to be
responsible for differences in groundwater discharge and is classified into coarse and fine
textured sediments, till, shallow bedrock, and organic deposits.

The modeling process estimates a value of base flow index for each of the geological
classifications, calculates the weighted averages of these values for each of the gauges based on
the extents of the classifications upstream of the gauges, and then modifies the weighted averages
as a function of the extent of surface water upstream of the gauges. A non-linear regression
algorithm was used to determine the values of base flow index for the geological classifications
and the parameter in the surface water modifier that correspond to the best match between the
values of base flow index calculated using hydrograph separation and the values predicted using
the model. The process was repeated for each of the six methods of hydrograph separation.

Extrapolation of base flow index from gauged to ungauged watersheds was performed using the
results of the modeling process. The ungauged watersheds consist of 67  tertiary watersheds in
Ontario and  102 eight-digit hydrologic unit code or HUC watersheds in the Great Lakes states.
The extents of surface water for the ungauged watersheds are shown in Figure 4 where the ranges
of values used in the legend match those used to  average the values of base flow index shown in
                         Draft for Discussion at SOLEC 2006

-------
                                               'AW^W^™' JIU III t"^™i"">  '""^*    a» "fe       T "
                                               #Mwrj^*^iv*^wH, ^F1 »w* " "*•'* i,^h(f*i/^,/* vr«7	
Figure 3. A component of base flow due to delayed flow through lakes and wetlands appears to
be likely over extensive portions of the Great Lakes basin. The distribution of the classifications
of geology is shown in Figure 5. Organic and fine textured sediments are not differentiated in this
rendering of the classifications because both classifications have estimated values of base flow
index due to groundwater discharge in the range of 0.0 to 0.1; however, organic deposits are of
very limited extent and represent, on average, less than 2% of the area of the ungauged
watersheds. The spatial variation of base flow index shown in Figure 5 resembles the variation
shown in Figure 2. However, it is important to note that the information shown in Figure 2
includes the influence of delayed flow through lakes and wetlands upstream  of the gauges while
this influence has been removed, or at least reduced, in the information shown in Figure 5.

Figure 6 indicates the values of the geological component of base flow index for the ungauged
watersheds obtained by calculating the weighted averages of the values for the geological
classifications that occur in the watersheds. This map therefore represents an estimate of the
length-of-record contribution of base flow due to groundwater discharge to total stream flow that
is consistent and seamless across the Great Lakes basin. The pie charts indicate the range of
values of the geological component of base flow index for the six methods of hydrograph
separation averaged over the sub-basins of the Great Lakes. Averaging the six values  for each of
the sub-basins yields contributions of base flow due to groundwater discharge of approximately
60% for Lakes Huron, Michigan, and Superior and 50% for Lakes Erie and Ontario. It is
important to note that there is frequently greater variability of this contribution within the sub-
basins than among the sub-basins as the result of variability of geology that is more uniformly
averaged at the scale of the sub-basins.

Mapping the geological component of base flow index, which is assumed to  be due to
groundwater discharge, across the Great Lakes basin in a  consistent and seamless manner is an
important accomplishment in the development of this indicator. Additional information is,
however, required to determine the extent to which human activities have impaired groundwater
discharge. There are various alternatives for the generation of this information. For example, the
values of base flow index calculated for the selection of stream flow gauges using hydrograph
separation can be  compared to the corresponding modeled values. If a calculated value is less
than a modeled value, and if the difference is not related to the limitations of the modeling
process, then base flow is less than expected based on physiographic factors  and it is possible that
discharge has been impacted by human activities.  Similarly, if a calculated value is greater than a
modeled value, then it possible that the increased base flow is the result of human activities such
as flow regulation and wastewater discharge. Time series of base flow can also be used to assess
these impacts. The previous report for this indicator illustrated the detection  of temporal change
in base flow using data for watersheds with approximately natural stream flow and with extensive
flow regulation and urbanization; however, no attempt has yet been made to  systematically assess
change at the scale of the  Great Lakes basin. Change in base flow over time may be subtle and
difficult to quantify (e.g.,  variations in the relation of base flow to climate) and may be
continuous (e.g., a uniform increase in base flow due to aging water supply infrastructure and
increasing conveyance losses) or discrete (e.g., an abrupt reduction in base flow due to a new
consumptive water use). Change may also be the result of cumulative impacts due to a range of
historical and ongoing human activities, and may be more pronounced and readily detected at
local scales than at the scales that are typical of continuous stream flow monitoring.
                         Draft for Discussion at SOLEC 2006

-------
Figure 7 is an alternative view of the data for the Grand River at Gait, Ontario that was previously
used to illustrate the impact of flow regulation on base flow. The cumulative depth of base flow
calculated annually as the total volume of flow at the location of the gauge during each year
divided by the area that is upstream of the gauge, is plotted relative to cumulative total flow. Base
flow index is, by definition, the slope of the accumulation of base flow relative to the
accumulation of total flow. The change in slope and increase in base flow index from a value of
0.45 prior to the construction of the reservoirs that  are located upstream of the gauge to 0.57
following the construction of the reservoirs clearly indicates the impact of active flow regulation
to mitigate low and high flow conditions. Calculating and interpreting diagnostic plots such as
Figure 7 for hundreds to thousands of stream flow gauges in the Great Lakes basin will be a large
and time consuming, but perhaps ultimately necessary,  task.

Improving the spatial resolution of the current estimates of base flow due to groundwater
discharge would be beneficial in some settings. For example, localized groundwater discharge has
important implications in terms of aquatic habitat and it is unlikely that this discharge can be
predicted using the current regional estimates of base flow. The extrapolation of base flow
information from gauged to ungauged watersheds described by Neff et al. (2005) is based on a
classification and therefore reduced resolution representation of the Quaternary geology of the
basin. Figure 8 compares this classification to the full resolution of the available  1:1,000,000
scale (Ontario Geological Survey 1997) and 1:50,000 scale (Ontario Geological Survey 2003)
mapping of the geology of the gauged portion of the Grand River watershed in southern Ontario.
Interpretation of base flow in terms of these more detailed descriptions of geology, where feasible
relative to the network of stream flow gauges, may result in an improved estimate of the spatial
distribution of groundwater discharge for input into functions such as aquatic habitat
management.

Estimation of base flow using low flow observations, single "spot" measurements of stream flow
under assumed base flow conditions, is another means of improving the spatial resolution of the
current prediction of groundwater discharge.  Figure 9 illustrates a series of low flow observations
performed within the watershed of Duffins  Creek above Pickering, Ontario where the
observations are standardized using continuous monitoring information and the drainage areas for
the observations following the procedure described by Gebert et al. (in press) and then classified
into quantile groupings of high, intermediate, and low values. The standardized values of low
flow illustrate the spatially variable pattern of groundwater discharge that results from the
interaction between surficial geology, the complex three-dimensional hydrostratigraphy,
topography, and surface water features. Areas of potentially high groundwater discharge may
have particularly important implications in  terms of aquatic habitat for cold water fish species
such as Brook Trout.

Finally, reconciling estimates of base flow generated using differing methods of hydrograph
separation, perhaps by interpreting the information in a relative rather than absolute manner, will
improve the certainty and therefore performance of base flow as an indicator of groundwater
discharge. It may also  be possible to assess the source of this uncertainty using chemical and
isotopic data in combination with the methods of hydrograph separation if adequate data is
available at the scale of the gauged watersheds. Figure  10 compares the values of base flow index
calculated for the selection of 960 stream flow gauges in Ontario and the Great Lake states using
                         Draft for Discussion at SOLEC 2006

-------
                                           f^^^M^iiii^MiSi^^^	
the PART (Rutledge 1998) and UKIH methods of hydrograph separation. The majority of the
values calculated using the PART method are greater than the values calculated using the UKIH
method and there is considerable scatter in the differences among the two methods. The average
of the differences between the two sets of values is  0.15 and is significant when measured relative
to the differences in the estimates of base flow index for the sub-basins of the Great Lakes, which
is on the order of 0.1.

Pressures
The discharge of groundwater to surface water features is the end-point of the process of
groundwater recharge, flow, and discharge. Human activities impact groundwater discharge by
modifying the components of this process where the time scale, and to some extent the severity,
of these impacts is a function of hydrogeological factors and the proximity of surface water
features. Increasing the extent of impervious surfaces during residential and commercial
development and installation of drainage to increase agricultural productivity are examples of
activities that may reduce groundwater recharge and ultimately groundwater discharge.
Withdrawals of groundwater as a water supply and during dewatering (pumping groundwater to
lower the water table during construction, mining, etc.) remove groundwater from the flow
regime and may also reduce groundwater discharge. Groundwater discharge may be impacted by
activities such as the channelization of water courses that restrict the motion of groundwater
across the groundwater and surface water interface. Human activities also have the capacity to
intentionally,  or unintentionally, increase groundwater discharge. Induced storm water
infiltration, conveyance losses within municipal water and wastewater systems, and closure of
local water supplies derived from groundwater are examples of factors that may increase
groundwater discharge. Climate variability and change may compound the implications of human
activities relative to groundwater recharge, flow, and discharge.

Management Implications
Groundwater has important societal and ecological  functions across the Great Lakes basin.
Groundwater is typically a high quality water supply that is used by a significant portion of the
population, particularly in rural areas  where it is often the only available source of water.
Groundwater discharge to rivers, lakes, and wetlands is also critical to aquatic species and habitat
and to in-stream water quantity and quality. These functions are concurrent and occasionally
conflicting. Pressures such as urban development and water use, in combination with the potential
for climate impacts and further contamination of the resource, may increase the frequency and
severity of these conflicts. In the absence of systematic accounting of groundwater supplies, use,
and dependencies; it is the ecological function of groundwater that is most likely to be
compromised.

Managing the water quality of the Great Lakes requires an understanding of water quantity and
quality within the inland portion of the basin, and this understanding requires recognition of the
relative contributions of surface runoff and groundwater discharge to stream flow. The results
described in this report indicate the significant contribution of groundwater discharge to flow
within the tributaries of the Great Lakes. The extent of this contribution has tangible management
implications. There is considerable  variability in groundwater recharge, flow, and discharge that
must be reflected in the land and water management practices that are applied across the basin.
The dynamics of groundwater flow and transport are different than those of surface water flow.
                         Draft for Discussion at SOLEC 2006

-------
Groundwater discharge responds more slowly to climate and maintains stream flow during
periods of reduced water availability; however, this capacity is known to be both variable and
finite. Contaminants that are transported by groundwater may be in contact with geologic
materials for years, decades, and perhaps even centuries or millennia. As a result, there may be
considerable opportunity for attenuation of contamination prior to discharge. However, the
lengthy residence times of groundwater flow also limit opportunities for the removal of
contaminants, in general, and non-point source  contaminants, in particular.

Comments from the author(s)
The indicated status and trend are estimates that the authors consider to be a broadly held opinion
of water resource specialists within the Great Lakes basin. Further research and analysis is
required to confirm these estimates and to determine conditions on a lake by lake basis.

Acknowledgments
Authors: Andrew Piggott, Environment Canada;
Brian Neff, U.S. Geological Survey; and
Marc Hinton, Geological Survey of Canada.
Contributors: Lori Fuller, U.S. Geological Survey and
Jim Nicholas, U.S. Geological Survey.

Base flow information cited in the report is a product of Groundwater and the Great Lakes: A Co-
ordinated Bi-national Basin-wide Assessment in Support of Annex 2001 Decision Making, which
was supported by the Great Lakes Protection Fund.

Data Sources
Gebert, W.A., Lange, M.J., Considine, EJ.,and Kennedy, J.L., in press, Use of streamflow data to
estimate baseflow/ground-water recharge for Wisconsin: Journal of the American Water
Resources Association.

Neff, B.P., Day, S.M., Piggott, A.R., Fuller, L.M., 2005, Base Flow in the Great Lakes Basin:
U.S. Geological Survey Scientific Investigations Report 2005-5217, 23 p.

Ontario Geological Survey, 1997, Quaternary geology, seamless coverage of the province of
Ontario: Ontario Geological Survey, ERLIS Data Set 14.

Ontario Geological Survey, 2003, Surficial geology of southern Ontario: Ontario Geological
Survey, Miscellaneous Release Data 128.

Piggott, A.R., Moin, S., and Southam, C., 2005, A revised approach to the UKIH method for the
calculation ofbaseflow: Hydrological Sciences Journal, v. 50, p. 911-920.

Rutledge, A.T., 1998, Computer programs  for describing the recession of ground-water discharge
and for estimating mean ground-water recharge and discharge form streamflow data - update:
U.S. Geological Survey Water-Resources Investigations Report 98-4148, 43 p.
                         Draft for Discussion at SOLEC 2006

-------
List of Figures
Figure 1. Hydrograph of observed total stream flow (black) and calculated base flow (red) for the
Nith River at New Hamburg during 1993.
Source: Environment Canada and the U.S. Geological Survey

Figure 2. Distribution of the calculated values of base flow index relative to the gauged (light
grey)  and ungauged (dark grey) portions of the Great Lakes basin.
Source: Environment Canada and the U.S. Geological Survey

Figure 3. Comparison of the calculated values of base flow index to the corresponding extents of
surface water. The step plot (red) indicates the averages of the values of base flow index within
the four intervals of the extent of surface water.
Source: Environment Canada and the U.S. Geological Survey

Figure 4. Distribution of the extents of surface water for the ungauged watersheds.
Source: Environment Canada and the U.S. Geological Survey

Figure 5. Distribution of the geological classifications. The classifications are shaded using the
estimated values of the geological component of base flow index shown in parentheses.
Source: Environment Canada and the U.S. Geological Survey

Figure 6. Distribution of the estimated values of the geological component of base flow index for
the ungauged watersheds. The pie charts indicate the estimated values of the geological
component of base flow index for the Great Lakes sub-basins corresponding to the six methods of
hydrograph separation. The charts are shaded using the six values of base flow index and the
numbers in parentheses are the range of the values.
Source: Environment Canada and the U.S. Geological Survey

Figure 1'. Cumulative base flow as a function of cumulative total flow for the Grand River at Gait
prior to (red), during (green), and following (blue) the construction of the reservoirs that are
located upstream of the stream flow gauge. The step plot indicates the cumulative storage
capacity of the reservoirs where the construction of the largest four reservoirs is labeled. The
dashed red and blue lines indicate uniform accumulation of flow based on data prior to and
following, respectively, the construction of the reservoirs.
Source: Environment Canada and the U.S. Geological Survey

Figure 8. Geology of the gauged portion of the Grand River watershed based on the classification
(A) and Ml resolution (B) of the 1:1,000,000 scale Quaternary geology mapping and the Ml
resolution of the 1:50,000 scale Quaternary geology mapping (C) where random colours are used
to differentiate the various geological classifications and units.
Source: Environment Canada and the U.S. Geological Survey

Figure 9. Distribution of the standardized values of low flow within the watershed of Duffins
Creek above Pickering.
Source: Environment Canada and the U.S. Geological Survey, Geological Survey of Canada, and
Ontario Ministry of Natural Resources
                         Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
Figure 10. Comparison of the values of base flow index calculated using the PART method of
hydrograph separation to the values calculated using the UKIH method.
Source: Environment Canada and the U.S. Geological Survey

Last updated
SOLEC 2006
  1000
   1993/Q1/Q1
1993/04/01
1993/07/01
  Date
1993/10/01
1993/12/31
Figure 1. Hydrograph of observed total stream flow (black) and calculated base flow (red) for the
Nith River at New Hamburg during 1993.
Source: Environment Canada and the U.S. Geological Survey
                       Draft for Discussion at SOLEC 2006

-------
                             State of the Great Lakes 2007 - Draft
                                                     51° N
94° W
                                        Base Flow Index
        0.0    0.2    0.4    0.6    0.8
                                                            1.0
Figure 2. Distribution of the calculated values of base flow index relative to the gauged (light
grey) and ungauged (dark grey) portions of the Great Lakes basin.
Source: Environment Canada and the U.S. Geological Survey
10
Draft for Discussion at SOLEC 2006

-------
    1.0
    0.8  -
X
0)
•
o
IS
at
to
03
0.6
0.4  -
    0.0
                                                  •»*   ' *
                                                   £». %•   •
                  0,001           0,01
                         Extent of Surface Water
                                              0.1
Figure 3. Comparison of the calculated values of base flow index to the corresponding extents of
surface water. The step plot (red) indicates the averages of the values of base flow index within
the four intervals of the extent of surface water.
Source: Environment Canada and the U.S. Geological Survey
                         Draft for Discussion at SOLEC 2006
                                                                                11

-------
                            State of the Great Lakes 2007 - Draft
                                                   51° N
94° W
                                   Extent of Surface Water
                            0.001
                0,01
0.1
1
Figure 4. Distribution of the extents of surface water for the ungauged watersheds.
Source: Environment Canada and the U.S. Geological Survey
12
Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
                                                       51° N
94° W
                                      Geological Classification
                                 Organic   Till
                                (0.00) and  (0.33)
                                Fine (0.10)
Bedrock Coarse
 (0.59)   (0.82)
Figure 5. Distribution of the geological classifications. The classifications are shaded using the
estimated values of the geological component of base flow index shown in parentheses.
Source: Environment Canada and the U.S. Geological Survey
                       Draft for Discussion at SOLEC 2006
                               13

-------
                              State of the Great Lakes 2007 - Draft
                           Lake Superior
                            (0,47 - 0.70)
                                                           N
                                              Lake Huron
                                             (0.47 - 0.70)
                                                           Lake Ontario
                                                           (0.38 - 0.60)
 Lake Michigan
  (0.51 -0.63)
94° W
                                0.0    0.2   0.4   0.6    0.8
                                        1.0
Figure 6. Distribution of the estimated values of the geological component of base flow index for
the ungauged watersheds. The pie charts indicate the estimated values of the geological
component of base flow index for the Great Lakes sub-basins corresponding to the six methods of
hydrograph separation. The charts are shaded using the six values of base flow index and the
numbers in parentheses are the range of the values.
Source: Environment Canada and the U.S.  Geological Survey
14
Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes  2007 - Draft
     15
&   10
03
VI
to
00
CO
'S3-
ca
D
E
                      Shand
                              Conestogo
                             Luther
                                           Guelph
                                              I
   200
                           10                 20
                          Cumuiative Total Flow (m)
30
E
o
X

u
Q-
o
Figure 7. Cumulative base flow as a function of cumulative total flow for the Grand River at Gait
prior to (red), during (green), and following (blue) the construction of the reservoirs that are
located upstream of the stream flow gauge. The step plot indicates the cumulative storage
capacity of the reservoirs where the construction of the largest four reservoirs is labeled. The
dashed red and blue lines indicate uniform accumulation of flow based on data prior to and
following, respectively, the construction of the reservoirs.
Source: Environment Canada and the U.S. Geological Survey
                        Draft for Discussion at SOLEC 2006
                  15

-------
                              State of the Great Lakes 2007 - Draft
Figure 8. Geology of the gauged portion of the Grand River watershed based on the classification
(A) and full resolution (B) of the 1:1,000,000 scale Quaternary geology mapping and the full
resolution of the 1:50,000 scale Quaternary geology mapping (C) where random colours are used
to differentiate the various geological classifications and units.
Source: Environment Canada and the U.S. Geological Survey
16
Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
   High
   Intermediate
   Low
Figure 9. Distribution of the standardized values of low flow within the watershed of Duffins
Creek above Pickering.
Source: Environment Canada and the U.S. Geological Survey, Geological Survey of Canada, and
Ontario Ministry of Natural Resources
                      Draft for Discussion at SOLEC 2006
17

-------
    !0
    0,8
ai,  0.6

X
0)
T3
u.
as
CO
    0.4
    0.2
    0.0
       0.0
0,2        0.4         0.6

        Base Flow Index (UKIH)
0.8
1.0
Figure 10. Comparison of the values of base flow index calculated using the PART method of

hydrograph separation to the values calculated using the UKIH method.

Source: Environment Canada and the U.S. Geological Survey
                        Draft for Discussion at SOLEC 2006

-------
                                              OF   THE   GREAT
                              2007
Groundwater Dependant Plant and Animal
Communities
Indicator #7103

Assessment:  Not Assessed
Note: This indicator report uses data from the Grand River
watershed only and may not be representative of groundwater
conditions throughout the Great Lakes basin. Additionally, there
is insufficient biological and physical hydrological data for most
of the streams in the Grand River watershed to report on many
of the selected species reliant on groundwater discharge, hence
this discussion focuses on brook trout (Salvelinus fontinalis) as
an indicator of groundwater discharge.

Purpose
  To measure the abundance and diversity as well as presence or
absence of native invertebrates, fish, plant and wildlife (includ-
ing cool-water adapted frogs and salamanders) communities that
are dependent on groundwater discharges to aquatic habitat;
  To identify and understand any deterioration of water quality
for animals and humans, as well as changes in the productive
capacity of flora and fauna dependant on groundwater resources;
  To use biological communities to assess locations of ground-
water intrusions; and
  To infer certain chemical and physical properties of ground-
water, including  changes in patterns of seasonal flow.

Ecosystem Objective
The goal for this indicator is to ensure that plant and animal
communities  function at or near maximum potential and that
populations are not significantly compromised due to anthro-
pogenic factors.

State of the Ecosystem
Background
The integrity of larger water bodies can be linked to biological,
chemical and physical integrity of the smaller watercourses that
feed them. Many of these small watercourses are fed by ground-
water. As a result, groundwater discharge to  surface waters
becomes cumulatively important when considering the quality of
water entering the Great Lakes. The identification of groundwa-
ter fed streams and rivers will provide useful information for the
development  of watershed management plans that seek to pro-
tect these sensitive watercourses.

Human activities can change the hydrological processes in a
watershed resulting in changes to recharge rates of aquifers and
discharges rates to streams and wetlands. This indicator should
serve to identify organisms  at risk because of human activities
can be used to quantify trends in communities over time.
256
 Status of Groundwater Dependent Plant and Animal
 Communities in the Grand River Watershed
 The surficial geology of the Grand River watershed is generally
 divided into three distinct regions; the northern till plain, central
 moraines with large sand and gravel deposits, and the southern
 clay plain (Figure 1).  These surficial overburden deposits are
 underlain by thick sequences of fractured carbonate rock (pre-
 dominantly dolostone).
                                         Generalized Geologic Units
                                             Clay
                                             Gravel
                                             Organic
                                             Sand
                                             Sandy Till
                                             Silly Till
                                             Water
 10   0    10    20 .,..    .
                  Kilometres
Figure 1. Surficial geology of the Grand River watershed.
Source: Grand River Conservation Authority
 The Grand River and its tributaries form a stream network hous-
 ing approximately 11,329 km of stream habitat. The Ontario
 Ministry of Natural Resources (OMNR) has classified many of
 Ontario's streams based on habitat type. While many streams
 and rivers in the Grand River watershed remain unclassified, the
 MNR database currently available through the Natural
 Resources and Values Information System (NRVIS) has docu-
 mented and classified about 22% of the watershed's streams
 (Figure 2). Approximately 19% of the classified streams are

-------
 cold-water habitat and therefore dependent on groundwater dis-
 charge. An additional 16% of the classified streams are consid-
 ered potential cold-water habitat. The remaining 65% of classi-
 fied streams are warm-water habitat.
                                      Stream Classification
                                          Not Classified
                                        — Coldwater
                                        — Potential Coldwater
                                          Warmwater Sportfish
                                        — Warmwater Baitfish
                                          High Recharge Area
          4-
                 Kilometres
Figure 2. Streams of the Grand River watershed.
Source: Grand River Conservation Authority
 A map of potential groundwater discharge areas was created for
 the Grand River watershed by examining the relationship
 between the water table and ground surface (Figure 3). This map
 indicates areas in the watershed where water well records indi-
 cate that the water table could potentially be higher than the
 ground surface. In areas where this is the case, there is a strong
 tendency toward discharge of groundwater to  land, creating
 cold-water habitats. Groundwater discharge appears to be geo-
 logically controlled with most potential discharge areas noted
 associated with the sands and gravels in the central moraine
 areas and little discharge in the northern till plain and  southern
 clay plain. The map suggests that some of the unclassified
 streams in Figure  2 may be potential cold-water streams, particu-
 larly in the central portion of the watershed where geological
 conditions are favourable to groundwater discharge.
                                          o Spawning Location
                                         Potential Ho*ght ol Water Table
                                         Above Ground Surface (moires)
                                            120.
                                            19-20
                                            18-19
                                          —| 17 18
                                          —116-17
                                          — 115-16
                                            1-1 - 15
                                            13-14
                                            12- 13
                                            11-12
                                            10-11
                   Kilometres
Figure 3. Map of potential discharge areas in the Grand River
watershed.
Source: Grand River Conservation Authority
 Brook trout is a freshwater fish species native to eastern Canada.
 The survival and success of brook trout is closely tied to cold
 groundwater discharges in streams used for spawning.
 Specifically, brook trout require inputs of cold, clean water to
 successfully reproduce. As a result, nests or redds are usually
 located in substrate where groundwater is upwelling into surface
 water. A significant spawning population of adult brook trout
 generally indicates a constant source of cool, good quality
 groundwater.

 Locations of observed brook trout redds are shown on Figure 3.
 The data shown are a compilation of several surveys carried out
 on selected streams in 1988 and 1989. Additional data from sev-
 eral sporadic surveys carried out in the 1990s are also included.
 These redds may represent single or multiple nests from brook
 trout spawning activity. The results of these surveys illustrate

                                                           257

-------
                                            '''il
                                                                             LAKES   2007
that there are significant high quality habitats in several of the
subwatersheds in the basin.

Cedar Creek is a tributary of the Nith River in the central portion
of the watershed. It has been described as containing some of the
best brook trout habitat in the watershed. Salmonoid spawning
surveys for brook trout were carried out over similar stretches of
the creek in 1989 and 2003 (Figure 4). In 1989 a total redd count
of 53  (over 4.2 km) was surveyed while in 2003 the total redd
count was 59 (over 5.4 km). In both surveys, many of the redds
counted were multiple redds meaning several fish had spawned
at the same locations. Redd densities in 1989 and 2003 were
12.6 redds/km and 10.9 redds/km respectively. From Figure 4 it
appears that in 2003 brook trout were actively spawning in
Cedar Creek in mainly the same locations as in 1989. While
redd density in Cedar Creek has decreased slightly, the similar
survey results suggest that groundwater discharge has remained
fairly constant and reductions in discharge have not significantly
affected aquatic habitat.
surface will decrease the geological protection afforded ground-
water supplies and may increase the temperature of groundwater.
Higher temperatures can reduce the moderating effect groundwa-
ter provides to aquatic stream habitat. At local scales the creation
of surface water bodies through mining or excavation of aggre-
gate or rock may change groundwater flow patterns, which in
turn might decrease groundwater discharge to sensitive habitats.

In the Grand River watershed, groundwater is used by about
80% of the watershed's residents as their primary water supply.
Additionally, numerous industrial and agricultural users also use
groundwater for their operations. Growing urban communities
will put pressure on the resource and if not managed properly
will lead to decreases in groundwater discharge to streams.
Development in some areas can also lead to decreased areas
available for precipitation to percolate through the ground and
recharge  groundwater supplies.
\V~ r'N] -^ •(
-"1989 ^1
\ " /
(32' \ i.
lj-~~~. . J;
( "t"
, ^'-'^
1

'/^^C~ " ^
' \ \ 1, >}
^-^,.X ^ f*

\ ''*••• } ''"••
\ ^\ ] \

i .' yL—-~ 	 "
\ \ ___--" 	 ""f" _^s-
\T'"~ 1 f
\ -?' /
'" \X f^-x ^
\ 1 \
\ \ '^-^^--^^
\ \ T
/ ^^
^' 1 nS^ ~~~~*" 	 "~
\ ^ _.*~~~~~~'"~~"
m _— • 	
XV f \
-^ "2003 '
/ \ (,-'''
-v-1''" .-^' 'f^*1 "" • "^
f**'"" '^- •'^/ $* \^~- — _\
>^v \ *=* [ ,X>=
' /' \y ^^"" 	 '
:^.<3T ^ \>
"~/^P\ "~" 7
"• ^ '*>• i J
1 i .^p^ ^r
1" - ^-~'~-  ( 	 _---
^ \ ~~\
\ ,' \
" 	 N
i
"i \ iJ«^- — """
A j.^--"'''y' _^--
-\~\ f /'
"\\ ^ /
r • ^v/^^v^
r' \ I N
\ \ y v~~--__-____J_-
^ \ i' \
\ jAdf1
\ f aB
\y
"""" A.. -


• Survey Endpoints

• Redd Location
/ \ / Roads
f\/ Streams
200 0 200 400






N
afe_
^^
3









Figure 4. Results of brook trout spawning surveys carried out in the Cedar Creek
subwatershed in 1989 and 2003.
Source: Grand River Conservation Authority
                                                                             Management Implications
                                                                             Ensuring that an adequate supply of cold ground-
                                                                             water continues to discharge into streams
                                                                             requires protecting groundwater recharge areas
                                                                             and ensuring that groundwater withdrawals are
                                                                             undertaken at sustainable rates. Additionally, an
                                                                             adequate supply of groundwater for habitat pur-
                                                                             poses does not only refer to the quantity of dis-
                                                                             charge but also to the chemical quality, tempera-
                                                                             ture and spatial location of that discharge. As a
                                                                             result, protecting groundwater resources is com-
                                                                             plicated and generally requires multi-faceted
                                                                             strategies including regulation, voluntary adop-
                                                                             tion of best management practices and public
                                                                             education.

                                                                             Acknowledgments
                                                                             Authors: Alan Sawyer, Grand River Conservation
                                                                             Authority, Cambridge, ON;
                                                                             Sandra Cooke, Grand River Conservation
                                                                             Authority, Cambridge, ON;
                                                                             Jeff Pitcher, Grand River Conservation Authority,
                                                                             Cambridge, ON; and
                                                                             Pat Lapcevic, Grand River Conservation
                                                                             Authority, Cambridge, ON.
Pressures
The removal of groundwater from the subsurface through pump-
ing at wells reduces the amount of groundwater discharging into
surface water bodies. Increasing impervious surfaces reduces the
amount of water that can infiltrate into the ground and also ulti-
mately reduces groundwater discharge into surface water bodies.
Additionally, reducing the depth to the water table from ground

258
Alan Sawyer's position was partially funded through a grant from
Environment Canada's Science Horizons internship program. The
assistance of Samuel Bellamy and Warren Yerex of the Grand
River Conservation Authority, as well as Harvey Shear, Nancy
Stadler-Salt and Andrew Piggott of Environment Canada is grate-
fully acknowledged.

-------
Sources
Grand River Conservation Authority. 2003. Brook Trout
(Salvelinus fontinalis) Spawning Survey - Cedar Creek.

Grillmayer, R.A., and Baldwin, RJ. 1990. Salmonid spawning
surveys of selected streams in the Grand River watershed 1988-
1989. Environmental Services Group, Grand River Conservation
Authority.

