>EPA
  !& §
  ill
  (A C 25
    S
           United States
           Environmental Protection
           Agency
            Environmental Research
            Laboratory
            Gulf Breeze, FL 32561
EPA/600/9-85/018
June 1985
           Research and Development
Proceedings of the
Workshop:
Biodegradation
Kinetics, Navarre
Beach, Florida,
18-20 October 1983
  o - o
  « o B

-------
                                                                       EPA/600/9-85/018
                                                                       June 1985
J)                                PROCEEDINGS OF  THE  WORKSHOP:
(irj                                   BIODEGRADATION KINETICS
L                                   NAVARRE BEACH,  FLORIDA
'^]                                     18-20 OCTOBER  1983
M
 0
\o
 VJ
^                                  Edited and Coordinated by
                               Al W. Bourquin and  P.  H.  Pritchard
                              U.S. Environmental Protection  Agency
                                Environmental  Research Laboratory
                                   Gulf Breeze,  Florida  32561
                                        William W. Walker
                                 Gulf Coast Research  Laboratory
                                        East Beach  Drive
                                Ocean  Springs,  Mississippi  39564
                                           Rod Parrish
                                       8330 Wild Lake Road
                                    Pensacola, Florida  32506
                                    U.S. Environmental Protection Agency
                                    Region 5, Library (PL-12J)
                                    77 West Jackson Bou!evar.d, 12th Floor
                                    Chicago, IL  60604-3590
                                ENVIRONMENTAL RESEARCH LABORATORY
                               OFFICE OF RESEARCH AND  DEVELOPMENT
                              U.S. ENVIRONMENTAL PROTECTION AGENCY
                                   GULF  BREEZE,  FLORIDA  32561

-------
                                DISCLAIMER

     The information in this document has been funded wholly or in part by
the U.S. Environmental  Protection Agency.  It has been subject to the
Agency's peer and administrative review and approved for publication as
an EPA document.  Mention of trade names or commercial products does not
constitute endorsement  or recommendation for use.
                                       11

-------
                                   FOREWORD

     The proliferation of  organic  chemicals,  as a result  of  the  need for new
and better agricultural and  industrial  chemicals,  increases  the potential for
environmental pollution,  particularly in  aquatic  habitats.    Inherent  in any
fate study  is the need  for  an accurate  assessment  of the potential  for the
microbial degradation of the pollutant.   However,  there  is little information
on  the  rates  at  which  biodegradation  will   occur  in  the  environment.
Furthermore, kinetic data from laboratory and field studies are often based on
assumptions  that  have  no consensus  or  sound  scientific basis.    Thus, the
microbiologist  currently  has  no  consistent guidelines  regarding  appropriate
techniques   for   the  determination   of  biodegradation  rates   in  natural
environments,  nor have  microbiologists,  in  general,  provided the  chemical
industry  and   regulatory  agencies   with  evaluations   of   the   effect  of
environmental  parameters  and  chemical  structures  on   biodegradation  of
processes mediated by microbial populations.

     This Workshop  on Biodegradation  Kinetics  is a  follow-up to  a workshop
held in  Gulf  Breeze,  Florida, and  sponsored  by  the Environmental  Research
Laboratory at Gulf Breeze.   As was  true for the first workshop, this workshop
on  biodegradation  kinetics   was  organized  to  offer  representatives  from
government,  academia, and  industry  a forum for the  examination of key issues
regarding the future  direction  and  focus of scientific investigations in this
field.    Participants addressed a number of questions of primary concern:  What
are the  differences and  similarities  between  biodegradation  rates in aquatic,
terrestrial,  and  laboratory  environments?  What  methodological  criteria must
be  established  to provide interchangeable degradation information?   What is
the potential  for a  particular environment or  its  laboratory  simulation to
dispose  of a polluting chemical, and at what rate will it occur?  We hope that
this  publication   of  the  workshop  proceedings  will  provide an  up-to-date
reference  for  professionals   concerned  with  the   fate,   regulation,  and
production of potential  environmental pollutants.
                                                 _   ^nos
                                            Director
                                            Environmental Research Laboratory
                                            Gulf Breeze, Florida
                                      Til

-------
                                   ABSTRACT

     This workshop,  held  October  18-20,  1983,  at Pensacola  Beach,  Florida,
focused  on  pertinent issues  related  to the  scientific investigation  of the
microbial  degradation   rates  of organic chemicals  in  natural  environments.
Participants  discussed  methodological  criteria  for these  investigations and
the need for concentrating on the kinetic aspects of biodegradation.  Position
papers dealing with the following topics were presented in open sessions:  (1)
Statistical and  Experimental  Requirements  for Modeling  Decay  Curves; (2) The
"Second  Order"  Approach  Assumption,  Limitations and Research   Needs;  (3)
Factors  Controlling  Biodegradation   Rates   in   Microbial   Communities;  (4)
Application of  Uptake  and Mineralization  Kinetics;  (5) Relationships between
Chemical  Structure  and   Biodegradation  Rates;   and   (6)   Extrapolation  of
Laboratory Biodegradation  Data  to  the Field.   Discussions  within each session
are summarized by the panel members in  reports that include a consensus of the
direction   and  extent   of   research   required   for  the   description  of
biodegradation rates of xenobiotic chemicals in  natural  environments.    These
proceedings conclude with a summary report and suggestions for future research
in  biodegradation   kinetics.    This  report   is  submitted  in  fulfillment  of
Contract No.  CR  810789  by Gulf  Coast  Research  Laboratory  in conjunction with
Georgia  State  University Cooperative Agreement  R809370 under  the  sponsorship
of the  U.S. Environmental Protection Agency.  This report covers a  period from
April 4, 1984 to May 13, 1984, and work was completed  as of June 30,  1984.
                                       i v

-------
                                   CONTENTS

                                                                      Page

Foreword         	i i i
Abstract         	iv
Acknowledgment   	vi

Workshop Organization	,	1
    A.   Objectives	1
    B.   Working Concept	1
    C.   Workshop Structure	1
    D.   EPA Perspective - By R. Brink,  U.S. EPA	2
    E.   Ad-Hoc Advisory Committee	5

Statistical and Experimental  Requirements for Modeling  Decay  Curves	6

The Second Order Rate Approach:  Assumptions, Limitations,  and
     Research Needs	38

Factors Controlling Biodegradation Rates in Microbial  Communities	72

Application and Uptake of Mineralization Kinetics	83

Relationships between Chemical Structure and Biodegradation Rates	117

Extrapolation of Laboratory Biodegradation Data to  the  Field	127

Summary and Research Plan for Biodegradation Studies	146

Appendix - Proposed Benchmark Chemical for Biodegradation Research....161

-------
                                ACKNOWLEDGMENT

     The assistance  of Ms. Gail  Seidler,  Gulf Breeze  Environmental  Research
Laboratory, and  the  Georgia State University  Cooperative  Agreement  Office at
Gulf  Breeze,  in  preparing  material   before the   Workshop  is  gratefully
acknowledged.
                                       VI

-------
I.   Workshop Organization

A.   Objectives

     1.  To evaluate the  utility  and  limitations  of approaches currently used
         for the derivation of environmentally meaningful biodegradation rates
         from laboratory studies.

     2.  To identify the  future  research required  for  the  improvement  of our
         approaches in the development  of a  strategy for the extrapolation of
         laboratory information  on  the  kinetics  of biodegradation to complex
         environmental  situations.

B.   Working Concept

     Biodegradation studies,  in  general, provide decay  curves  for a chemical
by using a particular method  and  inoculum.   Decay curves vary under different
inocula or different experimental conditions.  Application of this laboratory-
derived  biodegradation  information  to  the  field  requires  that  degradation
rates  be  calculated from the  decay  curves.   This  is  usually accomplished by
developing mathematical  rate  equations or kinetic  expressions  that  model  the
decay  curves.   Because  of  uncertainties about  the factors  that  control  the
metabolic  diversity  and   degradative  potential  of  natural  communities  of
microorganisms,  certain  assumptions  about   the  activity,  concentration,  and
substrate  affinity  of  the biological  catalysts  within  these communities  and
about  the  effect  of environmental  parameters (sorption, temperature, multiple
substrates, etc.) on their activity must be  made if useful kinetic expressions
are to be developed and the key,  we believe,  to measuring biodegradation rates
in  the  laboratory  that  can   be  used  to  predict  the  biological  fate of  a
chemical in the field.   A critical evaluation of the logic associated with the
development  of  each  assumption  and  the  research  needed  to  validate  the
assumptions was the working theme for this workshop.

C.   Workshop Structure

     The workshop was  divided  into six sessions:

     1.  Statistical and Experimental  Requirements  for Modeling Decay Curves
     2.  The "Second Order" Approach:   Assumptions, Limitations,  and Research
         Needs
     3.  Factors Controlling Biodegradation  Rates in Microbial Communities
     4.  Application of Uptake and Mineralization Kinetics
     5.  Relationships Between Chemical Structure and Biodegradation Rates
     6.  Extrapolation of Laboratory Biodegradation Data to the Field

     Each  session  was  initiated with  several   10-minute talks  that  briefly
summarized  the  status,   limitations,   and   research   needs   of  a  particular
research approach used to study biodegradation kinetics.  The talks, presented
below,  were  intended  to  provide  representative data  upon  which subsequent
discussion was based.   Position statements and suggestions were encouraged.

     The objective of each session was  to provide the session chairperson and
a  panel  of  3-5  individuals  with  interpretations,  strategies,  theories,
recommendations, and priorities that they could use to prepare a report to the

-------
workshop  coordinator.    Session  reports  were  presented  to  the  workshop
participants  for  further  discussion  and  revision  in  a  general  plenary
session.  The final  panel  reports are presented below.

D.   An EPA Perspective by Robert Brink, U.S. EPA

     My  comments will  be  directed  to  the  needs  of  the  Office of  Toxic
Substances (OTS) of EPA, although I am sure that any advances that can be made
in the ability to estimate environmentally realistic biodegradation  rates will
be of value to many others in EPA and elsewhere.

     The  Toxic  Substances   Control  Act  (TSCA),  under  which  OTS  operates,
applies to  all  chemical substances  not  covered by prior  acts  and allows for
the  regulation  of   those  chemical  substances  that  may  or  do   present  an
unreasonable risk of injury to human health or the environment.

     An evaluation  of  the  risk  or  likelihood that  a  chemical may cause an
adverse effect  involves a determination  of  the hazard or adverse effects that
may  occur   if  there  is   exposure  and  an   assessment  of  the  exposure
possibilities.

     An exposure assessment  considers  the sources,  pathways, concentrations
and  potentially  exposed populations.   Chemical  fate aspects  of the exposure
assessment evaluate  the transport and  transformation  possibilities following
release  of  the chemical  to the  environment  and the  fate  conclusions  are
important  elements  of  the  pathways  and  concentrations  part  of  the exposure
assessment.

     An assessment of  the fate of  a  chemical  released to  the environment will
depend,  in  part,  upon  laboratory  data that  can  be  used  to  evaluate  the
properties  and  processes that  influence  transport  and  transformations.   In
actual  practice, fate  assessment  includes  a  consideration  of  existing data
from  reliable  literature citations  and estimation techniques  that should be
considered  and  that  may  be  used  either to  avoid unnecessary  testing  or to
guide  the proper selection  of  laboratory  testing.   The  overall  approach is
summarized  in  Table  1  which lists,  from top  to  bottom,  the  properties and
processes  for  which  data may be  needed and,  from left  to right, the various
levels  of data  input.   In this  scheme, laboratory data  at  one level  will be
required  only when data from a lower level are  insufficient to  answer relevant
exposure  assessment   questions.    The  testing  may begin  at  any  appropriate
spot.     For  example,  if  it is  already  known  that  the  only   significant
transformation processes will be biodegradation  studies  in  soil  and  water, the
laboratory  testing  might begin  with biodegradation  studies that will provide
kinetic  information  and data on potential  intermediates of  concern.   It is
also highly  unlikely  that any given  chemical  would require testing  for  all or
even  a high percentage of  the  properties  or  processes.   Existing knowledge
estimation  techniques and the early base-set  tests  will  provide  information
that will preclude much of the other testing.

     Environmentally  realistic  estimations   of   chemical  concentrations  of
nonpersistent  substances require valid kinetic  data with  calculations of half-
lives  under  relevant environmental  scenarios.   For  hydrolysis, the data  from
the  base-set will usually be sufficient  to calculate environmental  half-lives,
although  there  may  be  circumstances,   such  as environmental  extremes  of pH,

-------
that  will  dictate  a  need for  upper tier  simulations  to obtain  better rate
data.    For  selected  environments,  it  may be  necessary to  conduct  kinetic
studies  at  an  upper  tier,   with   simulations   of  particular  environmental
variables,  to  obtain  better  rate   values for   leaching,  evaporation,  and
photodegradation.   It  is  unlikely that  the base-set  biodegradation tests will
provide  sufficient  information  for  valid extrapolation  to environmental  rates
and  it  is  anticipated  that  any  determination  of  environmentally  relevant
biodegradation  rate  data  will  require upper tier  kinetic studies.  It should
be emphasized that upper tier tests for rate determinations are justified only
for  significant  transfer  transformation  processes.   For the organic compounds
evaluated  by  EPA, transformations  in  soil  and  water   by  micro-organisms  are
often the most significant transformation processes.

     The  base set  of fate  tests  will  indicate  the dominant transformation
processes that a chemical  may undergo in natural  environments.  In some cases,
there  will  be  indications   of whether  an  organic   substance   is  degraded
essentially to inorganic compounds (e.g. carbon dioxide and water) plus normal
metabolic intermediates.  If a substance is extremely persistent,  this will be
evident from the lack of reaction by any of the transformation processes.  For
all  of  the  situations  where  a  chemical  substance is  observed  neither  to
persist  nor  to clearly degrade  to  innocuous  inorganics  and  metabolites,  the
investigator   should   be  concerned  with   the  question   of  persistent  and
potentially toxic  transformation  products.   In some cases,  a  good background
in chemistry  and microbiology will  enable  the investigator to  make  informed
guesses about  the  potential  products,  but  testing will  be required to confirm
those  guesses.    The  purposes  of  the  degradation   pathways   testing  are
twofold.   One  question  to  be  answered  is whether a  toxic  substance,  when
degraded,  shows  concomitant   loss  of  that toxicity.    This  is   a  combined
transformation  and  effects   problem.    The  other  question,  which  involves
analytical  chemistry,  deals  with   the  identification  of  the   products  of
environmental transformation.

     It is very  important to  us,  in  OTS, to be able  to do better  in assessing
the  biodegradation  rates  of  organic  chemicals.   This must be accomplished by
the  use of  a  series  of  site-specific microcosms  that   have been  shown  to
replicate,  to  some   necessary  precision,  the  transformation  rates  in  the
simulated  environments.    I  hope  that  won't  be  necessary.    It would  be
preferable to  be  able  to use  some relatively  simple,  generic,  bench-scale
laboratory procedures  that  would provide  biodegradation   rate constants that
could  be translated  into  rates for  specific  environments  by  mathematical
manipulations  that  account  for  the  relevant environmental  factors that will
influence  the   rates.    We   hope   that   this  workshop   will  consider  the
possibilities and provide recommendations on the  best  research to pursue over
the next few years to achieve that goal.

-------
               TABLF I
CHEMICAL FATE TESTING SCHEME
                         Data  Source
Property or Process
                                           to
                                           .P
                                           (0
                                           D
               Q) -p
               ,H W

               ra QJ
               -p -p
               •H -H
               U i4
Physical/Chemical  Properties

Absorption, uv/visible spectra
Boiling Temperature
Density/Relative Density
Dissociation  Constants in Water
Henry's Law Constant
Melting Temperature
Particle  Size Distribution/Fiber Lngth.
Partition Coefficient (n-Octanol/Water)
pH of Water Soln or  Suspension
Water Solubility
Vapor Pressure


Transport Processes

Evaporation from Water
Evaporation from Soil
Adsorption/Desorption
Uptake by Biota


Transformation

Biodegradation
Photodegdn.-Sunlight Photolysis
            -Oxidn  by OH radical or 0-,
            -Indirect
Complex Formation
Hydrolysis
S-l
o

w
c
o
•H 03
-P Q)
(0 -P

^ 6
O -H
<-\ 4J
fO 03
U W
4J
0)
W
I  cn
Q> -P
03 U3
(ti (U
CQ H
                                                      -y
                                                      «

                                                      o —
                                                      •r-t U3
                                                      •<-> Q)
                                                      Q> 4-1
                                                      C fi
                                                      •H «
                                                          <1) (0
                                                          Q (X
C o^
O £
•H \n
4J O
(0 U
rH O
3 >-i
e o
•H -rH
W 2

0) .C
-p -p
•H -H
   Inappropriate and unnecessary data  elements are indicated
   by  X marks within the boxes.


 * Methods in the August, 1982 et  seq.  OTS Guidelines,
   published by NTIS.

-------
E.   AD-HOC ADVISORY COMMITTEE
Al  W. Bourquin, Co-chairman
U.S. Environmental Protection Agency
Environmental Research Laboratory
Sabine Island
Gulf Breeze, FL  32561
(904) 932-5311

Hap Pritchard, Co-chairman
U.S. Environmental Protection Agency
Environmental Research Laboratory
Sabine Island
Gulf Breeze, FL  32561
(904) 932-5311

Donald Ahearn
Goergia State University
Research Office
131 Sparks Hall
Atlanta, GA  30303
(404) 658-4350

Martin Alexander
Department of Agronomy
Cornell University
Ithaca, NY   14853
(607) 256-3267

Robert Boethling
U.S. Environmental Protection Agency
Office of Pesticides and Toxic
   Substances (TS-798)
Washington,  DC  20460
(202) 382-3913

Peter Chapman
Department of Biochemistry
University of Minnesota
St. Paul, MN  55108
(612) 373-1303
Robert Hodson
Department of Microbiology
University of Georgia
Athens, GA  30602
(404) 542-1434
Harvey Holm
U.S. Environmental Protection Agency
Environmental Research  Laboratory
College Station Road
Athens, GA  30601
(404) 250-3103

Carol Litchfield
Haskell Laboratories
DuPont Corporation
Elkton Road
Newark, DE  19711
(302) 366-5486

Jim Spain
Georgia State University
Cooperative Agreement Office
Environmental Research  Laboratory
Sabine Island
Gulf Breeze, FL  32561
(904) 932-5311

William Walker
Gulf Coast Research  Laboratory
Coast Beach Road
Ocean Springs, MS  39564
(601) 875-2244

-------
     STATISTICAL AND EXPERIMENTAL REQUIREMENTS  FOR MODELING DECAY  CURVES

     Objective:   To   discuss   the    limitations    and    possibilities   for
                 statistically   verifying    that    a   kinetic    expression
                 appropriately describes a decay curve or set  of decay curves.

     Biodegradation rates  are  commonly determined from  decay  curves assessed
by using certain statistical  treatments of the data.   Improper or insufficient
use  of statistical  methods  and  incomplete  data  sets,  however,  frequently
results  in  equivocal  assessments  of  degradation  rates.   What  statistical
analysis  is  necessary  to  verify  that  a kinetic  expression  appropriately
describes a decay  curve?   What experimental  criteria  (replication, sampling,
data points, precision,  time course, etc.) should be established to assure the
completeness  of  data   sets   used   for statistical  analysis?    When  should
nonlinear  or  linear  regressions  be  applied,  and  what  advantages do  they
offer?   At  what  point  does the  number of outlying  data  points  mean  that a
model  does  not  fit the  data?  What  logic  and discriminative  framework should
be used to select models that fit the experimental  data?


Panel Members:   Joe Robinson (chairperson), Montana State University
                 David Currie, McGill  University
                 Walter Maier, University of Minnesota

-------
                 PROBLEMS IN NON-LINEAR PARAMETER ESTIMATION

                                 David Currie

                                 Introduction


     In recent years the use of nonlinear models has become increasingly
common in biological studies.  The statistical  treatment of such models
presents several difficulties which are not encountered in ordinary linear
regression analysis.  As a result, uncautious treatment of nonlinear models
can often produce very poor parameter estimates and model  predictions.   It
is the purpose of this paper to describe three pitfalls in fitting nonlinear
models which are often observed in biological studies.  I  shall  concentrate
specifically on the problem of obtaining the best estimates of the model
parameters (i.e., constants).  A good, readable introduction to the basics of
regression analysis and some of its other potential problems is given by
Gujarati (1978) and more advanced treatments will be found in Draper and
Smith (1981) and Beck and Arnold (1977).

     In seeking the "best" estimates of model parameters,  I specifically mean
estimates which are both accurate (i.e., unbiased) and precise (i.e., having
low variance).  The precision of estimates almost entirely determines one's
ability to test hypothesis.  Imprecision undermines the ability to reject the
null hypothesis and increase the probability of retaining  the null hypothesis
when it is in fact false (Type II error).  More insidious  yet is the effect
of inaccuracy (i.e., bias) estimation.  Bias is the systematic tendency to
overestimate or underestimate a population mean value, and can cause differences
among hypothesis to appear to be statistically significant, when they are in
fact entirely spurious.  In the discussion which follows,  I shall consider
three common practices which tend to yield inaccurate and  imprecise parameter
estimates:  poor experimental design, inappropriate fitting techniques, and
inappropriate model choice.


                           POOR EXPERIMENTAL DESIGN

     The design matrix, that collection of values of the independent variable
(x) at which the experimenter decides to take observations, is rarely given
much explicit consideration before an experiment is carried out.  It can be
shown (Draper and Smith 1981) that, for any model, the variance of parameter
estimates will depend upon two things:  the magnitude of the experimental
error (which the experimenter often cannot control) and the design matrix
(which usually can be controlled).

     The effect of the design matrix is most easily illustrated with a  linear
relationship (Figure 1).  If one wishes to estimate the slope of a straight
line (the heavy line in the center of the figure) with a given amount of
experimental error around the line (depicted by the vertical bars), then
taking observations over a relatively narrow range of x values (interval A)

-------
will  lead  to a  much  broader confidence  interval  around the  estimate  of the
slope than will  a wider spread of observations (interval B).

     The  situation  is  consideraly  less  obvious  when  a model  is  nonlinear.
This is due to the fact that estimates of nonlinear parameters depend upon the
placement  of observations  with  respect  to  specific  points  on  the  curve.
Consider, for example, a simple exponential decay model:
where y  is usually the measured amount  of a decaying substance,  and x is an
independent  variable  such  as time  or  distance.   Suppose  that a,  the  initial
quantity of substance y, is known or fixed by the experimenter, and one wishes
to estimate the decay constant b.  To what extent does the choice of  values of
x influence the estimate of b?

     Figure  2  depicts  this  decay  curve  and  offers  several  conceivable
arrangements  of  values  of x  at  which  to measure  y.    The  variance  of the
estimate of b [i.e., var(b)] which results from each of these  designs is shown
in  Figure  2 as a  multiple of the  experimental  error o .   Var(b)  is seen to
change by a factor of three over the possible designs shown  here.  Intuitively
appealing  designs  that  spread the  observation  over a wide  range  of x values
yield estimates of  b  which are 50-100%  more  variable than the estimate found
by taking  all  of  the observations at x  =  1/b.   It  can  be shown that this i.s,
in fact, the design which minimizes the variance* of b.

     Each  observation  that  one  takes  contributes  a  specifiable  amount of
information  toward  estimating  b.    The  amount  of  information  conveyed  by  a
given observation  y,  depends  entirely upon  the  value of  x^  at  which y^ is
measured.    This  relationship  is   shown  in  Figure  3.    In  brief,  for the
exponential  function,  the  most information  is  contained  in an  observation
taken at x  =  1/b.   The  farther from x =  1/b an observation  is taken, the  less
information it contributes toward estimating b.

     In  general,  minimum variance estimates  of the  parameters  of any fairly
simple model  will  be obtained  by  concentrating all  of  the  observations at  a
number of  specific  points  equal  to  the number of parameters in  the model.  In
many cases  (such  as a  straight line), one or more of the optimum observation
points  will  be  at extremes  of the range  of  x.    The  exact  points  may be
determined  by optimization of the variance-covan'ance matrix (Draper  and Smith
1981 discuss  the  variance-covariance matrix,  and Box and Lucas  1959  deal  with
the  optimization  thereof),  but it  is  not  easy  going.  Alternatively,  one may
use  an  approximate  method  to find these  points as  outlined  in Appendix 1 and
discussed by Beck and Arnold (1977).

     The obvious  drawback  of  optimal  designs is that they provide no means of
checking  the  fit  of  one's  model   to the data.    In  practice,  it  is often
desirable   to  sacrifice  a  certain  amount  of  precision  for the   sake of
ascertaining that one's model is adequate.   It  is still possible in such cases
to  evaluate  the  effect of  the  design  matrix on  parameter  estimates.   In
general,  if observations  are  spread out  in the neighborhood of  the  optimal
design points, a minimum of precision is sacrificed.  For any  given model, one
may  generate  a  sensitivity  relationship  such  as  Figure   3  and   use   this
information  to  design one's  experiments (or  better still,  use the covariance
matrix approach).

-------
      Consider  a  second  example,  the  familiar  Michaelis-Menten  model   (or
alternatively, the formally equivalent Monod  function or Langmuir  isotherm):
                                        Vs
                                  v  = 	
                                       K +  s
where v  is the rate of  reaction,  s is the substrate  concentration,  and  V  and  K
are  fitted constants  (representing  a maximum  reaction velocity  and a  half-
saturation constant,  respectively).   The  most precise  parameter estimates  are
obtained  by  taking half  of  the observations at  s = K,  and the other  half at
s »K (i.e., as high as possible) (Currie 1982).

      Figure 4  (taken from  Currie  1982) appraises  the effect  of  using geometric
sequences  (4A  and 4B)  and linear sequences (4C and 40) of  seven  observations
to estimate K and V.  The  linear  sequences  of observations  s^ are  defined  by:

                     s1 =  i_a_                    for i =  1, 2,..., 7

(for  example,  when a = 0.5,  the  observations were  taken at x  values  of  0.5,
1.0,  1.5,  2.0, 2.5, ~3~.0,  3.5) and the geometric  sequences by:

                     s  = a_ 21"1                 for i =  1, 2,..., 7

(for  a_  = 0.5, the observations were taken  at  0.5,  1.0,  2.0,  4.0,  8.0,  16.0,
32.0).   In essence,  _a_  is  a  factor  which  serves to  telescope the  observations
such  that when   a_ is  small,  the  observations  are concentrated  nearer  the
origin, and when _a_ is large, the  observations are more  spread out.

     The  lower panels  of   Figure  4  (4B  and  4D)  show the variance  of estimates
of K  (i.e.,  K)  as a function of  the design matrix  as  characterized  by a_.  It
is immediately evident  that the  linear  sequences of observations  in FigUre 40
yield much higher variance estimates of K  than do the geometric  sequences in
Figure 4B.  Moreover,  among  each  of these  classes there are obvious minima in
the  variance  of   K  which  are obtained.   The variance  of  parameter  estimates
using the  best  geometric  sequence  of observations  is  approximately twice as
great as that obtained with the optimal design.   However, the geometric spread
of observations  does allow  one to  verify  the  fit  of the  model  while still
producing  reasonably  precise  parameter estimates.   The catch,  of  course, is
that  one must have  some  preliminary estimate  of   K  in  order  to   design  the
experiment (similarly,  for the exponential decay model,  it is  necessary to
have some preliminary estimate of b).  Given this information (usually  from an
exploratory experiment), one may then optimize the experimental  design.

     Not  only  is  the  precision  of  parameter estimates  affected  by   the
experimental  design, the  accuracy is affected  as well.   In Figure  4 panels A
and  C,  the ratio  of  the  estimated  half-saturation  constant (K)  to its  true
value (K)  is  shown.   This is  a measure of  bias:   a  ratio  of 1.0 indicates an
unbiased estimate  of K,  whereas  deviations  from 1.0   indicate  less and  less
accurate estimates.   It is evident   that some  sequences of  observations yield
badly biased  estimates of K.   The  geometric sequences of observations (4A) are
seen to yield estimates of K  which  are  also more accurate than those from the
linear sequences  (4C).

     The effect of experimental design  on bias  in parameter estimates  is more
difficult to  determine  than  its  effect  on  variance; in the  present case, it

-------
was determined  by  Monte Carlo simulations,  which  are fairly involved.   As  a
rule of thumb (which may not  invariably  be  correct),  it  appears that the more
precise parameter estimates also  tend to be  reasonably accurate.


                        INAPPROPRIATE  FITTING TECHNIQUE

     The  most  common   means   of   fitting   nonlinear  models  is  to  apply  a
transformation which linearizes the  relationship  and  then to fit  the data by
the   ordinary   least-squares  method.      Generally   speaking,   this   is
inappropriate.   Least-squares regression  involves the  assumption   that  the
error in the dependent  variable is normally  distributed and of equal magnitude
at  all  x.   When a  dependent  variable  is transformed,  the error  around that
variable is  also transformed  and,  depending upon the original  distribution,
the  final  error distribution  will  often  be something  very  far  from normal
which varies with x.

     Figure 5 shows the bias  and  variance of estimates  of K in  the Michealis-
Menten model  using the same  data as  in  Figure 4.   However,  in the  case of
Figure 4, the untransformed  data  were fitted with a  direct,  nonlinear least-
squares method (Bliss  and  James 1966; Hanson e^aj_. 1967), whereas in  Figure 5
the data were subjected to a  Woolf linearizing  transformation  first, and were
then fitted by ordinary linear least-squares.  A comparison of the two figures
reveals that  the Woolf transformation produces parameter  estimates which are
strongly biased  and which  have very much higher variance  than those found with
the untransformed  procedure   (Figure  4).    The  Woolf  transformation  has been
compared  to  other  linearizing   transformations  of   the  Michaelis-Menten
function, and it has been  shown to be the  best  of the possibilities (Dowd and
Riggs  1965;  Atkins  and Nimmo  1975).   As a  rule,  linearizing  transformations
will yield  high  biased,   high  variance parameter  estimates,   whereas direct
untransformed  fits  of  nonlinear  functions  will  be  much  more  reliable.
Nonlinear fitting  routines are now  available   in  virtually all of the  major
main-frame  statistical packages   (e.g.,  NLIN  in  SAS,   NONLINEAR   in  SPSS).
Linearizing transformations may be appropriate  in  cases where a  transformation
is  applied  in  order   to  normalize  the  error   distribution.    Obviously,  the
rationale here  is  that attention must  be  paid to  the  assumptions  of  one's
fitting technique.
                              INAPPROPRIATE  MODEL

     A third common  practice  which  leads to poor  parameter  estimates is that
of fitting an  inappropriate  function  for the sake  of  ease  of computation.   A
frequent example  is  the  case  in which a  straight  line is  fitted to something
called  "the   initial   linear   portion"   or  the  "first-order  portion"  of  a
nonlinear function.   In  most  cases, theoretical  considerations or examination
of the  residuals  around the  fitted  function   (e.g.,  Draper  and  Smith  1981)
should  be  sufficient to  show  that  there  is no initial   linear  portion.
Application  of  a  linear  model  to  an  "approximately  linear"  curve  will
invariably lead to biased parameter estimates.

     As an example,  consider^the  data shown in  Figure 6,  which  show the time
course  of  incorporation  of    P  orthophosphate  in  a  bacterial  culture.   The
goal   of  this   experiment   was   to  estimate  the  rate   of  orthophosphate


                                      10

-------
incorporation,  which  is known to  decline as orthophosphate  is  depleted from
solution.   A  very common  procedure  would  be  to fit  a  straight  line  to  an
initial  "linear"  portion of  the  curve and  to  take  its slope as  the rate of
phosphorus  incorporation.   On cursory  examination of the data, the assumption
of  initial  linearity  does  not  seem  unreasonable.    However, it  is unclear
exactly what portion of the curve should be  considered as linear.  The numbers
in  the  upper  left portion of  the  figure  represent estimates of the  slopes of
straight  lines  fitted  to  progressively  shorter  initial   segments of  the
curve.    It  can  be  seen  that as  the  latest  observations  are  sequentially
eliminated  from  consideration,  the estimated  slope  of the  "initial  linear
portion"  steadily  increases.   This  is  due to the fact that indeed there is no
linear  portion   of   the  curve.     As   a   result,   the  estimated  rate  of
orthophosphate  uptake   is  systematically  underestimated,  to  an  extent  which
depends upon the  length of the period  of observation.

     It  is  evident  that  the  initial  rate of  uptake  is  the  biologically
interesting  parameter  in  this  example.    The  solution  to  this  fitting
difficulty  is  to choose  a  function   which eliminates  the  pattern in  the
residuals,  and  which  is  therefore  insensitive  to  the  duration  of  the
experiment.   A  second  degree polynomial was found to fulfill these  criteria,
and yielded the  rate estimate in  the  lower  right portion of the figure.  This
rate  represents  an unbiased  estimate  of  the initial  rate  of orthophosphate
uptake.

     It is  worth  mentioning  in  passing that the straight line fits these data
very well,  as  judged  by the  correlation  coefficient  or an F-test.   Moreover,
the second  degree term of the polynomial  fit was never significant  according
to a partial-F test.   The second degree term was nonetheless retained in order
to satisfy the assumption (of any regression method) that the  error around the
function  is  unbiased   (i.e.,  that the  expected  value of  the  residuals  is
zero).   Failure  to  satisfy   this  assumption is evidence  of  an inappropriate
model, which causes parameter estimates to be biased.
                                   SUMMARY

     Fitting nonlinear  models  to experimental  data  poses a  number  of unique
statistical  problems   in  addition  to  those  encountered  in  ordinary linear
regression.  A  large  number of biological  studies  involving nonlinear models
make fundamental  statistical  errors  which  cause the  parameter  estimates and
model  predictions  to  be  biased and  highly variable.   Many  of  these errors
result from  neglect of the basic assumptions of  the  fitting techniques used.
Three  especially  common  inappropriate  procedures  are:   (1) neglect  of the
experimental design matrix  (that set  of values  of the independent variable at
which  the   experimenter   determines   to measure  the  dependent  variable);
(2) fitting  data  which have been  transformed to remove  nonlinearity without
consideration of  the  effect on  the  error  distribution;  and  (3)  fitting the
wrong function  for the sake of ease  of computation  (e.g.,  fitting a straight
line to the  "initial   linear portion"  of a nonlinear function).  Statistically
appropriate alternatives to these practices  are  readily available.
                                      11

-------
                               LITERATURE CITED

1.  Atkins, G.  I. and  I.  A.  Nimmo.   1975.  A  comparison  of  seven  methods for
      fitting the Michaelis-Menten equation.   Biochemical  Journal  149:775-777.

2.  Beck, J. V. and K. J. Arnold.   1977.   Parameter Estimation in  Engineering
      and Science.  New York:   Wiley.   501 pp.

3.  Bliss, C.  I.  and  A.  T.  James.  1966.   Fitting  the  rectangular hyperbola.
      Biometrics 22:573-602.

4.  Box, G. E.  P. and  H.  L.  Lucas.   1959.  Design of experiments in nonlinear
      situations.  Biometrika  46:77-90.

5.  Currie,  D.  J.   1982.    Estimating  Michaelis-Menten  parameters:    Bias,
      variance and experimental  design.  Biometrics  38:907-919.

6.  Dowd,  J.   E.  and  D.  S.  Riggs.   1965.    A  comparison  of estimates  of
      Michaelis-Menten    kinetic    constants     from     various     linear
      transformations.   Journal  of Biological  Chemistry 240:863-869.

7.  Draper,  N.  and  H. Smith.    1981.    Applied  Regression  Analysis  (Second
      edition).  New York:  Wiley.  709 pp.

8.  Gujarati, D.  1978.  Basic Econometrics.  New York:  McGraw Hill.  462 pp.

9.  Hanson,  K.  R.,  R.  Lung,  and E.   Haver    1967.    A computer  program for
      fitting   data  to   the  Michaelis-Menten  function.    Biochemical  and
      Biophysical Research Communications 29:194-197.
                                   APPENDIX


Consider a model in which y is a function of x and a parameter 8:

                                 y = f(x, G )

The variance of the estimated parameter 0 is approximated by:


                                        7    n    7
                              var(9)  =  a /   \   b
                                           1=1

where

                                    A   5f
                                    A -69

     2                                    2
and a  is the error variance.  Therefore A  evaluated at a given value of x  is
a  measure of  the  amount  of  information   that  that  observation  contributes
toward estimating the model parameter (Beck and Arnold 1977).
                                      12

-------
     When a model involves two or  more  parameters,  the situation becomes more
complex  since  each  observation contributes  information to the  estimation  of
every parameter.  An exact solution to the problem of designing the experiment
is given by  Box and Lucas (1959;  or Appendix  I  in  Currie  1982).  However,  in
most  cases  the  method  given  above,  carried  out  for  each  parameter  in  the
model, gives an adequate indication of how to design an experiment.
                                      13

-------
                           B
Figure  1.  h schematic representation of the confidence interval
  around the estimate of the siooe of a straight line, when
observations are taken over a narrow rannp (,\) versus a
broader ranqe (B).
                        14

-------
 c

 d
 e
                             —»    2.86 a'

                             -*-    2.28 a'

                             	    2.50 a:
1.86 a

1.48 a2
                                    4.10 a'
                                      .-bx
 0
                                      4/b
Figure   2.  The effect of the experimental design on
     the precision of estimation of the  decay constant b.
     Each array of diamonds (a-f) represents a series of five
     observations spread in various ways across the range
     of x.  The resulting variance of the estimate of b
     is shown as a multiple of the error variance of a2.
                  15

-------
     1.0  i
     0.75-
O)
CD
_CC

CD

oc
     0.50
      0.25
                               y. = ae
                               3 1
                                       -bX:
            0
1/b
2/b
3/b
                             X-
         Figure   3.  Tne relative amount of information which

              an  observation y.  taken at a given  x. contribute

              to  the estimate of b.
                      16

-------
    14

o
>
    1-0
    0-75
    050
   025 -
                                 B
      0
0-5
1-0    1-5
  a
0     0-5    1-0     1-5
              a
    Figure  1.   The bias (K/K) and variance (var(K)) of estimates
         of the Michaelis-Menten K, as a function of the design matrix.
         Panels A and B represent the effect of geometric sequences of
         observations; panels C and D show the effect of linear
         sequences.   In both cases, a_ serves to telescope out the observations
         such that then a is small, observations are concentrated near
         the origin,  and when a_ is large, observations are more spread out.
         A nonlinear  fitting technique was used.  See text.
                                  17

-------
     20-
     1-5
     1-0
    2-0
           o
            V.
     0
0-5     10
       a
                                                  -i	r
                                          \
0     0-5
                                                 1-0
                                                   a
—I—
 1-5
Figure  5.  Same as  Figure  4,  using a linear fit to Woolf transformed
            data.
                                 18

-------
                                                      %  PARTICULATE
•a c r+ -s
O "O O fD
3
O
3
   7T -5
   fD O
     CO
   O -5
     tt>
     to
QJ  C
—' -5
   < co  fD
cr ro -"•
-J. .  <  fD
c+    fD  CO
   a>
cr
 ii  -i-
O 3

b H-'
CO -"•
CO D»
   o
  T3
   ro
   a>
     CO  fl)
     3- <-t-
     O  fD
     -J  Q-
     rt-
     fD  to

        O
     -J.T3
     3  fD
   Q.
   fD
   fD
   -S
        CO
     CO  (-f-
     fD  T
     LQ  Ql

       CO
     3  3"
     r+  rt-
     oo

     O  -••
     -h  3
        fD
     CU  CO
Ol
0 O
1 1 1



o "
CO
m
§ "-
0 o
^L.
O
CO
CO
0 '




^ifc.
•"? M> j o- i cr
J II II
o Jo
o b
CO CO
> > ro ro .
00 o>


, »


cr
II
p
b
CO
CO
(O
» >
o
o
cr
II
o
b
ro
O)


-
cr
II
o
b
ro
a>











»• .
cr
II
0
b
_k.
00












   cu "0
        fD
        Q.

-------
             EXPERIMENTAL REQUIREMENTS  FOR MODELING DECAY CURVES

                                 W. J.  Maier


                                   ABSTRACT

           The objective in analyzing and  correlating  data  on  the decay
      of organic  substrates  is  to  develop mathematical models  that  can
      be used to  characterize  rates  of decay.  Mathematical models  are
      needed  to  extrapolate   the  available  information  to  different
      environments  or  to   predict  the time  required  to  reach  some
      predetermined concentration.

           Mathematical  curve  fitting per se  is  not   adequate  for  this
      purpose if  the physical-chemical-biological  significance  of  the
      coefficients has  not  been defined.  A  prerequisite  for  a  useful
      model  is  that  the  equations'   parameters correspond  to  the  rate
      controlling reactions.  The parameters  can  be adjusted to reflect
      different  environmental  conditions.
                              UNSTRUCTURED MODELS
     Biochemically  mediated  decay  of  organics  stems  from  the  metabolic
processes of the  microflora.   Rates are closely  related  to the concentration
of  organisms   as  an   important   description   even   if   the   microorganism
concentrations remain constant with time.  The framework for modeling is given
by:
                    ds _ px
                    dt    y
(1)
                    dt
                            - k .x
(2)
where    s   =  substrate concentration
         x   =  concentration of microorganisms
         u   =  growth rate coefficient
         y   =  yield coefficient
         K.  =  endogenous decay coefficient

For those situations where the decay rate is mediated by extracellular enzymes
(Eg), it may be necessary  to  relate  its  concentration to x.  The parameters u
ana y must be defined in terms of the important process variables.

     A number of alternate expressions have been proposed for y and y as shown
in Table 1.  The most appropriate expression to use for a specific application
depends  on  the range  of substrate concentration  of  the decay  curve  and the
presence of other competing organic substances.
                                      20

-------
                                   TABLE 1
                      Alternate Expressions Proposed for
                the Growth Rate (y)  and yield (y)  Coefficients
     y = y
          m
                                       Exponential  growth, unlimited
y =
            VHT
                                            Monod growth equation
     y =
     y =
ym
K
s


S
+ S + S2
Ki
y S
m
                   Ki
                        f(Sj)
                                       Monod   equation   with   substrate
                                       inhibition coefficient Ki
                                       Monod   equation   with   terms   to
                                       account  for  substrate  inhibition
                                       and  alternate  substrate  effects
                                       due to Sj substrates
     u =
       K
           Mm, S
             In n
        s,n
                        V + f(SJ)
Summation    of    growth    rates
represents  concurrent  utilization
of alternate substrates i
    Y = constant

    r-    Y
        1 + kd/y
    Y  =  f(T)
                                       Yield coefficient  is  dependent on
                                       rates  of   endogenous  decay  and
                                       growth

                                       Yield  coefficient  is  temperature
                                       dependent
* y  is a function of temperature, pH, and ionic strength

The first two  relationships  in Table 1  for p are  widely  used.   The advantage
of the  simpler models  is that  the  resulting  equation  can be  linearized  or
solved  explicitly.    Cell mass  concentration  can  be expressed  in terms  of
equivalent  substrate concentration,  thereby eliminating   it  as  a  dependent
variable.   This  allows  calculating  the numerical  values  of y  and  K   from
measurements of  substrate decay  curves  providing  an independent  measure  of
initial cell mass concentration is available.
     Solution of the  third  equation in Table  1  which  incorporates the effect
of substrate inhibition requires making some  simplifying  assumptions.   K  can
be evaluated  from data  at  low  concentration by neglecting  the  Ki  term and
conversely, Ki  can  be evaluated from  high substrate  concentration  data.   It
is,  however,  necessary to  factor  in  the time  rate  of  change of  cell  mass
concentration.    Application  of  the last  two relationships  for  growth  rate
                                      21

-------
expressions  in  Table 1  requires  the concurrent  analysis  of several  sets  of
decay  curves  to  define  the  interaction  effects,  namely,  inhibition  by
competing   substrate  and   growth   enhancement   by   concurrent   substrate
utilization.  The  equations  cannot  be solved explicitly.   One  approach is  to
correlate  the  data  using  the differential  forms  and  cross correlate to
describe the effects of  alternate  substrates.   Cell mass  concentration  is a
variable and must be factored into the analysis.

     It  is  important  to note  that  cell  mass  concentration  (x)  is  a  time
dependent  variable  regardless  of  the relationship used to  define  y.   In  some
situations  (when the initial cell mass  concentration is  large compared to the
rate  of  additional  cell  production),  it  is   acceptable  to  treat  x  as  a
constant,  thereby  simplifying  the  data  analysis procedures.   It  is, however,
necessary to obtain an independent measurement of the initial active cell  mass
concentration  in  order  to  characterize  the  substrate   decay curve  or  to
calculate the numerical value of y from a decay curve.

     In the  more general situation  where  cell  mass  concentration  changes with
time,  numerical  integration  of  equations  1  and 2 can be used  to simulate
substrate decay curves.

     Computerized  subroutines  for  solving coupled  differential  equations  are
readily available.  They can be used to predict decay curves from known values
of y , K ,  k ,,   Y  and  they  can  be  used  to  estimate  the  values  of  these
parameters   from  measured  substrate  decay  curves.    Parameter  estimation
subroutines  are  available  to calculate  the  best fit  values  for  simple  decay
curves  as  well  as  for  a  sequence  of  decay  curves starting  with different
initial concentration of cell mass or substrate.

     Examples  of the  use  of numerical  analysis  of  laboratory test data  are
illustrated  in the following figures.

     Figure  1  shows decay  curves  of  2,4-D  using  acclimated  inoculum at  two
different  initial  concentrations.   The  data  (points) from  both test runs were
fitted  by  nesting  the  data.   The  kinetic parameters were  estimated  using a
fitting subroutine.

     Figure  2  shows 2,4-D  decay  curves and  attendent  increase in cell  mass.
Initial cell mass  concentration was calculated  using  fitting  subroutine.

     Figure  3  shows  an  application  of parameter  estimation  to  fit  3,5-DCB
decay curve.  Lines  illustrate the effects of changing the  numerical values of
the  parameters  y , K$, XQ.   Best  fit  calculations  become progressively more
time consuming arVti  less  reliable for estimating multiple parameters.

     Figure  4  shows measured decay curves  of 2,4-D with  and without  addition
of  alternate substrate (nutrient  broth).   2,4-D acclimated inoculum was  added
at  time zero.    Filtered inoculum had  lower  active cell   mass  and much  lower
protozoa concentrations.

     Figure  5  shows measured decay of  2,4-D  (same  as Figure 4a-4b) and  model
calculations of   substrate  decay  curves.    Model  calculations   incorporate
effects   of  substrate  inhibition,  substrate  interaction,   and   concurrent
substrate  utilization.
                                       22

-------
                            SEMI-STRUCTURED MODELS

     A  major drawback  of  unstructured  models  is  that  they  are  limited  to
describing the  kinetics  of balanced  growth,  sometimes referred  to as steady
state growth kinetics.  Time dependent growth phenomena have been described in
a  variety  of situations.   The  most  obvious cases occur  when  a  new substrate
that requires acclimation of enzyme systems is suddenly made available.

     Recent  advances  in  modeling time dependent  growth  phenomena are largely
based on  advances  in understanding  of  the regulatory  processes  that control
enzyme  production  and  activity.   Recognition that some  catabolic enzymes are
constitutive, e.g.,  formed at  constant  rates and  amounts, while  others are
inducible,   e.g.,   affected  by  concentration   of   specific   substrates  or
metabolites,  has   been  particularly   helpful  in  explaining and  modeling the
growth  behavior of  unacclimated  cultures.   Changes  in growth rate  are seen to
be  the  result  of  changes  in  enzyme  activity  levels;  different  enzyme
controlled reactions can become the rate limiting step in the overall sequence
of  growth  related  metabolism;   hence,   the  need  for   a  structured  model.
However, it  has also  become evident  that complete stepwise modeling of growth
processes  is extremely  complicated  and  probably  not  feasible  for  routine
use.  As a  result,  model  development has focused on semi-structured models as
a  compromise between  rigor and  practicality.   Semi-structured  models  have
found  favor  because  they  can  be  used  to  describe  time-dependent  growth
phenomena  without  exceeding   practical   limits  on  the  number  of  rate
coefficients, equilibrium constants,  and rate equations needed to describe the
system.

     One version  of a two-enzyme  model  has been described to illustrate its
application  for analyzing  and  correlating  lag  phase behavior.    The   model
considers two rate controlling steps, namely, substrate transport by permeases
followed by  intracellular  metabolism.   The data represent the  observed lag
phase when a 2,4-D acclimated culture is challenged with mixtures of 2,4-D and
glucose.  The model assumes that transport and conversion of 2,4-D  and glucose
are mediated  by their respective permease enzyme systems E-^  and E^; P} and
?2 are  internally  available intermediates and are used  for growth and energy
metabolism by their  respective  enzyme systems, E2} and  Ep^.   The  presence of
P-^ and  P~  can  affect  permease  enzyme  production or  activity;  repression  of
permease fay  the other intermediate was  assumed.  The model is schematized in
Table 2.

     The corresponding  mathematical   formulations for  calculating  the   time-
concentration changes of  the  two substrates (Sj  and  S-), their corresponding
intracellular products (P^  and  P-),  enzymes (E,,, E,p,  iLi,  E-p) and overall
cell   mass   formation  (X)   subject   to   material  balance   considerations  are
summarized in Table 3a; the symbols are defined in Table 3b.