Holysh, S., Pitcher, J., and Boyd, D. 2001. Grand River
Regional Groundwater Study. Grand River Conservation
Authority, Cambridge, ON. 78pp. + figures and appendices.

Scott, W.B., and Grossman, EJ. 1973. Freshwater fishes of
Canada. Bulletin 184, pp. 208-213. Fisheries Research Board of
Canada, Ottawa, ON.
Authors' Commentary
This report has focused on only one species dependent on
groundwater discharge for its habitat. The presence or absence of
other species should be investigated through systematic field
studies.

Last Updated
State of the Great Lakes 2005
                                                                                                                  259

-------
                                              OF    THE   GREAT
                             2007
Area, Quality and Protection of Special Lakeshore
Communities - Alvars
Indicator #8129 (Alvars)

Assessment: Mixed, Trend Not Assessed

Purpose
  To assess the status of Great Lakes alvars (including changes
in area and quality), one of the 12 special lakeshore communities
identified within the nearshore terrestrial area;
  To infer the success of management activities; and
  To focus future conservation efforts  toward the most ecologi-
cally significant alvar habitats in the Great Lakes.

Ecosystem Objective
The objective is the preservation of the area and quality of Great
Lakes alvars, individually and as an ecologically important sys-
tem, for the maintenance of biodiversity and the protection of
rare species. This indicator supports Annex 2 of the Great Lakes
Water Quality Agreement.

State of the Ecosystem
Background
Alvar communities are naturally open habitats occurring on flat
limestone bedrock. They have a distinctive  set of plant species
and vegetative associations, and include many species of plants.
molluscs, and invertebrates that are rare elsewhere in the basin.
All 15 types of alvars and associated habitats are globally imper-
iled or rare.

A four-year study of Great Lakes alvars completed in 1998 (the
International Alvar Conservation Initiative-IACI)  evaluated con-
servation targets for alvar communities, and concluded that
essentially all of the existing viable occurrences should be main-
tained, since all types are below the minimum threshold of 30-60
viable examples. As well as conserving these ecologically dis-
tinct communities, this target would protect populations of
dozens of globally significant and disjunct species. A few
species, such as lakeside daisy (Hymenoxis herbacea) and the
beetle Chlaenius p. purpuricollis, have nearly all of their global
occurrences within Great Lakes alvar sites.

Status of Great Lakes Alvars
Alvar habitats have likely always been sparsely distributed, but
more than 90% of their original extent has been destroyed or
substantially degraded by agriculture and other human uses.
Approximately 64% of the remaining alvar area occurs within
Ontario, with about 16% in New York  State, 15% in Michigan.
4% in Ohio, and smaller areas in Wisconsin and Quebec.
Data from the IACI and state/provincial alvar studies were
screened and updated to identify viable community occurrences.
Just over two-thirds of known Great Lakes alvars occur close to
the shoreline, with all or a substantial portion of their area within
one kilometre of the shore.

No. of alvar sites
No. of community occurences
Alvar area (ha)
Total in Basin
82
204
1 1 ,523
Nearshore
52
138
8,097
Table 1 . Number of alvar sites/communities found
nearshore and total in the basin.
Source: Ron Reid, Bobolink Enterprises
Typically, several different community types occur within each
alvar site. Among the 15 community types documented, six
types show a strong association (over 80% of their area) with
nearshore settings. Four types have less than half of their occur-
rences in nearshore settings.

The current status of all nearshore alvar communities was evalu-
ated by considering current land ownership and the type and
severity of threats to  their integrity. As shown in Figure 1, less
than one-fifth of the nearshore alvar area is currently fully pro-
tected, while  over three-fifths is at high risk.
  Limited 11.9%
                            Partly 9.1%
                                            Fully 18.8%
At High Risk 60.2%
Figure 1. Protection status of nearshore alvar area (2000).
Source: Ron Reid, Bobolink Enterprises
The degree of protection for nearshore alvar communities varies
considerably among jurisdictions. For example, Michigan has
66% of its nearshore alvar area in the Fully Protected category.
while Ontario has only 7%. In part, this is a reflection of the
much larger total shoreline area in Ontario, as shown in Figure
2. (Other states have too few nearshore sites to allow compari-
son).

Each location of an alvar community or rare species has been
documented as  an "element occurrence" or EO. Each alvar com-
260

-------
Acres of Alvar





L nnn

!000 -











	


1
Ontario
• At High Risk
Partly Protected

















^^i
Michigan









H Limited
• Fully Protected


Figure 2. Comparison of the protection status of nearshore
alvars (in acres) for Ontario and Michigan.
Source: Ron Reid, Bobolink Enterprises
 munity occurrence has been assigned an "EO rank" to reflect its
 relative quality and condition ("A" for excellent to "D" for
 poor). A and B-ranks are considered viable, while C-ranks are
 marginal and a D ranked occurrence is not expected to survive
 even with appropriate management efforts. As shown in Figure
 3, protection efforts to secure alvars have clearly focused on the
 best quality sites.
                         AB          B
                          EO  Rank
          BC&C
       ]  Partly Protected
Fully Protected
Figure 3. Protection of high quality alvars. EO Rank = Element
Occurrence (A is excellent, B is good and C is marginal).
Source: Ron Reid, Bobolink Enterprises
 Documentation of the extent and quality of alvars through the IACI
 has been a major step forward, and has stimulated much greater
 public awareness and conservation activity for these habitats. Over
 the past two years, a total of 10 securement projects have resulted in
 protection of at least 2140.6 ha of alvars across the Great Lakes
 basin, with 1353.5 ha of that within the nearshore area. Most of the
 secured nearshore area is through land acquisition, but 22.7 ha on
 Pelee Island (ON) are through a conservation easement, and 0.6 ha
 on Kelleys Island (OH) are through state dedication of a nature
reserve. These projects have increased the area of protected alvar
dramatically in a short time.

Pressures
Nearshore alvar communities are most frequently threatened by
habitat fragmentation and loss, trails and off-road vehicles, resource
extraction uses such as quarrying or logging, and adjacent land uses
such as residential subdivisions.  Less frequent threats include graz-
ing or deer browsing, plant collecting for bonsai or other hobbies,
and invasion by non-native plants such as European buckthorn and
dog-strangling vine.

Acknowledgments
Authors: Ron Reid, Bobolink Enterprises, Washago, ON; and
Heather Potter, The Nature Conservancy,  Chicago, IL.

Sources
Brownell, V.R.,  and Riley, J.L. 2000. The alvars of Ontario: signifi-
cant alvar natural areas in the Ontario Great Lakes Region.
Federation of Ontario Naturalists, Toronto, ON.

Cusick, A.W. 1998. Alvar landforms and plant communities in
Ohio. Ohio Department of Natural Resources, Columbus, OH.

Oilman, B. 1998. Alvars of New York: A Site Summary Report.
Finger Lakes Community College, Canandaigua, NY.

Lee, Y.M., Scrimger, L.J., Albert, D.A., Penskar, M.R., Comer, P.J.,
and Cuthrell, DA. 1998. Alvars  of Michigan. Michigan Natural
Features Inventory, Lansing, MI.

Reid, R. 2000. Great Lakes alvar update, July 2000.  Prepared for
the International Alvar Conservation Initiative Working Group.
Bobolink Enterprises, Washago,  ON.

Reschke, C., Reid, R., Jones, J.,  Feeney, T., and Potter, H. 1999.
Conserving Great Lakes alvars: final technical report of the
International Alvar Conservation Initiative. The Nature
Conservancy, Chicago, IL.
                           Authors' Commentary
                           Because of the large number of significant alvar communities at
                           risk, particularly in Ontario, their status should be closely watched
                           to ensure that they are not lost. Major binational projects hold great
                           promise for further progress, since alvars are a Great Lakes
                           resource, but most of the unprotected area is within Ontario.
                           Projects could be usefully modeled after the 1999 Manitoulin Island
                           (ON) acquisition of 6880 ha through a cooperative project of The
                           Nature Conservancy of Canada, The Nature Conservancy,
                           Federation of Ontario Naturalists, and Ontario Ministry of Natural
                           Resources.
                           Last Updated
                           State of the Great Lakes 2001
                                                                                                                          261

-------
Area, Quality and Protection of Special Lakeshore
Communities - Cobble Beaches
Indicator #8129 (Cobble Beaches)

Assessment: Mixed, Deteriorating

Purpose
  To assess the status of cobble beaches, one of the 12 special
shoreline communities identified within the nearshore terrestrial
area. To assess the changes in area and quality of Great Lakes
cobble beaches;
  To infer the success of management activities; and
  To focus future conservation efforts toward the most
ecologically significant cobble beach habitats in the Great
Lakes.

Ecosystem Objective
The objective is the preservation of the area and quality of
Great Lakes cobble beaches, individually and as an ecolog-
ically important system, for the maintenance of biodiversi-
ty and the protection of rare species. This indicator sup-
ports Annex 2 of the Great Lakes Water Quality
Agreement.

State of the Ecosystem
Background
Cobble beaches are shaped by wave and ice erosion. They
are home to a variety of plant species, several of which are
threatened or endangered provincially/statewide, globally,
or both making them one of the most biodiverse terrestrial
communities along the Great Lakes shoreline. Cobble beaches
serve as seasonal spawning and migration areas for fish as well
as nesting areas for the piping plover, a species listed in the U.S.
as endangered.

Status of Cobble Beaches
Cobble beaches have always been a part of the Great Lakes
shoreline. The number and area of these beaches, however, is
decreasing due to shoreline development. In fact,  cobble shore-
lines are becoming so scarce that they are considered globally
rare.

Lake Superior has the most cobble shoreline of all the Great
Lakes with 958 km of cobble beaches (Figure 1); 541  km on the
Canadian side and 417 km on the U.S.  side. This constitutes
20% of the whole Lake Superior shoreline  (11.3% on the
Canadian side and 8.7% on the U.S. side).

Lake Huron has the 2nd most cobble shoreline with approximate-
ly 483 km of cobble shoreline; 330 km on the Canadian side and
153 km on the U.S. side. Most of the cobble beaches are  found
along the shoreline of the Georgian Bay (Figure 2). This consti-
262
      tutes approximately 9% of the whole Lake Huron shoreline
      (6.1% on the Canadian side and 2.8% on the U.S. side).

      Approximately 164 km of the Lake Michigan shoreline is cob-
      ble, representing 6.1% of its shoreline. Most of these beaches are
      located at the northern end of the lake in the state of Michigan
      (Figure 3).

      Lake Ontario has very little cobble shoreline of about 35 km,
      representing only 3% of its shoreline (Figure 4).
                                                 \
Figure 1. Cobble beaches along Lake Superior's shoreline (red = cobble
beach locations).
Source: Lake Superior Binational Program, Lake Superior LaMP 2000,
Environment Canada, and Dennis Albert
      Figure 2. Cobble beaches along Lake Huron's shoreline (red =
      cobble beach locations).
      Source: Environment Canada

-------

Figure 3. Cobble beaches along Lake Michigan's shoreline
(red = cobble beach locations).
Source: Albert 1994a, Humphrys et al.  1958
                                                        Figure 5. Cobble beaches along Lake Erie's shoreline (red = cobble
                                                        beach locations).
                                                        Source: Environment Canada
Lake Superior's large cobble shoreline provides for several rare
plant species (Table 1) some of which include the Lake Huron
tansy and redroot. It is also home to the endangered heart-leaved
plantain, which is protected under the Ontario Endangered
Species Act.
Figure 4. Cobble beaches along Lake Ontario's shoreline (red =
cobble beach locations).
Source: International Joint Commission (IJC) and Christian J.
Stewart
Lake Superior
Common Name
Bulrush sedge
Great northern aster
Northern reedgrass
Purple clematis
Northern grass of Parnassus
Mountain goldenrod
Narrow-leafed reedgrass
Downy oat-grass
Pale Indian paintbrush
Butterwort
Pearlwort
Calypso orchid
Lake Huron tansy
Redroot
Heart-leaved plantain
Scientific Name
Carex sclrpoldea
Aster modestus
Calamagmstis lacustris
Clematis occldentalls
Pamassia palustris
Solldago decumbens
Calamagmstis stricta
Trisetum spicatum
Castllleja septentrlonalls
Plnguicula vulgaris
Sagina nodosa
Calypsa bulbosa
Tanacetum humnense
Lachnanthes carol/ana
Plantago cordata
Table 1 . Rare plant species on Lake Superior's cobble
shoreline.
Source: Lake Superior LaMP, 2000
  Lake Erie has the smallest amount of cobble shoreline of all the
  Great Lakes with only 26 km of cobble shore. This small area
  represents approximately 1.9% of the lake's shoreline (Figure 5).

  While the cobble beaches themselves are scarce, they do have a
  wide variety of vegetation associated with them, and they serve
  as home to plants that are endemic to the Great Lakes shoreline.
Lake Michigan and Lake Huron's cobble shorelines are home to
Houghton's goldenrod and the dwarf lake iris, both of which are
endemic to the Great Lakes shoreline (Table 2, Table  3). Some
other rare species on the Lake Michigan shoreline include the
Lake Huron tansy and beauty sedge (Table 2).

Not many studies have been conducted on the cobble shorelines
of Lake Ontario and Lake Erie because these areas are so small.
The report author was unable to find any information about the
                                                       263

-------
vegetation that grows there.
Lake Michigan
Common Name
Dwarf lake iris
Houghton's goldenrod
Slender cliff-brake
Lake Huron tansy
Beauty sedge
Richardson's sedge
Scientific Name
Iris lacustris
Solidago houghtonii
Cryptogramma stelleri
Tanacetum huronense
Carex concinna
Carex richardsonii
Table 2. Rare plant species along Lake Michigan's
cobble shoreline.
Source: Dennis Albert

Lake Huron
Common Name
Dwarf lake iris
Houghton's goldenrod
Scientific Name
Iris lacustris
Solidago houghtonii
Table 3. Rare plant species along Lake Huron's cobble
shoreline.
Source: Environment Canada
Pressures
Cobble beaches are most frequently threatened and lost by
shoreline development. Homes built along the shorelines of the
Great Lakes cause the number of cobble beaches to become lim-
ited. Along with the development of homes also comes increased
human activity along the shoreline resulting in damage to rare
plants in the surrounding area and ultimately a loss of terrestrial
biodiversity on the cobble beaches.

Acknowledgments
Author: Jacqueline Adams, Environmental Careers Organization,
on appointment to U.S. Environmental Protection Agency, Great
Lakes National Program Office.

Sources
Albert, D. 1994a. Regional landscape ecosystems of Michigan,
Minnesota, and Wisconsin: a working map and classification.
Michigan Natural Features Inventory, Lansing, MI.

Albert, D., Comer, P., Cuthrell, D., Penskar, M., Rabe, M., and
Reschke, C. 1994b. Bedrock shoreline surveys of the Keweenaw
Peninsula and Drummond Island in Michigan s Upper
Peninsula. Michigan Natural Features Inventory, Lansing, MI.

Albert, D.A., Comer,  P.J., Corner, R.A., Cuthrell, D., Penskar,
M., and Rabe, M. 1995. Bedrock shoreline survey of the
Niagaran escarpment in Michigan s Upper Peninsula: Mackinac
County to Delta County. Michigan Natural Features Inventory,
Lansing, MI.
Environment Canada.  1994a. Environmental Sensitivity Atlas for
Lake Erie (including the Welland Canal) and the Niagara River

264
shorelines. Environment Canada, Ontario Region, United States
Coast Guard, and the United States National Oceanic and
Atmospheric Administration (NOAA).

Environment Canada. 1994b. Environmental Sensitivity Atlas for
Lake Huron s Canadian shoreline (including Georgian Bay).
Environment Canada, Ontario Region.

Humphrys, C.R., Horner, R.N., and Rogers, J.H. 1958. Shoretype
Bulletin Nos. 1-29. Michigan State University Department of
Resource Development, East Lansing, MI.

International Joint Commission (IJC) 2002. Classification of
shore units. Coastal working group. Lake Ontario and Upper St.
Lawrence River. Environment Canada and U.S. Environmental
Protection Agency (USEPA).

Lake Superior Binational Program. 2000. Lake Superior
Lakewide Management Plan (LaMP) 2000. Environment Canada
and U.S. Environmental Protection Agency (USEPA).

Michigan's Natural Features Inventory (MNFI). Rare Plant
Reference Guide. Michigan State University Extension.
http://web4.msue.msu.edu/mnfi/data/rareplants.cfm. last
accessed October 5,  2005.

Stewart, CJ. 2003. A revised geomorphic, shore protection and
nearshore classification of the Canadian and United States
shorelines of Lake Ontario and the St. Lawrence River. Christian
J. Stewart Consulting, British Columbia, Canada.
Authors' Commentary
Not much research has been conducted on cobble beach commu-
nities; therefore, no baseline data have been set. A closer look
into the percentage of cobble beaches that already have homes
on them or are slated for development would yield a more accu-
rate direction in which the beaches are headed. Also, a look at
the percentage of these beaches that are in protected areas would
provide valuable information. Projects similar to Dennis Albert's
Bedrock Shoreline Surveys of the Keweenaw Peninsula and
DrummondIsland in Michigan's Upper Peninsula (1994) for the
Michigan Natural Features Inventory, as well as the International
Joint Commission's Classification of Shore Units Coastal
Working Group: Lake Ontario and Upper St. Lawrence River
(2002), would be very useful in determining exactly where the
remaining cobble beaches are located and what is growing and
living within them.

Last Updated
State of the Great Lakes 2005

-------

                                                     ^ ._
                                  ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''
Extent, Condition and Conservation Management of Great Lakes Islands
Indicator #8129
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
The Framework for Binational Conservation of Great Lakes Islands
will be completed in 2007. The following results reflect detailed
analysis from Canadian islands and preliminary results from the US.
This project has created the first detailed binational map Great Lakes
islands. This includes the identification of 31,407 island polygons with
a total coastline of 15,623 km.

This project has established baseline information that will be used to
assess future trends.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Good
Undetermined
Detailed analysis for Canada only. Total (Canada and US) of 2,591 island
polygons.  St. Mary's River has 630 island polygons.

Canadian islands in Lake Superior have the lowest threats score in the
basin. A high proportion of these islands are within protected areas and
conservation lands. Overall condition is good.  These islands include a high
number of disjunct plant species.
Not Assessed
Undetermined
Detailed analysis not completed.  Total of 329 island polygons.
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
Detailed analysis for Canada only. Total (Canada and US) of 23,719 island
polygons (includes Georgian Bay).

These islands tend to be more threatened in the south compared to the north.
A large number of protected areas and conservation lands occur in the
northern region. Southern regions are more developed, and under
increasing pressures from development.  These islands include high number
of globally rare species and vegetation communities.
                        Draft for Discussion at SOLEC 2006

-------
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
Detailed analysis for Canada only. Total (Canada and US) of 1,724 island
polygons. Other island polygons with Lake St. Clair/ St. Clair River (339),
Detroit River (61) and Niagara River (36).

These islands include a mix of protected areas and private islands. Islands
in the western Lake Erie basin have some of the highest biodiversity values
of all Great Lakes islands.
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
Detailed analysis for Canada only. Total (Canada and US) of 2,591 island
polygons (including upper St. Lawrence River).

Many of these islands have high threat index scores and a long history of
recreational use. One of the highest building point counts. Few areas have
been protected.
Purpose
•To assess the status of islands, one of the 12 special lakeshore communities identified within the
nearshore terrestrial area.

Ecosystem Objective
To assess the changes in area and quality of Great Lakes islands individually, and as an
ecologically important system; to infer the success of management activities; and to focus future
conservation efforts toward the most ecologically significant island habitats in the Great Lakes.

State of the Ecosystem
Background
There are 31,407 islands that have been idnetified in the Great Lakes (Figure 1). The islands
range in size from no bigger than a large boulder to the world's largest freshwater island,
Manitoulin, and often form chains of islands known as archipelagos. Though not well known, the
Great Lakes contain the world's largest freshwater island system, and are globally significant in
terms of their biological diversity. Despite this, the state of our knowledge about them as a
collection is quite poor.

By their very nature, islands are vulnerable and sensitive to change. Islands are exposed to the
forces of erosion and accretion as water levels  rise and fall. Islands are also exposed to weather
events due to their 360-degree exposure to the  elements across the open water. Isolated for
perhaps tens of thousands of years from the mainland, islands in the past rarely gained new
species, and some of their resident species evolved into endemics that differed from mainland
varieties. This means that islands are especially vulnerable to the introduction of non-native
species, and can only support a fraction of the number of species of a comparable mainland area.
                         Draft for Discussion at SOLEC 2006

-------

                                                        ^ ._
                                    ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''
Some of the Great Lakes islands are among the last remaining wildlands on Earth. Islands must
be considered as a single irreplaceable resource and protected as a whole if the high value of this
natural heritage is to be maintained. Islands play a particularly important role in the "storehouse"
of Great Lakes coastal biodiversity. For example, Michigan's 600 Great Lakes islands contain
one-tenth of the state's threatened, endangered, or rare species while representing only one-
hundredth of the land area. All of Michigan's threatened, endangered, or rare coastal species
occur at least in part on its islands. The natural features of particular importance on Great Lakes
islands are colonial waterbirds, neartic-neotropical migrant songbirds, endemic plants, arctic
disjuncts, endangered species, fish spawning and nursery use of associated shoals and reefs and
other aquatic habitat, marshes, alvars, coastal barrier systems, sheltered embayments, nearshore
bedrock mosaic, and sand dunes.  New research indicates that nearshore island areas in the
Ontario waters of Lake Huron account for 58% of the fish spawning and nursery habitat and thus
are critically important to the Great Lakes fishery. Many of Ontario's provincially rare species
and vegetation communities can be found on islands in the Great Lakes.

Pressures
By their very nature, islands are more sensitive to human influence than the mainland and need
special protection to conserve their natural values. Proposals to develop islands are increasing.
This is occurring before we have  the scientific information about sustainable use to evaluate,
prioritize, and make appropriate natural resource decisions on islands. Island stressors include
development, invasive species, shoreline modification, marina and air strip development,
agriculture and forestry practices, recreational use, navigation/shipping  practices, wastewater
discharge, mining practices, drainage or diversion systems, overpopulation of certain species such
as deer, industrial discharge, development of roads or utilities, abandoned landfills, and disruption
of natural disturbance regimes.

Management Implications
Based on the results of assessments of island values, biological significance, categorization, and
ranking, the Binational Collaborative for the Conservation of Great Lakes Islands will soon
recommend management strategies on Great Lakes islands to preserve the unique ecological
features that make islands so important. In addition, based on a proposed threat assessment to be
completed in 2005, the Collaborative will recommend management strategies to reduce the
pressures on a set of priority island areas.

Comments from the author(s)
The Great Lakes islands provide a unique opportunity to protect a resource of global importance
because many islands still remain intact. The U.S. Fish and Wildlife Service's Great Lakes Basin
Ecosystem Team (GLBET) has taken on the charge of providing leadership to coordinate and
improve the protection and management of the islands of the Great Lakes. The GLBET island
initiative includes the coordination and compilation of island geospatial data and information,
developing standardized survey/monitoring protocols, holding an island workshop in the fall of
2002 to incorporate  input from partners  for addressing the Great Lakes Island indicator needs,
and completion of a Great Lakes  Island  Conservation Strategic Plan.

A subset of the GLBET formed the Binational Collaborative for the Conservation of Great Lakes
Islands. Recently, the Collaborative received a habitat grant from the Environmental Protection
                         Draft for Discussion at SOLEC 2006

-------
Agency's Great Lakes National Program Office (GLNPO) to develop a framework for the
binational conservation of Great Lakes islands. With this funding, the team has developed:
1) An island biodiversity assessment and ranking system (based on a subset of biodiversity
parameters) that will provide a foundation to prioritize island conservation;
2) A freshwater island classification system; and
3) A suite of indicators that can be monitored to assess change, threats, and progress towards
conservation of Great Lakes islands biodiversity.

To date, the Collaborative has tentatively proposed ten state, five pressure, and two response
indicators. We anticipate developing additional response indicators and may be able to
incorporate existing Great Lakes response indicators. The island indicators are still being
evaluated and are not final. Final selection of indicators will take place in 2005-2006,  and will be
based on relevance, feasibility, response variability, and interpretation and utility. The
Collaborative is currently drafting the Framework for the Binational Conservation of Great Lakes
Islands, which is expected to be submitted for public and peer review in the fall of 2006.

The information conveyed by a science-based suite of island indicators will help to focus
attention and management efforts to best conserve these unique and globally significant Great
Lakes resources.

Acknowledgments
Authors: Richard H. Greenwood, U.S. Fish and Wildlife Service, Great Lakes Basin Ecosystem
Team Leader and Liaison to U.S. Environmental Protection Agency, Great Lakes National
Program Office, Chicago, IL;
Dr. Karen E. Vigmostad, Great Lakes Policy Analyst Ecosystem Team, Northeast-Midwest
Institute, Washington, DC;
Megan M. Seymour, Wildlife Biologist, U.S. Fish and Wildlife Service, Great Lakes Basin
Ecosystem Team Island Committee Chair, Ecological Services Field Office, Reynoldsburg, OH;
Dr. Francesca Cuthbert, Dept. of Fisheries, Wildlife, and Conservation Biology, University of
Minnesota, St. Paul, MN;
Dr. David Ewert, Director of Conservation Science, Great Lakes Program, Nature Conservancy,
Lansing, MI;
Dan Kraus, Coordinator of Conservation Science, Ontario Region of Nature Conservancy of
Canada, Guelph, ON; and
Linda R. Wires, Research Associate, Dept. of Fisheries, Wildlife, and Conservation Biology,
University of Minnesota, St. Paul, MN.

Data Sources
Susan Crispin, Director, Montana Natural Heritage Program, Helena, MT. Ph: 406-444-5434,
scrispin@state.mt.us.

Bruce Manny and Greg Kennedy, U.S. Geological Survey, Great Lakes Science Center, 1451
Green Road Ann Arbor, MI 48105-2807. Ph: 734-214-7213, bruce_manny@usgs.gov or
gregory_kennedy@usgs.gov.
                         Draft for Discussion at SOLEC 2006

-------

                                ! w              4jrs:;:,- •  ,  ^^fcw-s5
                                w'=ttv?9*"i«'""'"*',/•'  "'  ' *•" ""   ,. '   ^Kp-1. :*•••,.
Dr. Judy Soule, Director, U.S. Network Partnerships, Nature Serve, East Lansing, MI. Ph: 517-
381-5310,
judy_soule@natureserve.org.

Dr. Karen E. Vigmostad, Great Lakes Policy Analyst, Northeast-Midwest Institute, Washington,
DC. 20003. Ph: 202-464-4016, kvigmostad@nemw.org.

List of Tables
Table 1. Biodiversity and Threats Scores for Great Lakes Islands (Canada only), by coastal
environment.
Source: Framework for Binational Conservation of Great Lakes Islands

List of Figures
Figure 1. Islands of the Great Lakes

Last updated
SOLEC 2006
                        Draft for Discussion at SOLEC 2006

-------
Costal
Environment
Georgian Bay 1
Georgian Bay 2
Georgian Bay 3
Georgian Bay 4
Georgian Bay 5
Georgian Bay 6
Lake Erie 1
Lake Erie 2
Lake Erie 3
Lake Erie 4
Lake Erie 5
Lake Erie 6
Lake Erie 7
Lake Erie 8
Lake Huron 1
Lake Huron 2
Lake Huron 3
Lake Ontario 1
Lake Ontario 2
Lake Ontario 3
Lake Ontario 4
Lake Ontario 5
Lake Superior 1
Lake Superior 2
Lake Superior 3
Lake Superior 4
Lake Superior 5
St. Clair 1
St. Clair 2
St. Clair 3
St. Clair 4
St. Clair 5
St. Lawrence 1
No.
Individual
Islands
3992
17615
38
36
290
225
0
15
2
66
2
1461
21
17
887
31
8
0
9
34
74
603
167
1228
495
77
246
21
234
53
1
41
337
No. Islands/
Complexes
595
848
22
18
90
119
0
15
2
13
2
30
18
4
173
19
5
0
7
13
32
171
117
459
160
28
45
11
25
11
1
14
111
Biodiversity Score
Mean
85.2
90.2
93.9
95.8
103.6
92.8
0
151.7
92.5
198.9
90.5
203.4
88.4
144.5
103.4
85.0
127.0
0
108.6
127.0
131.5
114.1
84.6
81.2
71.7
97.2
93.6
119.7
162.2
160.3
116
162.1
92.4
Range
0-345
0-290
57-244
47-195
39-300
46-401
0
87-388
91-94
154-340
87-94
81-333
57-143
96-164
39-490
57-137
114-145
0
90-148
86-190
83-231
44-302
39-290
37-288
40-195
57-253
49-275
84-187
92-336
102-239
116
79-231
44-211
Threat Score
Mean
1.3
11.8
8.2
5.7
4.0
9.7
0
11.2
1.0
4.8
2.0
9.7
7.7
2.3
8.2
3.4
2.8
0
2.3
7.0
3.3
3.7
2.2
2.0
2.4
3.3
8.8
22.1
9.2
6.0
2
11.5
19.5
Range
0-65
0-52
1-46
1-33
1-44
1-581
0
1-88
1
1-32
1-3
1-41
1-42
1-6
1-179
1-22
1-4
0
1-5
1-27
1-22
1-143
1-25
1-40
1-28
1-26
1-138
1-46
1-68
1-36
2
1-36
1-81
Table 1. Biodiversity and Threats Scores for Great Lakes Islands (Canada only), by coastal
environment.
Source: Framework for Binational Conservation of Great Lakes Islands
                        Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
                                                         Framework for the
                                                      Binational Conservation
                                                       of Great Lakes Islands

                                                           Lakes and Connecting
                                                               Channels

Figure I. Islands of the Great Lakes.
Source: Framework for the Binational Conservation of Great Lakes Islands
                     Draft for Discussion at SOLEC 2006

-------
Extent of Hardened Shoreline
Indicator #8131

Assessment: Mixed, Deteriorating

Purpose
  To assess the extent (in kilometres) of hardened shoreline
along the Great Lakes through construction of sheet piling, rip
rap, or other erosion control structures.

Ecosystem Objective
Shoreline conditions should be healthy enough to support aquat-
ic and terrestrial plant and animal life, including the rarest
species.

State of the Ecosystem
Background
Anthropogenic hardening of the shorelines not only directly
destroys natural features and biological communities, it also has
a more subtle but still devastating impact. Many of the biologi-
cal communities along the Great Lakes are dependent upon the
transport of shoreline sediment by lake currents. Altering the
transport of sediment disrupts the balance of accretion and ero-
sion of materials carried along the shoreline by wave action and
lake currents. The resulting loss of sediment replenishment can
intensify the effects of erosion, causing ecological and economic
impacts. Erosion of sand spits and other barriers allows
increased exposure of the shoreline and loss of coastal wetlands.
Dune formations can be lost or reduced due to lack of adequate
nourishment of new sand to replace sand that is carried away.
Increased erosion also causes property damage to shoreline
properties.