     Published information  on  kinetic coefficients for  transport enzymes and
overall  catabolic  enzyme systems was  used.   Test  data using single substrates
were used to calculate coefficients for the acclimated culture.  The system of
differential  equations was solved by  numerical  integration using a  Runge-Kutta
subroutine available through the University of Minnesota Computer Center.  The
initial  values of  substrate and cell  mass are measured.
                                      23

-------
                               TABLE  2

           2-enzyme model, concurrent substrate utilization
          Version II - Different  intracellular intermediates

     Case I - Glucose permease repressed by 2,4-D intermediate
              2,4-D permease repressed by glucose intermediate
Uptake:
                1?
                                              ?2
                                               - -
                                                             Growth
The model:
Uptake:
!2
                            1?
                            Lf-
                                    c
                                  12 2
                                               c    + P
                                               C12    2
Growth:
           F   + p
           L21   P
               j. p
                 V
                       '
                       1?
                       ie-
                                    3?
                                     •
                                          21
                                 24

-------
                                  TABLE  3a
                               Rate Equations
               [Ell.TOT][SlJ
dt  ~ "31 S7   k   + S
            X    o -i
dt
             [EH.TOT][S1]   .    el
                           " K
dt  '  31 bl   k$  + Sx    "  31 f1   kp  vX
             >
" K                    " K
"dT" " K32 b2    k$  + S2   " K32 f2   kp vX


dX          el [E?1 Tni][Pl]          e? ^E?? TOT^PO]
5£ = k '  v   —   dL> IUI    i  + \e ' v   -A   ^2, TUT    2     ,
dt   K31yP1 fj   kp vX + P!  + K32yP2 f2   kp vX -HT^  "


dEHJOT _     Cl      dX
   dt       .    -. Dm , cTf
           dl + ]1P2 X

dE12,TQT =      C2     dX
    dt      .  .  ,  m0  "dT
           d2 + I


dE21,TOT      dX
   dt    ' Bl dt
dE22,TOT _    ^X
   dt      ^2 dt
                                     25

-------
                        TABLE 3b
                Explanations and Symbols
Sj:            glucose concentration, mg/1
S2'            2,4-D concentration, mg/1
Pj_:            glucose Intracellular intermediate, mg/1
P2:            2,4-D intracellular intermediate, mg/1
X:             biomass, mg/1
Eij JQ-J-:       total glucose transport enzyme, mg/1
Ej2 JOT:       total 2,4-D transport enzyme, mg/1
Eoj YQT:       total glucose Intermediate metabolic enzyme, mg/1
E22 TOT:       total 2,4-D Intermediate metabolic enzyme, mg/1
ko :           Michaelis constant for S^, 1=1,2; mg/1
 31                                    '
kp :           Michaelis constant for P.., 1=1,2; mg/1
K3.j, k' .:     enzyme rate constants, 1=1,2; hrs~*
kd:            biomass death rate coefficient, hrs"1
yp :           yield coefficient, mg biomass/mg P^; 1=1,2
a,, bi, e^, f ^:   molecular weights for S^, E^., P^, E2l-
v:             specific volume of biomaterial, 1/mg
                           26

-------
   MULTIPLE COMPONENT MODELS AND STEPWISE DEGRADATION OF COMPLEX SUBSTRATES

     Stepwlse  degradation  of   complex  substrates  can  Involve  one  or  more
species  of  microorganisms  acting  concurrently.    Oxidation  of ammonia  to
nitrite  and  nitrate by  nitrosommas  is a classical  example of  a  two species
process.   Measurements  of monochlorobiphenyl  degradation  show  evidence  of
sequential blodegratlon.

     Models that  explicitly simulate  substrate  utilization with formation and
utilization of intermediates by independently acting microorganism populations
have been  defined.   These models  have  the  same form as  the single substrate
models  described in  previous  sections.    The  only  difference  is  that  the
Inventory  of  each microorganism  population  is calculated using  a  series  of
coupled differential equations.  The availability of the second step substrate
(product  from  first step)  is  dependent on the  Sites of  utilization  of both
species.

     The principal criterion for  determining  whether  it  is necessary to use a
muIt1 component   model    1s   the   formation   and   accumulation   of  product
Intermediates.   If the measured concentrations of intermediates are  negligibly
small,  1t 1s  reasonable to  use  a  single  culture  model.    If  intermediate
products tend to accumulate at  a  slow rate,  the use of a  multicomponent model
is recommended.
                          PROTOZOA PREDATION EFFECTS

     The  role  of  protozoa  1n  controlling  bacterial  populations  has  been
recognized  and   documented  by   a   number  of  investigators  in  wastewater
treatment.  For example,  the rate  of 2,4-D utilization with 3 micron filtered
Inoculum resulted in higher maximum  rates of substrate removal.  This has been
ascribed  to  the  effect  of   removing   protozoa   thereby   reducing predation
effects.   Figure 7  illustrates  the  phenomenology for  2,4-D  utilization and
attendent population dynamics of bacteria and protozoa.

     Protozoa predation has usually  been modeled indirectly by introducing the
endogenous  decay  term kd  shown  in  equation 2.   This approach  appears  to be
satisfactory  for  analyzing continuous  flow  reactor systems.    However,  it is
not  adequate  for  simulating  the  phenomenology  described  in  Figure  7  which
represents batch test data and is typical of shock  loading  situations.

     Explicit  modeling  of protozoa predation  has  been  defined.   A  prey
predator  model  has  been  used  to  simulate  the  data in  Figure 7.   Kinetic
coefficients  that  describe the  growth  of protozoa and bacterial  decay  were
calculated and give a reasonable fit of the data.

     The  methods   for  measuring   microbial  concentrations   (bacteria  and
protozoa) are time  consuming  and relatively few  studies have  been published.
As  a  result,  1t  has not  been  possible  to  validate  prey   predator  models
adequately.    Additional  test  data  that  are  representative of  conditions in
natural  bodies  of  water  are  needed to  define  the  potential  importance of
protozoa and  other bacterial  dieoff mechanisms  that  could affect population
dynamics.
                                       27

-------
                                   SUMMARY

     It 1s  recommended  that the measurement  and modeling of  substrate decay
phenomena be broadened to include characterization of the accompanying changes
in active  cell  mass or  enzyme  concentrations.    Quantitative  descriptions  of
cell  mass  are  essential  in order  to obtain precise information  on  rates and
allow extrapolation of data to other environments.

     The form of the rate equation that is best suited for analyzing test data
depends on  the complexity  of  the environmental  system to  be characterized.
Substrate inhibition effects must  be incorporated at elevated concentrations;
the critical concentrations vary considerably from  100+  milligrams  per liter
to 300+ micrograms  per liter for 2,4-D and  PCP? respectively.  The effects of
alternate substrates can  be an  important factor  in  regard  to the utilization
of the  target  substrate and in  enhancing the proliferation  of  active cells.
The effects of alternate  substrates  should  be  considered at  all  concentration
levels.  This also implies that  the decay curves  of alternate substrate should
be measured concurrently with target substrate decay rates.

     It  is  further  recommended  that   semi-structured  models  be  used  for
analyzing  the   correlating  data  that   represent  transients  or  acclimation
periods.   The  greater complexity  of these models  is  compensated for  by the
advantages that accrue from more detailed mechanistic descriptions of the rate
controlling steps.  For  example, the semi-structured modeling approach can be
used to simulate the effects of  specific enzyme systems thereby giving insight
into  mechanisms.    Simulation  studies  can  also  be used to specify  enzyme
activity  measurements   that  are   needed  to   describe  the  transients  or
acclimation process.
                                        28

-------
                                CONCENTRATION,  mg/l
CQ
C
-5
n
rr

o

o
T3

n>
3
O
X
<<
Cu
O
n>
rf
o
	14

Q.


Q.


O
                                                         X ?^
                                                        o   ex
                                                                    7; -< r:
                                                                  ~s.   •  3
                                                                     35
                                                     H   3—33--

                                                     co   \ n i  -^ -^J"
n
c
-5
<
CD
                                                          O O
                                                            O  O   O i— K^

                                                            O        00 ho

                                                            M        OC — I

-------
CO
o
                               C
                               -5
(D
3 n
Q. -i-
rt> O
3 3"
0
0 -5
ro o
— ' 3"
n>
3 3
CL> O
I/) X
co <<
OJ
o o
o o>
3 c~*"
O -J.
(D O
3
<-+ CD
-5 O









H
2
m
^
5
c
^•D
CO


                             QJ
                             rt Q-

                             O Q-
                             3 fD
                             (/i O
                             •  CD
                               O
                               c
                               -5
                                            O
                                             H-
         • * SI
S,-2,4-D  CONCENTRATION, mg/l
X, = ACTIVE  CELL MASS CONCENTRATION, _mg/I
                           00
          *
 o
—t—
                                                                 0= X
                            o
                             -
                                                                                           O
                                                                                           O

-------
0.0
            25
                                                    3,5-DCB TEST DATA
                                                    MODEL CALCULATIONS
                                                  \
                                                    \C
                                                     \
                                              0.412    0.257    0.257    0.405

                                      \ Ks   24'6     12.7     24.6     24.6

                                                             0.000218
50
75       100      125
 TIME, HOURS
150       175
          Figure 3.  3,5 Dichlorobenzoate decay curves.

-------
                                              CONCENTRATION, mg/I
                                                                                                              CONCENTRATION,  mg/
CO
                       Q> IQ
                       13 C
                       O- -S
                          ro
—' c:
o -s
3 fD
n>

CU Q.
3 •
Q.

s; ro
3- I
   o
3
c:  cr
c4- -J.
-S  O
_.. CL

3 
H
m
i
CO






S
o
8-
§•

M -
o


r/
f
il
%




£J
                                     £-
                                    » pt
                                    *Ul
                                      5"n c ~n
                                    ,S?q
                                    5sS»
                                    s«s°
                                    SiEU
                                    „ 3  5
                                    IzSg
                                                             (D /
                                                    01   00
                                            ro   +»  O)   uu   (-,   rvj   -»»
                                        OoOOOOOO
                                     Of)
                                                                CU U3
                                                                3 C
                                                                Q. -S
                                                                  ro
                                                                                      <-+ Cu
                                                                                      c-l- "
                                                                                      -S -f*
                                                                                      — '•  I
                                                                                      n> o
                                                                                      3
                                                                                      (-1- cr

                                                                                      cr o
                                                                                      -s Q-
                                                                                      O CD
                                                                                         -s
                                                                                         EU
                                                                                         O-
                                                                                         Cu
                                                                                         rt
                                                                  Cu

                                                                  O
                                                                  3
                                                                  ro
                                                                                      —i C
                                                                                      o -s
                                                                                      Co O
                                                                                      3 •
                                                                                      Q.
                                                                3
                                                                c cr
                                                                <-+ -••
                                                                -s o
                                                                _•. o.
                                                                n> rt>
                                                                3 U3
                                                                rt- -s
                                                                  o>
                                                                cr O.
                                                                -S Cu
                                                                O r+
                                                                <-t- -".
                                                                rr O
                                                                •  3

-------
                         2,4-D CONCENTRATION, mg/l
                                                                    2,4-D CONCENTRATION, mg/l
CO
oo
           c
           -s
           fD
           O
           O
           O
           3
           O
           c
           -s

           ro
           3
           Q)
           O
           3
           O
           -h
no
           o>
           3
           Q.
           fD
           13
           CO


           O
           t-h
           3"
O N
b oi
b en
	 1 i —
v_>
6 r
                           NUTRENT  BROTH, mg/l_
                            ro     01      ->i     O
                            Ot     O      01     O
       ro -
       O
       S

o
t-o
Ui


K>
m
o


o
o
-p-
r-o






c
o
o
hO





o >—
^O t-o
0 0


1 — ' K>
O ^-n
J> ^-J
O


C 0
1 — ' 1 — '
oo ro
CO — 1


0 0
.£> UJ
l-o
                                                     UD
                                                     c

                                                     ro

                                                     tn
                                                     cr
                                                     o
                                                     o
                                                     3
                                                     O
                                                     c
                                                     -s

                                                     CO
                                                     N
                                                     EU
                                                     rl-

                                                     o'
                                                     3

                                                     O
                                                     -h

                                                     ro
                                                     DJ
                                                     3
                                                     CL
                                                                -S

                                                                fD
                                                     CO

                                                     O
 ISJ
 O
CD
O
8
NUTRIENT BROTH, mg/l_
                                                                                                   t—  NJ

                                                                                                   O  l_n

                                                                                                     J>- --J C
                                                                                                           ^   00

-------
     Figure 6 shows a comparison of the test data and model  calculations for a
batch test with 2,4-D acclimated inoculum using  a  mixture  of glucose and 2,4-
D.   The  latter is  utilized  immediately;  glucose  utilization  starts  after
approximately  20   hours.    The  20-hour  lag  phase  in  glucose  utilization,
followed  by   concurrent  utilization,  and  enhanced  overall   rates  of  2,4-D
utilization  as  cell  mass  concentrations increase,  was correctly  simulated.
The  model  predicts  a  gradual   increase  in  glucose  transport and  catabolism
until  a  balanced  enzyme  activity  distribution  that  describes  concurrent
substrate  utilization  is achieved.   The latter condition  represents  pseudo
balanced growth with  concurrent utilization  of both substrates.  The excellent
agreement between  data and model simulation  is noteworthy because the findings
of these studies are  in accord with basic research using defined systems.
                                    34

-------
                   Tfgure  6
Model Calculations, Version II, Case I
A : Experimental data on glucose
Q : Experimental data on 2,4-0
                                            80.
                      35

-------
                                                        CONCENTRATION — mg/l
GO
CTi
                            in
                            c
                            en
                          CT •
                          EU
                          n
                          r+ ro
                          fD >»
                          -5 -P»
                          -<• I
                          o> n

                          Q. !-•
                            O
                          •a o
                          rt-ia
                          o •>-,
                            03

                            o
                            a.
                            n>
                            IQ

                            cu
                            a.
                            cu
                            rt-

                            o'
                            =5

                            Cu
                            3
                            O-
                            c

                            cr
                            0)
                            •s
                                               NUMBERS OF  BACTERIA 8  PROTOZOA

-------
                                 PANEL REPORTS

     Our  recommendations  on statistical  techniques to  be applied  to  future
studies  in  biodegradation  must  be  divided  into  two  sets.    One  set  of
recommendations  applies   to  those  individuals   (Group   I)   who  feel   that
estimating biodegradation  rate coefficients to within an order of magnitude of
their correct  value  is  acceptable.  This  acceptance  largely  results from the
level  of  noise   (error)   in  the data   from  which  these  parameters  are
estimated.   A different  set of  recommendations  applies  to  the  second  group
(Group  II),   which  is  interested  in  more  precise  parameter  estimates  and
mechanistic model-bullding.  Presumably,  individuals  in  the latter  group work
with data several-fold more precise than the data to which the first group has
access.
RECOMMENDATIONS FROM GROUP I

    1.  Never  propose  a   model  that  is  more  complex  than  your  data  will
        support.  This becomes especially important when the level of error is
        high  (e.g.,  when  the standard deviation  of  replicate  measurements of
        the  dependent  variable  exceeds   10%  of  the  average  value  of  the
        dependent variable).

    2.  If  more  complex models are  to be used,  then  the  addition of higher-
        order terms  (e.g.,  nonlinear terms) should be  tested  for by using an
        F-test.

    3.  Use  nonlinear  parameter  estimation  (NPE)  techniques  to  estimate
        microbial   rate   coefficients,  when  possible,  but  recognize  that
        estimating  these  coefficients by  using  linearized  forms of nonlinear
        models probably will not produce disastrous results.   Fitting the data
        directly to  a nonlinear  model, rather than to a linearized form, can
        buy the investigator protection even when the level of noise is high.
RECOMMENDATIONS  FROM GROUP  II

    1.  Given  a  data set,  propose  several  competing models  to represent the
        underlying  biological  process.   Differences among these models may be
        initially  assessed via  simulation.    The simulated  decay  curves may
        subsequently be used to aid the design of an experimental program.

    2.  Estimate all parameters for the proposed models  via NPE.  The  standard
        errors of the parameter should also be approximated.   Explicitly  state
        the  updating method used  to fit  the data  directly  to the nonlinear
        models.   No single method for estimating  the  parameters of nonlinear
        models  is  best  and there are several  algorithms available.   The most
        commonly used NPE method -- typically available  on main-frame  computer
        systems  —  is the  Levenberg  Marquardt algorithm.  The  Gaussian method
        is another  popular  NPE technique.

    3.  Select the  "best" of the proposed models  by  using an  F-test  or similar
        model-building test.
                                      37

-------
    4.   Once the appropriate  model  has been selected,  additional  experiments
        may  be  run  to  obtain  even  more  precise  estimates of the microbial  rate
        coefficients.   In these cases, optimal  design criteria  should  be  used
        to select the  range  of measurement  of  the dependent  variable.   This
        will   ensure  that  the  variance  of  the  estimated   parameters  is
        minimized.

             As a simple example, assume  that  the first-order  model  has  been
        found to best  represent a  given  data  set  by  using  an F-test.   Now
        additional  experiments are to  be  run with the  goal  of obtaining  more
        precise  estimates  of  K,  the  first-order decay   coefficient.    The
        optimal design for these experiments is to center  the  measurements of
        the   dependent  variable (presumably, substrate  concentration)  on the
        point at which  the  dependent  variable  is  37%  of  its  starting value.
        Of course,  data obtained for the optimally designed experiments should
        be analyzed by using  an  NPE method,  again  stating the technique of
        choice.

    5.   Robust  regression (e.g., bi-weight regression) may be used  to minimize
        the   influence   of  outliers  when  data are   fitted to  a  particular
        nonlinear model, but only when  it  is known a  priori that the data are
        best represented by  the chosen  nonlinear mode~"L
                              GENERAL REFERENCES

     The  following  general  references  may  be  used  for  details  on  fitting
nonlinear  models  to  data.     Further,  some  of  these  references  provide
information on optimally designing experiments  for  parameter estimation.  The
textbook by Beck and Arnold  pays  a great  deal  of attention to the statistical
assumptions necessary for the correct application of least-squares analysis.

1.  Beck  and  Arnold   (1977).     Parameter   Estimation   in  Engineering  and
      Science.  John Wiley & Sons, Inc.

2.  Draper and Smith (1981).   Applied Regression Analysis, 2nd ed.  John Wiley
      & Sons, Inc.

3.  Wonnacott and Wonnacott  (1971).  Econometrics.  John Wiley & Sons, Inc.


                         THE "SECOND ORDER" APPROACH:
                 ASSUMPTIONS, LIMITATIONS, AND RESEARCH NEEDS
    Objective:      To   review   the  conceptual   framework   and  assumptions
           associated  with  deriving  biodegradation  rates  from  decay curves
           which appear first order with respect to chemical concentration and
           whose slopes  are linearly  related  to catalyst  concentration.   To
           develop a strategy for future research.

     Many  chemicals  appear to  degrade  exponentially when  exposed to  inocula
taken  directly  from  the  field.    Plotting  the  logarithm  of  the   chemical
concentration  against  time  produces a straight  line.   The  slope of this line


                                      38

-------
is  generally considered  to  be  a  rate constant  which  is  first  order with
respect  to  chemical   concentration  if the  catalyst  concentration  remains
constant  during  the experiment  and the chemical  concentration  is  below the
half saturation  constant  (K$).   Moreover,  it is  assumed that differences in
the first order rates from one experiment to another are  due to differences in
the catalyst  concentration.    How  valid are these assumptions?   Do they need
further experimental verification and how?  Are there special cases where they
apply?    What  range  of  chemical   concentrations  needs  to  be  tested?   Are
catalyst  concentrations  being  determined  properly  and effectively?    Are
biodegradation rate constants  for any particular chemical  the same, regardless
of the  source  or composition  of the  biological  catalysts?   What experimental
evidence  is  needed  to  substantiate or  refute  this  concept?   What alternative
mathematical  expressions  can  be  used  or  developed to  describe exponential
decay?

    Panel Members:  Bob Boethling (Chairperson), U.S. EPA
           John Rogers, Battelle Northwest Laboratory
           David Lewis, U.S. EPA
           Joe Suflita, University of Oklahoma
           John Wilson, U.S. EPA
           John Rodgers, North Texas State University
           Rick Cripe,  U.S.  EPA
                                      39

-------
                THE  "  SECOND  -  ORDER  " MODEL  - AN OVERVIEW

                             Robert  Boethling



                                  THEORY

     Kinetic models  to predict  rates  of biodegradation  of chemical
substances have generally relied  upon the  empirically derived Monod
equation (Monod 1949).  According to  Monod:
           d[B]
           dt
u
          max
                   CS]
                            (1)
                                 Kc  +  [S]
where [B] is the microbial  "concentration," V  m^x  the  maximum  rate  of
growth of the population, [S] is the concentration of  the  growth  -  limiting
substrate, and Ks is [S] when the growth rate  is  1/2 V niax.  It  is evident
that the Monod equation is  an expression of the rate of growth  of the
population, d[B]/dt, in terms of population size  and substrate  concentration,
To derive an expression for the rate of biodegradation of  a  substrate S,
equation (1) may be modified by addition of a  yield coefficient,
                            Y = d[B]/d[S]
                                          (2)
which describes the efficiency of conversion of substrate into microbial
biomass.  By multiplying equation (1) by 1/Y adding a negative sign to
convert equation (1) to an expression for loss  of S, one obtains equation
(3).
           d[B]
           dt
d[S]
don
                         max
                          KS
                         [B] [S]
                         + [S]
          -dLS]
          dt
 [B]LS]
;+  CS.1
                                          (3)
The expression -d[SJ/dt is equivalent to the rate of disappearance of the
growth  -  limiting substrate.

      To make use of equation  (3), one must assume that the rate of
population  growth is  limited  by the substrate S over the expected range
of [S].  This is not  likely in environmental situations, however, because
many  other  compounds  that can serve as sources of carbon and energy are
commonly  available.
                                     40

-------
The difficulty of studying bacterial  growth  kinetics  at  environmentally
relevant concentrations of S, where S is truly growth -  limiting,  is
amply illustrated by the work of Shehata and Marr (1971)  with  pure cultures
of Escherichia coli.  To remove traces of utilizable  carbon  that  would
have precluded growth limitation by the added substrate  glucose  at parts  -
per - billion (ppb)  levels of that sugar, procedures  such as distillation
of water from permanganate solution and exhaustive washing of  glassware
with chromi acid were required.

     A closely related problem arises from consideration  of  the  yield
coeeficient Y.  When substrate is growth limiting, Y  approaches  zero as
substrate is utilized for growth and energy, and [S]  approaches  zero.
Although not obvious from equation (2), this phenomenon  is well  documented
and is a reflection  of the diversion of substrate metabolism from growth
to the need for "energy of maintenance" under conditions  of  carbon and
energy needs may be  provided by metabolism of those substrates.   The
practical outcome is that the "correct" value of Y to use in equation  (3)
cannot easily be determined.

     If one is willing to disregard these theoretical difficulties,
equation (3) can still be used to derive first-and second-order  rate
expressions for biodegradation.  If the concentration of  substrate, [S],
is much greater than Ks, Ks + [S] approximates [S] and equation  (3)
reduces to:

          -d[S]                 =    Umax    [B]                 (4)

          dt                         Y

     In this case the rate of substrate disappearance is  proportional to
the microbial concentration, [R], but not to substrate concenration.
This is the usual situation in laboratory studies of  microbial  growth  in
pure culture, where  substrate is plentiful.   Environmentally,  we are
usually more concerned with situations in which [S] is lower.   If [S]  is
much lower  than Ks, Ks + [Sj approximates Ks and equation (3)  reduces to:


          -dLS]               =  Umax[B]  [S]                    (5)

          dt                     YKS

     Equation  (4) is a first-order kinetic  expression because only one
term is present as an independent variable.   Equation (5) is a second-
order expression because substrate disappearance is dependent  upon two
variables, [B] and [S].

     The kinetic constants U max, and Ks and yield coefficient Y  (assumed
to be constant) may  be expressed as a single constant K:
                              = U,nax                             (6)
                                YKC
                                     41

-------
Equation (5) then becomes:

           -d[S]               =    K  [B]  [S]                      (7)
           dt

     Equation (7) is the second-order rate  expression  used  EPA's  Athens
laboratory, and K is the second-order rate  constant.   Because of  experimental
difficulties in measuring U max,  Y  and Ks for  many  organic  chemicals,  K
is usually determined from equation (7) rather than equation  (6).
Rearrangement and integration (assuming  [B]  is  constant) of  equation  (7)
yields equation (8):

                       In LS] = -K[B]t +  In LS0]                 (8)

where [S0] is the concentration of  substrate  at  the start of  the  experiment.
A plot of In [S] vs. t should yield a straight line having  a  slope  of  -K[B]
and a y intercept of In [S0].  The  second-order rate constant,  K,  is
obtained from the positive slope of equation  (8) by simply  dividing by
the microbial concentration, [B].  The positive slope  of equation (8),
K[B], is also referred to as the pseudo first-order rate constant.   It is
clear from equation (8) that only disappearance of  parent compound  and
microbial concentraion must be determined experimentally.

     Unfortunately, the assumption  that [S] is typically much lower than
Ks, necessary to derive equation (7), may be a major theoretical  stumbling
block.  Values of Ks for naturally  occurring organic compounds  such as
sugars and amino acids are usually  in the ppb range or lower  (Hodson
1977, Wright and Hobbie 1965).  Fewer Ks values have been obtained  for
xenobiotic chemicals, but existing  data do suggest  that ppb values  of  Ks
are not uncommon.   For example, Button et al. '(1981) recently obtained Ks
values for acetylene, benzene, and  ethylacetate degradation in  seawater
samples of 0.25 to  2.8 ppb.   In the same study, Vmax for degradation of
toluene in seawater was reached at  a substrate concentration  less than or
equal to 52.7 ppb,  suggesting that  Ks was significantly lower.   Moreover,
using estuarine water samples, Bartholomew and Pfaender (1983)  have
observed Ks  values  well below 10 ppb for several chemicals, including  m-
cresol and chlorobenzene.  Baughman et al. (1980) state that  Ks values
typically  range  from 0.1 to  10 parts per million (ppm) in  natural waters,
but provide  no  references  in  support of  that conclusion.  Since concentrations
of xenobiotics  in natural waters are often in the  low ppb  range (Sheldon
and Hites  1979), it is clear  from  the  foregoing that [S] cannot be assumed
to be much  lower than Ks.  Thus, it seems  likely that equation (3) will
in many cases not be approximated  by equation (5).   In those cases the
simple second-order rate model represented by equations (5) through (8)
is invalid.