Status of Hardened Shorelines in the Great Lakes
The National Oceanic and Atmospheric Administration (NOAA)
Medium Resolution Digital Shorelines dataset was compiled
between 1988 and 1992. It contains data on both the Canadian
and U.S. shorelines, using aerial photography from 1979 for the
state of Michigan and from 1987-1989 for the rest of the basin.

From this dataset, shoreline hardening has been categorized for
each Lake and connecting channel (Table  1). Figure 1 indicates
the percentages of shorelines in each of these categories. The St.
Clair, Detroit, and Niagara Rivers have a higher percentage of
their shorelines hardened than anywhere else in the basin.

Of the  Lakes themselves, Lake Erie has the highest percentage
of its shoreline hardened, and Lakes Huron and Superior have
the lowest (Figure 2). In 1999, Environment Canada assessed
change in the extent of shoreline hardening along about 22 kilo-
metres of the Canadian shoreline of the St. Clair River from
1991-1992 to 1999. Over the eight-year period,  an additional 5.5
             All 5 Lakes
All Connecting
  Channels
Entire Basin
              i 0-15% Hardened
              i 40-70% Hardened
     n  15-40% Hardened
     •  70-100% Hardened
 Figure 1. Shoreline hardening in the Great Lakes compiled
 from 1979 data for the state of Michigan and 1987-1989 data
 for the rest of the basin.
 Source: Environment Canada and National Oceanic and
 Atmospheric Administration
kilometers (32%) of the shoreline had been hardened. This is
clearly not representative of the overall basin, as the St. Clair
River is a narrow shipping channel with high volumes of Great
Lakes traffic. This area also has experienced significant develop-
ment along its shorelines, and many property owners are harden-
ing the shoreline to reduce the impacts of erosion.

zb
0)
= on
o
•= m-
(0 10
T3
2 m
c 10
0)
1 5
« 0
a? o



—
•=,_•=,_ n



1 	
i — i
h
///s////
V 
-------
                                                                                           2007
Lake / Connecting
Channel
Lake Superior
St. Marys River
Lake Huron
Lake Michigan
St. Clair River
Lake St. Clair
Detroit River
Lake Erie
Niagara River
Lake Ontario
St. Lawrence Seaway
All 5 Lakes
All Connecting Channels
Entire Basin
70 - 100%
Hardened
3.1
2.9
1.5
8.6
69.3
11.3
47.2
20.4
44.3
10.2
12.6
5.7
15.4
7.6
40 - 70%
Hardened
1.1
1.6
1.0
2.9
24.9
25.8
22.6
11.3
8.8
6.3
9.3
2.8
11.5
4.6
15-40%
Hardened
3.0
7.5
4.5
30.3
2.1
11.8
8.0
16.9
16.7
18.6
17.2
10.6
14.0
11.3
0 - 15%
Hardened
89.4
81.3
91.6
57.5
3.6
50.7
22.2
49.1
29.3
57.2
54.7
78.3
54.4
73.5
Non-structural
Modifications
0.03
1.6
1.1
0.1
0.0
0.2
0.0
1.9
0.0
0.0
0.0
0.6
0.3
0.5
Unclassified
3.4
5.1
0.3
0.5
0.0
0.1
0.0
0.4
0.9
7.7
6.2
2.0
4.4
2.5
Total
Shoreline
(km)
5,080
707
6,366
2,713
100
629
244
1,608
184
1,772
2,571
17,539
4,436
21 ,974
Table 1 . Percentages of shorelines in each category of hardened shoreline. The St. Clair, Detroit and Niagara
Rivers have a higher percentage of their shorelines hardened than anywhere else in the basin. Lake Erie has the
highest percentage of its shoreline hardened, and Lakes Huron and Superior have the lowest.
Source: National Oceanic and Atmospheric Administration
Pressures
Shoreline hardening is generally not reversible, so once a section
of shoreline has been hardened it can be considered a permanent
feature. As such, the current state of shoreline hardening likely
represents the best condition that can be expected in the future.
Additional stretches of shoreline will continue to be hardened,
especially during periods of high lake levels. This additional
hardening in turn will starve the downcurrent areas of sediment
to replenish that which eroded away, causing further erosion and
further incentive for additional hardening. Thus, a cycle of
shoreline hardening can progress along the shoreline. The future
pressures on the ecosystem resulting from existing hardening
will almost certainly continue, and additional hardening is likely
in the future. The uncertainly is whether the rate can be reduced
and ultimately halted. In addition to the economic costs, the eco-
logical costs are of concern, particularly the percent further lost
or degradation of coastal wetlands and sand dunes.

Management Implications
Shoreline hardening can be controversial, even litigious, when
one property owner hardens a stretch of shoreline that may
increase erosion of an adjacent property. The ecological impacts
are not only difficult to quantify as a monetary equivalent, but
difficult to perceive without an understanding of sediment trans-
port along the lakeshores. The importance of the ecological
process of sediment transport needs to  be better understood as an
incentive to reduce new shoreline hardening. An educated public
is critical to ensuring wise decisions about the stewardship of the
Great Lakes basin  ecosystem, and better platforms for getting
understandable information to the public is needed.
Acknowledgments
Authors: John Schneider, U.S. Environmental Protection
Agency, Great Lakes National Program Office, Chicago, IL;
Duane Heaton, U.S. Environmental Protection Agency, Great
Lakes National Program Office, Chicago, IL; and
Harold Leadlay, Environment Canada, Environmental
Emergencies Section, Downsview, ON.

Sources
The National Geophysical Data Center, National Oceanic and
Atmospheric Administration (NOAA). Medium resolution digital
shoreline, 1988-1992. In Great Lakes Electronic Environmental
Sensitivity Atlas, Environment Canada, Environmental Protection
Branch, Downsview, ON.
Authors' Commentary
It is possible that current aerial photography of the shoreline will
be interpreted to show more recently hardened shorelines. Once
more recent data provides information on hardened areas,
updates may only be necessary basin-wide every 10 years, with
monitoring of high-risk areas every 5 years.

Last Updated
State of the Great Lakes 2001
270

-------
Contaminants Affecting Productivity of Bald
Eagles
Indicator #8135

Assessment: Mixed, Improving

Purpose
  To assess the number of territorial pairs, success rate of nest-
ing attempts, and number of fledged young per territorial pair as
well as the number of developmental deformities in young bald
eagles;
  To measure concentrations of persistent organic pollutants and
selected heavy metals in unhatched bald eagle eggs and in
nestling blood and feathers; and
  To infer the potential for harm to other wildlife caused by eat-
ing contaminated prey items.

Ecosystem Objectives
This indicator supports annexes  2, 12, and 17 of the Great Lakes
Water Quality Agreement.

State of the Ecosystem
As the top avian predator in the  nearshore and tributary areas of
the Great Lakes, the bald eagle integrates contaminant stresses,
food availability, and the availability of relatively undeveloped
habitat areas over most portions of the Great Lakes shoreline. It
serves as an indicator of both habitat quantity and quality.
Concentrations of organochlorine chemicals are decreasing or
stable but still above No Observable Adverse Effect
Concentrations (NOAECs) for the primary organic contami-
nants, dichlorodiphenyl-dichloroethene (DDE) and polychlori-
nated biphenyls (PCBs). Bald eagles are now distributed exten-
sively along the shoreline of the Great Lakes (Figure 1). The
number of active bald eagle territories has increased markedly
from the depths of the population decline caused by DDE
(Figure 2). Similarly, the percentage of nests producing one or
more fledglings (Figure 3) and the number of young produced
per territory (Figure 4) have risen. The recovery of reproductive
output at the population level has followed similar patterns in
each of the lakes, but the timing has differed between the vari-
ous lakes. Lake Superior recovered first, followed by Erie  and
Huron, and most recently, Lake Michigan. An active territory
has been reported from Lake Ontario. Established territories in
most areas are now producing one or more young per territory
indicating that the population is healthy and capable of growing.
Eleven developmental deformities have been reported in bald
eagles within the Great Lakes watershed; five of these were from
territories potentially influenced by the Great Lakes.
MINNESOTA
 Figure 1. Approximate nesting locations of bald eagles (in red) along
 the Great Lakes shorelines, 2000.
 Source: W. Bowerman, Clemson University, Lake Superior LaMPs,
 and for Lake Ontario, Peter Nye, and N.Y. Department of
 Environmental Conservation
        200
        180
        160
        140
     o  120
     Q.
        100
                                                                       80
                                                                       60
                                                                       40
                                                                       20
                                                                       0
                                                                                                Year
                                                                                 -Superior-*-Michigan-^- Huron-*- Erie A Ontario
                                                                   Figure 2. Average number of occupied bald eagle territories per
                                                                   year by lake.
                                                                   Source: David Best, U.S. Fish and Wildlife Service; Pamela
                                                                   Martin, Canadian Wildlife Service; and Michael Meyer,
                                                                   Wisconsin Department of Natural Resources
    Pressures
    High levels of persistent contaminants in bald eagles contin-
    ue to be a concern for two reasons. Eagles are relatively rare
    and contaminant effects on individuals can be important to
    the well-being of local populations. In addition, relatively
    large habitat units are necessary to support eagles and con-
    tinued development pressures along the shorelines of the

                                                       271

-------
                                                OF   THE   GREAT
                                                                                         2007
2
I
3
a.
90

80

70

60

50

40

30

20

10

 0
         Superior    Michigan
                                Huron
                                           Erie
                                                      Ontario
D 1962-1 966
D 1967-1 971
D 1972-1 976
D 1977-1 981
D 1982-1 986
D 1987-1991
D 1992-1 996
D 1997-2001
Figure 3. Average percentage of occupied territories fledging at
least one young.
Source: David Best, U.S. Fish and Wildlife Service; Pamela Martin.
Canadian Wildlife Service; and Michael Meyer, Wisconsin
Department of Natural Resources
1.6
1.4
j? 1.2
•§ 1.0
% 0.8
| 0.6
Z 0.4
0.2
0







,-r
n



rJl


r





n




rl









I -



















Superior Michigan Huron Erie Ontario
• 1962-1966 D 1972-1 976
D1967-1971 D1977-1981




• 1982-1986 D 1992-1 996
D1987-1991 D1997-2001


Figure 4. Average number of young fledged per occupied
territory per year.
Source: David Best, U
S. Fish
and Wildlife Service;
Pamela Martin, Canadian Wildlife Service; and Michael
Meyer, Wisconsin Department
of Natural Resources
  Great Lakes constitute a concern. The interactions of contami-
  nant pressures and habitat limitations are unknown at present.
  There are still several large portions of the Great Lakes shore-
  line, particularly around Lake Ontario, where the bald eagle has
  not recovered to its pre-DDE status despite what appears to be
  adequate habitat in many areas.

  Management Implications
  The data on reproductive rates in the shoreline populations of

  272
Great Lakes bald eagles imply that widespread effects of persist-
ent organic pollutants have decreased. However, there are still
gaps in this pattern of reproductive recovery that should be
explored and appropriate corrective actions taken. In addition.
information on the genetic structure of these shoreline popula-
tions is still lacking. It is possible that further monitoring will
reveal that these populations are being maintained from surplus
production from inland sources rather than from the productivity
of the  shoreline birds  themselves. Continued expansion of these
populations into previously unoccupied areas is encouraging and
might  indicate several things; there is still suitably undeveloped
habitat available, or bald eagles are adapting to increasing alter-
ation of the available habitat.

Acknowledgments
Authors: Ken Stromborg, U.S. Fish & Wildlife Service;
David Best, U.S. Fish & Wildlife Service;
Pamela Martin, Canadian Wildlife Service;  and
William Bowerman, Clemson University.

Additional data were contributed by:  Ted Armstrong,  Ontario
Ministry of Natural Resources; Lowell Tesky, Wisconsin
Department of Natural Resources; Cheryl Dykstra, Cleves, OH;
Peter Nye, New York Department of Environmental
Conservation; Michael Hoff, U.S. Fish and Wildlife Service.
John Netto, U.S. Fish  & Wildlife Service assisted with computer
support.
                                                            Authors' Commentary
                                                            Monitoring the health and contaminant status of Great Lakes
                                                            bald eagles should continue across the Great Lakes basin. Even
                                                            though the worst effects of persistent bioaccumulative pollutants
                                                            seem to have passed, the bald eagle is a prominent indicator
                                                            species that integrates effects that operate at a variety of levels
                                                            within the ecosystem. Symbols such as the bald eagle are valu-
                                                            able for communicating with the public. Many agencies continue
                                                            to accomplish the work of reproductive monitoring that results
                                                            in compatible data for basin-wide assessment. However, the
                                                            Wisconsin Department of Natural Resources and Ohio
                                                            Department of Natural Resources programs are diminished as
                                                            the result of budgetary constraints, while Michigan Department
                                                            of Environmental Quality, New York State  Department of
                                                            Environmental Conservation and Ontario Ministry of Natural
                                                            Resources programs will continue for the near future. In the very
                                                            near future, when the bald eagle is removed from the list of
                                                            threatened species in the United States, existing monitoring
                                                            efforts may be severely curtailed. Without the required field
                                                            monitoring data, overall assessments of indicators like the bald
                                                            eagle will be impossible. Part of the problem with a lessened
                                                            emphasis on wildlife monitoring by governmental agencies is
                                                            the failure of initiatives such as the State of the Lakes Ecosystem

-------
                               OF   T H
Conference (SOLEC) process to identify and designate programs
that are essential in order to ensure that data continuity is main-
tained. Two particular needs for additional data also exist. There
is no basin-wide effort directed toward assessing habitat suitabil-
ity of shoreline areas for bald eagles. Further, it is not known to
what degree the shoreline populations depend on recruiting sur-
plus young from healthy inland populations to maintain the cur-
rent rate of expansion or whether shoreline populations are self-
sustaining.

Last Updated
State of the Great Lakes 2005
                                                                                                                       273

-------
Population Monitoring and Contaminants
Affecting the American Otter
Indicator #8147

Assessment: Mixed, Trend Not Assessed

Purpose
  To directly measure the contaminant concentrations found in
American otter populations within the Great Lakes basin; and
  To indirectly measure the health of Great Lakes habitat,
progress in Great Lakes ecosystem management, and/or concen-
trations of contaminants present in the Great Lakes.

Ecosystem Objective
As a society we have a moral responsibility to sustain healthy
populations of American otter in the Great Lakes/St. Lawrence
basin. American otter populations in the upper Great Lakes
should be maintained, and restored as sustainable populations in
all Great Lakes coastal zones, lower Lake Michigan, western
Lake Ontario, and Lake Erie watersheds and shorelines. Great
Lakes shoreline and watershed populations of American otter
should have an annual mean production of >2 young/adult
female; and concentrations of heavy metal and organic contami-
nants in otter tissue samples should be less than the No
Observable Adverse Effect Level found in tissue
sample from mink. The importance of the American
otter as a biosentinel is related to International Joint
Commission Desired Outcomes 6: Biological
Community Integrity and Diversity, and 7: Virtual
Elimination of Inputs of Persistent Toxic Chemicals.
State of the Ecosystem
A review of State and Provincial otter population
data indicates that primary areas of population sup-
pression still exist in southern Lake Huron water-
sheds, lower Lake Michigan and most Lake Erie
watersheds.  Data provided from New York
Department of Environmental Conservation
(NYDEC) and Ontario Ministry of Natural
Resources (OMNR) suggest that otter are almost
absent in western Lake Ontario (Figure  1). Most
coastal shoreline areas have more suppressed popu-
lations than interior zones.

Areas of otter population suppression are directly
related to human population centers and subsequent
habitat loss, and also to elevated contaminant con-
centrations associated with human activity. Little
statistically-viable population data exist for the
Great Lakes populations, and all suggested popula-
tion levels illustrated were determined from coarse
population assessment methods.
274
            Pressures
            American otters are a direct link to organic and heavy metal con-
            centrations in the food chain. It is a relatively sedentary species
            and subsequently synthesizes contaminants from smaller areas
            than wider-ranging organisms, e.g. bald eagle. Contaminants are
            a potential and existing problem for many otter populations
            throughout the Great Lakes. Globally, indications of contaminant
            problems in otter have been noted by decreased population lev-
            els, morphological abnormalities (i.e. decreased baculum length)
            and decline in fecundity. Changes in the species population and
            range are also representative of anthropogenic riverine and
            lacustrine habitat alterations.

            Management Implications
            Michigan and Wisconsin have indicated a need for an independ-
            ent survey using aerial survey methods to index otter popula-
            tions in their respective jurisdictions. Minnesota has already
            started aerial population surveys for otter. Subsequently, some
            presence-absence data may be available for Great Lakes water-
            sheds and coastal populations in the near future. In addition, if
            the surveys are conducted frequently, the trend data may become
            useful. There  was agreement among resource managers on the
            merits of aerial survey methods to index otter populations,
            although these methods are only appropriate  in areas with ade-
            quate snow cover. NYDEC, OMNR, Federal jurisdictions and
                                                  Stable
                                                  Non-stable
                                                  Almost Absent
                                                  Extirpated
Figure 1. Great Lakes shoreline population stability estimates for the American
otter.
Source: Thomas CJ. Doolittle, Bad River Band of Lake Superior Tribe of
Chippewa Indians

-------
Tribes on Great Lakes coasts indicated strong needs for future
assessments of contaminants in American otter. Funding, other
than from sportsmen, is needed by all jurisdictions to assess
habitats and contaminant levels, and to conduct aerial surveys.

Acknowledgments
Thomas CJ. Doolittle, Bad River Band of Lake Superior Tribe
of Chippewa Indians, Odanah, WI.

Sources
Bishop, P., Gotie, R., Penrod, B., and Wedge, L. 1999. Current
status of river otter management in New York. New York State
Department of Environmental Conservation, Otter management
team, Delmar, New York.

Bluett, R.D. 2000. Personal Communication. Illinois Department
of Natural Resources, Springfield, IL.

Bluett, R.D., Anderson, E.A., Hubert, G.F., Kruse, G.W., and
Lauzon, S.E. 1999. Reintroduction and status of the river otter
(Lutra canadensis) in Illinois. Transactions of the Illinois State
Academy of Science 92(1 and 2):69-78.

Brunstrom, B., Lund, B., Bergman, A., Asplund, L.,
Athanassiadis, L, Athanasiadou, M., Jensen, S., and  Orberg, J.
2001. Reproductive toxicity in mink (Mustela vison) chronically
exposed to environmentally relevant polychlorinated biphenyl
concentrations. Environ. Toxicol. Chem. 20:2318-2327.

Chapman, J.A., and Pursley,  D. (eds.). Worldwide furbearers
conference proceedings. Worldwide Furbearer Conference, Inc.
Frostburg, MD, pp.1752-1780.

Dawson, N. 2000. Personal Communication. Ontario Ministry of
Natural Resources, Northwest Region. Thunder Bay, ON.

Dwyer, C.P. 2000a. Personal Communication. Ohio  Division of
Wildlife, Oak Harbor, OH.

Dwyer, C.P. 2000b. Population assessment and distribution of
river otters following their reintroduction into Ohio. Crane
Creek Wildlife Experiment Station, Ohio Division of Wildlife,
Oak Harbor, OH.

Foley, F.E., Jackling, S J., Sloan, R.J., and Brown, M.K. 1988.
Organochlorine and mercury residues in wild mink and otter:
comparison with fish. Environ. Toxicol. Chem. 7:363-374.

Friedrich, P.D. 2000. Personal Communication. Michigan
Department of Natural Resources. East Lansing, MI.

Halbrook, R.S., Jenkins, J.H., Bush, P.B., and Seabolt, N.D.
1981. Selected environmental contaminants in river otters (Lutra
canadensis) of Georgia and their relationship to the possible
decline of otters in North America. In Proc. Worldwide
Furbearer Cong., pp. 1752- 1762, Worldwide Furbearer
Conference, Inc.

Hammill, J. 2000. Personal Communication. Michigan
Department of Natural Resources. Crystal Falls, MI.

Henny, C.J., Blus, L.J., Gregory, S.V., and Stafford, C.J. 1981.
PCBs and organochorine pesticides in wild mink and river otters
from Oregon. In Proc. Worldwide Furbearer Cong., pp. 1763-
1780.

Hochstein, J., Bursian, S., and Aulerich, R. 1998. Effects of
dietary exposure to 2,3,7,8-tetrachlorodibenzo-p-dioxin in adult
female mink (Mustela vison). Arch. Environ. Contam. Toxicol.
35:348-353.

Johnson, S. 2000. Personal Communication.  Indiana Department
of Natural Resources. Bloomington, IN.

Johnson, S.A., and Berkley, K.A. 1999. Restoring river otters in
Indiana. Wildlife Society Bull. 27(2):419-427.

Johnson, S.A., and Madej, R.F. 1994. Reintroduction  of the river
otter in Indiana - a feasibility study. Indiana Department of
Natural Resources, Bloomington, IN.

Kannan, K., Blankenship, A., Jones, P., and Giesy, J.  2000.
Toxicity reference values for the toxic effects of polychlorinated
biphenyls to aquatic mammals. Human Ecological Risk
Assessment 6:181-201.

Kautz, M. 2000. Personal Communication. New York
Department of Environmental Conservation, Delmar,  NY.

Leonards,  P., de Vries, T., Minnaard, W., Stuijfzand, S., de
Voogt, P., Cofino, W., van  Straalen, N., and van Hattum, B.
1995. Assessment of experimental data on PCB induced repro-
duction inhibition in mink, based on an isomer- and congener-
specific approach using 2,3,7,8-tetrachlorodibenzo-p-dioxin
toxic equivalency. Environ. Toxicol. Chem. 14:639-652.

Mason, C. 1989. Water pollution and otter distribution: a review.
Lutra 32:97-131.

Mason, C., and Macdonald, S. 1993. Impact of organochlorine
pesticide residues and PCBs on otters (Lutra  lutra): a  study from
western Britain. Sci. Total Environ. 138:127-145.
Mayack, D.T. 2000. Personal Communication. New York
                                                                                                                     275

-------
                                                                                     E S   2007
Department of Environmental Conservation, Gloversville, NY.

Michigan Department of Natural Resources. 2000a, Distribution
of otter harvest by section 1998-99. East Lansing, MI.

Michigan Department of Natural Resources. 2000b. River otter
reproductive and harvest data 1995-1999. East Lansing, MI.

New York State Department of Environmental Conservation.
1998-99. Furbearer harvest by county and region. Albany, NY.

Ohio Division of Wildlife. 1999-2000. Watersheds with river
otter observations. Oak harbor, OH.

Olson, J. 2000. Personal Communication. Furbearer Specialist,
Wisconsin Department of Natural Resources, Park Falls, WI.

Ontario Ministry of Natural Resources. 2000. Ontario furbearer
population ranks through trapper questionnaires by Wildlife
Assessment Unit. Thunder Bay, ON.

Roos, A., Greyerz, E., Olsson, M., and Sandegren, F. 2001. The
otter (Lutra lutra) in Sweden? Population trends in relation to
3DDT and total PCB concentrations during 1968-99. Environ.
Pollut. 111:457-469.

Route, W.T., and Peterson, R.O.1988. Distribution and abun-
dance of river otter in Voyageurs National Park, Minnesota.
Resource Management Report MWR-10. National Park Service,
Omaha, NE.

Sheffy, T.B., and St. Amant, J.R. 1982. Mercury burdens in
furbearers in Wisconsin. J.  Wildlife Manage. 46:1117-1120.

Wisconsin Department of Natural Resources. 2000a. Distribution
of otter harvest by management unit 1998-99. Madison, WI.

Wisconsin Department of Natural Resources. 2000b. Otter popu-
lation model statewide 1982-2005. Madison, WI.

Wisconsin Department of Natural Resources. 1979-1998.
Summary of otter reproductive information. Madison, WI.

Wren,  C.  1991. Cause-effect linkages between chemicals and
populations of mink (Mustela vison) and otter (Lutra canadensis)
in the Great Lakes basin. J. Toxicol Environ. Health 33:549-585.
or provincial-wide scales. Most coarse population assessment
methods were developed to assure that trapping was not limiting
populations and that otter were simply surviving and reproduc-
ing in their jurisdiction. There was little work done on finer spa-
tial scales using otter as an indicator of ecosystem heath.
In summary, all state and provincial jurisdictions only marginal-
ly index Great Lakes watershed populations by presence-absence
surveys, track surveys, observations, trapper surveys, population
models, aerial surveys, and trapper registration data.

Michigan has the most useful spatial data that could index the
largest extent of Great Lakes coastal populations due to their
registration requirements. Michigan registers trapped otter to an
accuracy of 1 square mile. However, other population measures
of otter health, such as reproductive rates, age and morphologi-
cal measures, are not tied to spatial data in any jurisdiction, but
are pooled together for entire jurisdictions. If carcasses are col-
lected for necropsy, the samples are usually  too small to accu-
rately define health of Great Lakes coastal otter verses interior
populations. Subsequently, there is a large need to encourage and
fund resource management agencies to streamline data for tar-
geted population and contaminant research on Great Lakes otter
populations, especially in coastal zones.

Last Updated
State of the Great Lakes 2003
Authors' Commentary
All State and Provincial jurisdictions use different population
assessment methods, making comparisons difficult. Most juris-
dictions use survey methods to determine populations on state-
276

-------

                                                   ^ ^  .,•_
                                  ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''  './ ''l3
                                                           ^!?
Biodiversity Conservation Sites
Indicator #8 164
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Not Assessed
                   Undetermined
                   Information on Biodiversity Conservation sites is limited at this time
                   making it difficult to assess the status and trend of this indicator.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Not Assessed
                   Undetermined
                   Not available at this time.
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Not Assessed
                   Undetermined
                   Not available at this time.
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Not Assessed
                   Undetermined
                   Not available at this time.
                   Not Assessed
                   Undetermined
                   Not available at this time.
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                   Not Assessed
                   Undetermined
                   Not available at this time.
                        Draft for Discussion at SOLEC 2006

-------
Purpose
• To assess and monitor the biodiversity of the Great Lakes watershed.

Ecosystem Objective
The ultimate goal of this indicator is to generate and implement a distinct conservation goal for
each target species, natural community type and aquatic system type within the Great Lakes
basin.  Through establishing the long-term survival of viable populations, the current level of
biodiversity within the region can be maintained, or even increased. This indicator supports
Great Lakes Quality Agreement Annexes 1, 2 and  11.

State of the Ecosystem
Background
In 1997, the Great Lakes Program of The Nature Conservancy (TNC) launched an initiative to
identify high priority biodiversity conservation sites in the Great Lakes region. Working with
experts from a variety of agencies, organizations, and other public and private entities throughout
the region, a collection of conservation targets was identified.  These targets, which represented
the full range of biological diversity within the region, consisted of globally rare plant and animal
species, naturally occurring community types within the ecoregion, and all aquatic system types
found in the Great Lakes watershed.

In order to ensure the long-term survival of these conservation targets, two specific questions
were asked: how many populations or examples of each target are necessary to ensure its long-
term survival in the Great Lakes ecoregion, and how should these populations or examples be
distributed in order to capture the target's genetic and ecological variability across the Great
Lakes ecoregion?  Using this information, which is still limited as these questions have not been
satisfactorily answered in the field of conservation biology, a customized working hypothesis, i.e.
conservation goal, was generated for each individual conservation target. Additionally, to
effectively and efficiently achieve these conservation goals, specific portfolio  sites were
identified. These sites, many of which contain more than one individual target, support the most
viable examples of each target, thus aiding in the preservation of the overall biodiversity within
the Great Lakes region.

With support from TNC, the Nature Conservancy of Canada has undertaken a similar initiative,
identifying additional targets, goals, and conservation sites within Ontario, Canada. However, as
the commencement of this project occurred some time after the U.S. counterpart, there is a wide
discrepancy in the information that is currently available.

Status of Biodiversity Conservation Sites in the Great Lakes Basin
Within the U.S. portion of the Great Lakes region,  208 species (51 plant species, 77 animal
species and 80 bird species) were identified.  Of these, 18 plant species and 28 animal species can
be considered endemic (found only in the Great Lakes region) or limited (range is primarily in the
Great Lakes ecoregion, but also extends into one or two other ecoregions).  Furthermore, 24
animals and 14 plants found within the basin are recognized as globally imperiled. Additionally,
274 distinct natural community types are located throughout the ecoregion: 71 of which are
endemic or largely limited to the Great Lakes, while 45 are globally imperiled. The Great Lakes
                         Draft for Discussion at SOLEC 2006

-------

                                       tf"j%*-l''|wSrjp3fe-™=*'  	; ,fei^ * i'i" ''  ./ 'l3™B*j,A
                                                               ^!?r?
watershed also contains 231 aquatic system types, all of which are inextricably connected to the
region, and thus do not occur outside this geographical area.

A total of 501 individual portfolio sites have been designated throughout the Great Lakes region:
280 of which reside fully within the U.S., 213 are located entirely in Canada, while the remaining
8 sites cross international borders.  However, there is an uneven distribution among the
conservation priority sites found in the U.S., as over half are completely or partially located
within the state of Michigan. New York State contains the second greatest number of sites with
56; Wisconsin, 29; Ohio, 25; and Minnesota, 20. Furthermore, 9 sites are located within the state
of Illinois, 7 sites in Indiana, while only 2 sites are found in the state of Pennsylvania (11 sites
cross state borders, while one international and one U.S. site cross more than one border). The
sizes of the selected portfolio sites have a wide distribution, ranging from approximately 60 to
1,500,000 acres; with three-fourths of the sites having areas which are less than 20,000 acres.

The currently established conservation sites provide enough viable examples to fully meet the
conservation goals for 20% of the 128 species and 274 community types described within the
Great Lakes conservation vision. Additionally, under the existing Conservation Blueprint, 80%
of the aquatic systems are sufficiently represented in order to meet their conservation goals.
However, these figures might not present an accurate depiction of the current state of the
biodiversity within the region. Due to a lack of available data for several species, communities,
and aquatic systems, a generalized conservation goal, e.g. "all viable examples" was established
for these targets.  As such, even though the conservation goals may have been met, there might
not be an adequate number of examples to ensure the long-term survival of these targets.

In order to sustain the current level of biodiversity, i.e. number of targets that have met their
conservation goals, attention to the health and overall integrity of the conservation sites must be
maintained.  While approximately 60% of these sites are irreplaceable, these places represent the
only opportunity to protect certain species, natural communities, aquatic  systems, or assemblages
of these targets within the Great Lakes region.  Only 5% of all U.S. sites  are actually fully
protected. Furthermore, 79% of the Great Lakes sites require  conservation attention within the
next ten years, while more than one-third of the sites need immediate attention in order to protect
conservation targets. These conservation actions range from changes in policies affecting land
use, i.e. specific land protection measures (conservation easements or changes in ownership), to
the modification of the management practices currently used.

Pressures
In the U.S., information was obtained from 224 sites regarding pressures associated with the
plants, animals, and community targets within the Great Lakes basin: from this data four main
threats emerged.  The top threat to biodiversity sites throughout the region is currently
development, i.e. urban, residential, second home, and road, as it is affecting approximately two-
thirds of the sites in the form of degradation, fragmentation, or even the complete loss of these
critical habitats.  The second significant threat, affecting the integrity of more than half the sites,
is the impact exerted by invasive species, which includes non-indigenous species such as purple
loosestrife, reed canary grass, garlic mustard, buckthorn, zebra mussels, and exotic fishes, as well
as high-impact, invasive, native species such as deer. Affecting almost half of the U.S. sites,
hydrology alteration, the third most common threat to native biodiversity, includes threats due to
dams, diversions, dikes, groundwater withdrawals, and other changes to the natural flow regime.
                          Draft for Discussion at SOLEC 2006

-------
Finally, recreation (boating, camping, biking, hiking, etc.) is a major threat that affects over 40%
of the sites.

Management Implications
A continuous effort to obtain pertinent information is essential in order to maintain the most
scientifically-based conservation goals and strategies for each target species, community and
aquatic system type within the Great Lakes basin.  Additional inventories are also needed in many
areas to further assess the location, distribution and viability of individual targets, especially those
that are more common throughout the region. Furthermore, even though current monitoring
efforts and conservation actions are being implemented throughout the watershed, they are
generally site-specific or locally concentrated.  A greater emphasis on a regional-wide approach
must be undertaken if the long-term survival of these metapopulations is to be ensured. This
expanded perspective would also assist in establishing region-wide communications, thus
enabling a more rapid and greater distribution of information. However, the establishment of
basin-wide management practices is greatly hindered by the numerous governments represented
throughout this region, (two federal governments,  100 tribal authorities, one province, and eight
states (each with multiply agencies),  13 regional and 18 county municipalities in Ontario, 192
counties in the US and thousands of local governments) and the array of land-use policies
developed by each administrations.  Without additional land protection measures, it will be
difficult to preserve the current sites and implement restoration efforts in order to meet the
conservation goals for the individual conservation targets.

Acknowledgments
Authors: Jeffrey  C. May, U.S. Environmental Protection Agency,  GLNPO Intern.
Contributors: Mary Harkness, The Nature Conservancy.

Data Sources
The Nature Conservancy, Great Lakes Ecoregional Planning Team. 1999.  Great Lakes
Ecoregional Plan: A First Iteration.  The Nature Conservancy, Great Lakes Program, Chicago, IL,
USA. 85pp.  + iv.

 The Nature Conservancy, Great Lakes Ecoregional Planning Team.  1999. Toward a New
Conservation Vision for the Great Lakes Region: A Second Iteration. The Nature Conservancy,
Great Lakes Program, Chicago, IL, USA. 12 pp.

List of Figures
Figure 1: Map of Biodiversity Conservation Sites within the Great Lakes Region.
http://www.nature.org/wherewework/northamerica/greatlakes/files/tnc great lakes  web.pdf

Last updated
SOLEC 2006
                         Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
Figure 1. Map of Biodiversity Conservation Sites within the Great Lakes Region.
http://www.nature.org/wherewework/northamerica/greatlakes/files/tnc great lakes web.pdf
                      Draft for Discussion at SOLEC 2006

-------

                                                      ^ .
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''  './ ''l3i™B*j,Arf(St';j(\f^*
                                                            ^!?r?lB*?
Forest Lands - Conservation of Biological Diversity
Indicator #8500

Note: This indicator includes four components that correspond to Montreal Process Criterion
#1, Indicators 1, 2, 3, and 5.

Indicator #8500 Components:
    Component (1) - Extent of area by forest type relative to total forest area
    Component (2) - Extent of area by forest type and by age-class or successional stage
    Component (3) - Extent of area by forest type in protected area categories
    Component (4) — Extent afforest land conversion, parcelization, and fragmentation (Still
    under development for future analysis; data not presented in this report)

Overall Assessment
           Status:  Mixed
           Trend:  Undetermined
   Primary Factors  There is a moderate distribution of forest types in the Great Lakes
      Determining  basin by age-class and serai stage.  Additional analysis is required by
  Status and Trend  forestry professionals.

Lake-by-Lake Assessment
Lake Superior
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  Data by individual lake basin was not available for the U.S. at this time.
      Determining
  Status and Trend

Lake Michigan
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  Data by individual lake basin was not available for the U.S. at this time.
      Determining
  Status and Trend

Lake Huron
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  Data by individual lake basin was not available for the U.S. at this time.
      Determining
  Status and Trend

Lake Erie
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  Data by individual lake basin was not available for the U.S. at this time.
      Determining
                        Draft for Discussion at SOLEC 2006

-------
  Status and Trend

Lake Ontario
            Status:
            Trend:
   Primary Factors
      Determining
  Status and Trend
Not Assessed
Undetermined
Data by individual lake basin was not available for the U.S. at this time.
Purpose
•To describe the extent, composition and structure of Great Lakes basin forests; and
•To address the capacity of forests to perform the hydrologic functions and host the organisms
and essential processes that are essential to protecting the biological diversity, physical integrity
and water quality of the watershed.

Ecosystem Objective
To have a forest composition and structure that most efficiently conserves the natural biological
diversity of the region

State of the Ecosystem
Component (1):
Forests cover over half (61%), of the land in the  Great Lakes basin. The U.S. portion of the basin
has forest coverage on 54% of its land, while the Canadian portion has coverage on 73% of its
land.

In the U.S. portion of the basin, maple-beech-birch is the most extensive forest type,  representing
7.8 million hectares, or 39%  of total forest area in the basin. Aspen-birch forests constitute the
second-largest forest type, covering 19% of the total. Complete data are available in Table 1 and
are visually represented in Figure 1.

The entire Canadian portion  of the basin is dominated by mixed forest, representing 39% of the
total forest area, followed by hardwoods, covering 23% of the total forest area analyzed from
satellite data, (see Table 2A). The most extensive provincial forest type is the upland mixed
conifer, representing 23% of the forested area available for analysis, followed by the
mixedwoods, tolerant hardwoods, white birch, and poplars, (see Figure 2 and Table 2B).

Implications for the health of Great Lakes forests and the basin ecosystem are difficult to
establish. There is no consensus on how much land in the basin should be forested; much less on
how much land should be covered by each forest type.  Generally speaking, maintenance of the
variety of forest types is important in species preservation, and long-term changes in forest type
proportions are indicative of changes in forest biodiversity patterns, (OMNR 2002).

Comparisons to historical forest cover, although of limited utility in developing landscape goals,
can illustrate the range of variation experienced within  the basin since the time of European
settlement. (See supplemental section entitled "Historical Range of Variation in the Great Lakes
Forests of Minnesota, Wisconsin and Michigan" in the  State of the  Great Lakes 2005 version of
                         Draft for Discussion at SOLEC 2006

-------

                                                        ^  ._
                                    ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''
this indicator report, #8500, for more information). Analysis of similar historical forest cover
data for the entire Great Lakes Basin over the past several years would be useful in establishing
current trends to help assess potential changes to ecosystem function and community diversity.

Component (2):
In the U.S. portion of the basin, the 41-60 and 61-80 year age-classes are dominant and together
represent about 41% of total forest area. Forests 40 years of age and under make up a further
30%, while those in the 100+ year age-classes constitute 7% of total forest area. Table  3 contains
complete U.S. data for age-class distribution as a percentage of forested area within each forest
type.

Because forests are dynamic and different tree species have different growth patterns, age
distribution varies by forest type. In the U.S. portion of the basin, aspen-birch forests tend to be
younger, being more concentrated than other forest types in age classes under 40 years, while the
Oak-Pine forests are more concentrated in the 41-60 and 61-80 year age classes, comparatively.
Spruce-fir and Oak-Hickory forests have a general distribution centered around 41-80 years, but
also have the highest amount of oldest trees, representing about 10% each of total forest area in
the  100+ year age class, (see Figure 3).

This age-class data can serve as a coarse surrogate for the vegetative structure (height and
diameter) of a forest, and can be combined with data from other indicators to provide insight on
forest sustainability.

U.S  data on the extent of forest area by successional or serai stage is not available. Although
certain tree species can be associated with the various successional stages, a standard and
quantifiable protocol for identifying successional stage has not yet been developed. It is expected,
however, that in the absence of disturbance, the area covered by early-successional forest types,
such as aspen-birch, is likely to decline as forests convert to late-successional types, such as
maple-beech-birch.

Canadian forest data for this component is available by serai stage. Ontario's forests have a
distribution leaning towards mature  stages, representing about 50% of the total forest area
analyzed.  Forests in the immature stage make up the next largest group with 20% of the total,
followed by those in late successional with 14%. Every Canadian forest type distribution follows
this general trend except for jack pine. Complete available data for Ontario  can be viewed in
Table 4 and is visually represented in Figure 4.

Although the implications of this age-class and serai stage data for forest and basin health overall
are unclear, some conclusions can be made.  In general, water quality is most affected during the
early successional stages after a disturbance to forest habitats. Nutrient levels in streams can
increase during these times until the surrounding forest is able to mature, (Swank et. al 2000).
The  trend towards mature forests in Canada would therefore mean that area  of the Great Lakes
basin has improved water quality. Alternately, forests with balanced forest type distributions and
diverse successional stages are generally considered more sustainable,  (USDA Forest Service et.
al 2003). The combined effect on ecosystem health resulting from the balance of these  opposing
forces  would need to be determined.
                          Draft for Discussion at SOLEC 2006

-------
Component (3):
In the U.S. basin, 7.8% of forested land is in a protected area category. Among major forest types,
8.9% of maple-beech-birch, 6.6% of aspen-birch and 9.2% of spruce-fir forests are considered to
have protected status. The oak-gum-cypress category has the highest protection rate, with 19.2%
of its forest area protected from harvest. Please refer to Table 1  for complete U.S. data.

In the entire Canadian portion of the basin, 10.6% of forest area, or 1.6 million hectares, are
protected, (see Table 2A). For the region of Ontario that has available forest type data, protection
rates range from 15.4% for red and white pine and 11% for white birch, to 6.4% for poplar and
5.7% for mixed conifer lowland forests, (see Table 2B).

It is difficult to assess the implications of the extent of protected forest area, since there is no
consensus on what the actual  proportion should be. National forest protection rates are  estimated
to be 8.4% in Canada (WWF  1999)  and 14% in the U.S. (USDA Forest Service 2004).  Despite
the fact that updated trend data for protected status is not available at this time for the Great Lakes
basin, earlier analyses have shown a recent general increase in protected areas, (see 2005 version
of this report).

As for the range of variation in protection rates by forest types, protected areas should be
representative of the diversity in forest composition within a larger area. However, defining what
constitutes this "larger area" is problematic. Policymakers often  have a different jurisdiction than
the Great Lakes basin in mind when deciding where to locate protected areas. Also, the tree
species  and forest types found on an individual plot of protected land can change over time due to
successional processes.

Differences among the U.S., Canadian and International Union for the Conservation of Nature
(IUCN) definitions of protected areas should also be noted. The  IUCN standard contains six
categories of protected areas - strict nature reserves/wilderness areas, national parks, natural
monuments, habitat/species management areas, protected landscapes/seascapes, and managed
resource protection areas. The U.S.  defines protected areas as forests "reserved from harvest by
law or administrative regulation," including designated Federal Wilderness areas, National Parks
and Lakeshores, and state designated areas (Smith 2004). Ontario defines protected areas as
national parks, conservation reserves, and its six classes of provincial parks - wilderness, natural
environment, waterway, nature reserve, historical and recreational (OMNR 2002). There is
substantial overlap among the specific U.S., Ontario and IUCN definitions, and a more consistent
classification system would ensure proper accounting of protected areas.

Common to the U.S., Ontario and IUCN definitions is that they only include forests in the public
domain. However, there are privately-owned forests similarly reserved from harvest by land
trusts, conservation easements and other initiatives. Inclusion of these forests under this indicator
would provide a more complete definition of protected forest areas.

Moreover, there is debate on how protected status relates to forest sustainability, water quality,
and ecosystem health. In many cases, protected status was conferred onto forests for their scenic
or recreational value, which may not contribute significantly to conservation or watershed
management goals. On the other hand, forests available for harvest, whether controlled by the
                          Draft for Discussion at SOLEC 2006

-------

                                                        ^ ._
                                    ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''
national forest system, state or local governments, tribal governments, industry or private
landowners, can be managed with the stated purpose of conserving forest and basin health
through the implementation of Best Management Practices and certification under sustainable
forestry programs. (For more information, refer to Indicator #8503, Forest Lands - Conservation
and Maintenance of Soil and Water Resources).

Component (4):
This component is still under development, as consensus still has not been reached on definitions
of forest fragmentation metrics and which ones are therefore suitable for SOLEC reporting. The
proposed structure is split into the forces that drive fragmentation, (land conversion and
parcelization,) and a series of forest spatial pattern descriptions based off of (as yet to be agreed
upon) fragmentation metrics.

Conversion of forest land to other land-use classes is considered to be a major cause of
fragmentation.  Proposed metrics to describe this include the percent of forest lands converted to
and from developed, agricultural, and pasture land uses. Both Canadian and U.S. data are
available and can be obtained from the Ontario Ministry of Natural Resources and the USDA
Natural Resources Conservation Service, Natural Resources Inventory, respectively.

Parcelization of forest lands into smaller privately owned tracks of land can lead to a disruption of
continuous ecosystems and habitats and therefore increased fragmentation. A proposed metric is
the average size of land holdings. Canada does not have available data for this metric, while the
U.S. data should be available through the USDA Forest Service, Forest Inventory and Analysis
Program and the National Woodland Owner Survey.

Data for various fragmentation metrics exists for both Canada and the U.S, but the way these
metrics are viewed is drastically different. According to sources that have compiled U.S. data,
fragmentation, "is viewed as a property of the landscape that contains forest... [as opposed to] a
property of the forest itself," (Riitters et. al 2002). That inconsistency aside, data exists for
Ontario for the following metrics: area, patch density and size, edge, shape, diversity and
interspersion, and core area.  U.S. data exists for patchiness, perforation,  connectivity, edge, and
interior or core forest, and is available from the USDA Forest Service and is also being compiled
by the U.S. EPA.  Substantial discussion is still required to refine these metrics before reporting
and analysis of this component can continue.

Pressures
Urbanization, seasonal home construction and increased recreational use, (driven in part by the
desire of an aging and more affluent population to spend time near natural settings,) are among
the general demands being placed on forest resources nationwide.

Additional disturbances caused by lumber removal and forest fires can also alter the structure of
Great Lakes basin forests.

Management Implications
Increased communication and agreement regarding the definitions and reporting methods for
forest type, successional stage, protected area category and fragmentation metrics between the
United States and Canada would facilitate more  effective basin-wide analyses.
                         Draft for Discussion at SOLEC 2006

-------
Reporting of U.S. forest data according to watershed as opposed to county would enable analysis
by individual lake basin, therefore increasing the data's value in relation to specific water quality
and biodiversity objectives.

Canadian data by forest type and serai stage for the entire Great Lakes basin in Ontario as
opposed to just the Area of the Undertaking (AOU), (see definition below in Comments section,)
would allow for a more complete analysis. This can only be accomplished if managers decide to
extent forest planning inventories into the private lands in the southern regions of the province.

Managing forest lands in ways that protect the continuity of forest cover can allow for habitat
protection and wildlife  species mobility, therefore maintaining natural biodiversity.

Comments from the author(s)
Stakeholder discussion will be critical in identifying pressures and management implications,
particularly those on a localized basis, that are specific to Great Lakes basin forests. These
discussions will add to  longstanding debates on strategies for sustainable forest management.

There are significant discrepancies within and between Canadian and U.S. data that made it
difficult to analyze the  data across the Great Lakes basin as a whole.  The most pervasive
problems are related to the time frame, frequency and location of forest inventories and
differences in metric definitions.

Canadian Great Lakes data for provincial forest type and serai stage is only available in areas of
Ontario where Forest Resources Planning Inventories occur. This region is commonly referred to
as the Area of the Undertaking (AOU) and only represents about 72% of Ontario's total Great
Lakes basin land area.  The remainder of Ontario's forests (and therefore Ontario as a whole) can
only be analyzed using satellite data, which is meant for general land use/land cover analysis and
does not have a fine enough resolution to allow for more detailed investigation.

Forest inventory time frames for the U.S. also have an effect on data consistency. Although the
2002 RPA assessment was used as the data source for the U.S. portion of this report, it actually
draws data from a compilation of numerous  state inventory years as follows: Illinois (1998),
Indiana (1998), Michigan (1993), Minnesota (1990), New York (1993), Ohio (1993),
Pennsylvania (1989), and Wisconsin (1996).  A re-analysis of U.S. Great Lakes basin forests with
data from the same time frame would be useful.

Also, the U.S. data provided for this report was compiled by county and not by watershed, so the
area of land analyzed is not necessarily completely within the Great Lakes basin and all related
values are therefore skewed. This factor also made it impossible to represent the  data by
individual lake basin. Additional GIS analysis of the raw inventory data would be required to
provide forest data by watershed.

Definition of forest type differs between the U.S. and Canada as well.  In the U.S., forest cover
type is done according to the predominant tree species and is divided into the nine major groups
represented in this report.  The Canadian provincial forest type classifications, (for which data
                         Draft for Discussion at SOLEC 2006

-------

was available for this report,) however, are based on a combination of ecological factors
including dominant tree species, understory vegetation, soil, and associated tree species, (OMNR
2002). The definitions of each provincial forest type are available in Table 5.  Standardization of
forest type definitions between the U.S. and Ontario would be necessary for analysis across the
entire Great Lakes basin.

As previously mentioned earlier in this report, the forest fragmentation component of this
indicator needs additional refining before it can be included for analysis.

Acknowledgments
Authors: This report was updated by Chiara Zuccarino-Crowe, Environmental Careers
Organization, on appointment to U.S. Environmental Protection Agency (US EPA), Great Lakes
National Program Office (GLNPO), zuccarino-crowe.chiara@epa.gov from the State of the
Great Lakes 2005 Indicator report #8500 written and prepared by associate Mervyn Han,
Environmental Careers Organization, on appointment to US EPA, GLNPO. (Available online at,
http://binational.net/solec/sogl2005  e.html)

Contributors:
Support in the preparation of this report was given by the members of the SOLEC Forest Land
Criteria and Indicators Working Group. The following members aided in the development of
SOLEC Forest Lands indicators, collection, reporting and analysis of data, and the review and
editing of the  text of this report:
Constance Carpenter, Sustainable Forests Coordinator, USDA Forest Service,  Northeastern Area,
State and Private Forestry, conniecarpenter@fs.fed.us;
Larry Watkins, Forest Analyst, Ontario Ministry of Natural Resources, Forest  Evaluations and
Standards Section, Forest Management Branch, larry.watkins@mnr.gov.on.ca;
Eric Wharton, USDA Forest Service, ewharton@fs.fed.us;
T. Bently Wigley, NCASI, wiglev@clemson.edu;

Mike  Gardner (Sigurd Olson Environmental Institute, Northland College), Dain Maddox (USDA
Forest Service), Ann McCammon Soltis (Great Lakes Indian Fish & Wildlife Commission),  Wil
McWilliams (USDA Forest Service), Bill Meades (Canadian Forest Service), Greg Nowacki
(USDA Forest Service), Teague Prichard (Wisconsin Department of Natural Resources), Karen
Rodriguez (US EPA, GLNPO), Steve Schlobohm (USDA Forest Service), and Chris Walsh
(Ontario Ministry of Natural Resources).

Data  Sources
Canadian Council of Forest Ministers. 2000. Criteria and Indicators of Sustainable Forest
Management  in Canada: National Status 2000. http://www.ccfm.org/ci/index e.php

Canadian Council of Forest Ministers. 2003.  Defining Sustainable Forest Management in
Canada: Criteria and Indicators, 2003.  http://www.ccfm.org/current/ccitf_e.php

Canadian Great Lakes Basin forest data source: Ontario Ministry of Natural Resources, Forest
Standards and Evaluation Section. Landsat Data based on Landcover 2002 (Landsat 7) classified
imagery, Inventory data based on Forest Resources Planning Inventories, and several common
                         Draft for Discussion at SOLEC 2006

-------
NRVIS coverages such as watersheds, lakes and rivers etc.  Data supplied by Larry Watkins,
Ontario Ministry of Natural Resources, larry.watkins@mnr.gov. on. ca .

Carpenter, C., Giffen, C., and Miller-Weeks, M. 2003. Sustainability Assessment Highlights for
the Northern United States. Newtown Square, PA: USDA Forest Service, Northeastern Area
State and Private Forestry. NA-TP-05-03. http://www.na.fs.fed.us/sustainability/pubs/pubs.shtm

Ontario Ministry of Natural Resources (OMNR). 2002. State of the Forest Report, 2001. Ontario,
Canada: Queen's Printer for Ontario.
http://ontariosforests.mnr.gov.on. ca/spectrasites/Viewers/showArticle.cfm?id=20661E52-EE91-
453D9BD475CE675F7DlA&method=DISPLAYFULLNOBARNOTITLE R&ObjectID=20661
E52-EE91-453D-9BD475CE675F7D1A

Riitters, K.H., Wickham, J.D., O'Neill, R.V., Jones, K.B., Smith, E.R., Coulston, J.W., Wade,
T.G., and Smith, J.H. 2002. Fragmentation of Continental United States Forests. Ecosystems 5:
815-822.

Smith, W.B. 2004.  United States 2003 Report on Sustainable Forests, Data Report: Criterion 1,
Indicators 1, 2, 3, 4, Conservation of Biological Diversity. U.S. Department of Agriculture
(USDA) Forest Service. FS-766A. 24pp. http://www.fs.fed.us/research/sustain/contents.htm

Swank, Wayne. 2000.  Effects of Vegetation Management on Water Quality: Forest Succession.
In Drinking Water from Forests and Grasslands: A Synthesis of the Scientific Literature, ed.
G.E. Dissmeyer, pp.103-119. Asheville, NC:  USDA Forest Service, Southern Research Station.
SRS-39.

U.S. Great Lakes Basin forest data source:  USDA Forest Service,  Forest Inventory and Analysis
National Program, 2002 Resource Planning Act (RPA) Assessment Database.
http://ncrs2.fs.fed.us/4801/fiadb/rpa tabler/webclass rpa tabler.asp .  Data supplied by Eric
Wharton, USDA Forest Service, ewharton@fs.fed.us .

USDA Forest Service. 2000. 2000 RPA Assessment of Forest and Range Lands. Washington
DC: USDA Forest Service. FS-687. http://www.fs.fed.us/pl/rpa/rpaasses.pdf

USDA Forest Service. 2004. National Report on Sustainable Forests - 2003. FS-766.
http://www.fs.fed.us/research/sustain/documents/SustainableForests.pdf

USDA Forest Service and Northeastern Forest Resource Planners Association. 2002.
Sourcebook on Criteria and Indicators of Forest Sustainability in the Northeastern Area.
Newtown Square, PA: USDA Forest Service, Northeastern Area State and Private Forestry. NA-
TP-03-02. http://www.na.fs.fed.us/spfo/pubs/sustain/crit_indicators/02/cover.pdf

USDA Forest Service and Northeastern Forest Resource Planners Association. 2003. Base
Indicators of Forest Sustainability: Metrics and Data Sources for State and Regional
Monitoring.  Durham, NH: USDA Forest Service, Northeastern Area State and Private Forestry.
                         Draft for Discussion at SOLEC 2006

-------

                                      tf"j%*-l''|wSrjp3fe-™=*' 	; ,fei^ * i'i" ''  ./ 'l3™B*j,A
                                                             ^!?r?
World Wildlife Fund (WWF). 1999. Canada's commitment to forest protected areas: Forests for
life. WWF Status Report. World Wildlife Fund Canada. Toronto, ON. 17 p. Cited in, Canadian
Council of Forest Ministers. 2000. Criteria and Indicators of Sustainable Forest Management in
Canada: National Status 2000. 7pp.

List of Tables
Table 1. Total forest area and protected area by forest type in U.S. Great Lakes basin counties
Caption: Non-stocked =
timberland less than 10% stocked with live trees
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database

Table 2. Total forest area and protected area by forest type in, A) Canadian Great Lakes basin, B)
AOU* portion of Ontario
Caption: * The Area of the Undertaking (AOU) land area represents 72% of the total land area
analyzed in Ontario's portion of the Great Lakes basin.
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages

Table 3. Age-class distribution as a percentage of area within forest type for U.S. Great Lakes
basin counties
Caption: Non-stocked = timberland less than 10% stocked with live trees
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database

Table 4. Serai stage distribution as a percentage of area within provincial forest type in AOU*
portion of Canadian Great Lakes Basin
Caption: * The Area of the Undertaking (AOU) land area represents 72% of the total land area
analyzed in Ontario's portion of the Great Lakes basin.
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages

Table 5. Description of Canadian provincial forest types
Source: Descriptions taken from, Forest Resources of Ontario 2001: State of the Forest Report,
Appendix 1, p.  41, (OMNR 2002).

List of Figures
Figure 1. Proportion of forested area by forest type in U.S. Great Lakes basin
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database

Figure 2. Proportion of forested area by provincial forest type in AOU* portion of Canadian
Great Lakes basin
Caption: * The Area of the Undertaking (AOU) land area represents 72% of the total land area
analyzed in Ontario's portion of the Great Lakes basin.
                         Draft for Discussion at SOLEC 2006

-------
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages

Figure 3. Age-class distribution as a percentage of forested area within forest type for U.S. Great
Lakes basin counties
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database

Figure 4. Serai stage distribution as a percentage of forested area within provincial forest type in
AOU* portion of Canadian Great Lakes Basin
Caption: * The Area of the Undertaking (AOU) land area represents 72% of total land area
analyzed in Ontario's portion of the Great Lakes basin.
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages

Last updated
SOLEC 2006
Forest Type
White-Red-Jack Pine
Spruce-Fir
Loblolly-Shortleaf
Pine
Oak-Pine
Oak-Hickory
Oak-Gum-Cypress
Elm-Ash-Cottonwood
Maple-Beech-Birch
Aspen-Birch
Nonstocked
Totals
Area (ha)
1,791,671
2,866,777
4,305
72,675
1,988,126
50,589
1,692,069
7,828,700
3,821,272
88,443
20,204,626
% of Total
Forest
Area
8.87%
14.19%
0.02%
0.36%
9.84%
0.25%
8.37%
38.75%
18.91%
0.44%

Protected
Area (ha)
168,737
263,216
0
4,178
129,431
9,730
45,564
692,600
252,443
4,677
1,570,576
%
Protected
9.42%
9.18%
0.00%
5.75%
6.51%
19.23%
2.69%
8.85%
6.61%
5.29%
7.77%
Table 1.  Total forest area and protected area by forest type in U.S. Great Lakes basin counties
Caption: Non-stocked =
timberland less than 10% stocked with live trees
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database
                         Draft for Discussion at SOLEC 2006

-------

                                                    ^ ^  .,•_
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''  './ ''l3
                                                             s^
A) Canadian Great Lakes Basin
Satellite Classes
Forest - Sparse
Forest - Hardwood
Forest - Mixed
Forest - Softwood
Swamp - Treed
Fen - Treed
Bog - Treed
Disturbed Forest - cuts
Disturbed Forest - burns
Disturbed Forest -
regenerating
Totals
Area (ha)
2,053,869
3,468,513
5,750,313
2,407,729
49,933
30,197
436,083
578,450
97,545
35,987
14,908,617
% of Total
Forest
Area
13.78%
23.27%
38.57%
16.15%
0.33%
0.20%
2.93%
3.88%
0.65%
0.24%

Protected
Area (ha)
245,118
361,147
649,342
268,753
1,413
3,726
28,128
8,973
18,628
381
1,585,608
%
Protected
11.93%
10.41%
1 1 .29%
11.16%
2.83%
12.34%
6.45%
1.55%
19.10%
1.06%
10.64%

B) AOU* Portion of Ontario
Provincial Forest Type
White Birch
Mixed Conifer Lowland
Mixed Conifer Upland
Mixedwood
Jack Pine
Poplar
Red & White Pine
Tolerant Hardwoods
Totals
Area (ha)
1,593,114
1,048,126
2,657,086
2,099,760
714,165
1,189,573
685,124
1,616,502
11,603,450
% of Total
Forest
Area
13.73%
9.03%
22.90%
18.10%
6.15%
10.25%
5.90%
13.93%

Protected
Area (ha)
175,261
60,192
239,194
194,682
54,991
75,538
105,682
108,993
1,014,533
%
Protected
11.00%
5.74%
9.00%
9.27%
7.70%
6.35%
15.43%
6.74%
8.74%
Table 2.  Total forest area and protected area by forest type in, A) Canadian Great Lakes basin,
B) AOU* portion of Ontario
Caption:  * The Area of the Undertaking (AOU) land area represents 72% of the total land area
analyzed in Ontario's portion of the Great Lakes basin.
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages
                         Draft for Discussion at SOLEC 2006

-------
Forest Type
White-Red-Jack
Pine
Spruce-Fir
Loblolly-Shortleaf
Pine
Oak-Pine
Oak-Hickory
Oak-Gum-Cypress
Elm-Ash-
Cottonwood
Maple-Beech-Birch
Aspen-Birch
Nonstocked
Total
Age Class (in years)
0-20
13.86%
8.84%
0.00%
7.08%
9.43%
4.47%
14.03%
9.25%
25.40%
63.98%
13.29%
21-40
27.04%
18.55%
47.96%
14.58%
10.13%
36.37%
24.29%
12.38%
19.91%
16.73%
16.85%
41-60
25.41%
21 .84%
0.00%
47.30%
18.14%
19.84%
23.21%
21 .96%
26.15%
2.97%
22.77%
61-80
1 1 .63%
1 7.96%
52.04%
18.29%
21.49%
8.75%
15.95%
20.87%
16.64%
1.71%
18.37%
81-100
7.47%
9.57%
0.00%
3.02%
14.14%
4.08%
8.58%
12.31%
3.85%
0.00%
9.65%
100+
4.32%
10.23%
0.00%
6.49%
10.06%
0.00%
6.17%
8.75%
1 .36%
1.14%
7.02%
Mixed
2.40%
0.33%
0.00%
3.18%
1 1 .38%
5.73%
5.21%
6.21%
0.45%
0.00%
4.33%
not
measured
7.87%
12.69%
0.00%
0.07%
5.22%
20.76%
2.56%
8.27%
6.25%
13.47%
7.72%
Table 3. Age-class distribution as a percentage of area within forest type for U.S. Great Lakes
basin counties
Caption: Non-stocked = timberland less than 10% stocked with live trees
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database
Provincial Forest
Type
White Birch
Mixed Conifer
Lowland
Mixed Conifer
Upland
Mixedwood
Jack Pine
Poplar
Red & White Pine
Tolerant Hardwoods
Totals
Serai Stage
Presapling
3.49%
13.81%
5.91%
4.60%
8.60%
6.60%
4.94%
1 .23%
6.00%
Sapling
4.52%
9.31%
13.12%
7.92%
31.96%
10.45%
3.77%
0.87%
10.14%
Immature
15.55%
13.38%
22.51%
26.06%
29.24%
18.97%
23.28%
6.40%
20.12%
Mature
63.58%
47.00%
42.11%
51 .03%
27.51%
52.55%
62.95%
60.13%
49.84%
Late
Successional
12.87%
16.50%
16.36%
10.39%
2.69%
1 1 .43%
5.06%
31.37%
13.91%
Table 4. Serai stage distribution as a percentage of area within provincial forest type in AOU*
portion of Canadian Great Lakes Basin
Caption: * The Area of the Undertaking (AOU) land area represents 72% of the total land area
analyzed in Ontario's portion of the Great Lakes basin.
                         Draft for Discussion at SOLEC 2006