      The  conclusion is not that this approach should be abandoned, but
rather that  more research  is  needed.   Particularly needed  is better
knowledge  of Ks  values for biodegradation  of organic chemicals by  natural
microbial  populations.   But  more fundamentally, we need better knowledge
of how microbial populations  degrade organic chemicals at  low, environmentally
 relevant  concentrations.   For example, evidence  (Rubin et al.  1982)
suggests  that the bacteria  responsible for biodegradation of organic
chemicals  at very low  substrate  concentrations  (oligotrophic bacteria)
                                      42

-------
may be different from those carrying out similar transformations  at  higher
substrate levels, so that rate constants determined in studies  using
substrate concentrations of 0.1-10 ppm may not accurately relect  degradation
kinetics at lower concentrations.
                                     43

-------
                                 PRACTICE
     Regardless of theoretical  difficulties,  the simple  second-order rate
model may be useful as long as  it "works."   But  inadequate  validation
remains a serious deficiency, and careless  manipulation  of  environmental
samples used in degradation studies can invalidate  conclusions.   With
respect to validation, the model  predicts that the  rate  of  biodegradation
of a substrate should be proportional  to the  microbial  concentration, and
that second-order rate constants  should therefore be  reproducible from
site to site and sample to sample.  Although  a general  relationship
between population size and biodegradation  rate  is  intuitively  reasonable,
few quantitative studies are available to support it.   Paris  et  al  (1981)
established such a relationship for three chemicals,  the butoxyethyl
ester of 2,4-dichlorophenoxyacetic acid (2,4-DBE),  malathion, and
chlorpropham (CIPC).  Site-to-site  reproducibility of  the  second-order
rate constants was excellent, with coefficients  of  variation  for each
chemical of less than 65% over  all sites.  Larson and Payne (1981)  recently
provided similar evidence for linear alkylbenzene sulfonates  (LAS)
degradation in samples of Ohio  River water.

     On the other hand, several studies have  failed to  demonstrate  a
significant relationship between  biodegradation  rate  and microbial
concentration, or have shown that other factors  may be  more important in
determining biodegradation rates.  In  a study of biodegradation  of  several
organic chemicals in fresh, marine, and estuarine waters, Bartholomew and
Pfaender (1983) found that measurements of  microbial  numbers  did not
correlate very well with measurements  of microbial  activity,  and in
particular, with the maximum rate of utilization of the  added substrate m-
cresol.  Nesbitt and Watson (1980a, b) studied the  biodegradation of 2,4-
dichlorophenoxyacetic acid in river water,  and showed that  degradation
was primarily dependent upon rnicrobial activity  rather  than the  total
number of microorganisms.  They also demonstrated that  microbial activity
was related to the concentration  of inorganic nutrients, the  suspended
sediment load, and the dissolved  organic carbon  content  of  the  water.

     Yordy and Alexander (1980) demonstrated  that the rate  of disappearance
of N-nitrosodiethoanolaniine in  samples were collected.   These authors
suggested that the presence of competent microorganisms  in  water samples
was more important than total population size.   Boethling and Alexander
(1979) observed apparent thresholds at ppb  concentrations for degradation
of certain organic chemicals in samples of  stream water.  These  observations
do not support a simple second-order rate model, since  at low substrate
concentrations biodegradadation rates  should  be  directly proportional to
the initial substrate concentration.

     Spain et al. (1980) showed that adaptation  of microbial  populations
by prior exposure to an organic chemical could have a marked  effect  on
biodegradation rates observed in  subsequent exposures to the  chemical.
Measurements of total microbial population  size  did not  reflect  the
altered biodegradation rates.  These and other studies  show that environemtal
factors may overshadow the effects of population size.   This  statement has
important implications for prediction of biodegradation  rates in natural
waters.  It says that a smiple  second-order rate constant determined
                                     44

-------
using a given water sample cannot be used to predict degradation rates  at
other sites, without consideration of factors other than total  population
size that may affect the rates.

     It is significant that (1) the compounds studies by Paris  et al.
(1981) are known to be biodegraded initially by hydrolysis,  catalyzed  by
enzymes that are probably constitutive, and (2) biodegradation  was followed
by measuring disappearance of parent compound.  The choices  of  test
compound and analytical method thus acted to maximize the probability
that biodegradation rates would be proportional to total population size.
It is reasonable to suppose that different results might be  obtained if
test compounds and conditions required degradation by less obvious routes.
Would degradation rates always be proportional to total  population size
if the transformation of interest were carried out by few bacteria, via
complicated, inducible pathways?  Would strict proportionality  be obtained
if the concerted action of several different bacterial  strains  were
required? These phenomena are very common.  Couldn't prior exposure to
the chemical of interest, or to chemicals having related structures, have
greater impact on degradation rates than total proulation size?

     Many of these problems could be avoided if there existed good methods
for measuring not the total microbial population, but only that segment
of the population active in carrying out the desired chemical transformation,
Unfortunately, such methods are not available for most chemicals.  This
is an area in which innovative research is badly needed.  In the absence
of methods for measurement of transformer populations or activity,
calculation of second-order rate constants must depend upon  measurements
of total population size.  Examples of methods commonly used by
microbiologists include plate counts, determination of total adenosine
triphosphate (ATP), and direct counting by epifluorescence microscopy.
Which method is used really doesn't matter, since they all suffer from
the same drawbacks relative to their application to the simple  second-
order rate model.

     The preceding discussion points to the need for much greater effort
in the validation of the model.  The most critical need is for  testing  of
more compounds, representing a wide variety of chemical  structures of
varying complexity.  The biodegradation of these chemicals should be
studied using samples of natural waters that would be expected  to present
a variety of adaptation possibilities: water from streams receiving
industrial effluents, pristine lakes, saltwater environments, etc.
Furthermore, biodegradation should be determined by measuring a parameter
such as C02 evolution in addition to disappearance of parent compound.
The approach of Larson and Payne (1981) is encouraging in this  regard,  as
these investigators used ring-labelled LAS and followed production of
  C02 from cleavage of the benzene ring.  Ring cleavage requires the
action of inducible enzymes.  The choice of LAS may be criticized from
the standpoint of model validation, however, since their use as detergents
has no doubt led to the widespread development in nature of  microbial
populations competent to degrade them.

     Another issue worthy of mention is related to the changes  in microbial
numbers and activity, population diversity, and nutrient concentrations,
that usually occur during long-term incubation of environmental samples.
                                     45

-------
It has long been known,  for example,  that  changes  in  population density
of 100 fold or more can  occur in  samples stored  for  long or even short
(less than 24 hours)  periods of time,  even when  the  samples are refrigerated
(ZoBell 1946).  Accordingly, certain  precautions must  be observed  in  any
biodegradation study in  which degradation  in  environmental samples  is
determined.  Such changes  are difficult  or impossible  to avoid in  long-
term biodegradation tests, but they can  be minimized  during the period
from sample collection to  the start of the test, and  changes  in microbial
numbers over time, at least, can  be easily followed.

     It is therefore unfortunate  that Paris et  al. (1981)  do  not indicate
how long or under what conditions wAter  samples  were  held  before they were
received in the laboratory.  Moreover, significant changes may have
occurred during the 48-hour period between receipt of the  samples  and the
start of biodegradation  tests.  Paris et al.(1981) state that bacterial
numbers did not vary significantly over  the course of their experiments.
Such behavior could be expected if one or  more  days  had elapsed between
sample collection and the  start of biodegradation  experiments, because
bacterial  numbers in stored water samples  typically  increase  rapidly, and
then decrease very slowly  over several days or  longer (ZoBell 1946).
Considering the fact that  water samples  were  collected from 40 sites  in
18 states, delays of one to several days would  seem inevitable.   If the
microbial  numbers thus determined do not reflect in situ values, second-
order rate constants calculated from those numbers cannot  have environmental
significance for the sites from which the  samples  were collected.
                         SUMMARY AND CONCLUSIONS

     The second-order model for prediction of biodegradation of organic
chemicals  in natural waters offers beguiling simplicity.  One needs only
to follow disappearance of parent compound and microbial  concentration
over time.  The pseudo first-order rate constant is determined from a plot
of In chemical concentration vs. time, and the second-order rate constant
is obtained from the pseudo first-order rate constant by dividing the
latter by the microbial concentration.  The second-order constant is
presumably site and sample independent and can be used, given a measurement
of mocrobial concentration for any site, to predict a half-life for the
chemical of interest at that site.  Structure-activity relationships can
be established by correlation of the second-order biodegradation rate
constants with other known physical/chemical or fate parameters, such as
alkaline hydrolysis rate constants (Wolfe et al, 1980).

     Numerous theoretical  and practical problems intervene in the
implementation of this model, however.  Theoretical problems include the
false assumption that the  test chemical is growth limiting, the uncertain
meaning of the yield coefficient Y in such cases, and probably most
important, the potentially false assumption that natural chemical
concentrations are much lower than Ks.  Foremost among practical problems
is that the model is as yet largely unvalidated.  If one is willing to
suspend reservations about their handling of water samples, Paris et al.
(1981) have indeed demonstrated the site  independence of second-order
rate constants.  But only  three chemicals have been extensively tested,
                                      46

-------
and the chemicals had structures for which some degree  of  site  independence
of the second-order rate constants  should not  be terribly  surprising.
Moreover, there exists a substantial body of evidence to suggest  that,
for many chemicals, various environmental factors can overshadow  the
effects of microbial  numbers, thus  precluding  the establishment of
significant correlations between total  numbers and degradation  rates.

     The most urgent  need is for more research to demonstrate the practical
usefulness of the model.  Many chemicals  having a variety  of structures
should be studied, and proper attention should be paid  to  storage and
handling of water samples.  Concomitantly, microcosm studies should be
initiated with the same chemicals,  and  the results from the two techniques
compared.  Until these experiments  have been performed, it cannot be
assumed that second-order rate constants  are site independent,  or even
independent of sample for samples collected over time at a given  site.
If such independence  cannot be shown, second-order rate constants offer
no advantage over first-order constants.
                             LITERATURE CITED

     Bartholomew,  G.W. and F.K.  Pfaender.  1983.   Influence  of  spatial  and
         temporal  variations on  organic pollutant  biodegradation  rates  in
         an estuarine environment.   Appl.  Environ.  Microbiol.  45:103-109.

     Baughman, G.  L.,  D.  F. Paris,  and W.  C.  Steen.   1980.  Chapter 6.
         Quantitative expression of  biotransformation.   Quantitative
         expression of biotransformation  rate.   In  A.  W.  Maki,  K.  L.
         Dickson,  and J.  Cairns, Jr. (eds),  Biotransformation  and fate  of
         chemicals in the  aquatic environment.   Washington,  DC:   American
         Society of Microbiology, pp.  105-111.

     Boethling, R. S. and  M. Alexander.  1979.   Effect  of concentration
         of organic chemicals on their  biodegradation   by natural microbial
         communities.  Appl. Environ. Microbiol.  37:1211-1216

     Button, D. K,  D. M.  Schel1, and B.R.  Robertson.   1981.   Sensitive
         and accurate methodology for  measuring   the  kinetics  of
         concentration-dependent hydrocarbon metabolism rates  in  seawater
         by microbial communities.   Appl.  Environ.  Microbiol.  41:936-941.

     Hodson, R. E.  1977.   On the role  of  bacteria  in  the cycling of
         dissolved organic matter in the  sea.  Ph.D. dissertation.
         University of California, San  Diego,  Scripps  Institution of
         Oceanography.

     Larson, R. J. and A.  G. Payne.  1981.   Fate  of  the  benzene  ring of
         linear alkylbenzene sulfonate  in  natural waters.   App. Environ.
         Microbiol. 4.h  621-627.

     Monod, J.  1949.  The growth of bacterial cultures.  Ann.  Rev.
         Microbiol . _3: 37-394.

     Nesbitt,  H. J. and J.  R. Watson. 1980.  Degradation  of  the herbicide
                                     47

-------
    2,4-D in river water  -  I.   Description of study area and survey of
    rate determining  factors.   Water  Res. 14:1683-1688.

Nesbitt, H.  J.  and J.  R.  Watson.   1980.  Degradation of the herbicide
    2,4-D in river water  -  II.  The role of suspended sediment,
    nutrients and water temperature.  Water Res. 14:1689-1694.

Paris, D. F.,  W. C.  Steen,  G.  L.  Baughman, and J. T. Barnett. 1981.
    Second-order model  to predict  microbial degradation of organic
    compounds in natural  waters.   Appl. Environ. Microbiol.  41:603-609.

Rubin, H. E., R. V. Subba-Rao,  and M. Alexander.   1982.
    Rates of mineralization  of  trace  concentrations of aromatic
    compounds in lake water  and sewage  samples.  Appl. Environ.
    Microbiol.   4_3: 1133-1138.

Shehata, T.  E.  and A.  G.  Marr.  1971.  Effect of nutrient concentration
    on the growth of Escherichia  coli.  J. Bacteriol.  107: 210-216.

Sheldon, L.  S.  and R.  A.  Hites.  1979.  Sources and movement of
    organic  chemicals in  the Delaware River.  Environ. Sci. Technol.
    U: 571-579.

Spain, J. C., P. H. Pritchard,  and A. W. Bourquin.  1980.  Effects
    of adaptation on biodegradation rates in sediment water cores
    from estuarine and freshwater environments.  Appl. Environ. Microbiol
    4^:762-734.

Wolfe, N. L., D. F. Paris,  W.  C.  Steen, and G. L.  Baughman.
    Correlation of microbial degradation rates with chemical structure.
    Environ. Sci. Technol.   _14_: 1143-1144.

Wright, R. T. and J. E. Hobbie.  1965.  The uptake of organic  solutes
    in lake water.  Limnol.  Oceanogr. 10:22-28.

Yordy, J. R. and  M. Alexander.  1980.  Microbial  metabolism of N-
    nitrosodiethanolamine in lake water and  sewage.  Appl. Environ.
    Microbiol.  J39:559-565.

ZoBell, C. E.  1946.  Marine microbiology.   Waltham, MA:   Chronica
    Botanica.
                                48

-------
               THE SECOND-ORDER APPROACH TO PREDICTIVE MODELING

                                John E. Rogers


                                   ABSTRACT

           To estimate the  rate  of  degradation  and thus the persistence
      of organic compounds  in  a  wide  range of environments will require
      an  in-depth  understanding  of the  microbiological  transformation
      pathways  (reaction  sequences) of  major compound classes and the
      effects of physical-chemical  (environmental) parameters as well as
      microbiological  population   dynamics  on   the   rates   of  these
      reactions.   Although  not  complete,  a wealth  of  information is
      available  in  the  open  literature  describing  the   metabolism of
      major  compound  classes,   especially  in  the   area  of  aerobic
      degradation.   Anaerobic  metabolism  has not  been  characterized to
      the   same   degree.       The   microbiological    degradation   and
      transformation rates  of  numerous  organics have  been determined in
      environmental  samples.   Unfortunately, much of this  data base is
      sample  specific  and  not transferable  to other  sites.    Only in
      recent years have attempts  been  made to quantitate  these rates so
      they can be  used to  estimate  the persistence of specific organics
      in  a  range  of  environments.    These studies  have  led to  the
      development  of  a  number  of  mathematical  models  to  estimate
      persistence of organics.

                       DEVELOPMENT OF PREDICTIVE MODELS

     The  development  of  viable  research  and   predictive  models  for  the
degradation of organic compounds  in the environment  requires the interactions
of the following two research areas.

    -  The    development    of    environmentally   applicable    mathematical
       representations  of  the  microbiological   processes  involved  in  the
       transformation and ultimate degradation of  organic compounds.

    -  The  design  and development  of  procedures  to  produce  laboratory data
       compatible with the mathematical representations.

     Previously these  areas have,  in  many  respects,  developed independently.
In   the   future,    as   microbiological   degradation  models   become   more
sophisticated,  it  is essential  that  they  be  investigated in  an  interactive
manner,  each  process in  turn  being directed by the  other  as  new information
and methodologies are developed.  This will ensure the most rapid and complete
development  of  useful  microbiological  models.    This   does  not  preclude
theoretical  studies.   However,   the  development  of  models  is generally  of
little practical value when no consideration is  given  to the availability of
laboratory data to exercise the model  or worse if  no consideration is given to
the development  of laboratory  techniques  to provide  the data.   This  is also
true of degradation  studies,  where the usefulness of  the  data,  as  input for
available  or developing  computer models, has not  been  considered.   Continued
efforts  to integrate  these two  areas  will  lead  to  better  microbiological
models  and  a  better  understanding  of  microbiological   degradation  in  the
environment.
                                      49

-------
     An  important  first  step in  this  direction  has  been to  use  a  pseudo
second-order rate  equation  to describe  the microbiological  transformation of
organics in  fresh  and  marine waters.   There  are two  constraints,  however, to
the  use of  this  equation:   the  substrate concentration  must  be  less  than
saturating  (assuming that  microorganisms can  be considered  catalysts)  and it
must not support  growth  of the microorganisms.  The  utility of this equation
has  been  demonstrated over  a large  number  of  waters for  several  compounds
which include a  number of phenol  derivatives  (Paris  et  al. 1982), malathion,
and  chloropropham  (Paris et  al.  1981)  as well  as  butoxyethyl  ester  of 2,4-
dichlorophenoxyacetic  acid   (Paris  et al.  1981; Rogers  et al.  1983).   The
pseudo  second-order   rate  constants   were reproducible  within ~65% for  the
different waters  examined within  a  given  laboratory.  This  reproducibility
offered significant encouragement for the prediction of the environmental fate
in water of organics  that met the constraints  listed above.

     Early  in  these   studies  it  was observed  that   certain   compounds  were
transformed  only  after  a   variable and concentration   dependent  lag  or
adaptation period.   In the laboratory pseudo  second-order  rate constants are
determined by  dividing the measured  pseudo  first-order rate  constant  by the
bacterial concentration.   The  pseudo first-order  rate constant  is obtained
from the slope of  a  log  (residual  organic) vs. time plot.   When adaptation to
a compound occurs, the pseudo first-order rate constant is  determined from the
linear  portion  of the  log  plot.    It  was   immediately  apparent  from these
studies  that  the  use of  second-order   rates  to  model  the   degradation of
compounds which first  require adaptation is only a  partial  model.   A complete
model   would  include  a   mathematical   expression   for  adaptation  and  the
transformation  rate.    For  situations   where  growth  is  occurring  during
adaptation,  the  curve can  be represented by the  Monod growth  equation for
example.  In situations  where growth  does not occur but the organic substrate
concentration  is  saturating,  the  resulting  data  can  be   represented  by the
Michaelis-Menten equation for enzyme kinetics.  We  have observed that for  some
compounds  the  adaptation  period  is  far greater than the  time  during which
degradation occurs.  In  this  situation one can simply state that the compound
resides  in  the  environment  at a specific  concentration  for a specific period
of time.
                        UTILITY OF SECOND-ORDER MODELS

     The potential utility  of  the  second-order models rests on the ability of
laboratory  studies to  determine  an  average pseudo second-order rate constant,
using a number  of  different waters  with different microbial populations.  The
average rate  constant  can then be used  to predict degradation rates at other
sites  if  the  microbial   population  and  organic  substrate  concentrations  are
known.  This  combined  methodology can also  be used when the degradation  rate
is  described  by alternative  mathematical  expressions.   For  example,  if  the
Michaelis-Menten  equation  (1)  were used,  one  would  simply determine  the
average values  for Km, Vmax and the  bacterial  population of a  number of waters
and  then  use  these average   values  to estimate  degradation  rates  in other
waters  where   the  substrate  concentration   and  bacterial   populations  are
known.  The bacterial population in  this case  is  used to set the value  of  Vma
from an average specific  activity  term  (sp.  act.  = Vmax/bacterial population)
determined  in the  initial evaluation studies.  Similar examples can be  derived
from other mathematical expressions  for the degradation rate.
                                      50

-------
     A continued research effort in  this  area,  which I  will  term the "second-
order approach to predictive modeling," should address the areas of adaptation
kinetics, reaction  kinetics,  and quantitation  of  biomass.   Biomass  has  been
included because  it  becomes important when extrapolating from  basic research
models to widely  applicable predictive models  which  are  normally constructed
to  use  readily  available  or  easily  attainable  data.     For  example,  the
potential for  the degradation of organic compounds cannot  be  quantitated  by
using  a  most  probable  number  technique  (MPN)  which  utilizes  14-C-labeled
substrates.    Using this  technique,  one  obtains  a compound  specific biomass
measure.  14-C-MPN  techniques  can  be expensive.   However,  this technique can
be  shown to  be  proportional  to  the  total   bacterial  colony-forming  units
obtained by standard water  quality methods;  for example,  the initially costly
14-C-MPN  data base  can  be expanded  cost effectively  to   a  wider  range  of
environments.   This  would make  the  best use  of 14-C-MPN  which  is  a  good
research  tool  and  standard   plate  count  methods  which  is   a  widely  used
nonspecific enumeration technique.


                                   SUMMARY

     A research effort  should  include  at  least the following set of  tasks and
subtasks.

    Task  1.   Identify  adaptation  phenomena for  different compound  classes  at
           different concentrations.

           Subtask 1.     Classify compound classes as to 1,  2, and 3  below.
              1.           No adaptation.
              2.           Adaptation  is approximately equal  to half  the time
                          required to degrade the compound.
              3.           Adaptations  are  markedly  longer than  time required
                          for degradation.

           Subtask 2.     Define  the  kinetics  of   the   adaptation  process
              (growth, adjustment to  compound,  induction, etc.)

           Subtask 3.     Define  computer  inputs  and mathematical  expression
              for  different  adaptation  processes  (e.g., Monod  kinetics which
              require input  of binding constants  and specific growth rates).

    Task 2.   Identification of specific transformation kinetic processes.

           Subtask 1.     Characterize  compound  classes  as  to  pseudo  first-
              order, pseudo  second-order,  and Michaelis-Menten kinetics.

           Subtask 2.     Define  computer  inputs and  mathematical expressions
              for different  kinetic processes.

    Task 3.   Develop site independent biomass measures.

           Subtask 1.     Develop  compound  specific  biomass measures   (e.g.,
              MPN).
                                      51

-------
           Subtask 2.     Develop  general   biomass  measures  which  relate  to
             specific  measures  (e.g.,  cfu).


                               LITERATURE CITED

1.   Paris, D.  F.,  W.  C.  Steen, G. L.  Baughman, and J.  T.  Barnett,  Jr.   1981.
      Second-order model  to predict microbial  degradation of organic compounds
      in natural  waters.   Appl. Environ. Microbiol.   41:603-609.

2.   Paris, D.  F., N.  L.  Wolfe,  and W.  C.  Steen.   1982.   Structure-activity
      relationships in microbial  transformation of  phenols.   Appl.  Environ.
      Microbiol.   44:153-158.

3.   Rogers,  J.  E., S. W. Li,  and L.  J. Felice.   1984.   Microbiological
    transformation kinetics of xenobiotics in the aquatic environment.
    KPA-Grant No.  CR810436-01-0.
                                       52

-------
                   USE OF FIRST- AND SECOND-ORDER MICROBIAL
            TRANSFORMATION RATE COEFFICIENTS FOR PREDICTIVE MODELS

                                David L. Lewis
                                   ABSTRACT

           In  our  research  at  the  Athens-ERL,  we  have  assumed  that
      microbial transformations  of  xenobiotic  chemicals  are a result of
      enzymatic  reactions,  thus  following  Michaelis-Menten  kinetics.
      Based  on  Michaelis-Menten  kinetics,   we  have  used  first-  and
      second-order  transformation   rate   coefficients   for  predicting
      transformation   rates   by  suspended,   attached,   and  sediment-
      associated  microbiota.    Our  second-order  transformation  rate
      coefficients  have  been  determined  using  either total  plate counts
      or transformer enumerations for suspended organisms (1), including
      blended  aufwuchs  (2)  and  ratios   of  surface  area  to container
      volume  for  surfaces,  including sediments, covered  with  attached
      organisms (3,4,5).

                APPLICATION OF TRANSFORMATION RATE COEFFICIENTS

     The  application   of   transformation  rate  coefficients  to  predicting
transformation  rates   in   field  situations  utilizes   numerous  assumptions
relating  to  (i) the  applicability  of the  Michaelis-Menten  equation  and its
derivative equations,   (ii) the  use  of  correlative  parameters  such  as plate
counts,  direct counts,  and  colonized surface areas for  determining  second-
order  transformation  rate  coefficients,  and  (iii) the  effects of  various
environmental   parameters   on  transformation  rates.     The  strengths  and
weaknesses of  these assumptions  have  been discussed  in  a recent manuscript on
the application of single- and multiphasic-kinetics to predictive modeling for
aquatic  ecosystems  (6).   An  analysis  of  these  assumptions  has  led  us  to
conclude that  in addition to further  research on kinetics, other areas need to
be investigated to  improve our predictive capabilities.   These areas  include
(i) suppression  of  microbial  transformations  of  xenobiotic  chemicals  by
diauxie,  biologically   produced  inhibitors,  and  other  xenobiotic  chemicals;
(ii)  adaptation; (iii) mass-transport effects; and (iv)  multiphasic kinetics.