-------

                                                   ^  ^ .,•_
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''  './ ''l3
                                                            s^
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages
Provicial Forest
Type
White Birch
Upland Conifers
Lowland Conifers
Mixedwood
Jack Pine
Poplar
White and Red Pine
Tolerant Hardwoods
Description
predominantly white birch stands
predominantly spruce and mixed jack
pine/spruce stands on upland sites
predominantly black spruce stands on low,
poorly drained sites
mixed stands made up mostly of spruce, jack
pine, fir, poplar and white birch
predominantly jack pine stands
predominantly poplar stands
all red and white pine mixedwood stands
predominantly hardwoods such as maple and
oak, found mostly in the Great Lakes forest
region
Table 5.  Description of Canadian provincial forest types
Source: Descriptions taken from, Forest Resources of Ontario 2001: State of the Forest Report,
Appendix I, p. 41, (OMNR 2002).
                        Draft for Discussion at SOLEC 2006

-------
                      Nonstocked East 0.44%
               [
       Aspen-Birch 18.91%
            Maple-Beech-Birch
                38.75%
                                                        White-Red-Jack Pine
                                                              8.87%
                                                                Spruce-Fir 14.19%
                                                 Loblolly-Shortleaf Pine
                                                       0.02%
                                                                            "~-—Oak-Pine 0.36%
                                                                     ^Oak-Hickory 9.84%
                                           ^^_ Oak-Gum-Cypress
                                                    0.25%
                                                           ,lm-Ash-Cottonwood
                                                               8.37%
Figure 1. Proportion of forested area by forest type in U.S. Great Lakes basin
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database
14
Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
                        Tolerant Hardwoods
                            13.93%
White Birch
 13.73%
        Red & White Pine
             5.90%
           Poplar _____
           10.25%
              Jack Pine
               6.15%
                                                                  Mixed Conifer Lowland
                                                                       9.03%
                  Conifer Upland
                  22.90%
                             18.10%
Figure 2.  Proportion of forested area by provincial forest type in AOU* portion of Canadian
Great Lakes basin
Caption: * The Area of the Undertaking (AOU) land area represents 72% of the total land area
analyzed in Ontario's portion of the Great Lakes basin.
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages
                        Draft for Discussion at SOLEC 2006
                                 15

-------
    60%
     0%
                                                                         -•-White-Red-Jack Pine

                                                                         -•-Spruce-Fir

                                                                            Loblolly-Shortleaf Pine

                                                                         -x- Oak-Pine

                                                                         -*-Oak-Hickory

                                                                         -•- Oak-Gum-Cypress

                                                                         —l— Elm-Ash-Cottonwood

                                                                         —— Maple-Beech-Birch

                                                                         — Aspen-Birch
            0-20
                      21-40
                                 41-60       61-8
                                 Age Class (in years)
                                                     81-100
                                                                 100+
Figure 3. Age-class distribution as a percentage of forested area within forest type for U.S. Great
Lakes basin counties
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database
16
Draft for Discussion at SOLEC 2006

-------
         State of the Great Lakes 2007 - Draft
    70%
    60%
    50%
    40%
    30%
  5 20%
  ai
  a.
    10%
                                                                         -White Birch
-Mixed Conifer
 Lowland

 Mixed Conifer
 Upland

 Mixedwood
                                                                         -Jack Pine
                                                                         -Poplar
                                                                         -Red & White Pine
                                                                         -Tolerant
                                                                          Hardwoods
                                                              Late
                                                           Successional
Figure 4.  Serai stage distribution as a percentage of forested area within provincial forest type in
AOU* portion of Canadian Great Lakes Basin
Caption: * The Area of the Undertaking (AOU) land area represents 72% of total land area
analyzed in Ontario's portion of the Great Lakes basin.
Source:  Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages
                         Draft for Discussion at SOLEC 2006
            17

-------

                                                     ^ .
                                  ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''  './ ''l3i™B*j,Arf(St';j(\f^*
                                                           ^!?r?lB*?
Forest Lands - Maintenance of Productive Capacity of Forest Ecosystems
Indicator # 8501

Note: This indicator includes three components and corresponds to Montreal Process Criterion
2, Indicators 10, 11, and 13.

Indicator #8501 Components:
    Component (1) - Area of forest land and area of forest land available for timber production
    Component (2) - Total merchantable volume of growing stock on forest lands available for
    timber production
    Component (3) - Annual removal of wood products compared to net growth, or the volume
    determined to be sustainable (proposedfor future analysis; data not presented in this report)

Overall Assessment
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  Additional discussion amongst forestry experts is needed for an
      Determining  assessment determination.
  Status and Trend

Lake-by-Lake Assessment
Lake Superior
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  U.S. data by individual lake basin is not available.
      Determining
  Status and Trend

Lake Michigan
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  U.S. data by individual lake basin is not available.
      Determining
  Status and Trend

Lake Huron
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  U.S. data by individual lake basin is not available.
      Determining
  Status and Trend

Lake Erie
           Status:  Not Assessed
           Trend:  Undetermined
   Primary Factors  U.S. data by individual lake basin is not available.
      Determining
                        Draft for Discussion at SOLEC 2006

-------
Not Assessed
Undetermined
U.S. data by individual lake basin is not available.
  Status and Trend

Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Purpose
    •  To determine the Great Lakes forests' capacity to produce wood products
    •  To allow for future assessments of changes in productivity over time, which can be
       representative of social and economic trends affecting management decisions and can
       also be related to ecosystem health

Ecosystem Objective
To maximize the productive capacity of Great Lakes forests while maintaining the ecosystem's
health and sustainability.

State of the Ecosystem
Component (1):
The total area of forest land analyzed in the Great Lakes basin for this report was 35,113,242
hectares. Of this area, about 89% (or a total of 31,194,790 hectares) can be considered as
available for timber production, as calculated from U.S. timber land estimates and Canadian
productive forests not restricted from harvesting. In the U.S. portion of the basin, the proportion
of land available for timber production increased to about 91%, while the value decreased to 86%
for the entire Canadian portion of the basin  and then rose to 91% for Ontario's managed forests.
Complete U.S. data broken down by state and Canadian data broken down by lake basin can be
viewed in Tables 1 and 2, respectively.

The amount of forest land available for timber production is directly related to the productive
capacity of forests for harvestable goods.  This proportion is affected by different types of
management activities, which provides an indication of the balance between the need for wood
products with the need to satisfy assorted environmental concerns aimed at conservation of
biological diversity.

Component (2):
In the analyzed area of Great Lakes basin forests available for timber production, 78% of the total
wood volume was merchantable.  This percentage of growing stock increased to 92% for the U.S.
portion of the basin and decreased to 61% for Ontario's managed forests in the Canadian part of
the  basin.  Complete U.S. data broken down by state and  Canadian data broken down by lake
basin can be viewed in Tables 3 and 4, respectively.

If the values of net merchantable volume are compared to the total area of forest land available
for timber production, a rough estimate of the forests' productive capacity can be obtained.  This
      Draft for Discussion at SOLEC 2006

-------

                                                       ^  ._
                                    ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''
puts U.S. forests' per-unit-area productivity at a value of 92.7 cubic meters per hectare (m3/ha),
and Canadian forests' at 90.2 (mVha).

Changes in productivity values can be indicative of the ecosystem's health and vigor, as a
lowered ratio of merchantable volume to available timber land can suggest reduced growth and
ability of trees to absorb nutrients, water and solar energy and increased disease and tree
mortality. Further assessment of productive capacity would require additional historical data and
analysis by forestry experts.

Component 3:
The growth to  removal ratio is often used as a course surrogate for the concept of sustainable
production in the U.S. Although exact data for this measure have not been compiled for this
report, nationwide U.S. studies have shown that timber growth has exceeded removals for several
decades, and Ontario's wood removals on managed timber land is supposedly done within
sustainable limits by definition of the forestry practices enacted in those areas.

Pressures
Fluctuating marketplace demands for wood products and increased pressures to reserve forest
lands for recreation, conservation of biodiversity and wildlife habitat can affect the volume of
timber available for harvest.

Disease and disturbance from fires or other events can also affect productivity capacity.

Management  Implications
Timber productivity can be increased through the use of timber plantations and sustainable
management of forests available for timber production.

Continued discussion of the meaning of sustainability and how it is affected by wood product
removal is crucial to the effectiveness of future management decisions.

Comments from the author(s)
It can be difficult to analyze forest areas and growing stocks for a set moment in time, because
inventory time frames can vary.  U.S. 2002 RPA data are  compiled from a range of different
years (1989-1998  for Great Lakes states) depending on when the most recent state inventories
were conducted. This issue should diminish as the FIA switches to an annualized survey cycle,
and future analyses should therefore incorporate this data.

Although Canadian data are available by watershed, U.S.  forest data are compiled by county for
this report, so the area of U.S. land analyzed is not necessarily completely  within the Great Lakes
basin. Corresponding data may be skewed. This factor makes it difficult to represent the data by
individual lake basin.  Additional GIS analysis of the U.S. raw inventory data would be required
to provide forest data by watershed.

Area of timber land in the U.S. is used as a proxy for the net area land available for timber
production in U.S. data calculations, but timber land area may include currently inaccessible and
inoperable areas or areas where landowners do not have timber production as an ownership
                         Draft for Discussion at SOLEC 2006

-------
«K
objective, and is therefore an overestimation of the net area available for timber production and
associated merchantable wood volumes.

Canadian data for growing stock is only available for Ontario's managed forests where Forest
Resources Planning Inventories occur.  This area is commonly referred to as the Area of the
Undertaking (AOU), and only represents 72% of Ontario's total Great Lakes basin land area and
78% of its total forest area. The rest of the Canadian part of the basin is restricted to satellite data
capabilities.

Data for annual removal of wood products as compared to net growth is available for Canada and
a few of the U.S. Great Lakes states, but was not prepared for the Great Lakes basin at the time of
this report.  This information should be compiled for future analyses when available, and is an
important ratio to monitor over time to ensure that wood harvesting is not reducing the total
volume of trees on timber land at larger spatial scales. Unfortunately this value does not add
much insight to the detailed ecological attributes of sustainability, and must be analyzed with
additional biological components to achieve this indicator's ecosystem objective.

Acknowledgments
Authors:  Chiara Zuccarino-Crowe, Environmental Careers Organization, on appointment to U.S.
Environmental Protection Agency (US EPA), Great Lakes National Program Office (GLNPO),
zuccarino-crowe.chiara@epa.gov , with assistance from the following:

Contributors:
Support in the preparation of this report was given by the members of the SOLEC Forest Land
Criteria and Indicators Working Group. The following members  aided in the development of
SOLEC Forest Lands indicators, collection, reporting and analysis of data, and the review and
editing of the text of this report:
Constance Carpenter, Sustainable Forests Coordinator, USDA Forest Service, Northeastern Area,
State and Private Forestry, conniecarpenter(Sjfs.fed.us;
Larry Watkins, Forest Analyst, Ontario Ministry of Natural Resources, Forest Evaluations and
Standards Section, Forest Management Branch, larry.watkins@mnr.gov.on.ca;
Eric Wharton, USDA Forest Service, ewharton(Sjfs.fed.us;
T. Bently Wigley, NCASI, wigley@clemson.edu;

Mike Gardner (Sigurd Olson Environmental Institute, Northland College), Dain Maddox (USDA
Forest Service), Ann McCammon Soltis (Great Lakes Indian Fish & Wildlife Commission),  Wil
McWilliams (USDA Forest Service), Bill Meades (Canadian Forest Service), Greg Nowacki
(USDA Forest Service), Teague Prichard (Wisconsin Department of Natural Resources), Karen
Rodriguez (US EPA, GLNPO), Steve Schlobohm (USDA Forest  Service), and Chris Walsh
(Ontario Ministry of Natural Resources).

Data Sources
Canadian Council of Forest Ministers. 2003. Defining Sustainable Forest Management in
Canada:  Criteria and Indicators, 2003. http://www.ccfm.org/current/ccitf e.php
                         Draft for Discussion at SOLEC 2006

-------
        |@|»IBiBI«BBi»l
Canadian Great Lakes Basin forest data source: Ontario Ministry of Natural Resources, Forest
Standards and Evaluation Section.  Landsat Data based on Landcover 2002 (Landsat 7) classified
imagery, Inventory data based on Forest Resources Planning Inventories, and several common
NRVIS coverages such as watersheds, lakes and rivers etc.  Data supplied by Larry Watkins,
Ontario Ministry of Natural Resources, larry.watkins@mnr.gov. on. ca .

Carpenter, C., Giffen, C., and Miller-Weeks, M. 2003. Sustainability Assessment Highlights for
the Northern  United States. Newtown Square, PA: USDA Forest Service, Northeastern Area
State and Private Forestry. NA-TP-05-03.
http://www.na.fs.fed.us/sustainability/pdf/front cover.pdf

Ontario Ministry of Natural Resources (OMNR). 2002. State of the Forest Report, 2001. Ontario,
Canada: Queen's Printer for Ontario.
http://ontariosforests.mnr.gov.on. ca/spectrasites/Viewers/showArticle.cfm?id=20661E52-EE91-
453D9BD475CE675F7DlA&method=DISPLAYFULLNOBARNOTITLE R&ObiectID=20661
E52-EE91-453D-9BD475CE675F7D1A

Smith, W.B.  2004. United States 2003 Report on Sustainable Forests, Data Report: Criterion
2, Maintenance of Productive Capacity of Forest Ecosystems.  USDA Forest Service. FS-766A.
http://www.fs.fed.us/research/sustain/documents/Indicator%2010/indicators%2010_14.pdf

USDA Forest Service and Northeastern Forest Resource Planners Association.  2003. Base
Indicators of Forest Sustainability: Metrics and Data Sources for State and Regional
Monitoring.  Durham, NH:  USDA Forest Service, Northeastern Area State and Private Forestry.

USDA Forest Service. 2004. National Report on Sustainable Forests - 2003. FS-766.
http://www.fs.fed.us/research/sustain/documents/SustainableForests.pdf

USDA Forest Service. 2000. 2000 RPA Assessment of Forest and Range Lands. Washington
DC:  USDA Forest Service. FS-687.  http://www.fs.fed.us/pl/rpa/rpaasses.pdf

U.S. Great Lakes Basin forest data source: USDA Forest Service, Forest Inventory and Analysis
National Program, 2002 Resource Planning Act (RPA) Assessment Database.
http://ncrs2.fs.fed.us/4801/fiadb/rpa tabler/webclass rpa tabler.asp . Data supplied  by Eric
Wharton, USDA Forest Service, ewharton(Sjfs.fed.us .

List of Tables
Table 1. Area of forest land available for timber production* in relationship to total area of forest
land in U.S. Great Lakes basin counties
Caption: * Area designated as timber land is used as a proxy for this value and may  include
inaccessible areas.  The presented data should therefore be considered an over-estimation of the
net area available for timber production.
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database

Table 2. Area of forest land available for timber production in relationship to total area  of forest
land in, A) Canadian Great Lakes basin, and B) the AOU* portion of Ontario
                         Draft for Discussion at SOLEC 2006

-------
Caption:  * The Area of the Undertaking (AOU) land area represents 72% of Ontario's total Great
Lakes basin land area and 78% of its total forest area.
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages

Table 3. Total volume of growing stock* in U.S. Great Lakes basin counties
Caption:  * Calculations do not take inaccessibility or inoperability of timber land into account, so
resulting values are skewed high
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database

Table 4. Total volume of growing stock in Canadian Great Lakes basin*
Caption:  * Data only available for Ontario's managed forests (AOU portion of Ontario)
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages

Last updated
SOLEC 2006
State
Illinois
Indiana
Michigan
Minnesota
New York
Ohio
Pennsylvania
Wisconsin
Total
Total Area of
Forest land
(ha)
29,322
198,351
7,802,663
3,345,320
4,775,982
742,161
223,904
3,086,921
20,204,626
Area of Forest
Land Available
for Timber
Production*
(ha)
5,634
182,287
7,533,587
2,818,676
3,928,686
668,190
210,992
3,033,084
18,381,137
% Available for
Timber
Production*
19.21%
91.90%
96.55%
84.26%
82.26%
90.03%
94.23%
98.26%
90.97%
Table 1.  Area of forest land available for timber production* in relationship to total area of forest
land in U.S. Great Lakes basin counties
Caption:  * Area designated as timber land is used as a proxy for this value and may include
inaccessible areas.  The presented data should therefore be considered an over-estimation of the
net area available for timber production.
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database
                         Draft for Discussion at SOLEC 2006

-------



A) Canadian Great Lakes Basin
Lake
Basin
Superior
Huron
Erie
Ontario
Totals
Total Area of
Forest Land
(ha)
7,061,238
6,162,419
322,317
1,362,643
14,908,617
Net area of Forest
Land Available for
Timber Production
(ha)
6,006,356
5,343,401
291,107
1,172,788
12,813,653
% Available for
Timber
Production
85.06%
86.71%
90.32%
86.07%
85.95%

B) AOU* Portion of Ontario
Lake
Basin
Huron
Ontario
Superior
Totals
Total Area of
AOU's Forest
Land (ha)
4,710,406
665,100
6,227,943
11,603,450
Net area of AOU
Forest Land Available
for Timber Production
(ha)
4,227,743
611,268
5,749,905
10,588,917
% Available for
Timber
Production
89.75%
91.91%
92.32%
91 .26%
Table 2.  Area of forest land available for timber production in relationship to total area of forest
land in, A) Canadian Great Lakes basin, and B) the AOU* portion of Ontario
Caption:  * The Area of the Undertaking (AOU) land area represents 72% of Ontario's total Great
Lakes basin land area and 78% of its total forest area.
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages
                         Draft for Discussion at SOLEC 2006

-------
State
Illinois
Indiana
Michigan
Minnesota
New York
Ohio
Pennsylvania
Wisconsin
Total
Total Live
Volume* (mA3)
on Forest Lands
Available for
Timber
Production
518,577
22,162,859
829,796,679
219,781,880
383,181,677
73,836,032
25,840,363
294,891,458
1,850,009,525
Net
Merchantable
Volume (mA3) of
Timber
Products
(Growing
Stock*)
500,423
18,342,594
754,964,965
199,559,859
365,098,413
71,466,897
24,880,573
269,125,981
1,703,939,705
Volume (mA3) of
Non-
merchantable
Timber Products
18,154
3,820,265
74,826,151
20,222,021
18,083,264
2,369,136
959,790
25,765,478
146,064,258
% Growing Stock*
(of Total Vol.
Available for
Timber
Production)
96.50%
82.76%
90.98%
90.80%
95.28%
96.79%
96.29%
91.26%
92.10%
Table 3.  Total volume of growing stock* in U.S. Great Lakes basin counties
Caption:  * Calculations do not take inaccessibility or inoperability of timber land into account, so
resulting values are skewed high
Source: USDA Forest Service, Forest Inventory and Analysis National Program, 2002 Resource
Planning Act (RPA) Assessment Database
Lake
Basin
Huron
Ontario
Superior
Totals
Total Volume
(mA3) on Forest
Lands Available
for Timber
Production
667,854,390
114,963,698
787,640,995
1,570,459,083
Net
Merchantable
Volume (mA3) of
Timber
Products
(Growing Stock)
421,077,634
72,717,983
461,410,679
955,206,296
Volume (mA3) of
Non-
merchantable
Timber Products
246,776,756
42,245,715
326,230,315
615,252,787
% Growing Stock
(of Total Vol.
Available for
Timber
Production)
63.05%
63.25%
58.58%
60.82%
Table 4.  Total volume of growing stock in Canadian Great Lakes basin*
Caption:  * Data only available for Ontario's managed forests (AOU portion of Ontario)
Source: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories, and NRVIS coverages
                         Draft for Discussion at SOLEC 2006

-------

                                                      ^ ._
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''
Forest Lands - Conservation and Maintenance of Soil and Water Resources
Indicator #8503

Note: This indicator includes two components and corresponds to Montreal Process Criterion 4,
Indicator 19

Indicator #8503 Components:
    Component (1) - Percent of forested land within riparian zones by watershed and percent of
    forested land within watershed by Lake basin
    Component (2) - Change in area of forest lands certified under sustainable forestry programs
    in Great Lakes states and Ontario
Overall Assessment
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
Trend information is not available for forested areas at this time. Data
for the area of certified forest lands can not be analyzed according to
Great Lakes Basin boundaries at this time, but the overall area of
certified lands is increasing across the region.
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Good
Undetermined
A large proportion of the basin's riparian zones and watersheds are forested.
Certification data does not exist specific to this individual lake basin.
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Improving, Unchanging, Deteriorating or Undetermined
Just over half of the basin's riparian zones and watersheds are forested.
Certification data does not exist specific to this individual lake basin.
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
Over half of the basin's riparian zones and watersheds are forested.
Certification data does not exist specific to this individual lake basin.
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
Poor
Undetermined
Only a small portion of the basin's riparian zones and watersheds are
forested. Certification data does not exist specific to this individual lake
                         Draft for Discussion at SOLEC 2006

-------
  Status and Trend  basin.
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Mixed
Undetermined
Just over half of the basin's riparian zones and watersheds are forested.
Certification data does not exist specific to this individual lake basin.
Purpose
•   To describe the extent to which Great Lakes basin forests aid in the conservation of the
    basin's soil resources and protection of water quality.
•   To describe the level of Great Lakes states' and Ontario's participation in sustainable forestry
    certification programs.

Ecosystem Objective
Improved soil and water quality within the Great Lakes basin.

State of the Ecosystem
Component (1):
Forests cover about 61% of the total land and 70% of the riparian zones (defined as the 30 meter
buffer around all surface waters) within the Great Lakes basin. This trend of a slightly greater
percentage of forested land by riparian zone as opposed to by overall watershed is repeated for
every major lake basin for the Great Lakes basin as a whole, (see Figure 2).

The U.S. portion of the basin (including the upper St. Lawrence River watersheds) has forest
coverage on 61% of its riparian zones (as of 1992), and the Canadian portion of the basin
(excluding the upper St. Lawrence River watersheds) has forest coverage on 76% of its riparian
zones (as of 2002), (see Table 1). Lake Superior has the best coverage overall, with forested
lands covering 96% of its riparian zones. Lakes Michigan (62%), Huron (74%) and Ontario
(61 %) all have at least half of their total riparian zones covered with forests, while Lake Erie has
only 30% coverage. The percentages of forested riparian zones by watershed are visually
represented in Figure 1 and are available summarized by Lake Basin in Figure 2.

While good water quality is generally associated with heavily forested or undisturbed watersheds,
(USDA 2004) the existence of a forested buffer near surface water features can also protect soil
and water resources despite the land use class present in the rest of the watershed, (Carpenter et.
al 2003). As the percentage  of forest coverage within a riparian zones increases, the amount of
runoff and erosion (and therefore nutrient loadings, non-point source pollution and sedimentation)
decreases and the capacity of the ecosystem to store water increases.  Studies show that heavy
forest cover is capable of reducing total runoff by as much as 26% as compared to treeless areas
with equivalent land-use conditions, (Sedell, et. al 2000) and that riparian forests can reduce
nutrient and sediment loadings by 30-90%, (Alliance for the Chesapeake Bay, 2004).

Biodiversity of aquatic species is further maintained in riparian areas with increased forest
coverage by an increase in the amount of large woody debris (which affects stream configuration,
                         Draft for Discussion at SOLEC 2006

-------

                                                       ^  ._
                                    ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''
regulation of organic matter and sediment storage, and aquatic habitat availability) and decreased
water temperatures, (Eubanks et. al 2002). A study completed in Pennsylvania by Lynch et. al in
1985 claimed that complete commercial clear cutting of a riparian zone allowed a 10 °C rise in
stream water temperatures, but the retention of a forested buffer strip only allowed an increase of
about 1 °C, (Binkley and MacDonald 1994).  This regulation of water temperatures  can be
critical to the maintenance of assorted cold-water fisheries like trout.

The lack of consensus on the desired percentage of forested land in the basin or riparian zone
(and the desired size of the riparian zone itself) makes it difficult to determine the specific
implications of the presented data. Comparisons to historical forest cover in riparian zones and
manipulative experiments would be useful for trend establishment.

Component (2):
Sustainable forestry management programs are designed to ensure timber can be grown and
harvested in ways that protect the local ecosystem.  Participation is often voluntary,  but once
certification is gained, compliance with management protocols is required. Data from three
different certification programs was analyzed for this report. It should be noted that their
numbers are not additive, as one area of land can be certified under more than one program at a
time.

The area of forest lands certified under the Sustainable Forestry Initiative (SFI®) program
increased by 855% from 2003 to 2005 across the Great Lakes region, (see Figure 3). Forest
landowners who only elect to enroll in the program, but not go through the formal certification
process, often choose to follow the forest management protocols, but are not required to do so
until they seek certification.  It is therefore possible that a much greater amount of forest lands are
being managed according to these sustainable practices than is represented by the given data.

Certification in two other sustainable forestry programs also grew in the U.S. Great Lakes states
over the past few years, (see Figure 4). The acres of forest lands certified by the  American Tree
Farm  System (ATFS) rose by 47% between 2004 and 2005. The  ATFS is a voluntary
certification program for non-industrial, private landowners, and states it's mission as, "To
promote the growing of renewable forest resources on private lands while protecting
environmental benefits and increasing public understanding of all benefits of productive
forestry," (American Forest Foundation, 2004).  The Forest Stewardship Council (FSC) is an
international body that accredits certification organizations and guarantees their authenticity.
Acres of forest lands certified under this organization grew by 50% between 2005 and 2006.

This rise in the area of certified forest lands under all three programs can be interpreted as a
greater commitment to sustainable forest management amongst forest industry professionals. The
assumption is that continued growth in sustainable management practices will lead to improved
soil and water resources in the areas where they are implemented.

Pressures
Component (1): The same pressures exerted on all forest resources also apply here.
Development of forest lands to other land use classes (such as developed, agricultural, or pasture)
decreases the amount of forest area across watersheds and in riparian zones. Urbanization and
                         Draft for Discussion at SOLEC 2006

-------
seasonal home construction can specifically impact riparian areas since they are among the most
desirable development locations.

Component (2):  Participation in sustainable forestry programs can be affected by marketplace
popularity. Political climate, status of the economy, and public opinion can all influence forest
managers decisions to gain certification.

Management Implications
Component (1):  Development of policy directed towards protecting forested lands within riparian
zones would help maintain forested buffers near surface waters, thereby leading to a possible
improvement of local ecosystem health regardless of the land use classification in the rest of the
watershed.

Component (2):  Increased reporting of certification data by watershed would make
corresponding analyses easier.  Greater participation in sustainable forestry certification programs
would ensure that all timberland is managed in a sustainable manner.

Comments from the author
Component (1):  For the purposes of this report, riparian zone was defined as 30 meters on each
side of a surface water feature. Research shows that a forested buffer of this size achieves the
widest range of water quality objectives, (Alliance for the Chesapeake Bay, 2004), and is the
standard value used in USGS Forestry Service, Northeastern Area.  Other sources quote different
amounts of forested buffer needed near surface water features to achieve the highest level of soil
and water resources protection, ranging anywhere from 8-150 meters from the water's edge,
(Illinois, Indiana, and Ohio Departments of Natural Resources, 2006). The ideal riparian zone
size can be affected by a variety of factors such as stream, vegetation and soil type,
geomorphology, slope of land, and season,  (Eubanks et. al, 2002).

The resolution of the US landcover dataset  used in this analysis was coarse enough to cause slight
inaccuracies, but the data was determined as suitable for summarization at the watershed scale.

Additional research of existing literature would be helpful in further quantifying the effects of
riparian forests on erosion, run-off, water temperatures, and nutrient and pollutant storage.
Although specific studies have been done on these topics, the differences in metrics and sample
locations complicate comparisons for the Great Lakes  Basin.

Component (2):  In subsequent analyses, data should be collected for the percent of forested
riparian zones that lie within areas also certified in sustainable forestry programs.  Presently,
certification data cannot be analyzed by watershed or riparian area,  and is therefore less useful for
any analyses other than assessment of changing trends in the programs' utilization.

Expanding this component to include rates  of compliance with Forestry Best Management
Practices (BMPs) would provide valuable information for additional analyses.  While certification
in sustainable forestry programs often includes the implementation of BMPs, not all forest lands
managed according to BMPs are also certified.  Forestry BMPs have been developed in all Great
                         Draft for Discussion at SOLEC 2006

-------

Lakes states and provinces, so obtaining the relevant audit data would provide a greater and more
detailed information base relating to the conservation of forest, soil and water resources.

Many BMPs are directed at reducing non-point source pollution and some states even have
monitoring data relating to issues such as water quality.  For example, Wisconsin's Forestry Best
Management Practices for Water Quality Report stated that, when BMPs were correctly applied
to areas where they were needed, 96% of the monitored area showed no adverse impact on water
quality, (Breunig et. al 2003). It is generally accepted that this trend exists in other states as well.
For although individual states' BMPs may differ, studies have shown that their correct
implementation results in effective protection of water quality overall.

Acknowledgments
Authors: Chiara Zuccarino-Crowe, Environmental Careers Organization, on appointment to U.S.
Environmental Protection Agency, Great Lakes National Program Office, zuccarino-
crowe.chiara@epa.gov , with assistance from the following:

Contributors:
Support in the preparation of this report was given by the members of the SOLEC Forest Land
Criteria and Indicators Working Group.  The following members aided in the development of
SOLEC Forest Lands indicators, collection, reporting and analysis of data,  and the review and
editing of the text of this report:
Constance Carpenter, Sustainable Forests Coordinator, USDA Forest Service, Northeastern Area,
State and Private Forestry, conniecarpenter@fs.fed.us;
Larry Watkins, Forest Analyst, Ontario Ministry of Natural Resources, Forest Evaluations and
Standards Section, Forest Management Branch, larry.watkins@mnr.gov.on.ca;
Rebecca L. Whitney, GIS Specialist, USDA Forest Service, Northeastern Area, State and Private
Forestry, rwhitney@fs.fed.us;
T. Bently Wigley, NCASI, wiglev@clemson.edu;
Jason Metnick, Manager, SFI Label and Licensing, Sustainable Forestry Board,
metnicki@aboutsfb.org;
Sherri Wormstead,  Sustainability Specialist, USDA Forestry Service, Northeastern Area,  State
and Private Forestry, swormstead@fs.fed.us;
John Schneider, Ecologist and GIS Specialist, U.S. EPA, Great Lakes National Program Office,
scneider.iohn@epa.gov;
Karen Rodriguez, Environmental Protection Specialist, U.S. EPA, Great Lakes National Program
Office, Rodriguez.karen@epa.gov;

Mike Gardner (Sigurd Olson Environmental Institute, Northland College), Dain Maddox (USDA
Forest Service), Ann McCammon Soltis (Great Lakes Indian Fish & Wildlife Commission), Wil
McWilliams (USDA Forest Service), Bill Meades (Canadian Forest Service), Greg Nowacki
(USDA Forest Service), Teague Prichard (Wisconsin Department of Natural Resources), Steve
Schlobohm (USDA Forest Service), Chris Walsh (Ontario Ministry of Natural Resources), and
Eric Wharton (USDA Forest Service).
                         Draft for Discussion at SOLEC 2006

-------
Data Sources
Alliance for the Chesapeake Bay. 2004. Riparian Forest Buffers, Linking Land and Water.
Chesapeake Bay Program, Forestry Workgroup, and USDA Forest Service.

American Tree Farm System. 2004. American Forest Foundation.
http://www.treefarmsystem.org/ (accessed August 15, 2006).

ATFS data  citation:  Program Statistics (January 2005), provided by Emily Chan, American
Forest Foundation, by e-mail on 11-4-2005, and reported via personal communication with Sherri
Wormstead, USDA Forest Service, swormstead@fs.fed.us .

Binkley, D. and L. MacDonald. 1994. Forests as non-point sources of pollution, and effectiveness
of best management practices. NCASI Technical bulletin No 672.
http ://www.warnercnr. colostate.edu/frws/people/facultv/macdonald/publications/ForestsasNonpoi
ntSourcesofPollution.pdf

Breunig, B., Gasser, D., and Holland, K. 2003. Wisconsin's Forestry Best Management
Practices for Water Quality, The 2002 Statewide BMP Monitoring Report. Wisconsin
Department of Natural Resources, Division of Forestry.  PUB-FR-252-2003.
http://dnr.wi.gov/org/land/forestrv/Usesof/bmp/2002MonitoringReport.pdf

Canadian Great Lakes Basin forest data source: Ontario Ministry of Natural Resources, Forest
Standards and Evaluation Section. Landsat Data based on Landcover 2002 (Landsat 7) classified
imagery, Inventory data based on Forest Resources Planning Inventories, and several common
NRVIS coverages such as watersheds, lakes and rivers etc.  Data supplied by Larry Watkins,
Ontario Ministry of Natural Resources, larry.watkins@mnr.gov. on. ca

Carpenter, C., Giffen, C., and Miller-Weeks, M. 2003.  Sustainability Assessment Highlights for
the Northern United States. Newtown Square, PA: USDA Forest Service, Northeastern Area
State and Private Forestry. NA-TP-05-03. http://www.na.fs.fed.us/sustainability/pubs/pubs.shtm

Eubanks, C.E. and Meadows, D. 2002. A  Soil Bioengineering Guide for Streambank and
Lakeshore Stabilization. San Dimas, CA: USDA  Forest Service, Technology and Development
Program. FS-683. http://www.fs.fed.us/publications/soil-bio-guide/

Forestry Best Management Practices for Illinois. August 8, 2000. Illinois DNR, Southern Illinois
University Carbondale, University of Illinois, and Illinois Forestry Development Council.
http://www.siu.edu/%7eilbmp/ (accessed August 10, 2006).

FSC data originally obtained from Will Price, The Pinchot Institute, and verified and edited from
FSC online database: http://www.fscus.org/certified_companies/ by Sherri Wormstead, USDA
Forest Service, swormstead@fs.fed.us .

Indiana DNR. "Forestry BMP's." July 28, 2006. Indiana Department of Natural Resources,
Division of Forestry, http://www.in.gov/dnr/forestry/  (accessed August 10, 2006).
                         Draft for Discussion at SOLEC 2006

-------

NCASI and UGA Warnell School of Forest Resources.  Forestry BMPs.
http://www.forestrybmp.net/ (accessed August 10, 2006).

Ohio DNR. 2006. Best Management Practices for Logging Operations, Fact Sheet. Ohio
Department of Natural Resources, Division of Forestry, Columbus, OH.
http://www.dnr.ohio.gov/forestry/landowner/pdf/BMPlogging.pdf

Ontario Ministry of Natural Resources. 2002. State of the Forest Report, 2001. Ontario, Canada:
Queen's Printer for Ontario.
http://ontariosforests.mnr.gov.on. ca/spectrasites/Viewers/showArticle.cfm?id=20661E52-EE91-
453D-
9BD475CE675F7DlA&method=DISPLAYFULLNOBARNOTITLE R&ObiectID=20661E52-
EE91-453D-9BD475CE675F7D1A

Sedell, J., Sharpe, M., Dravnieks Apple, D., Copenhagen, M. and Furniss, M..  2000.  Water and
the Forest Service. Washington, DC: USDA Forest Service, Policy Analysis. FS-660.
http://www.fs.fed.us/publications/policv-analysis/water.pdf

SFI data supplied via personal communication with Jason Metnick, SFI Label and Licensing,
Sustainable Forestry Board, metnickj(Sjaboutsfb.org , June 30, August 1 and 15, 2006.

Stednick, J.D. 2000.  Effects of Vegetation Management on Water Quality: Timber Management.
In Drinking Water from Forests and Grasslands:  A Synthesis of the Scientific Literature, ed. G.E.
Dissmeyer, pp.103-119. Asheville, NC: USDA Forest Service, Southern Research Station.
SRS-39.

USDA Forest Service. 2004. National Report on Sustainable Forests - 2003. FS-766.
http://www.fs.fed.us/research/sustain/documents/SustainableForests.pdf

U.S. Great Lakes Basin forest data source: USDA Forest Service, Northeastern Area State and
Private Forestry, Information Management and Analysis. 2005. Riparian Area Land Cover
Types based on  the 1992 National Land Cover Dataset.  Data supplied by Rebecca Whitney,
USDA Forest Service, rwhitney@fs.fed.us

USDA Forest Service, Northeastern Area State and Private Forestry, Information Management
and Analysis. 2006.  Forest land by Watershed. Data supplied by Rebecca Whitney, USDA
Forest Service, rwhitney@fs.fed.us .

List of Tables
Table 1. Percent of Land Forested within U.S. and Canadian Great Lakes Watersheds and
Riparian Zones by Lake Basin.
Caption for Table 1:  * = Including Upper St. Lawrence, ** = Not including Upper St. Lawrence
Data Sources:
US data:  USDA Forest Service, Northeastern Area State and Private Forestry, Information
Management and Analysis. 2005. Riparian Area Land Cover Types based on the 1992 National
Land  Cover Dataset. Lake Basin boundaries refined by U.S.  EPA, Great Lakes National Program
Office.
                         Draft for Discussion at SOLEC 2006

-------
Canadian data: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories andNRVIS watershed coverage (1994).

List of Figures
Figure 1. Percent Forested Land within Riparian Zones by Watershed in the Great Lakes Basin.
Area is technically part of the St. Lawrence River drainage, but included in the Great Lakes basin
by definition in the Clean Water Act and Great Lakes Water Quality Agreement.
Data Sources:
USGS National Hydrography Dataset (1999); USGS 1992 National Cover Dataset (1999); USGS
8-digit Watersheds (Hydrologic Unit Code; 1994); Riparian Areas created by the USDA Forest
Service North Central Research Station (2005).
Ontario Ministry of Natural Resources - NRVIS Watershed Coverage (1994); Landcover (2002);
Riparian Areas created by Forest Evaluation Section
Map data from USDA Forest Service, Information Management and Analysis Group, Durham,
NH and U.S. EPA, Great Lakes National Program Office.
Map created by U.S. EPA, Great Lake National Program Office, Technical Assistance and
Analysis Branch

Figure 2. Percent of Land Forested within Great Lakes Watersheds and Riparian Zones by Lake
Basin.
Caption for figure 2: * = Upper St. Lawrence data only available for U.S.
Data Sources:
US data: USDA Forest Service, Northeastern Area State and Private Forestry, Information
Management and Analysis. 2005. Riparian Area Land Cover Types based on the 1992 National
Land Cover Dataset. Lake Basin boundaries refined by U.S. EPA, Great Lakes National Program
Office.
Canadian data: Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories and NRVIS watershed coverages.

Figure 3. Forest Lands Certified Under SFI in the Great Lakes region (U.S. States and province
of Ontario), 2003-2005.
Data Source:
Personal communication with Jason Metnick, SFI Label and Licensing, Sustainable Forestry
Board, 2006.

Figure 4. Forest Lands Certified Under ATFS and FSC in the Great Lakes States (U.S.  only).
Data provided by Sherri Wormstead of the USDA Forestry Service (swormstead@fs.fed.us)
using following sources:
FSC data originally obtained from Will Price, the Pinchot Institute and verified and edited from
FSC online database: http://www.fscus.org/certified_companies/
ATFS data source:  Program Statistics (January 2005) (provided by Emily Chan, American Forest
Foundation, by e-mail on 11 -4-2005)
                         Draft for Discussion at SOLEC 2006

-------

                                                   ^ ^ .,•_
                                  ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''  './ ''l3
                                                           s^
Last updated
SOLEC 2006
Basin
Lake Superior
Lake Michigan
Lake Huron
Lake Erie
Lake Ontario
St. Lawrence
River
Totals
U.S. (1992)
% Forested
(Entire
Watershed)
87.73%
51.54%
55.07%
22.90%
52.15%
84.10%
53.13%*
% Forested
(Riparian
Areas)
88.44%
61.90%
54.28%
36.24%
63.25%
87.03%
60.43%*
Ontario (2002)
% Forested
(Entire
Watershed)
98.60%

74.65%
14.30%
49.99%

73.05%**
% Forested
(Riparian
Areas)
98.05%

77.04%
19.95%
59.28%

75.67%**
Table 1.  Percent of Land Forested within U.S. and Canadian Great Lakes Watersheds and
Riparian Zones by Lake Basin.
Caption for Table 1: * = Including Upper St. Lawrence, ** = Not including Upper St. Lawrence
Data Sources:
US data:  USDA Forest Service, Northeastern Area State and Private Forestry, Information
Management and Analysis. 2005. Riparian Area Land Cover Types based on the 1992 National
Land Cover Dataset. Lake  Basin boundaries refined by U.S. EPA, Great Lakes National Program
Office. Canadian data: Ontario Ministry of Natural Resources, Forest Standards and Evaluation
Section. Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data
based on Forest Resources Planning Inventories and NRVIS watershed coverage (1994).
                        Draft for Discussion at SOLEC 2006

-------
                               State of the Great Lakes 2007 - Draft
     % forested land
     within riparian zones
     by watershed
         ]l5-<25%
         I 25 - <40%
          40 - <60%
          75 - 85%  C
         ) St. Lawrence*
Figure 1. Percent Forested Land within Riparian Zones by Watershed in the Great Lakes Basin.
*The area within the St. Lawrence River drainage does not actually drain into the Great Lakes
basin, but is still included in the Great Lakes basin by definition in the Clean Water Act and the
Great Lakes Water Quality Agreement.
Data Sources:
USGS National Hydrography Dataset (1999); USGS 1992 National Cover Dataset (1999); USGS
8-digit Watersheds (Hydrologic Unit Code; 1994);  Riparian Areas created by the USDA Forest
Service North Central Research Station (2005).
Ontario Ministry of Natural Resources - NRVIS Watershed Coverage (1994); Landcover (2002);
Riparian Areas created by Forest Evaluation Section
Map data from USDA Forest Service, Information Management and Analysis Group, Durham,
NH and U.S. EPA, Great Lakes National Program Office.
Map created by U.S. EPA, Great Lake National Program Office, Technical Assistance and
Analysis Branch
10
Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
               D Within Entire Watershed
I Within Riparian Zones
   100%
    90%
  -g 80%
  ro 70%
  -3 60%
  •2 50%
  2 40%
  o
  "• 30%  H
  ^ 20%
    10%
      0%  -
           95% 96%
              Lake        Lake     Lake Huron   Lake Erie  Lake Ontario St. Lawrence
             Superior    Michigan                                         River

                                          Basin

Figure 2. Percent of Land Forested within Great Lakes Watersheds and Riparian Zones by Lake
Basin.
Caption for figure 2: * = Upper St. Lawrence data only available for U.S.
Data Sources:
US data: USDA Forest Service, Northeastern Area State and Private Forestry, Information
Management and Analysis.  2005. Riparian Area Land Cover Types based on the 1992 National
Land Cover Dataset. Lake Basin boundaries refined by U.S. EPA, Great Lakes National Program
Office.
Canadian data:  Ontario Ministry of Natural Resources, Forest Standards and Evaluation Section.
Landsat Data based on Landcover 2002 (Landsat 7) classified imagery, Inventory data based on
Forest Resources Planning Inventories  and NRVIS watershed coverages
                       Draft for Discussion at SOLEC 2006
                              11

-------
         30,000,000

         25,000,000
     •D
     i^  20,000,000
     '€
     O  15,000,000
     (/)
     p
     £3  10,000,000
     <
          5,000,000
                   2003
2005
Figure 3. Forest Lands Certified Under SFI in the Great Lakes region (U.S. States and province
of Ontario), 2003-2005.
Data Source:
Personal communication with Jason Metnick, SFI Label and Licensing, Sustainable Forestry
Board, 2006.
                       Draft for Discussion at SOLEC 2006

-------
        State of the Great Lakes 2007 - Draft
                                      ATFS--- FSC
    16,000,000

    14,000,000

"§  12,000,000

£  10,000,000


-------
                                               OF   THE   GREAT
                                                                2007
Acid Rain
Indicator #9000

Assessment: Mixed, Improving

Purpose
  To assess the pH levels in precipitation;
  To assess the critical loads of sulfate to the Great Lakes basin;
and
  To infer the efficacy of policies to reduce sulfur and nitrogen
acidic compounds released into the atmosphere.

Ecosystem Objective
The  1991  Canada-U.S. Air Quality Agreement (Air Quality
Agreement) pledges the two nations to reduce the emissions of
acidifying compounds by approximately 40% relative to 1980
levels. The 1998 Canada-Wide Acid Rain Strategy for Post-2000
intends to further reduce emissions to  the point where deposition
containing these compounds does not  adversely impact aquatic
and terrestrial biotic systems.

State of the Ecosystem
Background
Acid rain, more properly called "acidic deposition", is caused
when two common air pollutants, sulfur dioxide (SO2) and nitro-
gen oxides (NOX), are released into the atmosphere, react and
mix with atmospheric moisture and return to the earth as acidic
rain, snow, fog or particulate matter. These pollutants can be  car-
ried over long distances by prevailing  winds,  creating acidic pre-
cipitation  far from the original source  of the emissions.
Environmental damage typi-
cally occurs where local
soils and/or bedrock do  not
effectively neutralize the
acid.
Lakes and rivers have been
acidified by acid rain.
directly or indirectly caus-
ing the disappearance of
invertebrates, many fish
species, waterbirds and
plants. Not all lakes
exposed to acid rain become
acidified, however. Lakes
located in terrain that is rich
in calcium carbonate (e.g.
on limestone bedrock) are
able to neutralize acidic
deposition. Much of the
acidic precipitation in North
America falls in areas
284
                                  around and including the Great Lakes basin. Northern Lakes
                                  Huron, Superior and Michigan, their tributaries and associated
                                  small inland lakes are located on the geological feature known as
                                  the Canadian Shield. The Shield is primarily composed of
                                  granitic bedrock and glacially derived soils that cannot easily
                                  neutralize acid, thereby resulting in the acidification of many
                                  small lakes (particularly in northern Ontario and the northeastern
                                  U.S.). The five Great Lakes are so large that acidic deposition
                                  has little effect on them directly. Impacts are mainly felt on veg-
                                  etation and inland lakes in acid-sensitive areas.

                                  A recent report published by the Hubbard Brook Research
                                  Foundation has demonstrated that acid deposition is still a sig-
                                  nificant problem and has had a greater environmental impact
                                  than previously thought (Driscol et al. 2001). For example, acid
                                  deposition has altered soils in the  northeastern U.S. through the
                                  accelerated leaching of base cations, the accumulation of nitro-
                                  gen and sulfur, and an increase in concentrations of aluminum in
                                  soil waters. Acid deposition has also contributed to the decline
                                  of red spruce trees and sugar maple trees in the eastern U.S.
                                  Similar observations have been made in eastern Canada (Ontario
                                  and eastward) and are reported in the 2004 Canadian Acid
                                  Deposition Science Assessment (Environment Canada 2005).
                                  The assessment confirms that although levels of acid deposition
                                  have declined in eastern Canada over the last two decades.
                                  approximately 21% of the mapped area currently receives levels
                                  of acid rain in excess of what the region can handle, and 75% of
                                  the area is at potential risk of damage should all nitrogen deposi-
                                  tion become acidifying, i.e. aquatic and terrestrial ecosystems
                                  become nitrogen saturated.
                Transportation  Other
                  4%
Industrial Sources
   53%
                                    Electric Utilities
                                      25%
                                                        Industrial Sources
                                                  Fuel Combustion
                                                    18%
                                      Fuel Combustion
                                        18%
                      Canada
                                                                    United States
Figure 1. Sources of Sulfur Dioxide Emissions in Canada and the U.S. (1999)
Source: Figure 4 of Canada - United States Air Quality Agreement: 2002 Progress Report.
http://www.epa.gov/airmarkets/usca/airus02.pdfand Environment Canada 1999 National Pollutant
Release Inventory Data and U.S. Environmental Protection Agency 1999 National Emissions Inventory
Documentation and Data

-------
                          Bectric Utilities
                            12%
 Transportation
                                 Fuel Combustion
                                    19%
                              Industrial Sources
                                  11%
                   Canada
                                                        United States
 Figure 2. Sources of Nitrogen Oxides Emissions in Canada and the U.S. (1999)
 Source:  Figure 6 of Canada - United States Air Quality Agreement: 2002 Progress
 Report, http://www.epa.gov/airmarkets/usca/airus02.pdfand Environment Canada 1999
 Pollutant Release Inventory Data and U.S. Environmental Protection Agency 1999
 National Emissions Inventory Documentation and Data
Sulfur Dioxide and Nitrous Oxides Emissions Reductions
SO2 emissions come from a variety of sources. The most com-
mon releases of SO2 in Canada are industrial processes such as
nonferrous mining and metal smelting. In the United States.
electrical utilities constitute the largest emissions source (Figure
1). The primary source of NOX emissions in both countries is the
combustion of fuels in motor vehicles, with electric utilities and
industrial sources also contributing (Figure 2).
Canada is committed to reducing acid rain in its south-eastern
region to levels below those that cause harm to ecosystems - a
level commonly called the "critical load" - while keeping
other areas of the country (where acid rain effects have not
been observed) clean. In 2000, total SO2 emissions in Canada
were 2.4 million tonnes, which is about 23% below the
national cap of 3.2 million tonnes reiterated under Annex 1
(the Acid Rain Annex) of the Air Quality Agreement.
Emissions in 2000  also represent a 50% reduction from 1980
emission levels. The seven easternmost provinces' 1.6 million
tonnes of emissions in 2000 were 29% below the eastern
Canada cap of 2.3 million tonnes reiterated under the Acid
Rain Annex.
                        By 2000, Canadian NOX emissions were
                        reduced by more than 100,000 tonnes
                        below the forecast level of 970,000 tonnes
                        (established by Acid Rain Annex) at
                        power plants, major combustion sources.
                        and smelting operations. In the U.S..
                        reductions in NOX emissions have signifi-
                        cantly surpassed the 2 million ton reduc-
                        tion for stationary and mobile sources
                        mandated by the Clean Air Act
                        Amendments of 1990. Under the Acid
                        Rain Program alone, NOX emissions for
                        all the affected sources in 2002 were 4.5
                        million tons, about 33% lower than emis-
                        sions from the sources in 1990. Overall
                        NOX emissions decreased by about 12% in
                        the U.S. from 1993 to 2002 and remained
                        relatively constant in Canada since 1990.
                        but they are projected to decrease consid-
   erably in both countries by 2010. For additional information on
   SO2 and NOX emission reductions, including sources outside the
   Acid Rain Program, please refer to indicator report #4176 Air
   Quality.

   Figure 3 illustrates the trends in SO2 emission levels in Canada
   and the United States measured from 1980 to 2000 and predicted
   through 2010. Overall, a 38% reduction in SO2 emissions is pro-
   jected in Canada and the United States from 1980 to 2010. In the
   U.S., the reductions are mainly due to controls on electric utili-
In 2002, all participating sources of the U.S. Environmental
Protection Agency's Acid Rain Program (Phase I & II)
achieved a total reduction in SO2 emissions of about 35%
from 1990 levels, and 41% from 1980 levels. The Acid Rain
Program now affects approximately 3,000 fossil-fuel power
plant units. These units reduced their SO2 emissions to  10.19
million tons  in 2002, about 4% lower than 2001 emissions.
Full implementation of the program in 2010 will result  in a
permanent national emissions cap of 8.95 million tons,  repre-
senting about a 50% reduction from 1980 levels.
Figure 3. Canada-U.S. sulfur dioxide emissions, 1980-2010
Source: Figure 3 of Canada - United States Air Quality Agreement:
2002 Progress Report, http://www.epa.gov/airmarkets/usca/airus02.pdf
and U.S. Environmental Protection Agency. Projection year emissions
data, http://www.epa.gov/otaq/models/hd2007/r00020.pdf
                                                                                                                     285

-------
                                               OF   THE   GREAT
                              2007
(  Five-Year Mean nssSO,' Wet Deposition (1980-1994) I

              -   •*      V    -   "
                                                                         Wet Deposition 11996-2000)
                                                                                           kgfha.'yr
        Five-Vear Mean NOj Wet Deposition (1990-1994)
 Figure 4. Five-year mean patterns of wet non-sea-salt-sulfate (nssS042-) and wet nitrate deposition for
 the periods 1990-1994 and 1996-2000.
 Source: Figures 9 through 12 of Canada - United States Air Quality Agreement: 2002 Progress Report.
 http://www.epa.gov/airmarkets/usca/airus02.pdf, and Jeffries, D.S., T.G.,  Brydges, PJ. Billion and W.
 Keller. 2003. Monitoring the results of Canada/U.S.A. acid rain control programs: some lake responses.
 J. of Environmental Monitoring and Assessment. 88:3-20
ties under the Acid Rain Program and the desulphurization of
diesel fuel under Section 214 of the 1990 Clean Air Act
Amendments. In Canada, reductions of SO2 are mainly attrib-
uted to reductions  from the non-ferrous mining and smelting
sector, and electric utilities as part of the 1985 Eastern Canada
Acid Rain Program that was completed in 1994. Further SO2
reductions will be  achieved through the implementation of the
Canada-Wide Acid Rain Strategy.
                                    Figure 4 compares wet
                                    sulfate deposition (kilo-
                                    grams sulfate per hectare
                                    per year) over eastern
                                    North America before and
                                    after the 1995 Acid Rain
                                    Program Phase I emission
                                    reductions to  assess
                                    whether the emission
                                    decreases had an impact
                                    on large-scale wet deposi-
                                    tion. The five-year aver-
                                    age sulfate wet deposition
                                    pattern for the years 1996-
                                    2000 is considerably
                                    reduced from that for the
                                    five-year period prior to
                                    the Phase I emission
                                    reductions (1990-1994).
                                    For example,  the large
                                    area that received 25 to 30
                                    kg/ha/yr of sulfate wet
                                    deposition in  the 1990-
                                    1994 period had almost
                                    disappeared in the 1996-
                                    2000 period. The shrink-
                                    age of the wet deposition
                                    pattern between the two
                                    periods strongly suggests
                                    that the Phase I emission
                                    reductions were success-
                                    ful at reducing the sulfate
                                    wet deposition over  a
                                    large section of eastern
                                    North America.
                                    Monitoring data from
                                    2000 through 2002 indi-
                                    cate that wet sulfate depo-
                                    sition continued to
                                    decrease, probably as a
                                    result of Phase II of the
                                    Acid Rain Program.
                                    However, if SO2 emis-
sions remain relatively constant after the year 2000,  as predicted
(Figure 3), it is unlikely that sulfate deposition will change con-
siderably in the coming decade. Sulfate deposition models  pre-
dict that in 2010, following implementation of the Phase II acid
rain program, critical loads for aquatic ecosystems in eastern
Canada will still be exceeded over an area of approximately
800,000 km2.

A somewhat different story occurs for nitrate wet deposition.
                                                          Five-Year Mean NO,'Wet Deposition (1996-2000)
286

-------
                STATE   OF
The spatial patterns shown in Figure 4 are approximately the
same before and after the Phase I emission reductions. This sug-
gests that the minimal reductions in NOX emissions after Phase I
resulted in minimal changes to nitrate wet deposition over east-
ern North America.

Pressures
As the human population within and outside the basin continues
to grow, there will be increasing demands on electrical utility
companies and natural resources and increasing numbers  of
motor vehicles.  Considering this, reducing nitrogen deposition is
becoming more and more important, as its contribution to acidi-
fication may soon outweigh the benefits gained from reductions
in sulfur dioxide emissions.

Management Implications
The effects of acid rain can be seen far from the source of SO2
and NOX generation, so the governments of Canada and the
United States are working together to reduce acid emissions. The
1991 Canada - United States Air Quality Agreement addresses
transboundary pollution.  To  date, this agreement has focused on
acidifying pollutants and significant steps have been made in the
reduction of SO2 emissions. However, further progress in the
reduction of acidifying pollutants, including NOX, is required.

In December 2000, Canada and the United States signed Annex
III (the Ozone Annex) to the Air Quality Agreement. The Ozone
Annex committed Canada and the U.S. to aggressive emission
reduction measures to reduce emissions of NOX and volatile
organic compounds.  (For more information on the Ozone Annex,
please refer to Report #4176 Air Quality).

The 1998 Canada-wide Acid Rain Strategy for  Post-2000 pro-
vides a framework for further actions,  such as establishing new
SO2 emission reduction targets in Ontario, Quebec, New
Brunswick and Nova Scotia. In fulfillment of the Strategy, each
of these provinces has announced a 50% reduction from its
existing emissions cap. Quebec, New Brunswick and Nova
Scotia are committed to achieving their caps by 2010, while
Ontario committed to meet its new cap by 2015.

Since the last State of the Lakes Ecosystem Conference
(SOLEC) report, there has been increasing interest in both the
public and private sector in a multi-pollutant approach to  reduc-
ing air pollution. On March  10, 2005, the U.S.  Environmental
Protection Agency (USEPA) issued the Clean Air Interstate Rule
(CAIR), a rule that will achieve the largest reduction in air pol-
lution in more than a decade. Through a cap-and-trade approach,
CAIR will permanently cap  emissions of SO2 and NOX across 28
eastern states and the District of Columbia. When fully imple-
mented, CAIR is expected to reduce SO2 emissions in these
states by 73% and NOX emissions by 61% from 2003 levels.
The Clear Skies Initiative, originally proposed by U.S. President
George W. Bush in February 2002, would require a similar level
of SO2 and NOX reductions as CAIR. Because Clear Skies would
be enacted through legislation rather than regulation, it would be
a more efficient, long-term mechanism to achieve multi-pollu-
tant reductions on a national scale. The USEPA is committed to
working with Congress to pass this legislation. However, if
Clear Skies is not passed, CAIR still remains in effect.

Acknowledgments
Authors: Dean S. Jeffries, National Water Research Institute,
Environment Canada, Burlington, ON;
Robert Vet, Meteorological Service of Canada, Environment
Canada, Downsview, ON;
Silvina Carou, Meteorological Service of Canada, Environment
Canada, Downsview, ON;
Kerri Timoffee, Manager, Acid Rain Program, Environment
Canada, Gatineau, QC; and
Todd Nettesheim, Great Lakes National Program Office, United
States Environmental Protection Agency, Chicago, IL.

Sources
Canada - United States Air Quality Committee. 2002. United
States - Canada Air Quality Agreement: 2002 Progress Report.
http ://www. epa. gov/airmarkets/ usca/airus02.pdf. last accessed
June 17, 2004.

Canadian Council of Ministers of the Environment (CCME).
2004. 2002 Annual progress Report on the Canada-Wide Acid
Rain Strategy for Post-2000. ISBN 0-622-67819-2.
http://dev.sitesl.miupdate.com/l/assets/pdf/2002_ar_annual_rpt_
e.pdf. last accessed June 21, 2004.

Canadian Council of Ministers of the Environment (CCME).
2002. 2001 Annual Progress Report on the Canada-Wide Acid
Rain Strategy for Post-2000. ISBN 0-662-66963-0.
http ://www.ccme.ca/assets/pdf/acid_rain_e.pdf. last accessed
July 16, 2004.

Driscoll, C.T., Lawrence, G.B., Bulger, A.J., Butler, T.J., Cronan,
C.S., Eagar,  C., Lambert, K.F., Likens, G.E., Stoddard, J.L., and
Weathers, K.C. 2001. Acid Rain Revisited: advances in scientific
understanding since the passage of the 1970 and 1990 Clean Air
Act Amendments. Hubbard Brook Research Foundation.  Science
LinksTM Publication Vol. 1, no. 1.

Environment Canada. 2005. 2004 Canadian Acid Deposition
Science Assessment: Summary of Key Results, http: //www. msc -
smc.ec.gc.ca/saib/acid/acid_e.html.

Environment Canada. 2004. 2002 National Pollutant Release
Inventory Data.
                                                                                                                   287

-------
-    *"5|fP
—stimffii
                                                                                          2007
http://www. ec. gc. ca/pdb/npri/npri_dat_rep_e. cfm#highlights. last
accessed June 29, 2004.

Environment Canada. 2003a. 2001 National Pollutant Release
Inventory: National Overview.
http://www.ee. gc.ca/pdb/npri/npri_dat_rep_e.cfm#annual2001.
last accessed June 29, 2004.
     Report. EPA-430-R-03-011.
     http ://www. epa. gov/airmarkets/cmprpt/arp02/2002report.pdf. last
     accessed July 16, 2004.

     U.S. Environmental Protection Agency (USEPA). 2003c. EPA's
     Draft Report on the Environment: Technical Document. EPA-
     600-R-03-050. http://www.epa.gov/indicators/.
Environment Canada. 2003b. Cleaner Air through Cooperation:
Canada - United States Progress under the Air Quality
Agreement 2003. ISBN 0-662-34082-5. http://www.epa.gov/air-
markets/usca/brochure/brochure.htm. last accessed June 17,
2004.

Environment Canada. 2003c. Environmental Signals: Canada's
National Environmental Indicator Series 2003.
http://www.ec.gc.ca/soer-
ree/English/Indicator_series/default.cfm#pic. last accessed June
29, 2004.
     U.S. Environmental Protection Agency (USEPA). 2003d. Latest
     Findings on National Air Quality: 2002 Status and Trends.
     Office of Air Quality Planning and Standards. EPA-454/K-03-
     001. http ://www. epa. gov/airtrends/2002_airtrends_final.pdf. last
     accessed June 17, 2004.