                             FUTURE  RESEARCH  NEEDS

     In  general  we  feel   that  transformation   rate  coefficients  based  on
Michaelis-Menten  kinetics  are most  applicable for  predicting  transformation
rates of xenobJotic chemicals in low environmental concentrations.  A K  value
as low  as  10~  M^ was observed for transformation  of  diethyl phthalate figure
1).   Therefore,  the maximum  concentration  of  xenobiotic chemicals  for which
microbial  transformation  rates  can   be   predicted  without  the  confounding
effects  of  saturating  some  or  all  of the  enzyme  systems may  be  as  low  as
10 _M_.  Transformation  rates  of  higher xenobiotic chemical  concentrations may
also  be  confounded  by multiphasic  kinetics,  toxic  effects of  the  xenobiotic
chemical, and adaptation.  Figure 2 shows the general divisions  of confounding
effects and applicable transformation rate coefficients  for various  xenobiotic
chemical concentration ranges.
                                      53

-------
     We believe  that  transformation  rate  coefficients may  be satisfactorily
estimated  using  SAR  (7,8,9),  or  by  using  relative  transformation  rate
coefficients compared to coefficients determined for microbial transformations
of reference chemicals (6).

     There  have   been   many  frustrations   associated   with  taking  simple
approaches to  predicting  microbial  transformation  rates.   These frustrations
have  tended  to  compel  us  toward  increasing  the  sophistication  of  our
approaches.    However,   we  believe   that  our  methodology   must  incorporate
analytical methods commonly available to industries and governmental agencies,
thereby preserving practicality.   We hope to demonstrate that even diverse and
complex environments  operate within  sufficiently  narrow  bounds and without an
inordinate  number   of   essential   principles  so   that   good  estimates  of
transformation rates can be made using relatively simple models.
                               LITERATURE CITED

1.  Paris, D.  F.,  W.  C.  Steen, G. L. Baughman,  and  J.  T.  Barnett, Jr.   1981.
      Second-order model  to predict microbial degradation of organic compounds
      in natural waters.   Appl. Environ.  Microbiol.  41:603-609.

2.  Lewis, D.  L.  and H.  W.  Holm.   1981.   Rates  of  transformation of methyl
      parathion  and  diethyl   phthalate by  aufwuchs  microorganisms.     Appl.
      Environ. Microbiol.  42:698-703.

3.  Lewis, D.  L.,  H.  P.  Kollig,  and  T.  L.  Hall.    1983.    Predicting  2,4-
      dichlorophenoxyacetic  acid   ester transformation  rates  in  periphyton-
      dominated ecosystems.  Appl. Environ. Microbiol.   46:146-151.

4.  Lewis, D.   L.,  R. B.  Kellogg,  and H.  W.  Holm.    1984.    Comparison  of
      microbial  transformation  rate   coefficients  of  xenobiotic  chemicals
      between  field-collected  and  laboratory  microcosm  microbiota.    ASTM
      (STP).    In press.

5.  Lewis, D.  L.,  H.  W.  Holm,  H.  P.  Kollig,  and T.  L. Hall.   1984.  Transport
      and  fate of  diethyl  phthalate  in aquatic  ecosystems.    Environ.  Tox.
      Chem.  In press.

6.  Lewis,  D.   L.  1984.   Application of single- and multiphasic Michaelis-
      Menton  kinetics  to  predictive modeling  for aquatic ecosystems.
      Environ.  Toxicol.  Chem.  j3(3) 	.

7.  Paris, D.  F., N.  L.  Wolfe,  and  W.  C.  Steen.   1982.   Structure-activity
      relationships  in  microbial   transformation  of  phenols.    App.   Environ.
      Microbiol.  44:153-158.

8.  Wolfe,  N.   L.,  D.  F.  Paris,  W. C.  Steen,  and G.  L.  Baughman.    1980.
      Correlation  of  microbial   degradation  rates  with  chemical  structure.
      Environ. Sci. Technol.   14:1143-1144.
                                      54

-------
9.  Paris,  D.  F.,  N.  L.  Wolfe,  and  W.  C.  Steen.    1984.    Microbial
      transformation  of esters  of  chlorinated  carboxylic  acids.    Personal
      Communication.
                                       55

-------
 CD
en
        o
                                   Time  (h)
     Figure  1.  Saturation  of Brevibacterium sp. with diethyl phthalate (DEP).
          Kffl was approximately 80 ug liter"1 (4 X 1C)-7M).
                                     56

-------
     1CT2-
 £  10~-
 o
 (D
 O
 £
 O
O
 cti
-i-j
£
'o
OH
     l(T6-d
     10"
                 Spill or
                Discharge
             Concentrations
   Optimum
  Analytical
Concentrations
             Environmental
             Concentrations
                  Microbial transformation rates are:
                  • First—order to enzyme concentration
                  • Zero—order to pollutant concentration
                  • Frequently subject to toxic effect of pollutant
                  • Frequently subject to population adaptation
                  Applicable rate coefficients are:a
Microbial transformation rates are:
• First-order to enzyme concentration
• Uultiphasic  (series of first-,mixed—,and
 zero-order) to pollutant concentration
• Sometimes subject to toxic effect of pollutant
• Sometimes subject to population adaptation
Applicable rate coefficients are:
                  Microbial transformation rates are:
                  • First-order to enzyme concentration
                  • First-order to pollutant concentration
                  • Rarely subject to toxic effect of pollutant
                  • Rarely subject to population adaptation
                  Applicable rate coefficients are:
           Transformation rate coefficients  are based on rates of  loss of parent
           compound and include kg. nag liter" * h~^ per biomass concentration
           or per ratio of periphyton-colonized surface area to volume; kj, h"1;
           kfo. liters cell"* h~*; and kA. liters m~^ n-l
       Figure   2.   Summary  of microbial  transformation rate  information
          based on Michaelis-Menten kinetics, experimental data, and
          various  assumptions used in  the Exposure  Analysis  Modeling
          System (EXAMS).
                                       57

-------
                     THE USE  OF PROGRESS CURVES TO OBTAIN
                     SECOND-ORDER RATE KINETIC ESTIMATES

                              Joseph M. Sufiita
                                   ABSTRACT

           The  second-order  rate  model   provides  a  firm  theoretical
      foundation upon which  predictions  can be made  regarding  the  rate
      of xenobiotic  biodegradation.   The  advantages  and  assumptions  of
      this  approach  are  well   known  and  already  explained  in   the
      preceeding papers.   Most  importantly the  kinetic expression  is
      formulated in terms of substrate decay which is  (and in my opinion
      should be) the parameter routinely measured.  Secondly, the theory
      is  valid  at  low  substrate  concentrations  which  is  the  likely
      scenario  in  the natural environment.   Lastly,  the  expression  is
      site independent so different habitats can be compared.

           The problems  of  this  approach are equally well known.   Most
      obvious are   the  practical  constraints  associated  with  (a)  the
      assay of actively metabolizing biomass, and (b)  the measurement of
      trace  concentrations  of  substrate.     In  this   paper  I  will
      demonstrate  an under-utilized  method of  analyzing kinetic  data
      from  biodegradation  experiments  and  show  its  application  to
      second-order rate expression.

                     INTRODUCTION TO  PROGRESS CURVE  THEORY

     Progress curves  are experiments designed  to follow the  depletion  of an
initially saturating amount of substrate  through  the  zero-,  mixed-  and first-
order  regions.    An  idealized   progress   curve  is shown  in  Figure  1.    By
appropriate analysis of progress  curves, one can obtain the various Michaelis-
Menten parameters (K  and V   )  as  well as the  first-order decay constants for
experiments  where  microbial  growth  over the course  of the experiment  is an
insignificant factor.

     Usually  the  calculation of Km and  V    are  performed  by  using  one of
several  linearizing  transformations  of the  Michaelis-Menten  equation  such as
the  Lineweaver-Burk  double reciprocal  plot.   This approach  requires  several
substrate-velocity  data  pairs  and  conclusions  are  usually drawn from  5  or 6
data points  at  best.  It  is  becoming  well  known  that the parameter estimates
using these techniques can be notoriously inaccurate.

     However, such  difficulty is  partially circumvented through the  use and
analysis  of  progress  curves.     Because  the  Michaelis-Menten  model   is  a
differential velocity equation,  its  integrated form  is  valid  over  the entire
course of the reaction.  The apparent K  and V    parameters can tse determined
by measuring  the  loss of  substrate  (or  production of  product)  several  times
during  the   course  of the experiment  incorporating  as  many  data  points as
desired.   The  information is then  plotted using  the  integrated  Michaelis-
Menten  equation  (Figure  2).    From  such  a  plot,  or  one  of  several   other
linearizations  of  the Michaelis-Menten  equation,  one  can  easily  derive the
Michael is  parameters (Km  and Vmax)  as  well  as  the first-order  decay   rate
  max  nr
                                      58

-------
             APPLICATION  OF  PROGRESS  CURVES TO EXPERIMENTAL DATA

     Figure 3 shows a progress curve for the anaerobic degradation of 4-amino-
3,5-dichlorobenzote by sediment microorganisms.   Using  the  same  data set, the
information  was  plotted  using  several  linearizations  of  the  integrated
Michaelis-Menten  equation.     The  various  equations   and   the  associated
regression coefficients are  shown in  Figure  3.   I purposely chose an equation
which did  not  give  the  highest R  ,  estimated  the kinetic parameters  K» and
V   , and  then  used a computer to simulate  a  progress curve based  on those
kinetic estimates.   The  solid  line is  computer  simulated,  and  the  data are
superimposed on  it.   As  can be  seen,  the  fit  is fajrly  good, giving  a  Km
estimate of 30ym and a Vmax estimate of 1.5  ymoles 1~ h" .

     Let's examine another progress curve,  Figure  4a,  this  time  following the
anaerobic  dehalogenation  of  3-chlorobenzote  by  an  enriched  methanogenic
consortium.  On  this  arithmetric plot we  can get  a  fairly  good  straight line
through most of  the data, suggesting that it is  zero-order.   If we are truly
in the  proper  substrate  range,  we  should  get the  curve shown  in Figure 4b, a
semi-log plot.   Note that if a  regression  line was  drawn through the  initial
data points,  we would still  get a fairly good  straight  line with  an R  of
about 0.98.  In this particular case it covers at least one order of magnitude
before the curve bends off going  to the true first-order region.  The  same is
true for  the degradation of  3,5-dichl orobenzote  (Figure 5).  The  same basic
pattern  is  seen:  again  quite  a  long  region  of  linearity  before  the curve
gradually enters the first-order region.

     We can analyze this  progress curve data and generate the Michaelis-Menten
parameters and  simulate  progress curves  based  on our  kinetic estimates.   As
can be  seen  in  Figure 6, we can estimate  K  's of  about 67  and 47  yM  for the
mono and  dichlorosubstrates  respectively  and V     values of  about  24  and 8.
The data in  this figure  are  superimposed on  the  computer simulated solid line
based on  the kinetic estimates.   As can  be  seen, the fit  is fairly good in
both  cases.    The  important  point  is  that  the kinetic  parameters  can  be
estimated from a single experiment.

     Several  questions  arise.    Can  this  approach  be  extrapolated to  more
complex  systems  other than  enrichments?  How does  it all   relate  to  second-
order rate theory?   Knowing  Km  and  Vm   we can easily calculate a first-order
decay  constant   which  is  simply  V  /K •    According  to  second-order  rate
theory,  we   need   to   normalize  tne   first-order   rate   by   the   biomass
concentration.   The  open question  is  what  do  we  use as  a  measure  of the
biomass which is actively involved  in substrate metabolism?

     In the  saturating  or zero-order  region  of  substrate  decay,  the  rate of
degradation  is  independent  of  substrate  concentration and dependent  on the
amount of catalyst.   Therefore,  the zero-order  rate  of substrate decay should
be an  accurate  reflection of  the quantity of active  catalyst  in the  system.
The  zero-order   rate  is   simply  the  Vmax term.   We  then  need experimental
verification of this model.  Ideally, we should be able to experimentally vary
biomass, run progress curves at initially saturating substrate concentrations,
and  normalize  the   observed  first-order decay rates  by the  zero-order decay
rates.
                                      59

-------
     This  can  be seen  in  an experiment  done with anoxic  sediment slurries.
The amount of biomass was  varied  by  dilution  with filter sterilized anaerobic
lake water.  Progress curves were performed on these samples with 4-amino-3,5-
dichlorobenzoate as  substrate and kinetic  estimates  were  obtained  using the
integrated Michaelis-Menten  equation.   As  can  be seen  in  Figure  7, dilution
had very  little  effect  on  the  biochemistry of the reaction;  the  apparent K
was estimated to be  about  30 yM in each  case.   However,  dilution  did have air?
affect  on  the   amount   of  active  biomass  and  therefore  the  V     of  the
reaction.  It was approximately halved in the diluted sediment slurry*.  If the
first-order  decay  rate  is  then  normalized  by  the  maximum velocity,  then
essentially the same second-order  proportionality  constant  is obtained.   This
evidence   supports   the   contention   that  the   rate   of  biodegradation  is
proportional   to  both   substrate  concentration   and   biomass  and   that  V
estimates  are a  reasonable measure  of  the concentration of  active catalytic
units.   It therefore  does  not matter if  the  active  catalyst  is  an exoenzyme,
individual cells, or a consortium of organisms.

     Table 1  illustrates  that  there  is  fairly  good  agreement  between  the
expected  and  measured  rates  of  3-chlorobenzoate metabolism  when  the biomass
content  of sediment slurries  was varied  by  dilution.   Thus, a  decrease in
biomass resulted in a corresponding drop in the zero-order rate of decay.


                             FUTURE RESEARCH  NEEDS

     In conclusion, it can  be seen that  progress  curves can be used to obtain
parameter  estimation  of the  salient  kinetic constants  including  the second-
order rate constant.  In addition, it  is  believed that the V a  of a reaction
provides a reasonable estimate of the concentration of active catalyst.

     The advantages of using progress curves are:

    a.   Kinetic estimations are derived from a single experiment.
    b.   With accurate Kp values established with a prior research effort, any
           lab  can  measure  V     and  make  predictions  at  any  substrate
           concentration; no need to go to low substrate concentrations.
    c.   Expressions  can  be  amended  mathematically   to  account   for  other
           factors  like  partitioning, heterogeniety, etc. (see below).
    d.   Conservative rate estimations can be made.

     The disadvantages of progress curve analysis are:

    a.   It may  be  impossible to get  saturation for some substrates.
    b.   There is no accounting for lag or acclimation.
    c.   Math is sensitive to "renegade"  points.

     It  is  also assumed  in  the  use of  progress curves  that a  decrease in
reaction  velocity  is  due only to  decreasing saturation of the  catalyst.   In
addition, Michaelis-Menten  kinetics  only describe substrate  consumption when
this  process  is either  (a)  unlinked  to  growth,  or  (b)  when the  amount of
growth occurring is  less  than that which  gives  sigmoidal  substrate depletion
curves.
                                      60

-------
     It is very important in performing progress analysis to:

    a.    Choose saturating initial  substrate concentrations (i.e., 2-4 K )  and
           follow degradation to first-order region (O.IK ).
    b.    Take at least 10 data points covering the range in #1.
    c.    Use the  linearized form of the  integrated  Michaelis-Menten equation
           or  nonlinear  least-squares  analysis   of  data  to  get  parameter
           estimations.
    d.    Simulate a  progress  curve based on  the  derived estimates  of  Km  and
           Vm,^ to test reliability of parameter estimates.
            max


                               LITERATURE CITED

Suflita,  J.  M.,  J.  A.   Robinson,  and  J.  M.  Tiedje.    1983.   Kinetics  of
    microbial  dehalogenation  of   haloaromatic  substrates  in  methanogenic
    environments.  Appl. Environ. Microbiol.  45:1466-1473.
                                      61

-------
                             IDEAL PROGRESS CURVE
CTi
ro
   c

   0>
   D.
   0>
N
ro
o.

-a

3
IQ

0)
(/)
to

O
C
-i

CD
             18
                        Zer°-°rder decay
nits
         ~   12-
         (D
         *-•
         CO
         u.
         +*
         (0
         X)
         3
         CO
           6-
                                                    Kinetic Parameters

                                                      Vmax= ko
                                       Mixed-order decay
                                                      First-order decay (k^
                                    30

                                Time (Units)

-------
 IDEALIZED PLOT OF INTEGRATED

  MICHAELIS-MENTEN  EQUATION



 «n S0/S)/t = (-1/Km)(S0-S/t) + Vmax/Km
CO
•»•».
 o
CO
                      Slope = -1/K
m
                                     max
                (Sft-S)/t
  Figure  2. Idealized plot of integrated Michael is-: lenton equation.
                  63

-------
                 140-1
CTi
     CO fD
     v
     Ul
     I  CO
     Q. •

     o
-s •
o n

CO O
&> fD
3 IQ

OO 01
       O
       3

       O
       i
       CU
       3
       O
                                   Un S0/S)/t

                                   t/P
                                         P/t
         a.
         >^
         W

         <
         cc

         CO
         m
         D
         CO
(-1/KmXSo-S/t)+Vmax/Km            .805

Km/Vmax  [
-------
    1200i
                  COOH
COOH
     1000
      800
  5

  >«»

  UJ
  CO
  m
  D
  CO
      600-
      400
      200
                                                       CH,
Figure 4a.  Anaerobic dehalogenation of 3-chlorobenzoate by an enriched
          methanogenic consortium.
                              65

-------
       0
10
20
3'0      40

TIME (MRS)
50
60
70
Figure 4b. The dehalogenation of 3-chlorobensoic acid by an enriched
           methanogenic community.
                            66

-------
   0    30    60    90   120   150   180   210   240   270

                          TIME (MRS)

Figure 5. The pattern of 3,5-dichlorobenzoic acid dehalogenation to 3-
          chlorobenzoate by an enriched methanogenic community.
                           67

-------
     x

     F

    O •
   CO  =1
    HI

    <
    cc

    CO
    m
    CO
 h* N.
  •  •
 CO N
 CM
          N. N.
          CO *fr
CO N
CO CO
*- o
   N
                                                                     o
                                                                     o
                                                                     CNJ
N
m «
CO 10

   CO*
                                                            o
                                                            (O
                                                            IO
                                                            CM
                                                           o
                                                           o
CO
cc
                                                                     10
                                                                     N.
                                                                     o
                                                                     to
                                                                     IO
                                                                     CM
                                3-LVdlSanS
Figure 6.   Simulated progress curves on mono and dichlorosubstrates determined

           by the kinetic parameters.
                                68

-------
          TABLE 1.
CTl
                   Zero-order rates of 3-Chlorobenzoate
                     degradation in acclimated sediment
                         diluted to various extents
Sediment             Degradation rate (p moles  i   h  )
(%)a
100
75
50
25
Actual
2.71 ±0.08
2.32 ±0.11
1.45 ±0.13
0.64 ±0.1 7
Expected
	
2.03
1.35
0.68
            a)  100% = 50mg (dry weight) ml

-------
"vj
o
       -5
       fD
3 •
10

r^ -o
3" 1
fD O

-•• -5
3 fD
<•+ 10
fD (/>

-5 O
QJ c:
n- -s
fD <
Q. fD
      o o
      3 3"

       O
       -S
       O
       cr
       fD
       3
       Nl
       O
       CXI
       r+
       fD
                140-n
               120
               100
Sediment
99
50
Km
CpM)
29.9
29.7
(AT moles •!
V1)
1.49
0.65
k1
0.05
0.022
(AI moles I )
0.0335
0.0338
3- -S
fD r+
w' ""
1 -£=•
31
*
fD DJ
3 3
H- ->•
fD 3
3 O
1
fD CO
-Q •
c (_n
O> 1
c-h Q.
HI
H
DC
H
eo
CO
CO





80-



60-





4O-
                                                     50% SEDIMENT

                                                      A
                               99% SEDIMENT
           20-
                                              200           300

                                                  TIME (MRS)

-------
                                 PANEL REPORT

     The  term  "second  order,"   previously  used  to  describe  a  particular
approach  to   site-independent   biodegradation  kinetics,  was   felt   to  be
misleading   and   should  be   changed   to    "site-independent   biodegration
kinetics."    The   term   "second  order,"  developed  from  initial  studies  by
utilizing a  second-order rate equation  derived from  Monad kinetics,  is  now
considered to be limited to  the  number of compounds  and sites for which it is
applicable.   A  site-independent  mathematical approach  should  examine a number
of potential  kinetics  expressions, letting the data dictate  which expression
is to be  used  by  testing the fit  of  these equations  by nonlinear regression.
In a practical  sense,  the  approach would use  a series of equations  for which
the appropriate constants  (Km,   V   ,  K2)  have been developed  in preliminary
laboratory  studies,  and  which  given  some  simple  measure  (active  biomass,
substrate concentration,  soil moisture,  etc.) for  the specific  site,  could
then be applied to a wide variety of environments.

     Research needs may be  divided into two major areas:

    1.   The range  of chemical structures  and habitats  studied to date has not
        been  adequate  to define the  usefulness  of this  approach.    This is
        particularly  true   with   respect  to   structure,   since  most  of  the
        compounds   for which  we have data  are degraded  initially by hydrolytic
        enzymes.    We  should  now  focus  on   compounds   such  as  chlorinated
        hydrocarbons and heterocyclics that are degraded by other routes.  Of
        general importance   is   that  the compounds  selected  do  not  closely
        resemble  common  naturally  occurring   compounds  or  other  compounds
        expected  to be  ubiquitous   in  aquatic  environments.     It   is  also
        important   that  second-order  rate constants  be  obtained  for  a  wide
        variety of aquatic,  terrestrial,  and   subsurface  environments.   The
        rate constants  should then be compared  not only within but also across
        environmental types to determine the  degree of site independence.

    2.   An important gap in our  knowledge concerns  the length  of the lag or
        acclimination  period  before  degradation  is   observed.     To  our
        knowledge,  there  is  no  firm  theoretical  or  mathematical  foundation
        that can  be  used to  predict  when a  particular  compound may start to
        degrade.   We  suggest that  research  be  directed  to  understand  the
        reasons   for  acclimation   periods    and   develop   the  appropriate
        descriptive  mathematical  equations.    It would  be  extremely  useful if
        the  site   independent  microbial  kinetic   approach could  be  used to
        predict  not  only   biotransformation   rates,   but  also  acclimation
        periods.   As an initial approach,  it  might  be worthwhile  to try to
        correlate  acclimation periods with some measure(s) of biomass.

     As general guidance we also offer the following suggestions:

    1.   The methods  used for measuring  catalyst  concentrations represent an
           area of  special  concern  in  any  study  designed  to  evaluate  the
           usefulness of the second-order approach.   We  suggest  that simple
           measures  of  total biomass  such  as   plate  counts or  ATP  levels be
           tried first.   It should be  recognized,  however,  that  measures of
           total biomass may not  be  appropriate   for  many chemicals.   Thus,


                                      71

-------
          measures   of   specific  degraders   or,   even   better,   of   degrader
          activity  are  preferred.   Possible  parameters include  V     values,
          specific  degrader MPN's,  and microbial  numbers  as  determined  by
          microautoradiography.   There  may  be other  parameters and  methods
          that  give even  better  results,  however, and  these  should  also  be
          explored.

    2.   Test  chemical  concentrations  should range  from  very low  to much higher
          levels.    Whereas  very low  concentrations  simulate  conditions  in
          many  aquatic  environments, high concentrations may be more relevant
          near  point  source discharges,  in  spill  situations,  or  in  areas
          receiving leachate from hazardous waste dumpsites.

    3.   The   analytical   regimen   used  to  measure  biodegradation  should  be
          versatile, so that (a)  the complete mineralization of the  substrate
          can  be detected and  quantified  should  that eventuality occur,  and
           (b)  parent compound depletion  and  product  formation  can be measured
          if complete mineralilzation  is not the  ultimate  fate  of the parent
          compound.  A  mass  balance  should be established in all  studies.

    4.   To estimate  variability,  one needs an  adequate number  of  replicates
          for   each  environment.    Usually   three   to  six  replicates  are
          considered adequate.

Panel  Members:   Martin Alexander,  Chairman
           Bob  Hodson, University  of  Georgia
           Dennis Focht, University  of  California
           Dennis Laskowski,  Dow Chemical
           Bill  Walker,  Gulf Coast Research  Laboratory
           Herb Fredrickson,  Georgia  State  University
           Tom Federle,  University of Alabama  in Birmingham
       FACTORS  CONTROLLING  BIODEGRADATION  RATES  IN  MICROBIAL  COMMUNITIES

Objective: To  examine  the  physical  and  biological factors  which  control  the
           rates  of biodegradation  and  discuss  how these  rate-controlling
           factors   can   be   incorporated   into   kinetic   expressions   of
           biodegradation.   To develop a strategy for future research.

     Biodegradation  studies  frequently produce  chemical  disappearance curves
which  reflect  rate limitations due  to such factors  as  changes in population
size  or  composition,  adaptation, competition  for  substrates, cometabolism,
sorption  to  surfaces,   and  inorganic nutrient concentrations.   What evidence
suggests  that  these factors  will be  important in  predicting  biodegradation
rates  in  the field?   Can  the  effect of these factors be  deciphered from the
shape  of  the  decay curve?   What experimental  techniques  will  provide more
quantitative  information  on  these  rate-controlling  factors?    Can  we assume
that  the  size  and activity  of  a  degrader  population  within  a  microbial
community is  constant  from site to  site?   Is  it important to  develop  kinetic
expressions  which  will accommodate  different  rate-controlling  factors?  What
research   is   required  to   correctly   model   these  special   aspects   of
biodegradation?


                                       72

-------
                     UPTAKE OF DISSOLVED ORGANIC COMPOUNDS
                  BY AQUATIC MICROHETEROTROPHIC POPULATIONS:
           MULTIPHASIC AND SIMPLE MICHAELIS-MENTEN KINETIC PATTERNS

                               Robert E.  Hudson
                                   ABSTRACT

           Concentrations of  individual  dissolved organic  compounds  in
      natural waters  are very  low,  in  the range  of picomolar to,  at
      most,  micromolar.    The  low  ambient  concentrations  are due  to
      efficient   uptake  of   the   organic   substrates  by  microorganisms
      possessing specific, high affinity  transport  systems.   When  these
      systems are   examined   individually,  most  are found  to  possess
      finite V    and  K,   values and,  thus,  to  exhibit simple Michaelis-
      Menten uptake  kinetics.  Since the mid-1960's,  when  radiotracer
      studies of the uptake  of dissolved  organic  compounds  from natural
      waters were  first  carried  out, and until  recently,  it  has  been
      assumed that  natural  mixed-species  assemblages of  microorganisms
      also  exhibited  simple  Michaelis-Menten  uptake  kinetics.    This
      assumption formed the  basis  for models used to predict the fate of
      both   natural  and  pollutionally-derived   organic   compounds  in
      freshwater and   marine  environments.   Recent  evidence,  however,
      indicates   that  natural assemblages  of  aquatic  microheterotrophs
      possess a  high  degree  of "kinetic  diversity"  when  presented  with
      dissolved  organic compounds.   Simple hyperbolic kinetic patterns,
      indicative of the entire population's having a single Kt and  V   ,
      are observed  less frequently than are complex, multiphasic kinetic
      patterns  indicative  of  the   presence   of   a wide   variety  of
      individual transport systems within the population.

                       INTERPRETATION OF KINETIC PATTERNS

     Wright  and  Hobbie (1966) were  among  the  first to study  intensively  the
relationships between  the concentrations  of  dissolved  organic  substrates  in
natural  waters  and  the rates at  which  these substrates were taken  up  by  the
natural  aquatic  microbial populations.    They  employed a  linearization  plot
derived  from studies of enzyme-substrate interactions to determine  graphically
kinetic  parameters  for the uptake  of organic compounds.   In this procedure  the
value t/f,  the   incubation time   (t)  divided  by the  fraction   (f)   of  labeled
substrate  taken  up  during   the   incubation,   is  plotted  versus   (A),   the
concentration of added substrate.   The  experimental  procedures  used initially
involved determining t/f over a relatively narrow range  (e.g., between one  and
ten micromolar)  of  substrate  concentrations.   The  resulting  data often fit a
straight  line  fairly  well,  a fact  that  was  interpreted as  indicating  that
uptake  was   via   a   single  transport   system.     With   the   Wright-Hobbie
representation of the  uptake  data,  the  slope of  the line  of best fit is taken
to be the inverse  of  the V   ; the  Y-intercept  is the in  situ  turnover  time
for  the  substrate,  and  the   X-intercept  is equivalent to  the  sum (Kt +  Sn)
where Sn is the  in  situ substrate  concentration.

     After  this  approach became  widely  used, the  data   were   nearly  always
treated  as   though   the relationship  between  t/f  and  (A)  was  theoretically


                                      73

-------
linear  no  matter  how  well  the  experimental  data  fit a  straight  line.   It
became generally assumed that  the microbial  population in  a particular body of
water  possessed  a  single  Kt  and  Vmax  for  uptake  of  a  given  compound.
Concentrations  of  microbially  utilizable  compounds  are  exceedingly  low.
Therefore, it was  also  assumed that the lack of  diversity with  regard to the
kinetic  parameters  was  the  result of  long-term  selective pressure  toward
acquisition of  lower and lower hL  values.   Gradually, it  was assumed, all the
organisms would  evolve  the  same low K£  (that is, the  same  high  affinity for
uptake).  These  assumptions are  themselves  based  on  the tacit assumption that
an  entire  aquatic  habitat  is uniform  with  respect  to concentration  of any
particular dissolved organic compound.   They  take no account of the potential
heterogeneity that could exist  within  a  water body  on  either a  spatial  or
temporal basis.   For selective pressure to result in only a single low Kt, for
example, the  organisms  must never  experience transiently  high  concentrations
of  substrate.   If  they  did, one  might  assume that some of the organisms would
have  opted for  exploitation   of  the  "rich  microzone" niche,  and  that this
adaptation would  be  reflected in  higher  Kt and V    values  than  those in
organisms adapted  to the ultra-low substrate concentrations of bulk water.

     Microzones  of organic  enrichment do occur  in natural  waters.  Such  zones
include  the  immediate  vicinity of  photosynthesizing phytoplankton (even those
cells  that are  healthy),  surfaces of  suspended  detrital  particles,  and the
guts  of  aquatic   animals.    These  regions  might   produce  sustained  high
concentrations  of  dissolved organic substrates.    In addition, transiently high
concentrations  might  result from the  fragmentation  of organisms such as  algae
and protozoans  when they are   fed  upon  by  zooplankton, and from the  excretion
of  feces  by  aquatic  animals.   These  considerations  lead  one  logically to
assume  that the  local concentrations of dissolved organic  compounds  in  natural
waters  may  at times greatly exceed the  overall  bulk concentration that  would
be  measured  after  filtration  (homogenization)  of a  water sample for  organic
analysis.       Instead   of   monotonously   low,   nanomolar   levels,   the
microheterotrophs  might  encounter,  at various times,  a  wide  range of  substrate
concentrations  from nanomolar  to perhaps many micromolar or  even millimolar in
enriched microzones.
                                RECENT FINDINGS

      We  have  examined  the uptake kinetics over a wide concentration  range  for
 a    number   of    naturally    occurring   organic   compounds   by    natural
 microheterotrophic  populations  in  freshwater  and marine  environments  (Azam  and
 Hodson   1981;  Kirchman  and   Hodson,   unpublished).     When  the   substrate
 concentration range used is  increased  to include  concentrations  that can  be
 expected to occur  in organically rich microzones and the data  are treated in a
 manner  similar to  that  of  Wright and  Hobbie (1966),  straight line  plots  are
 rarely   obtained.   Instead,  the  data  plotted  on the  "linearization"  graph
 produce  a hyperbola-like curve.  No single 1C or  V   can  be  derived from  the
 curves,  but rather the curves suggest that a  wide  range of Kt and  Vmax values
 are represented in the  population.   Some components of  the population appear
 to be adapted  to  using the  ultra-low  levels of organic substrate typical  of
 bulk  water,  whereas  other  components  are adapted to take advantage  of very
 high  concentrations where  and  when  they  might  occur.   The overall  effect  is
 the failure  of  the substrate uptake rate to saturate.    The  population as a
 whole can rapidly respond  to a  range  of  substrate  concentrations  that
 stretches over many orders  of magnitude.
                                       74

-------
                            FUTURE RESEARCH NEEDED

     Many   pollutants   are   taken   up   (and   subsequently   degraded)   by
microorganisms  because  of  the  similarity  of their  molecular  structure  to
naturally  occurring  substrates  that  the microbes  are  adapted  to  utilize.
Models to  describe  the rate of removal of such compounds from natural waters
have often  assumed  that  the kinetics of  uptake of the pollutant will reflect
those of the natural analog.  And the assumption for the natural compounds has
previously been that  they are  removed  via simple Michaelis-Menten kinetics by
the natural mixed  populations.  These models  predicted  that uptake  (removal)
rates  would saturate  at some finite concentration  of  the compound  in  the
water.   What  we  now  know  about  the  multiphasic  nature of  uptake  kinetics
exhibited  by  many  natural  populations suggests  that pollutant  uptake  rates
might not  saturate  but continue to  increase  as higher and higher V ax uptake
mechanisms  come  into  play.     At  this   time  the  presence   or  absence  of
multiphasic  kinetics  in  the  uptake  of   xenobiotic  compounds  has   not  been
examined.  However,  the  implications of the  possible  presence of this kinetic
pattern with regard to modeling the  fate of pollutants in aquatic environments
warrant concentrating some research  in this area.


                               LITERATURE  CITED

1.  Azam,  F.  and  R.  E.  Hodson.   1981.    Multiphasic kinetics  for   D-glucose
      uptake  by  assemblages  of  natural   marine bacteria.   Mar.  Ecol.  Prog.
      Series.  6:213-222.
    Wright,  R.  T. and  J.  E.  Hobbie.   1966.   Use  of glucose
      bacteria and algae in aquatic ecosystems.  Ecology 47:447-
2.  Wright, R.  T.  and  J.  E. Hobbie.   1966.   Use  of glucose  and  acetate by
                 '   -     -        '                -  -     	-464.
                                      75

-------
              APPLICATION OF 3/2 ORDER KINETICS FOR SUBSTRATE
                          BIODEGRADATION  IN SOIL

                                Dennis  D.  Focht
                                   ABSTRACT

           The kinetics  of  mineralization  of carbonaceous substrates is
      explained by a  deterministic model which  is  applicable for either
      growth or non-growth  conditions  in soil.   The  mixed  order nature
      (referred to as  "3/2  order")  does not  require  a  priori  decisions
      about  reaction  order,  discontinuity  periods of  lag or stationary
      phases, or corrections for endogenous mineralization rates.

                               MODEL  FORMULATION

     The differential  equation is
            HS          I
               = K,S + K SE                                   (1)
When K E is expressed as a function of time we have

          - $ = K1S + K2St
for linear growth (i.e., K'E = K2t) and

          - f = Kis + EoeVlts

for exponential  growth, (i.e., K'E = EQeut). The rate of C02 product formation

          dp -  ds + K
          dt-~dt + Ko
where K0is the zero-order rate constant for mineralization of indigenous soil
organic matter or, in  the  case  of   C0? evolution,  the rate of mineralization
from  residual product that has become part of the new humus.  Upon integration of
equations 21, 2b, and 3 and by substitution we nave

                    -K,t - (K?t2)/2
          P = SoCl-e               ] + K0t                   (4a)

                    -K,t - ~(eut - 1)
                                      ] + K0t                (4b)


for linear and exponential  growth, respectively.


                           APPLICATION OF THE MODEL

       C02 evolution  data  from  soil  were  fit  to  the above  equation  by  a
nonlinear regression  analysis  with  the  "NLIN"  program by  SAS  on an  IBM 750
computer using  the  Marquardt algorith for  stepwise  iteration.    This  program
gives a  rapid convergence  even  when the  initial  estimates are  considerably


                                      76

-------
wrong.  Linear growth (equation 2a)  is applicable when initial  cell  densities
are relatively high and do not increase more than 100 fold,  while exponential
growth (equation 2b) is applicable to low cell  densities which  increase by
several orders of magnitude in response to substrate additions.  Because the
computer program might seem formidable and not  accessible to non-mathematics,
a linear transformation of equation 4a can be used after the values for S0
and K0 are determined by a simple linear regression analysis of the zero-
order portion of the reaction  (see Figure 1).  The transformation to the linear
form:

        Y = -iq - K2t/2                                              (5a)

is then possible in which
             ln[(So - P + K t)/S ]                                   (5b)
         Y =
                      t

This equation was found to be  an adequate approximation for equation 4a if
there are sufficient data points in the linear portion of the curve to
determine SQ and K0 accurately.  (It should be noted here that  S0 is not
strictly defined as the initial substrate concentration, but rather the
fraction of substrate that is  readily converted to C02«  This redefinition obviates
the problem of accounting for  substrate which goes into humus or biomass.


                            FUTURE RESEARCH NEEDS

     The advantages of the 3/2 order kinetics to the Monod equation are as
follows:  (1) only  2  (K]_, K^)  rather than 4 (Vmax, Ks, Y, X0) interdependent
constants have to be determined by nonlinear regression analysis (in lieu of
equations 5a,b); (2) the required initial estimate can be easily oFFained
from a linearized form (equations 5a, b) rather than from an interval estimate
of a differential equation; (3) substrate or product formation  can be expressed
as an explicit function of time; (4) biomass concentration does not have to
be known.

     The immediate  question regarding the model is:  what intrinsic
characteristics do  the rate constants possess?  The zero-order constant K0
represents the indigenous gross metabolic activity of the soil.  Certainly
this value will differ among  soils, so it represents a base value:  i.e., a
soil with a higher  K0 would probably also effect more rapid degradation of
added substrates than one with a lower K0.  Thus, S0 the fraction of readily
mineralized carbon  would be dependent upon K0.  Compartive degradation rates
of any given substrate might  be made on this basis.   In the absence of growth,
the meaning of Kj,  the "first-order" rate constant becomes very clear, and
one would thus be justified in using a "half-life" expression for the degradation
rate.  The intrinsic meaning  of l<2, the "second-order" rate constant, is not
clear at this time.   However,  it is probaly more yermaine to address the
question of how one expresses  the nonlinear rate process of a sigmoidal curve.
Since we are concerned with expressing some quantitative meaning to the upper
and lower rates of  degradation, it is justifiable then to define the kinetic
parameters Vm, the  maximum rates of mineralization, and tm, the time at which
this occurs, this can be determined by taking the second derivative of equation
2a to express the change in velocity.
                                        77

-------
d
dt
2  =
          _
        dt
                   K2t dS + K2S = d2P
                                     '
and when d2S/dt2 =
                       dt
                                                                     6)
                                  dt
                   0, dS/dt = Vm, so that
               K2t
                  m
                                                                   (  7)
where Sm is the substrate concentration (i.e., the saturation  constant)
                                                                     8)
This clearly states that K2
when there is no inflection
Thus the tm value is useful
length of the achieved.

                              K^; otherwise, t[n   0,  a situation that develops
                            point (i.e., the rate is essentially first order).
                            quantitiatively in telling us something about the
     When exponential growth gives a better fit than linear growth, the KI term
has always been found to be negligible.  Thus equations 2b and 4b are reduced
to
           dt
              = Eo e
             1 - e
                                    + K
                                                                    (10)
 If we take the second derivative of equation 9 to find the inflection point
 as we did in the development of equations 6-8, we obtain for exponential
 growth
                                -Eo/
                                                                    (11)
                      Sm - V
                                                                    (12)
                           In ( /E0)
The application of the linear and exponential growth models to the mineralization
of biphenyl in soil can be seen in Fig. 1  Linear growth gives a better fit
with  inoculation, while exponential growth gives a better fit in lieu of
inoculation:  presumably this is due to the low indigenous population density
of biphenyl degraders.
                                        78

-------
       200-
         0
          0
                                                        50
Fig.  1.   Fitting of the  3/2  order  kinetic model  to CO? evolution
         from biphenyl additions  (0.33%) to  soil.  Addition of
         1Q5 cells/g of  the  biphenyl-degrader  Acinetobacter P6
         (Furukawa) (Q)  and  no  inoculation  (•)  are respectively
         best fit for linear (S0  = 17,  KI =  0.0050, K? = 0.045)
         and exponential  (S0 =  15, y  =  0.18, E0  = 0.0059) growth
         The linear portion  of  both curves  (K0  = 0.090) is the
         same as the control  (Q)  without biphenyl.  Complete
         mineralization  of biphenyl  would bring  about 240 pmoles
         C02/g.
                                 79

-------
                FACTORS CONTROLLING DEGRADATION RATES IN SOILS

                               D. A. Laskowski


                                   ABSTRACT

           In keeping  with  the workshop  theme,  this paper  attempts  to
      summarize the  needs  for characterizing  degradation  kinetics  of
      chemicals in soils.   Kinetics is becoming  increasingly important
      because of its key role  in any evaluation  of  environmental  impact
      of chemicals.   Modeling is  definitely emerging as  the premier tool
      for fate  analysis  and degradation kinetics are mandatory  in such
      exercises.

                     KINETIC ANALYSIS OF SOIL DEGRADATION

     Unfortunately  there   is  not  much   kinetic   data   available   on  soil
degradation, even though much  effort has  been  spent in  studying chemical fate
in soil.  Few experiments have been  carried  out  that quantify the statistical
variability  in  a  rate  constant  when  switching from  one  soil   to  another.
Effects of climate on the rate constant have seldom  been addressed.  It is not
known if  kinetic  data  from  soil  can be used  in  aquatic environments  and vice
versa.   Researchers  believe the  differences are  so  great  between aquatic and
terrestrial   systems  that  kinetic data  from  one  are not transportable to the
other.   But  where  is the evidence?  Do rate  constants  from an aquatic system
truly  belong  to  a  different  population  when  temperature,  moisture,  and
inherent variability within  water and soil populations are factored out?


                             FUTURE  RESEARCH  NEEDS

     Our assessment of the  progress  in  learning  about  degradation kinetics of
chemicals  is that there  has  not  been much  learned in  the  last  ten years.
There are  many questions but  few  answers.   Questions  that  we ourselves have
arrived at over the years are  summarized below:

    1.  What  effect  does   climate  (soil   temperature   and  moisture)   have  on
        degradation rate and does this effect vary  from one  soil to another?

    2.  What is the relationship between initial concentration and degradation
        rate?   Soils seem to have a finite capacity  to  degrade chemicals.  How
        much does this vary from one soil  to another?

    3.  What  influence does   time  of  contact  have on  degradation  kinetics?
        There   is   suggestion  that   chemicals   become  less  available  for
        degradation as residence time in soil increases.

    4.   Finally,  how much  variation is encountered  in  changing from one soil
        to  another?  How  is this variation distributed statistically?   Can  it
        be   reduced  by  factoring  out correlations  between  rate   and soil
        property data  that  can be  obtained readily?
                                      80

-------
     We embarked on  a  program to address some of these  questions  since  it is
not possible  to  model  fate effectively  without  this information.   Soils  are
being  collected  from  across  the  country  and  will   be  used  to  examine
relationships   between   degradation    rate   and   the   variables'   initial
concentration, soil temperature, soil moisture,  and  soil.   We  hope to have at
least some answers from our experiments over the  next couple of years.
                                 PANEL REPORT

     Environmental factors  govern  the rate and extent  of  biodegradation and,
indeed,  may  determine  whether  or  not  biodegradation  occurs.    In  addition,
environmental    factors    determine   the   differences    observed    between
biodegradation rates  in  laboratory tests and in the  field,  between  different
sites  in natural  ecosystems,  and also  at  the  same  site  but  at  different
times.   Nevertheless,  inadequate  information  exists  on  which  environmental
factors in nature govern the rate and on the quantitative relationship between
the ecologically  significant  factors  and changes in  rates.   This information
is essential  for meaningful  predictions of rates in  natural  ecosystems based
on laboratory tests and of rates in different environments.

     In  some  instances  there  is  general  agreement  on  the  importance  of
individual   environmental   factors,  yet,   even   in   these   instances,  the
quantitative  impact  of increasing intensity of that  factor  on biodegradation
rate is uncertain.  In other instances only a general perception exists that a
particular  factor may  be  significant and, in  these  cases,  the  role  of that
factor  must  be established   and  then  its   quantitative   impact   must  be
established.     Not  all  physical,  chemical,   and  biological  properties  or
characteristics  of natural environments will have a significant effect, or any
effect,  on  the  rates  of biodegradation.   Yet widespread agreement  does not
exist on which factors should be ruled out as being unimportant.

     Thus,  we  recommend  that  the environmental factors  that are  of primary
importance in governing biodegradation rates be characterized.  The factors of
primary  importance  are:   (a) temperature,  (b)  aeration,  (c) concentration of
xenobiotic,  (d)  history  of  prior  exposure to xenobiotic or related chemicals,
(e) salinity,  (f) surfaces   (primarily  in  flowing  streams),   (g) inorganic
nutrients  (primarily  in  aquatic  environments), (h) moisture  level  of soils,
and  (i)  soil and  sediment  type   (including  organic  matter  content,  caution
exchange  capacity, texture,  and clay type).   Other  factors  that probably are
of lesser importance or that are important under specific circumstances or for
particular chemicals are  (a) sorption,  (b) suspended solids (aquatic systems),
and (c) pH.

     The  significance  of grazing  on  the  biodegrading  species  by protozoa and
other  organisms,  the   importance   of  the   sediment-water   and  air-water
interfaces,  and  the  use  of rates  of community metabolism as  a  predictor of
biodegradation   rates  should  also  be  considered.     In  each   instance  the
quantitative   relationship   between   the   intensity  of   the   factor  and
biodegradation  should  be  established.    Prime attention  should be  given to
assessing which  of these  are the major factors affectiny biodegradation  rates,
which are minor,  and which have no detectable influence.
                                      81

-------
     Representative chemicals  should  be chosen  as  benchmark compounds  to be
used  in   biodegradation  studies.    They  should represent  xenobiotics  that
intrinsically have different rates of biodegradation,  and  whose metabolism is
initiated by  markedly  different types of  enzyme reactions.   These  compounds
should be chosen with care.

     To facilitate the comparison  of  biodegradation rates under  a  variety of
environmental conditions,  we feel  that  a set of  standard  reference  compounds
(representative of various class of xenobiotics)  be examined in a manner that
produces  easily comparable  data.    Using  these various  test compounds,  we
should  compare  controlled  laboratory experiments  to  environmental  studies.
Individual factors should be manipulated in the laboratory and their effect on
biodegradation  assessed.   The   natural spatial  and  temporal variability in a
geographically  and trophically  diverse  set of systems  should  be  exploited by
using regression studies to correlate biodegradation with these factors.

     Environmental   factors   may  have   different   effects   with   different
compounds.   For example, the Q-.Q for the biodegradation of one compound  may be
2, while the biodegradation of another may be 20, and thus they have different
temperature  coefficients.    Or,  with  respect  to  aeration,  the  rate  of
hydrocarbon  degradation  would  be drastically  affected  (from relative rates of
100%  to  0%  while  the   rate   of  glucose degradation  would  be  relatively
unchanged,  (e.g.,  100-75%) by  varying aeration  status.   Hence,  we recommend
that  research on  the  role of  environmental  factors be  designed to evaluate
their  impacts  on  compounds whose  biodegradation  will  probably  be  affected
differently  or  to  different extents by changes in the same factor.

     The  confidence  intervals   for the  quantitative effects of  each  of these
environmental  factors  must  be defined.   Then,  the significance of  each of
these  factors  and the  interactions  among  them must  be validated  by field
trials.   If  these  field  trials suggest  that factors other than those proposed
here  are important, these factors should  be defined  and  their quantitative
impact evaluated.

     The  apparent  lag  phase  or period for acclimation is important because it
may be a  significant part  of the time that the compound persists.  Yet,  almost
nothing  is  known  of factors  affecting  the length  of  this  phase or its cause
(i.e.,   time  for  population  growth,   induction   or  enzymes,  or   genetic
changes).    Therefore,  research is needed  to establish  these factors  and to
determine the reason for this  apparent lag phase.
                                       82

-------
              APPLICATION  OF UPTAKE  AND  MINERALIZATION  KINETICS
    Objective:  To   evaluate   the   potential   application   of   uptake   and
                mineralization kinetics  to  the  biodegradation  of xenobiotic
                chemicals by natural microbial  communities.

     The biological transformation of naturally occurring organic materials at
concentrations  typically   found  in  receiving  waters   or   soil   has  been
extensively  studied  by kinetic  analysis  of the  rates at  which radiolabeled
substrates are taken up and/or mineralized  to  carbon dioxide by heterogeneous
microbial  communities.   These  studies  are based  on  the  demonstration  of
saturation-type  kinetics  from  which  estimations   of  specific  and  maximum
turnover  rates,  half-saturation  constants,  and specific  activity  indices can
be  derived.    Only recently  has this  kinetic  approach  been applied  to the
transformation of xenobiotic chemicals.   Is this a reasonable application, and
what  new information  or  insights  can  we expect to gain  from  this  approach
which  cannot be  obtained  by  other approaches?   Can  saturation  kinetics  be
demonstrated for most  xenobiotic chemicals?   Can  specific activity indices be
used  as  a  method for estimating  the fraction  of  the  microbial  population
involved  in  biodegradation?   Must xenobiotic  chemicals  be used as  carbon
and/or energy sources  by microorganisms before these kinetic  techniques can be
applied?
Panel Members:  Carol Litchfield, Dupont
                Fred Pfaender, University of North Carolina
                Bob Larson, Proctor and Gamble
                Don Button, University of Alaska
                Roy Ventullo, University of Dayton
                                      83

-------
                 HISTORICAL BACKGROUND TO THE MEASUREMENT OF
              HETEROTROPHIC POTENTIAL IN  THE  AQUATIC ENVIRONMENT

                               C.  D.  Litchfield
                                   ABSTRACT

           My purpose here is to present an historical  background to the
      measurement    of    heterotrophic    activity   in    the    aquatic
      environment.   This will enable all of us to have  the same frame of
      reference  for  evaluating  the application  of this technique  to
      biodegradation  studies.   These applications will be  described by
      the other members of this panel.

                             HISTORICAL  BACKGROUND
     In  1951  Steeman-Nielsen  published  the  first paper  on the  use  of    C
labeled carbon  to  estimate the production  of  organic matter in the  ocean by
phytoplankton (5).   The next year  he  applied this  method to field  studies and
described the technique which  is still used, essentially unchanged since that
paper  was  published  (6).    In   1962  Parsons  and  Strickland  applied  the
radiocarbon  labeled  concept  to the  measurement  of  heterotrophic  processes
using   C-labeled organic  substrates.  They noted  that  a Hanes  transformation
plot (1) of the data derived using Michaelis-Menten kinetics usually resulted
in  a  straight line.    From the  slope and  intercepts  of that  line  they then
calculated their estimations of  heterotrophic  potential (4).   In  that paper,
however,  they  cautioned,   "the  method is  rapid and  convenient and  can give
values of relative heterotrophic potential.  As  with  the radiochemical method
for  measuring  marine  photosynthesis,  the  exact  interpretation  of  results
presents many problems...." (4) (the emphases  are mine).