     U.S. Environmental Protection Agency (USEPA). 2003e.
     National Air Quality and Emissions Trends Report: 2003 Special
     Studies Edition. Office of Air Quality Planning and Standards.
     EPA-454/R-03-005. http://www.epa.gov/air/airtrends/aqtrnd03/.
     last accessed June 17, 2004.
Environment Canada. National Atmospheric Chemistry
Database and Analysis Facility. Meteorological Service of
Canada, Downsview, ON.

Jeffries, D.S., Clair, T.A., Couture, S., Dillon, P.J., Dupont, J.,
Keller, W., McNicol, D.K., Turner, M.A., Vet, R., and Weeber,
R. 2003. Assessing the recovery of lakes in southeastern Canada
from the effects of acidic deposition. Ambio. 32(3):176-182.

National Atmospheric Deposition Program. A Cooperative
Research Support Program of the State Agricultural Experiment
Stations (NRSP-3) Federal and State Agencies and Non-
Governmental Research  Organizations.
http ://nadp. sws .uiuc. edu/.
     U.S. Environmental Protection Agency (USEPA). 2002.
     Procedures for developing base year and future year mass and
     modeling inventories for the heavy-duty engine and vehicle stan-
     dards and highway dies el fuel (HDD) rulemaking. EPA420-R-
     00-020. http://www.epa.gov/otaq/models/hd2007/r00020.pdf. last
     accessed September 29, 2005.

     U.S. Environmental Protection Agency (USEPA). Clean Air
     Interstate Rule, http://www.epa.gov/cair/. last accessed June 8,
     2004.

     U.S. Environmental Protection Agency (USEPA). The Clear
     Skies Initiative, http ://www. epa. gov/clearskies/.
Ontario Ministry of the Environment (OMOE). 2004. Air
Quality in Ontario 2002 Report. Queen's Printer for Ontario.
http://www.ene.gov.on.ca/envision/techdocs/4521 eO 1 .pdf. last
accessed June 28, 2004.

Ontario Ministry of the Environment (OMOE). 2003. Air
Quality in Ontario 2001 Report. Queen's Printer for Ontario.
http://www.ene.gov.on.ca/envision/air/AirOuality/2001 .htm, last
accessed June 17, 2004.

U.S. Environmental Protection Agency (USEPA). 2003a. 1999
National Emissions Inventory Documentation and Data.
http ://www. epa. go v/ttn/chief/net/1999inventory.html.

U.S. Environmental Protection Agency (USEPA). 2003b. Clean
Air Markets Programs. In Acid Rain Program: 2002 Progress
     Authors' Commentary
     While North American SO2 emissions and sulfate deposition lev-
     els in the Great Lakes basin have declined over the past 10 to 15
     years, rain is still too acidic throughout most of the Great Lakes
     region, and many acidified lakes do not show recovery (increase
     in water pH or alkalinity). Empirical evidence suggests that there
     are a number of factors acting to delay or limit the recovery
     response, e.g. increasing importance of nitrogen-based acidifica-
     tion, soil depletion of base cations, mobilization of stored sulfur,
     climatic influences, etc. Further work is needed to quantify the
     additional reduction in deposition needed to overcome these lim-
     itations and to accurately predict the recovery rate.

     Last Updated
     State of the Great Lakes 2005
288

-------

                                                      ^  ._
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''
Non-native Species - Aquatic
Indicator #9002

Overall Assessment
           Status:  Poor
           Trend:  Deteriorating
   Primary Factors  NIS continue to be discovered in the Great Lakes. Negative impacts of
      Determining  established invaders persist and new negative impacts are becoming
  Status and Trend  evident
Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Fair
Unchanging
Lake Superior is the site of most ballast water discharge in the Great Lakes,
but supports relatively few NIS. This is due at least in part to less
hospitable environmental conditions.
Poor
Deteriorating
Established invaders continue to exert negative impacts on native species.
Diporeia populations are declining.
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend

Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Poor
Deteriorating
Established invaders continue to exert negative impacts on native species.
Diporeia populations are declining.
Poor
Deteriorating
Established invaders continue to exert negative impacts on native species.
A possible link exists between waterfowl deaths  due to botulism and
established NIS (round goby and dreissenid mussels)
Poor
Deteriorating
Native Diporeia populations are declining in association with quagga
mussel expansion.  Condition and growth of lake whitefish, whose primary
food source is Diporeia, are declining. A possible link exists between
waterfowl deaths due to botulism and established NIS (round goby and
dreissenid mussels).
                         Draft for Discussion at SOLEC 2006

-------
Purpose
•To assess the presence, number and distribution of nonindigenous species (NIS) in the
Laurentian Great Lakes; and
•To aid in the assessment of the status of biotic communities, because nonindigenous species can
alter both the structure and function of ecosystems.

Ecosystem Objective
The goal of the U.S. and Canada Great Lakes Water Quality Agreement is, in part, to restore and
maintain the biological integrity of the waters of the Great Lakes ecosystem. Minimally,
extinctions and unauthorized introductions must be prevented to maintain biological integrity.

State of the Ecosystem
Background
Nearly 10% of NIS introduced to the Great Lakes have had significant impacts on ecosystem
health, a percentage consistent with findings in the United Kingdom (Williamson and Brown
1986) and in the Hudson River of North America (Mills et al. 1997). In the Great Lakes,
transoceanic ships are the primary invasion vector. Other vectors, such as canals and private
sector activities, however, are also utilized by NIS with potential to harm biological integrity.

Status of NIS
Human activities associated with transoceanic shipping are responsible for over one-third of NIS
introductions to the Great Lakes (Figure 1). Total numbers of NIS introduced and established in
the Great Lakes have increased steadily since the 1830s (Figure 2a). Numbers of ship-introduced
NIS, however, have increased exponentially during the same time period (Figure 2b). Release of
contaminated ballast water by transoceanic ships has been implicated in over 70% of faunal NIS
introductions to the Great Lakes since the opening of the St. Lawrence Seaway in 1959
(Grigorovich et al. 2003).

During the 1980s, the importance of ship ballast water as a vector for NIS introductions was
recognized, finally prompting ballast management measures in the Great Lakes. In the wake of
Eurasian ruffe and zebra mussel introductions, Canada introduced voluntary ballast exchange
guidelines in 1989 for ships declaring "ballast on board" (BOB) following transoceanic voyages,
as recommended by the Great Lakes Fishery Commission and the International Joint
Commission. In 1990, the United States  Congress passed the Nonindigenous Aquatic Nuisance
Prevention and Control Act, producing the Great Lakes' first ballast exchange and management
regulations in May  of 1993. The National Invasive Species Act (NISA) followed in 1996, but this
act expired in 2002. A stronger version of NISA entitled the Nonindigneous Aquatic Invasive
Species Act has been drafted and awaits Congressional reauthorization.

Contrary to expectations, the reported invasion rate has increased following initiation of
voluntary guidelines in 1989 and mandated regulations in 1993 (Grigorovich et al. 2003, Holeck
et al. 2004). However, >90% of transoceanic ships that entered the Great Lakes during the 1990s
declared "no ballast on board" (NOBOB, Colautti et al. 2003; Grigorovich et al. 2003; Holeck et
al. 2004) (Figure 3) and were not required to exchange ballast, although their tanks contained
residual sediments and water that would be discharged in the Great Lakes. Recent studies suggest
that the Great Lakes may vary in vulnerability to invasion in space and time. Lake Superior
                         Draft for Discussion at SOLEC 2006

-------

                                                        ^ ._
                                    ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''
receives a disproportionate number of discharges by both BOB and NOBOB ships, yet it has
sustained surprisingly few initial invasions (Figure 4); conversely, the waters connecting lakes
Huron and Erie are an invasion 'hotspot' despite receiving disproportionately few ballast
discharges (Grigorovich et al. 2003). Ricciardi (2001) suggests that some invaders (such as
Dreissena spp.) may facilitate the introduction of coevolved species such as round goby and the
amphipod Echinogammarus.

Other vectors, including canals and the private sector, continue to deliver NIS to the Great Lakes
and may increase in relative importance in the future. Silver and bighead carp escapees from
southern U.S. fish farms have been sighted below an electric dispersal barrier in the Chicago
Sanitary and Ship Canal, which connects the Mississippi River and Lake Michigan. The
prototype barrier was activated in April 2002, to block the transmigration of species between the
Mississippi River system and the Great Lakes basin. The U.S. Army Corps of Engineers
(partnered by the State of Illinois) completed construction of a second, permanent barrier in 2005.

Second only to shipping, unauthorized release, transfer, and escape have introduced NIS into the
Great Lakes. Of particular concern are private sector activities related to aquaria, garden ponds,
baitfish, and live food fish markets. For example, nearly a million Asian carp, including bighead
and black carp, are sold annually at fish markets within the  Great Lakes basin. Until recently,
most of these fish were sold live. All eight Great Lakes states and the province of Ontario now
have some restriction on the sale of live Asian carp. Enforcement of many private transactions,
however, remains a challenge. The U.S. Fish and Wildlife Service is considering listing several
Asian carp as nuisance species under the Lacey Act, which would prohibit interstate transport.
Finally, there are currently numerous shortcomings in legal safeguards relating to commerce in
exotic live fish as identified by Alexander (2003) in Great Lakes  and  Mississippi River states,
Quebec, and Ontario. These include: express and de facto exemptions for the aquarium pet trade;
de facto exemptions for the live food fish trade; inability to proactively enforce import bans; lack
of inspections at aquaculture facilities; allowing aquaculture in public waters; inadequate
triploidy (sterilization) requirements; failure to regulate species of concern, e.g., Asian carp;
regulation through "dirty lists" only, e.g., banning known nuisance species; and failure to regulate
transportation.

Pressures
NIS have invaded the Great Lakes basin from regions around the globe (Figure54), and
increasing world trade and travel will elevate the risk that additional species (Table 1) will
continue to gain access to the Great Lakes. Existing connections between the Great Lakes
watershed and systems outside the watershed, such as the Chicago Sanitary and Ship Canal, and
growth of industries such as aquaculture, live food markets, and aquarium retail stores will also
increase the risk that NIS will be introduced.

Changes in water quality, global climate  change, and previous NIS introductions also  may make
the Great Lakes more hospitable for the arrival of new invaders. Evidence indicates that newly
invading species may benefit from the presence of previously established invaders. That is, the
presence of one NIS may facilitate the establishment of another (Ricciardi 2001). For  example,
round goby and Echinogammarus have benefited from previously established zebra and quagga
mussels. In effect, dreissenids have set the stage to increase the number of successful invasions,
particularly those of co-evolved species in the Ponto-Caspian assemblage.
                         Draft for Discussion at SOLEC 2006

-------
Management Implications
Researchers are seeking to better understand links between vectors and donor regions, the
receptivity of the Great Lakes ecosystem, and the biology of new invaders in order to make
recommendations to reduce the risk of future invasion. To protect the biological integrity of the
Great Lakes, it is essential to closely monitor routes of entry for NIS, to introduce effective
safeguards, and to quickly adjust safeguards as needed. Invasion rate may increase if positive
interactions involving established NIS or native species facilitate entry of new NIS.  Ricciardi
(2001) suggested that such a scenario of "invasional meltdown" is occurring in the Great Lakes,
although Simberloff (2006) cautioned that most of these cases have not been proven.
To be effective in preventing new invasions, management strategies must focus on linkages
between NIS, vectors, and donor and receiving regions. Without measures that effectively
eliminate or minimize the role of ship-borne and other, emerging vectors, we can expect the
number of NIS in the Great Lakes to continue to rise, with an associated loss of native
biodiversity and an increase in unpredicted ecological disruptions.

Comments from the author(s)
Lake by lake assessment should include Lake St. Clair and connecting channels (Detroit River,
St. Clair River). Species first discovered in these waters were assigned to Lake Erie for the
purposes of this report.

Acknowledgments
Authors: Edward L. Mills, Department of Natural Resources, Cornell University, Bridgeport, NY;
Kristen T. Holeck, Department of Natural Resources, Cornell University, Bridgeport, NY; and
Hugh Maclsaac, Great Lakes Institute for Environmental Research, University of Windsor,
Windsor, ON, Canada

Data Sources
Alexander, A. 2003. Legal tools and gaps relating to commerce in exotic live fish: phase 1 report
to the Great Lakes Fishery Commission by the Environmental Law and Policy  Center.
Environmental Law and Policy Center, Chicago, IL.

Colautti, R.I., Niimi, A.J., van Overdijk, C.D.A., Mills, E.L., Holeck, K.T., and Maclsaac, HJ.
2003. Spatial and temporal analysis of transoceanic shipping vectors to the Great Lakes. In
Invasion Species: Vectors and Management Strategies, eds. G.M. Ruiz and J.T. Carlton, pp. 227-
246. Washington, DC: Island Press.

Grigorovich, I.A., Colautti, R.I., Mills, EX., Holeck, K.T., Ballert, A.G., and Maclsaac, HJ.
2003. Ballast-mediated animal introductions in the Laurentian Great Lakes: retrospective and
prospective analyses. Can. J. Fish. Aquat. Sci. 60:740-756.

Holeck, K.T., Mills, E.L., Maclsaac, H.J., Dochoda, M.R., Colautti, R.I., and Ricciardi, A. 2004.
Bridging troubled waters: understanding links between biological invasions, transoceanic
shipping, and other entry vectors in the Laurentian Great Lakes. Bioscience 54:919-929.
                         Draft for Discussion at SOLEC 2006

-------

                                                       ^ .
                                   ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''  './ ''l3i™B*j,Arf(St';j(\f^*
                                                             ^!?r?lB*?
Kolar, C.S., and Lodge, D.M. 2002. Ecological predictions and risk assessment for alien fishes in
North America. Science 298:1233-1236.

Mills, E.L., Leach, J.H., Carlton, J.T., and Secor, C.L. 1993. Exotic species in the Great Lakes: A
history of biotic crises and anthropogenic introductions. J. Great Lakes Res. 19(l):l-54.

Mills, E.L., Scheuerell, M.D., Carlton, J.T., and Strayer, D.L.  1997. Biological invasions in the
Hudson River. NYS Museum Circular No. 57. Albany, NY.

Ricciardi, A. 2006. Patterns of invasions in the Laurentian Great Lakes in relation to changes in
vector activity. Diversity and Distributions 12: 425-433.

Ricciardi, A. 2001. Facilitative interactions among aquatic invaders: is an"invasional meltdown"
occurring in the Great Lakes? Can. J. Fish. Aquat. Sci. 58:2513-2525.

Ricciardi, A., and Rasmussen, J.B. 1998. Predicting the identity and impact of future biological
invaders: a priority for aquatic resource management. Can. J. Fish. Aquat. Sci. 55:1759-1765.

Rixon, C.A.M., Duggan, I.C., Bergeron, N.M.N., Ricciardi, A., and Maclsaac, HJ. 2004.
Invasion risks posed by the aquarium trade and live fish markets on the Laurentian  Great Lakes.
Biodiversity and Conservation (in press).

Simberloff, D. 2006. Invasional meltdown 6 years later: important phenomenon, unfortunate
metaphor, or both? Ecology Letters (in press).

Stokstad, E. 2003. Can well-timed jolts keep out unwanted exotic fish? Science 301:157-158.

Williamson, M.H., and Brown, K.C. 1986. The analysis and modeling of British invasions.
Philosophical Transactions of the Royal Society of London, Series B. 314:505-522.

List of Tables
Table 1. Nonindigenous species predicted to have a high-risk of introduction to the Great Lakes.
Source: Ricciardi and Rasmussen 1998; Kolar and Lodge 2002; Grigorovich et al. 2003; Stokstad
2003; Rixon et al. 2004

List of Figures
Figure 1. Release mechanisms for aquatic nonindigenous (NIS) established in the Great Lakes
basin since the 1830s.  Source: Mills et al. 1993; Ricciardi 2001; Grigorovich et al. 2003;
Ricciardi 2006

Figure 2. Cumulative number of aquatic nonindigenous (NIS) established in the Great Lakes
basin since the 1830s attributed to (a) all vectors and (b) only the ship vector.
Source: Mills et al. 1993; Ricciardi 2001; Grigorovich et al. 2003; Ricciardi 2006

Figure 3. Numbers of upbound transoceanic vessels entering the Great Lakes from  1959 to 2002.
Source: Colautti et al. 2003; Grigorovich et al. 2003; Holeck et al. 2004
                         Draft for Discussion at SOLEC 2006

-------
Figure 4. Lake of first discovery for NIS established in the Great Lakes basin since the 1830s.
Discoveries in connecting waters between Lakes Huron, Erie and Ontario were assigned to the
downstream lake.

Figure 5. Regions of origin for aquatic NIS established in the Great Lakes basin since the 1830s.
Source: Mills et al. 1993; Ricciardi 2001; Grigorovich et al. 2003; Ricciardi 2006

Last updated
SOLEC 2006
 Lake/Basin of First
 Discovery                    Fauna       Flora
 Unknown/Widespread                 33                        9
 Multiple                               4                        1
 Ontario                              24                       33
 Erie                                 16                       21
 Huron                                4                        3
 Michigan                             11                       16
 Superior                               3                        4
                                      95                       87         182

Table 1. Nonindigenous species predicted to have a high-risk of introduction to the Great
Lakes.
Source: Ricciardi and Rasmussen 1998; Kolar and Lodge 2002; Grigorovich et al. 2003; Stokstad
2003 ;Rixon et al. 2004
                        Draft for Discussion at SOLEC 2006

-------

                                                                                Total = 182
         Accidental   Aquarium   Shipping
          release     release
Cultivation   Deliberate  Natural means Railroads and  Solid ballast   Unknown
 release     release             highways
Primary mechanism
Figure 1. Release mechanisms for aquatic nonindigenous (NIS) established in the Great Lakes
basin since the 1830s.
Source: Mills et al. 1993; Ricciardi 2001; Grigorovich et al. 2003; Ricciardi 2006
                            Draft for Discussion at SOLEC 2006

-------
         1830s
                  1850s
                          1870s
                                   1890s
                                            1910s     1930s
                                              Decade
                                                              1950s
                                                                      1970s
                                                                               1990s
Figure 2a. Cumulative number of aquatic nonindigenous (NIS) established in the Great Lakes
basin since the 1830s attributed to all vectors.
Source: Mills et al. 1993; Ricciardi 2001; Grigorovich et al. 2003; Ricciardi 2006
    80
    70
    60
  I
    50
    40
    30
   E
   5
  o
    20
    10
                                                            Total = 70
      1850
               1870
                         1890
                                   1910
                                             1930
                                             Year
                                                       1950
                                                                 1970
                                                                           1990
                                                                                     2010
Figure 2b. Cumulative number of aquatic nonindigenous (NIS) established in the Great Lakes
basin since the 1830s attributed to the ship vector.
Source: Mills et al. 1993; Ricciardi 2001; Grigorovich et al. 2003; Ricciardi 2006
                         Draft for Discussion at SOLEC 2006

-------


     1600 -r
     140°
I
     1200
   8 1000
      600
   |  400

   .Q
   Q.

   D  200




        0
                                                                         • BOB


                                                                         DNOBOB
         1959
                  1964
                          1969
                                   1974
                                           1979
                                                    1984
                                                            1989
                                                                     1994
                                                                             1999
Figure 3. Numbers of upbound transoceanic vessels entering the Great Lakes from 1959 to 2002.

Source: Colautti et al. 2003; Grigorovich et al. 2003; Holeck et al. 2004
                        Draft for Discussion at SOLEC 2006

-------
       Unknown/Widespread   Multiple
                                  Ontario        Erie         Huron
                                    Lake/Basin of first discovery
                                                                     Michigan
                                                                                Superior
Figure 4. Lake of first discovery for NIS established in the Great Lakes basin since the 1830s.
Discoveries in connecting waters between Lakes Huron, Erie and Ontario were assigned to the
downstream lake.
Source: Grigorovich et al. 2003
                          Draft for Discussion at SOLEC 2006

-------
                                                                        Total = 182
      Africa/Asia        Atlantic         Eurasia       Mississippi        Pacific        Unknown
                                                  basin
                                            Donor region

Figure 5. Regions of origin for aquatic NIS established in the Great Lakes basin since the 1830s.
Source: Mills et al. 1993; Ricciardi 2001; Grigorovich et al. 2003; Ricciardi 2006
                          Draft for Discussion at SOLEC 2006

-------

                                                    ^  ._
                                  ^ fttf"j%*-l5''|wS^rjp3fe-™=*' 	'; ,fe\i^ * if'/i" ''
Non-native Species - Terrestrial
Indicator #9002
      Determining
  Status and Trend
Overall Assessment
           Status:   Mixed
           Trend:   Deteriorating/Undetermined
   Primary Factors   Terrestrial Non-indigenous species are pervasive in the Great Lakes
                   basin. Although not all introductions have an adverse effect on native
                   habitats, those that do pose a considerable ecological, social, and
                   economic burden. Historically, the Great Lakes Basin has proven to be
                   particularly vulnerable to non-indigenous species, mainly due to the
                   high volume of transboundary movement of goods and people,
                   population, and industrialization. Improved monitoring of non-
                   indigenous species is needed to adequately assess the status, trends, and
                   impacts of non-indigenous species in the region.

Lake-by-Lake Assessment
Lake Superior
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                  Not Assessed
                  Undetermined
                  Not available at this time.
Lake Michigan
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                  Not Assessed
                  Undetermined
                  Not available at this time.
Lake Huron
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                  Not Assessed
                  Undetermined
                  Not available at this time.
Lake Erie
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
                  Not Assessed
                  Undetermined
                  Not available at this time.
                        Draft for Discussion at SOLEC 2006

-------
Lake Ontario
           Status:
           Trend:
   Primary Factors
      Determining
  Status and Trend
Not Assessed
Undetermined
Not available at this time.
Purpose
    •  To evaluate the presence, number, and impact of terrestrial non-indigenous species in the
       Great Lakes Basin.
    •  To assess the biological integrity of the Great Lakes Basin ecosystems.

Ecosystem Objective
The ultimate goal of this indicator is to limit, or prevent, the unauthorized introduction of non-
indigenous species, and to minimize their adverse affect in the Great Lakes Basin. Such actions
would assist in accomplishing one of the major objectives of U.S. and Canada Great Lakes Water
Quality Agreement, which is to restore and maintain the biological integrity of the waters of the
Great Lakes ecosystem.

State of the Ecosystem
Globalization, i.e. the movement of people and goods, has led to a dramatic increase in the
number of terrestrial non-indigenous species (NIS) that are transported from one country to
another. As a result of its high population density and high-volume  transportation of goods, the
Great Lakes Basin (GLB) is very susceptible to the introduction of such invaders.  Figure 1
depicts this steady increase in the number of terrestrial NIS introduced into the GLB and the rate
at which this has occurred, beginning in the 1900s.  In addition, the degradation, fragmentation,
and loss of native ecosystems have also made this region more vulnerable to these invaders,
enabling them to become invasive (non-indigenous species or strains that become established in
native communities or wild areas and replace native species). As such, the introduction of NIS is
considered to be one of the greatest threats to the biodiversity and natural resources of this region,
second only to habitat destruction.

Monitoring of NIS is largely locally based, as a region-wide standard has yet to be established.
As  a result, the data that is generated comes from a variety of agencies and organizations
throughout the region, thus providing some difficulty when attempting to assess the overall
presence and impact these species are having on the region.  Information provided by the World
Wildlife Fund of Canada indicates that there are 157 exotic plants and animals located within the
GLB, which includes:  95 vascular plants, 11 insects, 6 plant diseases, 4 mammals, 2 birds, 2
animal diseases, 1 reptile, and 1 amphibian. However, the Invasive Plant Association of
Wisconsin has identifies 116 non-native plants within the state, while over one hundred plants
have been introduced into the Chicago region (Chicago Botanic Garden). Even though these
figures are greater then the one provided by the WWF of Canada, they do not compare to the over
900 non-native plants that have been identified within the state of Michigan by the Michigan
Invasive Plant Council.
                         Draft for Discussion at SOLEC 2006

-------

                                                       ^  ._
                                    ^ fttf"j%*-l5''|wS^rjp3fe-™=*'  	'; ,fe\i^ * if'/i" ''
The impact NIS have on the areas in which they are introduced can vary greatly, ranging from
little or no affect to dramatically altering the native ecological community.  Figure 2 shows the
degree to which each taxonomic group has had an impact on the ecoregion. The WWF of Canada
has listed 29 species, 19 of which are vascular plants, as having a "severe impact" on native
biodiversity.  These species, which were generally introduced for medicinal or ornamental
purposes, have become problematic as they continue to thrive due to the fact that they are well
adapted to a broad range of habitats, have no native predators, and are often able to reproduce at a
rapid rate. Common buckthorn, garlic mustard, honeysuckle, purple loosestrife, and reed canary
grass are several examples of highly invasive plant species, while the Asian longhorn beetle,
Dutch elm disease, emerald ash borer, leafy spurge, and the West Nile virus are other terrestrial
invaders that have had a significant impact of the GLB.

One type of terrestrial non-native species not covered in this report is genetically modified
organisms (GMOs). Although GMOs are typically cultivated for human uses and benefits, the
problem arises when pollen is moved from its intended site (often by wind or pollinator species)
and transfers genetically engineered traits, such as herbicide resistance and pest resistance, to
wild plants. This outward gene flow into natural habitats has the potential to significantly alter
ecosystems and create scenarios that would pose enormous dilemmas for farmers. Both Canada
and the U.S. are major producers  of genetically modified organisms (GMOs). Although GMO
crops are monitored for outward gene flow, no centralized database describing the number of
GMO species, or land area covered by GMOs in the Great Lakes Basin currently exists.

There are currently numerous policies, laws and regulations within the GLB that address NIS;
however, similar to NIS monitoring, they originate from state, provincial and federal
administrations and thus have similar obstacles associated with them. As such, strict enforcement
of these laws, in  addition to continuous region-wide mitigation, eradication and management of
NIS is needed in order to maintain the ecological integrity of the GLB.

Pressures
The growing transboundary movement of goods and people has heightened the need to prevent
and manage terrestrial NIS. Most cases of invasiveness can be linked to the intended or
unintended consequences of economic activities (Perrings, et al., 2002). For this reason, the GLB
has been, and will continue to be, a hot bed of introductions, unless preventive measures are
enforced. The growth in population, threats, recreation and tourism all contribute to the number
of NIS affecting the region. Additionally, factors such as the increase in development and human
activity, previous introductions and climate change have elevated the levels of vulnerability.
Because this issue has social, ecological, and economic dimensions it can be assumed that the
pressure of NIS will persist unless it is addressed on all three fronts.

Management Implications
Since the early 1800s, biological invasions have compromised the ecological integrity of the
GLB. Despite an elevated awareness of the issue and efforts to prevent and manage NIS in the
Great Lakes, the area remains highly vulnerable to both intentional and non-intentional
introductions. Political  and social motivation to address this issue is driven not only by the effects
on the structure and function of regional ecosystems, but also by the  cumulative economic impact
of invaders, i.e. threats  to food supplies and human health.
                         Draft for Discussion at SOLEC 2006

-------
Managers of terrestrial NIS in the GLB recognize that successful management strategies must
involve collaboration across federal, provincial and state governments, in addition to non-
governmental organizations. Furthermore, improved integration, coordination and development
of inventories, mapping, and mitigation of terrestrial invasive species can be used to adapt future
strategies and examine trends in terrestrial NIS at a basin-wide scale. Although current
monitoring programs in Canada are fragmented at best, a number of initiatives involving broad-
stakeholder participation and government collaboration are being developed to  determine future
priorities. This information will be applied to risk analysis, predictive science, modeling,
improved technology for prevention and management of NIS, legislation and regulations,
education and outreach, and international co-operation to encompass the multi-faceted aspect of
this ecological, social, and economic issue.

Comments from the author(s)
Currently, there is no central monitoring site for terrestrial NIS in Canada. In 1997 the Canadian
Botanical Conservation Network put together a database on invasive plant species for Canada, but
the information has not since been updated. In 2000 the World Wildlife Fund of Canada amassed
information about 150 known NIS in Canada in a centralized database, based on books, journal
articles, websites, and consultation with experts. The  author of the chapter acknowledges  that a
lack of centralized data was a limitation of the project. The information contained in this indicator
is based on the WWF-C database and has been updated with several more recent insect invaders
present in the GLB.

Acknowledgments
Authors:  Katherine Balpataky, Program Officer, Environment Canada - Ontario Region,  867
Lakeshore Road, Burlington, Ontario, Canada, L7R 4A6, (905) 336-6271.
Jeffrey C. May,  U.S. Environmental Protection Agency, GLNPO Intern. 77 W. Jackson Blvd (G-
17J) Chicago, Illinois 60604, May.Jeffrey@epa.gov

Contributors: Haber, Erich, National Botanical Services (author of the WWF-C report), Ottawa,
ON, K2A 3A8.
Hendrickson, Ole, Environment Canada, Biodiversity Convention Office, Gatineau, QC, K1A
OH3.
Morgan, Alexis, World Wildlife Fund-Canada, 245 Eglinton Ave. East, suite 410, Toronto, ON,
M4P3J1.
Wallace, Shaun, Plant Pest Surveillance Unit, Canadian Food Inspection Agency, 3851
Fallowfield Rd., Nepean, Ontario, K2H 8P9.
Canadian Food Inspection Agency. Plant Health Division. Proposed Action Plan for Invasive
Alien Terrestrial Plants and Plant Pests Phase 1,  2004.
http://www.cbin.ec.gc.ca/primers/ias_plants.cfm?lang=e, last viewed 28 August 2006.

Environment Canada. Biodiversity Convention Office. An Invasive Alien Species Strategy for
Canada. 2004. http://www.cbin.ec.gc.ca/primers/ias.cfin, last viewed 28 August 2006.

Food and Agricultural Organization. (2001). The state of food and agriculture 2001. Rome, Italy.
Available on the World Wide Web: http://www.fao.org/docrep/003/x9800e/x9800el4.htm.
                         Draft for Discussion at SOLEC 2006

-------


Gwenaelle Dauphina,, Stephan Zientaraa, Herve Zellerb, Bernadette Murgue. West Nile:
worldwide current situation in animals and humans. G. Dauphin et al. / Comp. Immun. Microbiol.
Infect. Dis. 27 (2004) 343-355.

Haack, Robert A. Intercepted Scolytidae (Coleoptera) at U.S. ports of entry: 1985-2000.
Integrated Pest Management Reviews 6 (2001) 253-282.

IJC, CMI, International Joint Commission, Commission mixte internationale. Then and Now:
Aquatic Alien Invasive Species. 2004. http://www.ijc.org/rel/pdf/ThenandNow_f.pdf.

Lavoie, Claude; Jean, Martin; Delisle, Fanny; Letourneau, Guy. Exotic plant species of the St.
Lawrence River wetlands: a spatial and historical analysis. Journal of Biogeography. 30: (2003)
537-549.