     As useful as the procedure was, they  did  not recognize, however, that the
majority  of  the carbon  was not  converted to  biomass.    In  fact, it  is  not
unusual for the percent converted to biomass to account for only 10-30% of the
total carbon used.   By 1966 it  had become  apparent that a major correction was
needed to account  for  the respired CO-, and such  corrections would obviously
be  important to any  interpretation of neterotrophic potential.   Two different
approaches  to solving  this problem  were  presented  by  Kadoata  and  his  co-
workers  (3)  and  by Williams and Askew (7).  These papers  provided  the basis
for  the  publication  in   1969  by   Hobbie and  Crawford  of  their  paper,
"Respiration  corrections  for  bacterial uptake  of  dissolved organic compounds
in natural waters"  (2).

     With  this  newer  information, then,  the   procedure began  to  be widely
applied  in  microbial   ecological   studies,  and  numerous   papers  have  been
published  discussing  the  assumptions,   applications  of  the  technique  to
different   ecosystems,   and   the   problems   with   the   interpretation   of
heterotrophic uptake/mineralization studies.
                                      84

-------
                        REVIEW OF THE BASIC TECHNIQUE

     I would  now  like  to  review  very  briefly  the basic  technique used  in
measuring heterotrophic  activity,  the calculations  which  are used  to  derive
"heterotrophic potential," and the assumptions inherent or  implied in adapting
Michaelis-Menten  kinetics  to  ecological  studies.   Typically  the  tests  are
performed in an enclosed sample bottle where  the  respired  CC^ is trapped on a
suspended  piece  of  filter  paper  moistened  with  phenethylamine  or  KOH.
Determination  of  the amount  of    C02  respired is  performed using  a  liquid
scintillation counter and the resulting data  are  used  to derive  the estimates
of heterotrophic  mineralization.   To determine the  heterotrophic  uptake,  the
incubation  mixture  is  filtered,  washed,  and  the  filter  also  counted.   The
amount  of  radiolabeled  carbon  in   or  on  the cells  is  the basis for  the
calculations of heterotrophic uptake.  Several concentrations of  the substrate
and several  time periods  (usually  five  minutes to  three  hours  or  less)  are
necessary unless one already knows the in  situ concentrations of  the substrate
and the biomass of organisms able to utilize fnat  particular substrate.

     There are many assumptions inherent  in  applying Michaelis-Menten kinetics
to such studies:

    o   The substrate is present at saturating concentrations.
    o   There  is  no  significant  change  in the  substrate concentration  during
        the incubation period.
    o   There is no change in the concentration of the biomass of the actively
        metabolizing population.
    o   There are no activators required  nor inhibitors present.
    o   In the equation  below, K3 is the  rate limiting step:
    Bacteria (B) + Substrate (S)    *nr~" BS -   B + Product(s)(P)
     Although not part  of  the kinetic assumptions, other  requirements  in  the
mechanics  of  these  studies are  that:   samples must  be incubated at  in  situ
temperatures and pressures, if applicable; added substrate concentrations  must
be less than the background or natural substrate level; and the oxygen, redox,
and pH conditions should approximate in situ conditions.

     If all  of  these assumptions  and  conditions  have been met,  the  data  can
then be plotted using the equation:
                                           K, +  S
                             t/f  =  1(s)  + -t	LL
                             WT    v         V
                                   max       max

     The  resulting  linear  plot  is  shown  in   Figure  1 where  t/f  is  plotted
against  the added  substrate  concentration,   and  the  slope  of  the line  is
l/Vmax, the y-intercept  is  a function of the  in  situ substrate concentration
plus an indication of the affinity of that population for the substrate (i.e.,
a  high  K^. will  require  a  high  substrate   concentration  to  activate  the
microbial population).
                                      85

-------
     If
invalid
data.
    all  of the  conditions  and
    to apply  Michaelis-Menten
              assumptions have not
              kinetic equations  to
         been met,  then  it  is
          the  analysis of  the
                             FUTURE RESEARCH  NEEDS

     These types  of studies  can  be  performed  using  substrates  not  commonly
found  in  the  environment,  xenobiotics, or  even  naturally  occurring organic
compounds.  We have been tracking changes  in the  heterotrophic mineralization
rates  for  several  naturally occurring  materials.   The  results  with  urea are
shown  in  Figure  2  where the  amount  of urea  C mineralized  per 24  hours is
plotted  for  different  zones  in  the  New  York  Bight  apex.    Using  the
heterotrophic mineralization/potential  approach, then, we are able to estimate
or compare relative microbial activities in different marine environments  over
a prolonged time.

     If the investigator remembers the  limitations  of this method and applies
the  resulting  data  appropriately,  heterotrophic  potential  measurements can
provide a relatively rapid indication of the ability of the existing  microbial
population  to   transform   and  metabolize  organic  compounds  entering  the
envi ronment.
1.
2.
3.
4.
5.
6.
 7.
Dixon, M.  and  E.
  Academic Press.
          LITERATURE  CITED

 C.  Webb.   1964.   The   nzymes,  2nd edition.   New  York
Hobbie, J.  E. and
  bacterial   uptake
  Limnol.  Oceanogr.
   C.  C.  Crawford.   1969.   Respiration corrections  for
   of  dissolved  organic  compounds  in  natural  waters.
    14:528-532.
Kadota, H., Y. Hatta, and  H.  Miyoshi.   1966.   A new method for estimating
  the mineralization  activity of lake water  and sediments.   Memoir Res.
  Inst. Food Sci.  Kyoto Univ.  27:28-30.

Parsons,  T.  R.  and  J.   D.  H. Strickland.   1962.   On the  production  of
  particulate  organic  carbon  by  heterotrophic processes  in  sea  water.
  Deep-Sea Res.  Oceanog Abst. 8^:211-222.
Steeman-Nielsen,  E.    1951.
  matter in the sea by means
              Measurement
            of carbon 14.
of  the  production
Nature  167:684.
of  organic
Steeman-Nielsen, E.   1952.
  organic production in the  sea.
  140.
            The  use  of radioactive  carbon  for measuring
                 J.  Cons.  Perm.  Int. Explor. Mer 18:117-
Williams,  P.  J.
  mineralization
  Deep-Sea Res.
 LeB.  and C.  Askew.    1968.   A  method of  measuring the
 by  microorganisms  of  organic  compounds  in  sea  water.
15:365-375.
                                      86

-------
                            STATION 2   0-IOcm

                                  15 MR   1974
                                            J
                                                  I  I
J	I
                                                   5        10
                                                 nm  [S]
                                14
Figure 1.    Typical Hanes plot of   C-mineralization data.   The plot shown is

            for one location in the 0-10 cm portion of a core  from the sediments

            at Sandy Hook Bay, Xew York.
                                   87

-------
           Seasonal Influence on the Amount of Urea-C
                Utilized Per 24 hr Day In Different
                Zones of the New York Bight Apex
                                                    'Surface
                                                     Layer
              •fay  July
              1§T7

                               CrulM
Figure 2.   Seasonal influence on the amount of Urea-C utilized per 24 hr
           day in different zones of the  New York Bight Apex.
                               88

-------
              APPLICATION OF UPTAKE AND MINERALIZATION KINETICS

                             Frederic K.  Pfaender


                                   ABSTRACT

           Over  the last  several  years we  have  been investigating the
      application  of heterotrophic  uptake kinetics to  the measurement of
      pollutant  biodegradation.   Various versions of  the technique  have
      been  used,  with  varying  degrees  of  success,  for many  years to
      assess  the  heterotrophic  potential of  natural  aquatic  bacteria,
      using substrates  like glucose,  acetate, and  amino  acids.   The
      approach  is  based on Michaelis-Menten kinetics, and was  developed
      to  describe  the  kinetics  of enzyme-substrate reactions.   The
      application   of   this  approach  to  mixed  populations  of  intact
      microorganisms has  often been questioned.   The  saturation response
      has been shown empirically in many, but not all, environments and
      for 3 variety Of  Substrates.   The fact that saturation is obtained
      for many pollutants is taken as an indication that this kindjsf kinetic
      approach can be  applied to describing the activities of microorganisms
      in environmental samples.

      APPLICATION OF MICHAELIS-MENTEN KINETICS TO POLLUTANT DEGRADATION

     The  use  of  Michaelis-Menten kinetics to describe pollutant biodegradation
rates requires  that certain  assumptions  be  made:

    1.  For the measured kinetic  parameters  to  be   accurate,  the amount  of
       enzyme  present  during the incubation period must  remain constant.   The
       extension  of  this assumption  to environmental samples  means that  the
       numbers  of organisms  and/or  the  amount of biomass  should  remain
       constant during  the  incubation.   This, of  necessity,  will require  a
        relatively short incubation period, since it  is  reasonably well  known
       that  the numbers  of  organisms  do change while confined  in a container.

    2.  It  is also required that  the concentration  of  substrate  not  change
       significantly   during the   course   of  the  measurement.    These  two
       assumptions can  only  be satisfied  if  the measurement   is  made  over  a
        relatively short  period  of time.  Long-term  incubations will result in
       both an increase in the  number  of  organisms  present and a decrease in
       the  concentration  of  the  substrate,  due  to  its  breakdown  by  the
       microorganisms.

    3. The  assumption  is   made  that  the  transport  systems  present  are
        responding only  to  the  compound added as substrate, and  not  to other
        compounds  present in the environment.

    4.  In  our particular case,  we make the assumption  that  concentration of
       the specific pollutant  substrate present in  the  environmental  sample
        is   negligible  compared  to  the amount  we   add  as  substrate.    This
        assumption allows us to  disregard  the  natural substrate concentration
        in  the  calculation of the kinetic parameters.
                                      89

-------
     We feel  that the  use of  this kinetic  approach  to measuring  pollutant
blodegradation has several  very positive features.   From past  experience  in
our  laboratory,  we know  that  confining  samples within  a  bottle  results  in
significant  changes  in   both   the  size  and  activity  of   the   microbial
community.   We also  know  that there  is  a period  of  approximately 12  to  16
hours before  these  changes become significant.  Therefore,  if biodegradation
rates can  be  measured during this  time  period,  a  rate can  be generated that
might be  quite close to  that  which  would  occur under  natural  environmental
conditions.    Since   the   assumptions  required  for  use  of  this  technique
necessitate short-term incubations, this  methodology appears  quite appropriate
for  answering  questions about  real world  rates.    In  addition, the  kinetic
approach  offers   the  advantage  of being  able  to  calculate  several  useful
biodegradation' parameters.  These include V    , the maximum potential
velocity that  the  indigenous  community canrafl:tain,  Km  or the  half-saturation
constant  which  gives  an  indication of the  concentration range  to  which the
community  is  adapted.   In  addition, a rate constant, which  we have called K,
and has the same  units  as  a first-order  rate  constant,  can be calculated from
the  kinetic  data in  several  different ways.   These include  the dividing of
V     by  Km,  or  taking  the slope  of  the  linear portion of the velocity vs.
concentration  plot.   Two  additional non-Michaelis-Menton parameters can  be calculated
from the data generated in these experiments.   The velocity of metabolism witK
any particular added  concentration  can be  obtained  from the  data generated by
multiplying  the   concentration  added   by  the  f/t  factor.   In  addition, the
turnover  time  for the  added  xenobiotic at any particular concentration  can
be measured directly  from  the  linearization of the kinetic data.


                    INTERPRETATION OF EXPERIMENTAL RESULTS

     We  can  calculate  several  parameters  from  the  uptake  and  mineralization
data, but  have questions  about which  one  is  most  useful.  V     may represent
the  maximum  potential velocity,  and  when  we look  at  the saturation plots we
see  that  for  some compounds saturation occurs  at  what  may be environmentally
realistic  concentrations  (cresol  Figure 1, and  chlorobenzene  Figure 2).   For
other compounds,  however,   (NTA)  many  hundreds  of mg/1  are  needed, which are
concentrations  far   in  excess  in  what  might   actually occur.   We   began
calculating K,  because  it  is derived  from  the  slope of the  linear portion of
the  curve, generally  in the  lower part  of the  concentration  range.   We  felt
that  it  gave  a realistic  additional measure of biodegradation.  When we
do   correlations   of   biodegradation   rate   measures  and   environmental
Characteristics   (Table 1),  V     gives  relationships  that make sense.   In
coastal  environments  where nutrient  concentrations  are  low,  we get a  strong
relationship  with nitrogen and phosphorous species  and chlorophyll, which is
the  principal  source  of nutrients.   In  the lake and river environments  where
concentrations   of    nutrients   are  many  times   higher than  the   coastal
environments,  nitrogen  and  phosphorous do not appear to  be strongly  related to
biodegradation  rates.  Although not shown  in  the  table, when K, and Vmax are
correlated with  one  another and with  environmental parameters,  they generally
Correlate  well with one another and with the  same  characteristics of the
environment.   While it  is  reassuring that these  two  measures of  biodegradation
appear  to be  describing similar  processes, it  does not  help  a  great deal in
determining   which  one  gives  the  most  realistic  or  useful   measure  of
biodegradation.   Since both  are obtainable  from the  same  set  of  laboratory
measurements,  it  is always  possible to calculate and report  both.


                                       90

-------
                  POTENTIAL  FOR  OBSERVING  SATURATION KINETICS

     We   have   always   been   able  to   obtain  saturation   if   appropriate
concentrations are  used.   As  mentioned  previously,  for some  compounds  this
requires  higher  concentrations  than are  likely  to  occur in  the  environment.
We are also aware of several reports in the literature of saturation not being
obtained  with many  natural  substrates.   Since we have  looked at  only four or
five pollutant compounds, in only  a few  environments,  it is difficult to make
generalizations  about  many different  classes  of  materials.   This  raises
questions  about  the conditions  under  which  saturation  may be obtained.   In
attempting to shed  light on this  question,  we have  used several techniques to
estimate  the  number of  degraders  of  specific  compounds.   Table 2  shows  the
results  of this  evaluation in  some  fresh water  samples  using  two methods,
plate counting and microautoradiography.  The important point of this table is
that the  specific degraders constitute  a relatively small  part  of  the total
community.   I  believe the   reason we are  able to obtain saturation  is because
this small community behaves in  a  manner more similar to a  pure enzyme system
than would the  larger, more diverse community that might be metabolizers of
naturally  occurring  substrates.   Since the technique we use  was  developed tc
measure the  activity  of pure  enzyme systems,  I  do  not  consider it  surprising
that we  are  able to obtain saturation when working  with a  very small part of
the  total assemblage  of microorganisms  present in an  environmental sample.
These questions, however, are still far from being answered.
                         EFFECTS  OF INCUBATION  PERIODS

     The  uptake  and  mineralization  technique  that we  have  used  to  assess
pollutant  biodegradation  requires  that  the   measurements   be  made  over  a
relatively short period  of time.   During  this  short  incubation  period only a
small proportion of the  label  added is  metabolized, usually from one to three
or  four  percent.   This  obviously  raises the question of  whether  this short-
term  rate can be  extrapolated  to the  longer  periods  of time necessary  to
achieve  complete  breakdown  of the  pollutant.    In an  attempt  to  gain  some
understanding  of  this  we  have  conducted  time course  experiments  in  which
extended  incubation  periods were  used  and  the amount metabolized  over  time
measured.    Table  3  shows  the  results  of  this  study   for  two  compounds,
chlorobenzene and meta-cresol.   We have  used KI to predict the percent of the
compound that would be degraded at different time periods and compared this to
what was observed in the time course study.  We have also used V    to predict
how many  micrograms  of  the  material  would have been  metabolizeci  at the  same
time periods  and  compared this  with  the micrograms actually converted to coz
and cells.  As you  can  see,  for  both compounds and both methods of estimating
biodegradation, there is good agreement between the two methods for an initial
period of several days.   However,  after  five to seven  days confinement in the
time  course   studies,  the  amount  of  degradation  slows  down,  yielding  an
apparent underestimation of the amount of pollutant that would be biodegraded,
as  predicted  by  the  kinetic  approach.    It is  interesting  to note  the  good
agreement  between   the  V     prediction  and   the  micrograms   actually
metabolized.   This would  tend  to indicate  that  in the  containers used for the
time course study the community  is functioning  at  close  to V   .   Since we do
not have  information  on  the long-term  degradation of these  compounds  in the


                                      91

-------
environment tested, we have no way to  determine which  method is providing the
most accurate  estimate  of persistence.   However,  good agreement  between  the
two techniques during the early stages  of Incubation and the fact that changes
are known  to occur  in  the container  after  some  period of  confinement would
tend to point to the kinetic approach,  giving a more realistic estimate of the
amount of time necessary for biodegradation.


                             FUTURE RESEARCH  NEEDS

     We are  relatively  confident that  the uptake  and  mineralization approach
has  potential  as  a  rapid and  useful  measure  of  pollutant  biodegradation.
However,  additional  research  needs  to  be  addressed  to   several  pertinent
questions:

    1.  To how broad  a  range of  compound and environment can this approach be
        successfully  applied?   At this  point we  have examined  four or five
        compounds  in  three coastal  and  four freshwater environments.   These
        sample sites  are  all in  one  geographic  part of the  United States, and
        certainly  do  not  represent the range of environmental characteristics
        that might be found throughout  even the United States.               f

    2.  Under  what environmental  conditions does  saturation  occur?   While we
        know that  it occurs for the compounds and environments we  have  tested,
        neither   represent  the  range   of   materials  or   sites  that  are
        available.    The  methodology   offers  the  opportunity  to   calculate
        several  degradation  parameters  (K,,  specific  velocities and  turnover
        times) even if saturation is not obtained.

    3.  How  valid  are extrapolations based on the uptake of only  a  small part
        of  the added compound?   In the  comparisons we have  done with time-
        course  studies,  the extrapolations  looked to  be  reasonably good,  at
        least  for  a period of several  days.   One  of  the  problems we  face  in
        testing  these extrapolations  is  in  determining what  to compare the
        extrapolations  to.   While  time-course studies  are extremely  useful  in
        determining   whether   a   compound  is  degradable  and   required  for
        identifying   degradation   products  and  mechanisms,  their   use  for
        estimating environmental  rates of biodegradation appear  questionable,
        but  there  are few  other ways to  do comparisons.

    4.  Does  adaptation  occur  at the low environmental concentrations  of most
        pollutants?    This  is  a  question   that   relates  not  only  to this
        technique,  but  to all  of the methods that  we  are discussing.   What  we
        really  need  to  know  is  whether  the  low  concentrations   of those
        pollutants that we  find  in  most  environments constitute a  selective
        pressure  for  the community of microorganisms  that  are present.   This
        raises  additional  questions  of  how  do  we  measure  adaptation  and  what
        factors  in  the  environment  might  be  influencing  whether  or not
        organisms  can adapt.

    5.  Why  do metabolic  rates  change  when  natural communities  are  confined?
        Are  the observed  changes  in  rate  due to  alterations in the  biomass
        present,  alterations  in  the species composition or  organisms present,
        or  the result of  changes in substrate  concentration that result  from
        microbial  activity.


                                       92

-------
Is the assumption  that  the microbes are  responding  to  only the added
substrate valid?   This question  is  important  for an  evaluation of the
uptake and  mineralization  technique,  since one  of the  assumptions we
make  is  that the  organisms respond  only to the  compound added.   I
think,  however,  this  question   has   broader  interest   in  terms  of
evaluating  the  specificity of the  enzymatic  mechanisms  available for
the degradation of pollutants.  It is quite well known from laboratory
studies that many  organisms  have  transport  systems  that function with
several  different  substrates.   In environments where  nutrients  are
scarce, it  would be  to  the organism's  advantage to be able to utilize
a  range  of compounds  with  as  few  enzymes  as  possible.    In  our
particular  case,   this  is  significant  because the  kinetics  may be
influenced  by  not  only the  substrate  we add,  but  by other compounds
that may be present in the environmental sample.
                              93

-------
       0.40--
                                                                                               . .  400
S-
o>
01
3.
QJ
       0.30--
0.20--
                                                                                                   300
                                                                                                   200
                                                                                                    T/F (
ra
4->
O
0.10..'
                                                                                                   100
               	1                   I                  I	h-
                            30                 60                90               120

                                              Concentration (yg-liter~ )


              Figure  1.  M-cresol uptake by Jordan Lake microbial community on 6/22/82,

                        (A—A) - velocity, (•— •) - linear transformation.

-------
    80 -
    60  -
    40  •
s-
CD
    20
                        15
30
                                                            -1
                                                                                    800
                                                                                 .   600
                                                                                  .  400
                                                 T/F
                                                 (h)
                                                                                     200
                                     Concentration (yg-liter  )


     Figure 2-   Total uptake of ch'borobenzene by Lake Michie microbial  community on 11/2/82,

                (A—A) - velocity, (D-D) - linear transformation.

-------
Table 1.   Correlation of nutrient concentrations and metabolic parameters
                      Coastal Systems
Lakes
Rivers
m-Cresol NTA ? Max Ami no

N03 + NO -N
NH4
P04
Tot Part. Wt.
Chlorophyl 1 a
+ = Correlated
* = Correlated
** = Correlated
*** = Correlated
Vm= vm,~ Uptake Acid Met
max max
.82*** .81** .76*** -.08
.69** .44 .59* -.22
.74** .80** .67** -.16
.08 -.11 .01 -.19
.85*** .06 .62** -.20
at 90? level
at 95? level
at 99? level
at 99.9? level
? Max Ami no ? Max Ami no
Uptake Acid Met Uptake Acid Het
.23 -.36 -.18 .07
.49 -.36 .11 .06
.49 -.39 .16 .01
- _
_ _




                                        96

-------
Table 2.  Enumeration of specific degraders for m-Cresol
          in North Carolina rivers and lakes
                                              Sample  Source
Characteristic                  Haw          EnoMichie      Jordan
Total heterotrophs
(CPU x 104/ml)                23.0        14.5        13.8         9.5

m-Cresol  degraders
(CPU x lOVml)                 3.30        1.36         .30        4.0

% m-Cresol degraders            .14         .09         .02         .42

Acridine Orange   fi
Direct Count (x 10b/ml)        4.3         5.7         8.2        16.0

m-Cresol  degraders microauto-
radiography (x KT/ml)        18.5        25.0        30.1        14.3

% m-Cresol degraders           4.3         4.3         3.7         0.89
                                      97

-------
Table 3.   Comparison of persistence predicted by  heterotrophic uptake and time
          course studies with Jordan Lake
                                % Metabolized                 ug Metabolized
                   Time         ~T7v"
Compound          (days)      Prediction   Observed        Prediction   Observed
Chlorobenzene        1           5.5        3.5-6              .033   .03- .07
                     4          22.1        18-39              .134   .06-.186
                     7          38.6        41-52              .235   .09-.237
                    14          77.2        32-57              .470   .12-.175

m-Cresol            .5           8.4         .5-8              .09   .037-.087
                     1          16.8         2-16              .18    .082-.21
                     4          67.2        50-62              .72     .34-5.6
                     7         117.2        45-62             1.26     .34-3.7
                    14          235         45-62             2.52    .34-3.87
                                      98

-------
   HYDROCARBON BIODEGRADATION:  KINETIC CONSTANTS AND THEIR APPLICATION

                O.K. Button, K.S. Craig and B.R. Robertson

     We have examined the persistence of hydrocarbons in aquatic environments
with focus on those hydrocarbons in the dissolved phase as they are
transformed and biodegraded by bacteria.  Our main experimental approach
has been to follow radiolabeled hydrocarbons through both laboratory and
environmental systems with data interpreted from biochemical and kinetic
points of view.  Our main interest is in the biological principles of
hydrocarbon transport mechanisms and kinetics.

     Early work led to the development of a sensitive technique for
following the rate of C0£ liberation from hydrocarbons (Button et al.,
1981a).  We were then able to locate sites of high degradative activity.
Rates were fast in Port Aransas, Texas, near industrialized Corpus Christi
(manuscript in preparation).  In Port Valdez, Alaska, rates were even
faster in a thin layer at 50 m depth due to ballast water discharge
(Button, et al., 1981b).

     In Resurrection Bay, Alaska, rates of toluene metabolism were
initially slow, but could be induced to higher values, perhaps due to
a microflora subsisting on conifer-derived terpenes (Button, Appl. Environ.
Microbiol., in press).  By exposing seawater samples to a range of toluene
concentrations, we found that induction proceeded at its maximal rate
after 48 h and at half its maximal rate (Kinc|) at a substrate concentration
of 1.9 ug per liter (manuscript in preparation).

     Microautoradiographic techniques, together with epifluorescence
microscopy, gave the original population of toluene oxidizers in this
pristine fjord.  About 8% of the bacteria (27 yg/liter) had the capacity
to metabolize toluene.  The process of induction raised the specific
affinity for toluene (aA) of this natural  population from a base of 4.2
liters/g cells x hours to 11.6.

     Induction raised the specific affinity of the marine isolate
Pseudomonas T2 from a base of 0.03 to 500.  The saturation constants kj
for natural assemblages of bacteria in seawater and for Ps.  T2 were
found to have values of 2 and 44 yg/liter, respectively, about the same
as the concentrations that gave half-maximal rates of induction, K-jnc|.
This similarity between Ky and K-jnc| is explained if toluene metabolism
proceeds according to constitutive systems with the above values of Kj
and then induces additional  enzyme (such as toluene dioxygenase) at the
rate-limiting step.  Then induction and transport are controlled by the
same process.  A peculiarity of this system is that while the values for
specific affinity are quite usual, the values of Kj are very small
(manuscripts in preparation).