Leung, B.; Finnoff, D.; Shogren, J.F.; Lodge, D. Managing invasive species: Rules of thumb for
rapid assessment. Ecological Economics. 55 (2005) 24-36.

Natural Resources Canada. 2006. Our Forests Under Threat. http://www.cfl.scf.rncan.gc.ca/CFL-
LFC/publications/activites/menace_e.html. last accessed August 28, 2006.

Maclsaac, H.J., LA. Grigorovich, and A. Ricciardi. Reassessment of species invasions concepts:
the Great Lakes basin as a model. Biological Invasions. 3: 405-416, 2001.

Mills, E.L., Leach, J.H., Carlton, J.T., Secor, C.L., Exotic species and the integrity of the Great
Lakes. Bioscience. 44, (1994) 666-676.

Mills, E.L., Holeck, K.T., Chrisman, J.R., 1999. The role of canals in the spread of non-
indigenous species in North America. In: Claudi, R., Leach, J. (Eds.). Non-indigenous Organisms
in North America: Their Biology and Impact, CRC Press LCL, Boca Raton, FL, pp. 345-377.

Midwest Natural Resources Group. 2006. Action Plan for Addressing Terrestrial Invasive Species
Within the Great Lakes Basin,  http://www.mnrg.gov. last viewed August 28, 2006.

Perrings, C., Williamson, M., Barbier, E., Delfino, D., Dalmazzone,  S., Shogren, J., Simmons, P.,
and Watkinson, A. (2002). Biological invasion risks  and the public good: An economic
perspective.  Conservation Ecology 6(1), 1. Available on the World Wide Web:
http://www.consecol.org/vol6/issl/artl.

Ricciardi, A. Patterns of invasion of the Laurentian Great Lakes in relation to changes in vector
activity. Diversity and Distributions 12: (2006) 425-433.

Wilkins, Pamela, and Del Piero Fabio. West Nile virus: lessons from the 21st century.
Journal of Veterinary Emergency and Critical Care  14(1) 2004, pp 2-14.
                         Draft for Discussion at SOLEC 2006

-------
List of Figures
Figure 1. A timeline of terrestrial introduction in the Great Lakes Basin by taxonomic group.
Data source: World Wildlife Fund-Canada's Exotic Species Database, and the Canadian Food
Inspection Agency.

Figure 2. Estimated impact of 124 known terrestrial NIS in the Great Lakes Basin.
Data source: World Wildlife Fund-Canada's Exotic Species Database.

Last updated
SOLEC 2006
    120 i
    100
  S1
  o
  •s
     40
     20
      1900
                   1920
                                1940
                                             1960
                                             Year
                                                           1980
                                                                        2000
                                                                                     2020
                  •Total Species •
                              -Total Insect
                                         Total Vascular plant
                                                         Total Bird •
                                                                 -Total Plant disease
Figure 1. A timeline of terrestrial introduction in the Great Lakes Basin by taxonomic group.
Data source: World Wildlife Fund-Canada's Exotic Species Database, and the Canadian Food
Inspection Agency.
                         Draft for Discussion at SOLEC 2006

-------



                              Impact on Ecosystem by taxonomic group
                              Slight
                                      Impact
Figure 2. Estimated impact of 124 known terrestrial NIS in the Great Lakes Basin by taxonomic
group.
Data source: World Wildlife Fund-Canada's Exotic Species Database.
                         Draft for Discussion at SOLEC 2006

-------

List of indicators by category




Contamination indicators
Status, Trend
Open Lake:
Mixed, Undetermined
Nearshore:
Poor, Undetermined
Mixed, Improving
SU, HU, ER, ON: mixed,
improving
MI:NA
Mixed, Improving
SU: good, improving
MI, HU, ER: mixed, improving
ON: poor, improving
Mixed,
Improving/Unchanging
Mixed, Undetermined
SU, MI, HU: fair,
undetermined
ER, ON: mixed, undetermined
Mixed,
Improving/Undetermined
Mixed, Improving
SU, MI, HU, ER, ON: fair,
improving
Poor, Unchanging
SU, MI, HU: undetermined
ER, ON: poor, unchanging
Good, Unchanging
Mixed, Undetermined

Mixed, Improving
Mixed, Improving
Mixed, Undetermined
SU, MI, HU: undetermined
ER, ON: mixed, undetermined
Undetermined
Progress Report
Mixed, Improving
Mixed, Undetermined
Mixed, Improving
Indicator Title (indicator number)
Phosphorus Concentrations and Loadings (111)
Contaminants in Young-of-the-Year Spottail Shiners (114)
Contaminants in Colonial Nesting Waterbirds (115)
Atmospheric Deposition of Toxic Chemicals (117)
Toxic Chemical Concentrations in Offshore Waters (118)
Concentrations of Contaminants in Sediment Cores (119)
Contaminants in Whole Fish (121)

External Anomaly Prevalence Index for Nearshore Fish (124)
Drinking Water Quality (4175)
Biologic Markers of Human Exposure to Persistent Chemicals
(4177)
Contaminants in Sport Fish (4201)
Air Quality (4202)
Contaminants in Snapping Turtle Eggs (4506)

Nutrient Management Plans (7061)
Wastewater Treatment and Pollution (7065)
Contaminants Affecting Productivity of Bald Eagles (8135)
Population Monitoring and Contaminants Affecting the American
Otter (8 147)
Acid Rain (9000)
Year
2006
2006
2006
2006
2006
2006
2006

2006
2006
2006

2006
2006
2006

2005
2006
2005
2003
2005
                      Draft for Discussion at SOLEC 2006

-------
Biotic Communities indicators
Status, Trend
Mixed, Improving
SU: fair, improving
MI: mixed, slightly improving
HU: fair, improving
ER: good, improving
ON: mixed, unchanging
Fair, Unchanging
Mixed, Deteriorating
SU: mixed, improving
MI, HU, ER, ON: mixed,
deteriorating
Undetermined
Mixed, Unchanging
SU: good, improving
MI: poor, declining
HU: mixed, improving
ER: mixed, unchanging
ON: mixed, declining
Mixed,
Unchanging/Deteriorating
SU: good, unchanging
MI, ER: mixed,
unchanging/deteriorating
HU, ON: mixed, unchanging
Mixed, Undetermined
Mixed, Improving
SU: good, improving
MI, HU, ER: mixed, improving
ON: poor, improving
Mixed, Undetermined
SU: good, unchanging
MI, HU, ER, ON: undetermined
Mixed, Improving
SU, MI, HU: poor, undetermined
ER: good/mixed,
improving/mixed
ON: undetermined
Mixed, Deteriorating
SU: mixed, unchanging
MI, HU, ER, ON: poor,
deteriorating
Mixed, Improving
SU, MI, HU: mixed,
improving/undetermined
ER: poor, undetermined
ON: mixed, improving
Progress Report
Undetermined
Mixed, Deteriorating
SU: undetermined
MI: poor, unchanging
HU, ER: mixed, deteriorating
ON: mixed, unchanging
Indicator Title (indicator number)
Salmon and Trout (8)

Walleye (9)
Preyfish Populations (17)
Native Freshwater Mussels (68)
Lake Trout (93)
Benthos Diversity and Abundance - Aquatic Oligochaete
Communities (104)
Phytoplankton Populations (109)
Contaminants in Colonial Nesting Waterbirds (115)
Zooplankton Populations (116)
Hexagenia (122)
Abundances of the Benthic Amphipod Diporeia spp. (123)
Status of Lake Sturgeon in the Great Lakes (125)
Coastal Wetland Invertebrate Community Health (4501)
Coastal Wetland Fish Community Health (4502)
Wetland-Dependent Amphibian Diversity and Abundance
(4504)
Year
2006
2006
2006
2005
2006
2006
2003
2006
2006
2006
2006
2006
2005
2006
2006
                      Draft for Discussion at SOLEC 2006

-------

                                flpflf^C^YaVy't:"*^^- '?•'''•'/'• '«
                                |p/Jst)*^Ji^ M.V-.VV.-I*' ; v
                                fe.aA-?  •••:
Biotic Communities indicators (continued)
Mixed, Deteriorating
SU: undetermined
MI, ER, ON: mixed, deteriorating
HU: poor, deteriorating
Mixed, Undetermined
SU: good, unchanging
MI, ER: mixed, unchanging
HU: mixed, deteriorating
ON: poor, unchanging
Undetermined
Mixed, Improving
Mixed, Undetermined
Mixed, Undetermined
Wetland-Dependent Bird Diversity and Abundance (4507)
Coastal Wetland Plant Community Health (4862)
Groundwater Dependant Plant and Animal Communities
(7103)
Contaminants Affecting Productivity of Bald Eagles
(8135)
Population Monitoring and Contaminants Affecting the
American Otter (8147)
Forest Lands-Conservation of Biological Diversity (8500)
2006
2006
2005
2005
2003
2006
Invasive Species indicators
Good/Fair, Improving
Poor, Deteriorating
SU: fair, unchanging
MI, HU, ER, ON: poor,
deteriorating
Mixed,
Deteriorating/Undetermined
Sea Lamprey (18)
Non-native Species — Aquatic (9002)
Non-native Species — Terrestrial (9002)
2005
2006
2006
                        Draft for Discussion at SOLEC 2006

-------
Coastal Zones indicators
Status, Trend
Progress Report
Undetermined
Mixed, Deteriorating
SU: undetermined
MI: poor, unchanging
HU, ER: mixed, deteriorating
ON: mixed, unchanging
Mixed, Undetermined
SU, MI, HU: undetermined
ER, ON: mixed, undetermined
Mixed, Deteriorating
SU: undetermined
MI, ER, ON: mixed,
deteriorating
HU: poor, deteriorating
Mixed, Deteriorating
Mixed, Undetermined
Mixed, Undetermined
SU: good, unchanging
MI, ER: mixed, unchanging
HU: mixed, deteriorating
ON: poor, unchanging
Progress Report
Mixed, Undetermined
Mixed, Deteriorating
Progress Report
Mixed, Undetermined
SU: good, undetermined
MI: undetermined
HU, ER, ON: mixed,
undetermined
Mixed, Deteriorating
Indicator Title (indicator number)
Coastal Wetland Invertebrate Community Health (4501)
Coastal Wetland Fish Community Health (4502)
Wetland-dependent Amphibian Diversity and Abundance (4504)
Contaminants in Snapping Turtle Eggs (4506)
Wetland-Dependent Bird Diversity and Abundance (4507)
Coastal Wetland Area by Type (4510)
Effect of Water Level Fluctuations (4861)
Coastal Wetland Plant Community Health (4862)
Land Cover Adjacent to Coastal Wetlands (4863)
Area, Quality, and Protection of Special Lakeshore
Communities — Alvars (8129)
Area, Quality, and Protection of Special Lakeshore
Communities — Cobble beaches (8129)
Area, Quality, and Protection of Special Lakeshore
Communities — Sand dunes (8129)
Area, Quality, and Protection of Special Lakeshore Communities
—Islands (8 129)
Extent of Hardened Shoreline (8131)
Year
2006
2006
2006
2006
2006
2005
2003
2006
2006
2001
2005
2005
2006
2001
                      Draft for Discussion at SOLEC 2006

-------


Aquatic Habitat indicators
Status/Trend
Open Lake:
Mixed, Undetermined
Nearshore:
Poor, Undetermined
Mixed, Improving
SU, MI, HU: fair, undetermined
ER, ON: mixed, undetermined
Mixed,
Improving/Undetermined
Undetermined
Undetermined
Mixed, Deteriorating
Undetermined
Mixed, Deteriorating
Indicator Title (indicator number)
Phosphorus Concentrations and Loadings (111)
Toxic Chemical Concentrations in Offshore Waters (118)
Concentrations of Contaminants in Sediment Cores (119)
Natural Groundwater Quality and Human-Induced Changes
(7100)
Groundwater and Land: Use and Intensity (7101)
Base Flow Due to Groundwater Discharge (7 1 02)
Groundwater Dependant Plant and Animal Communities (7103)
Extent of Hardened Shoreline (8131)
Year
2006
2006
2006
2005
2005
2006
2005
2001
Other sources of aquatic habitat information
Additional information on spatial and temporal trends in toxic contaminants in offshore waters
can be found in:
Marvin, C., S. Painter, D. Williams, V. Richardson, R. Rossmann, and P.Van Hoof. 2004.
Spatial and temporal trends in surface water and sediment contamination in the Laurentian Great
Lakes.  Environmental Pollution. 129(2004):  131-144.
Kannan, K., J. Ridal, and J. Struger.  2006. Pesticides in the Great Lakes.  Heidelberg
Environmental Chemistry 5(N): 151-199.
Great Lakes Binational Toxics Strategy. 2002 Progress Report. Environment Canada and US
EPA.
Great Lakes Binational Toxics Strategy Assessment of Level 1 Substances Summary.  Great
Lakes Binational Toxics Strategy (December 2005). U.S. EPA, Great Lakes National Program
Office and Environment Canada.
Additional information on base flow can be found in:
Neff, B.P., Day, S.M., Piggot, A.R., Fuller, L.M. 2005.  Base Flow in the Great Lakes Basin:
U.S. Geological Survey Scientific Investigations Report 2005-5217, 23p.
                        Draft for Discussion at SOLEC 2006

-------
Resource Utilization indicators
Status/trend
Undetermined
Mixed, Undetermined
SU: Mixed, Undetermined
MI, HU, ER, ON: undetermined
Mixed, Unchanging
Mixed, Undetermined
Undetermined
Poor, Deteriorating
Progress Report
Indicator Title (indicator number)
Commercial/Industrial Eco-Efficiency Measures (3514)
Economic Prosperity (7043)
Water Withdrawals (7056)
Energy Consumption (7057)
Solid Waste Disposal (7060)
Vehicle Use (7064)
Wastewater Treatment and Pollution (7065)
Year
2003
2003
2005
2005
2006
2006
2006
Land Use - Land Cover indicators
Status/Trend
Progress Report
Mixed, Undetermined
Undetermined
Mixed, Undetermined
Mixed, Improving
Undetermined
Progress Report
Undetermined
Undetermined
Mixed, Undetermined
Mixed, Deteriorating
Mixed, Undetermined
SU: good, undetermined
MI: undetermined
HU, ER, ON: mixed, undetermined
Progress Report
Undetermined
(Proposed Indicator)
Mixed, Undetermined
Undetermined
Mixed, Undetermined
Indicator Title (indicator number)
Land Cover Adjacent to Coastal Wetlands (4863)
Urban Density (7000)
Groundwater and Land: Use and Intensity (7101)
Land Cover/Land Conversion (7002)
Brownfields Redevelopment (7006)
Sustainable Agricultural Practices (7028)
Ground Surface Hardening (7054)
Nutrient Management Plans (7061)
Integrated Pest Management (7062)
Area, Quality and Protection of Special Lakeshore
Communities - Alvars (8129)
Area, Quality and Protection of Special Lakeshore
Communities - Cobble Beaches (8129)
Area, Quality and Protection of Special Lakeshore
Communities - Islands (8129)
Area, Quality and Protection of Special Lakeshore
Communities - Sand Dunes (8129)
Biodiversity Conservation Sites (8164)
Forest Lands - Conservation of Biological Diversity (8500)
Forest Lands - Maintenance of Productive Capacity of
Forest Ecosystems (8501)
Forest Lands - Conservation and Maintenance of Soil and
Water Resources (8503)
Year
2006
2006
2005
2006
2006
2005
2005
2005
2005
2001
2005
2006
2005
2006
2006
2006
2006
                      Draft for Discussion at SOLEC 2006

-------


 Human Health indicators
Status-Trend
Good, Unchanging
Mixed, Undetermined
Mixed, Unchanging
SU: good, undertermined
MI, ER, ON: fair, undetermined
HU: good, unchanging/undetermined
Mixed, Improving
Mixed, Improving
Indicator Title (indicator number)
Drinking Water Quality (4175)
Biological Markers of Human Exposure to Persistent
Chemicals (4177)
Beach Advisories, Postings and Closures (4200)
Contaminants in Sport Fish (4201)
Air Quality (4202)
Year
2006
2006
2006
2006
2006
 Other sources of human health information:
 Lake Wide Management Plans http ://www. epa. gov/glnpo/gl2000/lamps/index.html
 Agency for Toxic Substances and Disease Registry http ://www. atsdr. cdc.gov/grtlakes/index.html
 Climate Change indicators
Mixed, Deteriorating
Climate Change: Ice Duration on the Great Lakes (4858)
2003
 Other sources of climate change information:
 http://www.usgcrp.gov/usgcrp/nacc/greatlakes.htm
 http://www.nrel.colostate.edu/projects/brd_global_change/proj_3 l_great_lakes.html
 http://www.geo.msu.edu/glra/assessment/assessment.html
 http ://www. glerl.noaa.gov/res/Programs/ccmain.html
 http ://www.ucsusa. org/greatlakes/
                         Draft for Discussion at SOLEC 2006

-------
6.0 Acronyms and Abbreviations
Agencies and Organizations
ATSDR
CAMNet
CCME
CDC
CIS
CORA
CWS
DFO
EC
ECO
EIA
GLBET
GLC
GLCWC
GLFC
GLNPO
IJC
IUCN
MDEQ
MDNR
NHEERL
NOAA
NRC
NRCS
NYSDEC
ODNR
ODW
OFEC
OMAF
OMOE
OMNR
OSCIA
ORISE
PDEP
USDA
USEPA
USFDA
USFWS
USFS
USGS
WBCSD
WDNR
WDO
WiDPH

Units of Measure
fg
ha
kg
km
kt
kWh
m
Agency for Toxic Substances and Disease Registry
Canadian Atmospheric Mercury Network
Canadian Council of Ministers of the Environment
Center for Disease Control (U.S.)
Canada Ice Service
Chippewa Ottawa Resource Authority
Canadian Wildlife Service
Canada Department of Fisheries and Oceans
Environment Canada
Environmental Careers Organization
Energy Information Administration (U.S.)
Great Lakes Basin Ecosystem Team (USFWS)
Great Lakes Commission
Great Lakes Coastal Wetlands Consortium
Great Lakes Fishery Commission
Great Lakes National Program Office (USEPA)
International Joint Commission
International Union for the Conservation of Nature
Michigan Department of Environmental Quality
Michigan Department of Natural Resources
National Health & Environmental Effects Research Laboratory (USEPA)
National Oceanic and Atmospheric Administration
Natural Resources Canada
Natural Resources Conservation Service (USDA)
New York State Department of Environmental Conservation
Ohio Department of Natural Resources
Ohio Division of Wildlife
Ontario Farm Environmental Coalition
Ontario Ministry of Agriculture and Food
Ontario Ministry of Environment
Ontario Ministry of Natural Resources
Ontario Soil and Crop Improvement Association
Oak Ridge Institute for Science and Education
Pennsylvania Department of Environmental Protection
U.S. Department of Agriculture
U.S. Environmental Protection Agency
U.S. Food and Drug Administration
U.S. Fish and Wildlife Service
U.S. Forest Service
U.S. Geological Survey
World Business Council for Sustainable Development
Wisconsin Department of Natural Resources
Waste Diversion Organization (Ontario)
Wisconsin Department of Public Health
femptogram, 10~15 gram
hectare, 10,000 square metres, 2.47 acres
kilogram, 1000 grams, 2.2 pounds
kilometre, 0.62 miles
kiloton
kilowat-hour
metre

-------
mg             milligram, 10~3 gram
mg/kg          milligram per kilogram, part per million
mg/1            milligram per litre
ml              milliliter, 10~3 litre
MWh           megawatt-hour
ng              nanogram, 10~9 gram
ng/g            nanogram per gram, part per billion
pg              picogram, 10~12gram
ppb             part per billion
ppm            part per million
ton             English ton, 2000 Ib
tonne           metric tonne:  1000 kg, 2200 Ib
|o,g              microgram, 10"6 gram
ug/g            microgram per gram, part per million
(o,g/m           microgram per cubic metre
|o,m             micrometer, micron, 10"6 metre

Chemicals
2,4-D           2,4-dichlorophenoxyacetic acid
2,4,5-T         2,4,5-trichlorophenoxyacetic acid
BaP             Benzo[a]pyrene
BFR            Brominated flame retardants
CO             Carbon monoxide
DDT           1,1,1 -trichloro-2,2-bis(p-chlorophenyl)ethane or dichlorodiphenyl-trichloroethane
ODD           l,l-dichloro-2,2-bis(p-chlorophenyl) ethane
DDE           l,l-dichloro-2,2-bis(chlorophenyl) ethylene or dichlorodiphenyl-dichloroethene
DOC           Dissolved organic carbon
HBCD          Hexabromocyclododecane
HCB           Hexachlorobenzene
a-HCH         Hexachlorocyclohexane
y-HCH         Lindane
HE             Heptachlor epoxide
MeHg          Methylmercury
NAPH          Naphthalene
NO2            Nitrogen dioxide
NOX            Nitrogen oxides
NTU           Nephelometric turbidity unit
PAH           Polynuclear aromatic hydrocarbons
PBDE          Polybrominated diphenyl ether
PCA           Polychlorinated alkanes
PCB            Polychlorinated biphenyls
PCDD          Polychlorinated dibenzo-p-dioxin
PCDF          Polychlorinated dibenzo furan
PCN           Polychlorinated naphthalenes
PFOA          Perfluorooctanoic acid
PFOS           Perfluoroctanyl sulfonate
PM10           Atmospheric particulate matter of diameter 10 microns or smaller
PM2 5           Atmospheric particulate matter of diameter 2.5 microns or smaller
SO2             Sulfur dioxide
SPCB           Suite of PCB congeners that include most of PCB mass in the environment
TCDD          Tetrachlorodibenzo-p-dioxin
TCE            Trichloroethylene
TDS            Total dissolved solids
TOC           Total organic  carbon
TRS            Total reduced sulfur
VOC           Volatile organic compound

-------
Other
AAQC         Ambient Air Quality Criterion (Ontario)
AFO           Animal Feeding Operation
AOC           Area of Concern
APF           Agricultural Policy Framework (Canada)
ARET         Accelerated Reduction/Elimination of Toxics program (Canada)
BEACH        Beaches Environmental Assessment and Coastal Health (U.S. Act of 2000)
BKD           Bacterial Kidney Disease
BMP           Best Management Practices
BOB           Ballast On Board
BOD           Biochemical Oxygen Demand
CAFO         Concentrated Animal Feeding Operations
C-CAP         Coastal Change and Analysis Program (land cover)
CC/WQR      Consumer Confidence/Water Quality Report (drinking water)
CPU           Colony Forming Units
CHT           Contaminants in Human Tissue program (part of EAGLE)
CMA          Census Metropolitan Area
CNMP         Comprehensive Nutrient Management Plan (U.S.)
CSO           Combined Sewer Overflow
CUE           Catch per Unit of Effort
CWS           Canada-wide Standard (air quality)
DWS           Drinking Water System (Canada)
EAGLE        Effects on Aboriginals of the Great Lakes program
DWSP         Drinking Water Surveillance Program (Canada)
EAPI           External Anomaly Prevalence Index
EFP           Environmental Farm Plan (Ontario)
EMS           Early Mortality Syndrome
FCO           Fish Community Objectives
FIA           Forest Inventory and Analysis (USDA Forest Service)
FQI           Floristic Quality Index
GAP           Gap Analysis Program (land cover assessment)
GIS           Geographic Information System
GLWQA       Great Lakes Water Quality Agreement
HUC           Hydrologic Unit Code
IACI           International Alvar Conservation Initiative
IADN         Integrated Atmospheric Deposition Network
IBI            Index of Biotic Integrity
IGLD          International Great Lakes Datum (water level)
IMAC         Interim Maximum Acceptable Concentration
IPM           Integrated Pest Management
ISA           Impervious Surface Area
LaMP         Lakewide Management Plan
LEL           Lowest Effect Level
MAC           Maximum Acceptable Concentration
MACT         Maximum Available Control Technology
MCL           Maximum Contaminant Level
MGD          Million Gallons per Day (3785.4 m3 per day)
MMP          Marsh Monitoring Program
MSA           Metropolitan Statistical Area
MSWG        Municipal Solid Waste Generation
NAFTA        North America Free Trade Agreement
NATTS        National Air Toxics Trend Site (U. S. network)
NEI           National Emissions Inventory (U.S.)
NHANES      National Health and Nutrition Examination Survey (CDC)
NIS           Nonindigenous species

-------
NLCD         National Land Cover Data
NMP          Nutrient Management Plan (Ontario)
NOAEC        No Observable Adverse Effect Concentrations
NOAEL        No Observable Adverse Effect Level
NOBOB        No Ballast On Board
NPDES        National Pollution Discharge Elimination System (U.S.)
NPRI          National Pollutant Release Inventory (Canada)
NRVIS         Natural Resources and Values Information System (OMNR)
ODWQS       Ontario Drinking Water Quality Standard
OPEP          Ontario Pesticides Education Program
PEL           Probable Effect Level
PBT           Persistent  Bioaccumulative Toxic (chemical)
PNP           Permit Nutrient Plans (U.S.)
PGMN         Provincial Groundwater-Monitoring Network (Ontario)
RAP           Remedial  Action Plan
SDWIS        Safe Drinking Water Information System (U.S.)
SOLEC        State of the Lakes Ecosystem Conference
SOLRIS        Southern Ontario Land Resource Information System
SQI           Sediment  Quality Index
SSO           Sanitary Sewer Overflow
SWMRS       Seasonal Water Monitoring and Reporting System (Canada)
TCR           Total Coliform Rule
TDI           Tolerable  Daily Intake
TEQ           Toxic Equivalent
TIGER         Topological Integrated Geographic Encoding and Reference (U.S. Census Bureau)
TRI           Toxics Release Inventory (U.S.)
UNECE        United Nations Economic Commission for Europe
WIC           Women Infant and Child (Wisconsin health clinics)
WISCLAND    Wisconsin Initiative for Statewide Cooperation on Landscape Analysis and Data
WTP          Water Treatment Plant (U.S.)

-------


7.0 Acknowledgments

The State of the Great Lakes 2007 preparation team included:

Environment Canada                         United States Environmental Protection Agency
Nancy Stadler-Salt, lead                      Paul Bertram, lead
Stacey Cherwaty-Pergentile                   Jackie Adams
Katherine Balpataky                          Karen Rodriguez
Tracie Greenberg                            Elizabeth Hinchey Malloy
Leif Matiland                                Paul Horvatin
                                            Chiara Zuccarino-Crowe
                                            Jeffrey May

This report contains contributions from dozens of authors and contributors to the indicator reports
and the Lake and River assessments, and their work is sincerely appreciated. Their voluntary time
and effort to collect, assess and report on conditions of the Great Lakes ecosystem components
reflects their dedication and professional cooperation. Individual authors and contributors are
recognized at the end of their respective report component.

Many governmental and non-governmental sectors were represented by authors and contributors.
We recognize the participation of the following organizations. While we have tried to be
thorough, any misrepresentation of oversight is entirely unintentional, and we sincerely regret any
omissions.

Federal
Department of Fisheries and Oceans Canada
Environment Canada
       Air Quality Research Branch
       Canadian Wildlife Service
       Centre St. Laurent
       Climate and Atmospheric Research Directorate
Environmental Conservation Branch
       Environmental Protection Branch
Integrated Programs Division
Toxic Prevention Division
       Meteorological Service of Canada
       National Indicators and Assessment Office
       National Water Research Institute
       Ontario Region
               Great Lakes Environmental Office
               Regional  Science Advisor's Office
       Quebec Region - Environmental Conservation Branch
Industry Canada
National  Oceanic and Atmospheric Administration
       Great Lakes Environmental Research Laboratory
       Illinois/Indiana Sea Grant
National  Park Service
                         Draft for Discussion at SOLEC 2006

-------
Natural Resources Canada
U.S. Department of Agriculture
       Forest Service
       Natural Resource Conservation Service
U.S. Department of Health and Human Services
       Agency for Toxic Substance and Disease Registry
U.S. Environmental Protection Agency
       Great Lakes National Program Office
       Office of Research and Development
       Region 2
       Region 5
U.S. Fish and Wildlife Service
       Alpena Fishery Resources Office
       Ashland Fishery Resources Office
       East Lansing Ecological Services Office
       Green Bay Ecological Services Office
       Green Bay Fishery Resources Office
       Lower Great Lakes Fishery Resources Office
       Marquette Biological Station
       Reynoldsburg Ohio Ecological Services Office
U.S. Geological Survey
       Biological Resources Division
       Great Lakes Science Center
              Lake Erie Biological Station
              Lake Ontario Biological  Station
              Lake Superior Biological Station
       Water Resources Discipline

Provincial and State
Illinois Department of Natural Resources
Illinois Environmental Protection Agency
Indiana Department of Natural Resources
Indiana Geological Survey
Michigan Department of Environmental Quality
Michigan Department of Natural Resources
Minnesota Department of Health
Minnesota Department of Natural Resources
Minnesota Pollution Control Agency
New York State Department of Environmental Conservation
Ohio Department of Natural Resources
Ohio Division of Wildlife
Ontario Ministry of Agriculture and Food
Ontario Ministry of Environment
       Environmental Monitoring and Reporting Branch
       Standards Development Branch
Ontario Ministry of Natural Resources
                         Draft for Discussion at SOLEC 2006

-------


Pennsylvania Department of Environmental Protection
Quebec
       Direction des ecosystems aquatiques
Ministere de la Securite publique du Quebec
Wisconsin Department of Health and Family Services
Division of Public Health
Wisconsin Department of Natural Resources Division of Wildlife

Regional and Municipal
City of Chicago
City of St. Catherines
City of Toronto
Grand River Conservation Authority
Northeast-Midwest Institute

Aboriginal
Bad River Band of Lake Superior Tribe of Chippewa Indians
Chippewa Ottawa Resource Authority
Haudenosaunee Environmental Task Force
Mohawk Council of Akwesasne

Academic
Brock University, ON
Cornell University, NY
Clemson University, SC
Grand Valley State University, MI
James Madison University, VA
Michigan State University, MI
Michigan Technical University, MI
Northern Michigan University, MI
University of Michigan, MI
University of Minnesota - Duluth, MN
University of Minnesota - St.  Paul, MN
University of Windsor, ON

Coalitions
Binational Collaborative for the Conservation of Great Lakes Islands
Great Lakes Coastal Wetlands Consortium
Great Lakes Environmental Indicators

Commissions
Great Lakes Commission
Great Lakes Fishery Commission
Great Lakes Indian Fish and Wildlife Commission
International Joint Commission

Environmental Non-Government Organizations
                        Draft for Discussion at SOLEC 2006

-------
Bird Studies Canada
Great Lakes Forest Alliance
Great Lakes United
The Nature Conservancy

Industry
American Forests and Paper Association
Council of Great Lakes Industries
National Council for Air and Stream Improvement, Inc.

Private Organizations
Bio-Software Environmental Data
Bobolink Enterprises
DynCorp, A CSC Company
Environmental Careers Organization
LURA Consulting
Oak Ridge Institute for Science and Education
Stream Benders

Private Citizens
                        Draft for Discussion at SOLEC 2006

-------