     The formation of large quantities of metabolic products (type
compounds shown in Fig. 1) appears to be a phenomenon only of very dilute
microbial populations (Fig.  2), such as those in seawater.  The recycling
step of product reaccumulation by the bacteria is subject to different
types of transport systems,  perhaps of different classes (see below) and
                                    99

-------
certainly with different kinetics.  This phenomenon of dilute solutions
necessitates a correction in the basic assumption that the kinetic
constants for biodegradation are biomass-independent.  The biomass dependency
can be demonstrated by numerically simulating steady state systems using
a continuous culture model and solving for biodegradation rates using the
kinetic constants for toluene utilization, together with the product
formation and re-utilization kinetics determined from 3-methylcatechol
metabolism experiments.

     Basic formulations for biodegradation kinetics and their relation  to
Monod kinetics are shown in Fig. 3 (Button, 1983).  Also shown is the
relationship between the kinetics and enzyme content of the organism at
the rate limiting step.

     Consideration of these kinetics has led to a model for the active
transport of hydrocarbons called vectorial partitioning.  It is only the
fourth transport mechanism known and is shown in Fig. 4.  Recent inhibitor
data support this mechanism.

     The understanding of biodegradation in aquatic systems has increased
dramatically over the last few years.  Four areas in need of additional
work are:

     1)  Concentration-dependent kinetics and specificity of induction,
particularly the extent to which hydrocarbons induce low-level metabolic
activity in the bulk of the natural  aquatic microflora when the inducers
are sustained at very low concentrations over long periods.  These data
are useful because they reflect the  concentrations of a pollutant which
aquatic organisms are experiencing but which are rather difficult to
measure.  Also data define the concentrations of one hydrocarbon that
would be expected to assist in priming a system for the biodegradation  of
another.

     2)  The nature and kinetics of  ampholite degradation.  While apparent
Michael is constants for the transport of hydrocarbons seem to differ
radically from those for oxygenated  substrates due to different mechanisms
of accumulation, many anthropogenic  hydrocarbons as well as their products
of biodegradation are only mildly polar.  Little is known of the decomposition
kinetics of these compounds as they  move from recognition by non-polar
transport mechanisms to being accumulated by transport systems for polar
substrates.  For example, if an ampholite is accumulated by vectorial
partitioning, it may be of little use to study its degradation at
concentrations above nanomolar levels.

     3)  The chemical  nature of organic products formed from hydrocarbon
biodegradation.  Apparently hydrocarbon biodegradation, as studied in the
laboratory, contains metabolic loops which proceed outside the cell.  In
natural aquatic environments the biomass is too low for product
reaccumulation to occur, so these products are lost, generating a range
of unusual organic chemicals which can only be biodegraded with what may
be rather low-affinity transport mechanisms.
                                    100

-------
     4)  The mechanism for active transport of non-polar substrates.  If
vectorial partitioning is the main mechanism for hydrocarbon transport,
knowledge of this mechanism is useful  for accurate anticipation of how
the process will proceed for new compounds without having to test each alone,
                                 101

-------
REFERENCES

     Button, O.K., Schell, D.M., and Robertson,  B.R.,  Appl.  Environ.
Microbiol.  42, 936-941 (1981a).

     Button O.K., Roberts, B.R., and Craig,  K.S., Appl.  Environ.
Microbiol.  _42,_ 709-719, (1981b).

     Button, O.K., Trends Biochem.  Sciences  8,  121-124 (1983).
                                    102

-------
                            3 -Methyl
                            catechol
                                            <*-Hydroxy-
                                               keto  acid
         o-Cresol
1.  Chemical nature of products formed from toluene by Pseudomonas T2.


0)
o
o:
c
o
•^
0
E
L_
.c
k_
Q>
V.
*Q>
C
Q>
_3
O
4


3



2

                       100
                                  200
300
                     Toluene (A), ^g/liter
2.
    Formation of organic products (P) and carbon dioxide (Q) as a function
    of initial toluene concentration by Pseudomonas T2.
                             103

-------
                       Polar Substrate



                       Non Polar



                       Michaelian
                      C = k —
                             X
                           V
                                        a   =
 X


max

KT

 V
                                                max
                                                    - v
                                                 K
                                                   T
                                        J_

                                         X
                       Monod            //     = V
                                        m 3 x    m 3 x
                                             	  i/
                                          M  ~  KT

                                        Y   = constant
3.  Formulation of biodegradation  kinetics based on  the specific affinity and
    second  order kinetics.
                  Chemiosmotic
                             A,
   Group
Translocotion
                  Ligand Taxi

                  Cout
                                  /
                                      .....

                                   'in
                   Vectprial   Dout-
                  Partitioning      \
                         -OH-^DOHin

4.  Model  for transport of  hydrocarbons into bacteria  by vectorial  partitioning
    together with other models  for active transport.
                                        104

-------
                A HETEROTROPHIC BIODEGRADATION POTENTIAL (HBP)
                 APPROACH FOR PREDICTING BIODEGRADATION RATES
                             IN  NATURAL  ECOSYSTEMS

                               Robert J. Larson
                                   ABSTRACT

           Numberous investigtors  (1,  4,  5,  10, 12-14)  have  shown  that
      the rate of  uptake and/or mineralization of  radiolabeled  natural
      substrates  in  aquatic systems  can   be  described by  a  hyberbolic
      enzyme-saturation  model   of  the   Michaelis-Menten   type   (9).
      Recently,  saturation  kinetic  models  have  also  been  applied  to
      biodegradation  of  various   xenobiotic  compounds  in  ground  and
      surface  water  systems  (6,   11).   The application  of  saturation
      models to xenobiotic biodegradation  has important implications for
      environmental  fate  research.   These  models  directly  relate  the
      rate of biodegradation of a  specified substrate to the activity of
      the degrading  microbial  population.   As such, they  may be useful
      in  normalizing  microbial  numbers   for  microbial  activity  and
      extrapolating  biodegradation rate data to different environmental
      systems.
                                    THEORY

     Tl]fi general equation  relating  the  rate  of uptake (and/or mineralization)
of a    C-labeled  substrate to the  concentration of  added  substrate (when the
background substrate concentration  is unknown) can be written as:
                        K + (Sb + A)                                      [1]
where v is the degradation  rate  (ng/liter/h),  V    is the maximum degradation
rate (ng/liter/h),  Su  is  the background substrate concentration (yg/liter), A
is the  added  substrate concentration  (pg/liter) and  K  is the half-saturation
constant where v = 0.5 V     (ng/liter).   When the initial degradation rate is
constant, v can also be defined as:


                   v =  f_ (Sb + A)

                        t                     -                            [2]
                                      105

-------
where  f  is  the  fraction  taken  up  and/or  mineralized  at  time  t  (h).
Traditionally,  equations  [1]   and   [2]  have  been  combined,  inverted,  and
rewritten as follows to estimate kinetic constants (8):


                   1 (Sb + A) .   Vmax (Sb + A)

                   t              K + (Sb + A)                            [3]


                   '  =     "may
                   t     (K + Sb) + A                                     [4]


                   1 =   l/Vmax (A) + (K + Sb)

                                       [5]
     A  plot  of  t/f  vs.  A  (equation  [5])  yields  a  straight  line  with  y-
intercept T  ,  x-intercept  (K + Sb), and a slope  of  1/V     (Figure 1).  Tp is
defined as the time  required for  complete  degradation  or a chemical at its in
situ  concentration,   V     is  the  extrapolated   maximum  degradation  rate  ~aT
infinite substrate concentration, and  (K + S, )  represents the sum of the half
saturation  constant   and  the  background   substrate  concentration.   Although
unweighted  linear  regression of t/f vs. A  (equation  [5]) is commonly used to
estimate  kinetic  constants,  direct  nonlinear   regression  analysis  of  the
uninverted  equation  (equation  [4])  can   also   be  used  to  estimate  these
constants.   Nonlinear techniques offer  special advantages over standard linear
techniques,  in that  they  avoid  many  of  the statistical  biases  inherent  in
linear transformations of nonlinear data (8).  They also  give better estimates
of  kinetic   constants  than  linear  techniques  and are   readily  available  to
anyone with  access to a microcomputer.

     Equations [4] and  [5]  have traditionally been  used to determine kinetic
constants when A > S. .  A special situation occurs, however, when  S^ is either
much greater than  A  or so low  as  to be considered negligible.   When  Sb » A,
turnover  times can be calculated  directly from  the ratio  t/f  using a single
concentration  of  A.   The  concentration of added  substrate  is  so low that it
does not  influence the in  situ degradation  rate  (3),  and  observed Tn values
represent  the actual  turnover  time  for  the background  level  of substrate
present.  Since  only one  substrate concentration is tested  by this technique,
however,  V  ,   and  K  + $b  values cannot be  calculated.   When  Sb «  A,  as would
be  the  case when testing xenobiotics  in previously  unexposed systems, $b can
be  considered zero for all  practical  purposes  and ignored when combining [1]
and [2].  Under these conditions, the  following equation  can  be written:
                    v  =  1.  A=    Vmax  •  A
                        t          K1  =  A                                    [6]
                                      106

-------
where the  half  saturation constant is written  as  K'  to differentiate it from
the  half  saturation  constant  (K)  of  equation [5] which  includes  an  S^ term.
Strictly speaking, turnover  times  have no meaning when  S^j-O.   Therefore, in
systems where the in situ concentration of xenobiotic is negligible, T  values
cannot be calculated, although a ratio t/f may be obtained.  However, equation
[6]  can  be used  to  determine the  empirical  kinetic constants, V    and K'.
These  constants  indicate  the  "potential"   for  biodegradation   of  specific
xenobiotics in  natural  ecosystems  and  thus  have use  in  defining  the overall
heterotrophic   biodegradation   potential   (HBP)  of   the  natural  microbial
population present.   When some background level  of  the substrate  is present,
T  can be calculated from equation  [4] by the relationship T  = K"/V   , where
K" = K + Sb (8) and Vmax has units  of ug/liter/h.


                        GENERATION  OF  KINETIC CONSTANTS

     The ability  to  generate HBP  kinetic  constants has important  implications
for  biodegradation testing of  xenobiotics.   Use of these constants with other
measurements  of  microbial  biomass such  as  colony   forming  units,  CPU  (2),
acridine-orange-direct-counts,  AODC   (15),   and  most-probable-number   (MPN)
determinations  (7) should  allow microbial  numbers and degradation  activity to
be correlated  directly  via  a  specific activity index  (SAI).   The concept of
SAI was initially proposed to characterize the activity of individual cells on
natural  substrates   (13).    Turnover  ratee  (T  = i/Tr.)  or  vmax  values  were
divided  by bacterial  numbers  (CPU,  AODC)  to yield  Tp  or  V     SAI's.   In
principle,  a   similar   approach   should   be  applicable  to  degradation  of
xenobiotic  compounds,   if  these compounds  are  mineralized  as  carbon  and/or
energy sources.

     Figures  2,  3,  and  4  illustrate  the  results  of  HBP  studies  on  three
detergent   chemicals,    linear    alkylbenzene    sulfonate    (LAS),    sodium
nitrilotriacetate  (NTA)  and  dodecylnony1ethoxylate  (CjoEg)   in  ground  and
surface  waters.     All   three  materials   were  testea  over  a  range  of
concentrations  and  rate  data   (v =  f/t  .  A)   and   were  fit   by  nonlinear
regression techniques to equation  [6]  to  estimate the  kinetic  parameters K'
and  vmax*  Vjnax was  then dlvided  b^ the number of viable bacteria (CPU/liter)
to generate Vmax  SAI's in the different environmental samples tested.

     As  the  data  in  Figures 2-4  indicate,  biodegradation rate data  for all
three  materials  are accurately   described   by  a   saturation  kinetic  model
(equation [6]).   Precise estimates  of the kinetic constants K1 and V    can be
obtained  (Table  1),   even  though   maximum  degradation   rates  vary  for  the
detergent chemicals and  natural water samples tested.  Moreover, normalization
of Vj7iax  for  CPU  resulted in  SAI  values  that  agreed  relatively  well,  both
within  and between compound   classes.     This   good   agreement  suggests  that
consistent kinetic results can be  obtained  if degradation data  are normalized
for both microbial numbers and microbial activity using an HBP/SAI  approach.


                             FUTURE  RESEARCH  NEEDS

     HBP  kinetic   approaches  which incorporate  both  microbial   numbers  and
microbial  activity  measurements  (SAI)  appear  to show  promise  for predicting


                                     107

-------
biodegradation rates  of  radiolabeled xenobiotics  in  natural ecosystems.   In
theory, these  approaches  should  be  applicable to  die-away  studies  in  which
disappearance of  parent  compound is measured  by  specific  analytical  methods.
More  research  on  different   xenobiotic  compounds  is  needed,  however,  to
establish the  applicability of  HBR  measurements to  analytical  studies.   In
addition, several  important  research  areas  need to be  addressed  to determine
the predictive value of an HBP/SAI approach in estimating biodegradation rates
in the environment.  These areas include (but are not  limited to)  the:
    -  effect of incubation period on kinetic measurements,

    -  significance of T  value when S^ ^0

    -  appropriate biomass measurements for generating SAI values,

    -  and comparability  of  rate data derived from  short-term HBP assays vs.
       long-term biodegradability die-away assays.
      Information  is  needed  in these areas not  only  to formulate a conceptual
framework  for  using microbial  activity measurements  in  environmental   fate
research but also to establish a practical basis for predicting biodegradation
rates of xenobiotics in natural ecosystems.
                               LITERATURE CITED

 1.  Azam,  F.  and 0.  Holm-Hansen.   1973.   Use of  tritiated substances in  the
       study of  heterotrophy  in seawater.  Limnol.  Oceanogr.  23:191-196.

 2.  Buck,  J.   F.    1979.   The  plate count  in  aquatic  microbiology.    Native
       Aquatic  Bacteria:   Enumeration,  Activity  and  Ecology,   ASTM  STP  695,
       J. W. Costerton and R. R.  Colwell, (Eds.),   American  Society  for  Testing
       and  Materials,  pp.  19-28.

 3.  Gocke,  K.   1977.  Comparison of  methods  for  determining the  turnover  times
       of dissolved  organic compounds.   Mar.  Biol.  42:131-141.

 4.  Hobbie,  J.  E.  and   C.  C.  Crawford.   1969.    Respiration corrections  for
       bacterial  uptake  of  dissolved  organic   compounds   in  natural   waters.
       Limnol.  Oceanogr.  _l£:528-532.

 5.   Ladd,  T.   I., R.  M. Ventullo,  P.  M.  Wallis,  and  J.  W.  Costerton.   1982.
       Heterotrophic  activity  and   biodegradation  of  labile  and  refractory
       compounds by  groundwater  and   stream  microbial  populations.    Appl.
       Environ.  Microbiol.  44:321-329.

 6.   Larson,  R.  J.    1983.   Kinetic  and  ecological  approaches  for  predicting
       biodegradation   rates   of  xenobiotic  organic  chemicals   in   natural
       ecosystems.   In M. J.  Kluq and C. A.  Reddy  (Eds.),  Current Perspectives
       in Microbial  Ecology.  American Society for  Microbiology,  pp. 677-686.
                                      108

-------
7.  Lehmicke,  L.  G. ,  R. T.  Williams,  and R.  L.  Crawford.    1979.   14C-most-
      probable-number   method  for   enumeration   of   active   heterotrophic
      microorganisms  in  natural  waters.   Appl .  Environ. Microbiol.   38:644-
      649.

8.  Li,  W.   K.  W.    1983.    Consideration  of errors  in  estimating  kinetic
      parameters  based  on  Michaelis-Menten  formalism  in  microbial  ecology.
      Limnol. Oceanogr.   j?8_:185-190.

9.  Michaelis,  M.  and  M.  L. Menten.   1913.   Kinetics  of  invertase  action.
      Biochem. Z. 49:333-369.

10. Parsons,  T.  R.  and  J.  D.  Strickland.    1962.    On  the  production  of
      particulate  organic   carbon  by  heterotrophic  processes  in  sea  water.
      Deep-Sea Res.  8:211-222.

11. Pfaender,  F.  K.  and  G.  W.  Bartholomew.   1982.    Measurement  of  aquatic
      biodegradation rates by determining heterotrophic uptake of radiolabeled
      pollutants.  Appl. Environ. Microbiol.   44:159-164.

12. Wright,  R.  T. and  J.  E.  Hobbie.   1966.   Use of  glucose  and  acetate by
      bacteria and algae in aquatic ecosystems.  Ecology  47:447-464.

13. Wright, R. T.  1978.  Measurement and significance  of specific activity in
      the heterotrophic bacteria of natural waters.   Appl. Environ. Microbiol.
14. Wright, R.  T.  and  B. K. Burnison.   1979.   Hetertrophic activity measured
      with  radiolabeled  organic substrates, pp.  140-155.   _Ir^ J. W.  Costerton
      and  R. R.  Colwell   (Eds.),  Native  Aquatic  Bacteria:    Enumeration,
      Activity,  and  Ecology, ASTM  Technical  Publication  No.  695,  Amer. Soc.
      Testing Material,  Philadelphia, PA.

15. Zimmermann,  R.  and  L.  Meyer-Reil.    1974.   A new  method  of fluorescence
      staining   of   bacterial   populations   on  membrane   filters.     Kiel.
      Meeresforsch 30:24-27.
                                      109

-------
Table 1.   HBP kinetic constants for biodegradation of detergent chemicals  in
           ground and surface waters.3
Compound
LAS
NTA
C12E9

v
Water (ng/liter/h)
surface 539
(429-649)
ground 138
(111-165)
ground 121
(105-136)
surface 8081
(4575-11587)
K1
(ng/liter)
902
(546-1259)
63
(32-94)
388
(256-520)
504
(98-1008)
SAI
r2 (ng/cell/h)
0.99 1.5 x 10"5
0.99 1.7 x 10"5
0.99 3.3 x 10"5
0.99 3.0 x 10"5
a   Values in parentheses are 95% confidence intervals of true means.
                                      110

-------
                                     1/Vmax (A)
 y-intercept = T
                                                      max
                                                                 max
               -10
                x-jntercept
Figure   1.   Linear transformation of hyperbolic saturation curve
            (equation (4)) as derived by  Wright and Hobbie (12).
                              Ill

-------
             LAS RAPID CREEK RIVER WATER
         Dotted Contours Are 95.0% Conf. Limits Of True Mean
    480-
    400-
    320-
O
o
LU
            SUBSTRATE CONCENTRATION (yg/l) x 101
   Figure  2.  Rate or degradation of LAS in river water as a function
            of LAS concentration.  Data have been analyzed by equation
            (6), and parameter estimates are summarized in Table 1.
                          112

-------
            NTA CANADIAN GROUND WATER

        Dotted Contours Are 95.0% Conf. Limits Of True Mean
O)
c
O
g

LU
    120-
    100-
               40
80
240
             SUBSTRATE CONCENTRATION
   Figure  3.  Rate of degradation of NTA in groundwater as a function

             of INTA concentration.  Data have been analyzed by equation

             (6), and parameter estimates are summarized in Table 1.
                           113

-------
              C12E9 OHIO RIVER WATER
    Dotted Contours Are 95.0% Conf. Limits Of True Mean
560
                                C E OHIO GROUND WATER
                          Dotted Contours Are 95.0% Conf. Limits Of True Mean
                         0    20     40     60    80     100    120
                             SUBSTRATE CONCENTRATION Cug//) x 101
  0
            20       40      60       80      100      120

         SUBSTRATE CONCENTRATION (fjg/l) x 101
  Figure  4.  Rate of degradation of Ci2E9 concentration.  Data
            have been analyzed by eqdation (6), and parameter
            estimates are summarized in Table 1.
                     114

-------
                                 PANEL REPORT

     The purpose of this panel  was to evaluate the applicability of the short-
term   (approximately   24  hours)   heterotrophic   potential   method   to  the
biodegradation of  xenobiotic  compounds.   Before  discussing  the applicability
of heterotrophic potential measurements,  we believe it is crucial that certain
terms be defined to avoid confusion and problems with semantics.

     The  technique commonly  referred  to  as  heterotrophic  potential  is  an
activity assessment based on  saturation  kinetics  in  which  metabolic velocity
is measured over a range of substrate concentrations.  In addition to yielding
a velocity at each concentration  added,  the maximum potential  velocity  (Vma  )
that the  natural  community may  attain  can be established.   The technique is
based  on  the assumptions  specified by Michaelis-Menten which  require short-
term  incubations   to  avoid significant  changes  in  substrate  concentration,
active  biomass,  and  the enzymatic  rate  1 imitating  step  during  the tes.ting
period.  Other useful  biodegradation parameters can also be obtained  from this
data set.  These parameters are defined below.

    1.  Specific  Activity  Index  is  a  measure   of  the  amount  of   microbial
        activity per  unit  biomass  in  which  the biodegradation rate is divided
        by the  number  of  organisms active on  that compound.   The  units are
        rate/unit biomass.

    2.  V      is   the   rate   of   biodegradation  at   saturating   substrate
        concentrations; the units are g substrate/liter'hr.

    3.  Turnover   time   is  the   reciprocal of turnover  rate, which  is the
         fraction of material biodegraded at subaaturating  substrate  concentrations
        divided by time,  in units of  time.

    4.  Kt is  the  concentration  of a compound at half  V    in  units of grams
         \f  - .                                           nid A
        per liter.


     There  are  several  advantages  to  the  heterotrophic  potential testing
method to survey the  biodegradabi1ity of xenobiotics.

    1.  The test is relatively simple to perform.

    2.  A  wide  range  of concentrations  of  the   compound  can  be  tested in  a
        relatively short period.

    3.  These  substrate  concentrations   approximate realistic   environmental
        concentrations.

    4.  The  test  is short-term  (approximately 24 hours); therefore changes in
        certain environmental  parameters (biomass, etc.) do not  interfere with
        subsequent data analyses.

    5.  There are  statistical  techniques (nonlinear regression)  available  for
        data analyses.
                                      115

-------
    6.  Several useful  biodegradation rate constants and kinetic constants can
        be calculated

    7.  The  procedure  is  cost effective  compared  to  long-term  testing now
        performed.
     As with any test method,  there  are  also disadvantages which can restrict
the uses of this method:

    1.  If "adaptation" by the microbial  community  is  required,  it may not be
        detected.

    2.  A  radiolabeled  compound  is usually  required  to perform  the  tests at
        environmentally realistic concentrations.

    3.  For the test to be valid, it must be run at a range of concentrations,
        including levels well in excess of K..

    4.  Kinetically diverse microbial assembleges are not strictly Michaelian.
     Further  research  is  needed to assess  the  applicability  of heterotrophic
biodegradation  rates  in  complex environmental  systems.   These research areas
are:

    1.  Expansion  of  the  data  base  by  application  of  the  method  to other
        ecosystems, in particular soil and subsurface environments, by  using a
        range    of   compounds    (benchmark    compounds)    with    different
        physical/chemical  properties.

    2.  Definition  of  the  range  of  compounds  for  which  the  method  is
        applicable by systematic testing of different classes of chemicals.

    3.  Determination of  the most  "appropriate"  measure  of microbial  biomass
        for  estimating  numbers  of  substrate  utilizers and Specific Activity
        Indices.
     Also, the panel recommends the following:

    1.  Data  analysis  should be  performed  by nonlinear regression techniques
        to estimate kinetic constants and associated statistical parameters.

    2.  In all testing the importance of the  dilutent must be emphasized.

    3.  Rate  constants  and kinetic constants obtained  from different  testing
        protocols should be compared.

    4.  Future  improvements  in  methods  and  theory should be incorporated  into
        the assessment of biodegradability as technology permits.
                                      116

-------
Panel  Members:  Peter Chapman, University of Minnesota
                Doris Paris, U.S. EPA
                Gerald Neimi,  University of Minnesota, Duluth
                Dave Gibson, University of Texas
                Bill Gledhill, Monsanto
                Phil Howard, Syracuse Research Corporation
       RELATIONSHIPS  BETWEEN  CHEMICAL  STRUCTURE  AND  BIODEGRADATION RATES

Objective:  To  discuss  concepts  for   relating  the  biodegradation  rate  of  a
            chemical  to its structure.

     Because of  the  extremely large  number  of  chemicals which  the Office of
Toxic  Substances  will  be  required  to  regulate,  it is  important  to  evaluate
structure-activity  relationships  (SAR)  as   a  potential  tool  for  applying
biodegradation  information  for   representative  chemicals  to  other  similar
chemicals for which  relatively little  fate  information  is  available.   Current
SAR  studies  use  BOD's  and  persistence  testing  data  to  compare  chemical
structure  and  chemical   properties   to  biodegradability.    These  data  are
frequently generated using sewage sludge  or  other environmentally unrealistic
inocula.    Can  biodegradation  rate   information  from  studies   using  natural
inocula  and  intact  microbial  communities  be   used   successfully   in  SAR
studies?   What  limitations and  disadvantages  are inherent in  this approach?
What  methods  and quantitative  analysis  of  the  data  should   be  used  for
obtaining  biodegradation   rate  information?    Is  the  information too  site-
specific?   What  chemical  classes should be tested?   What  aspects of  chemical
structure or chemical properties  should be considered?
                                     117

-------
                   STRUCTURE-ACTIVITY RELATIONSHIPS (SARs)

                                P. J. Chapman


                                   ABSTRACT

           In   attempting   to    relate    activity   to   structure   in
      biodegradation  studies, the  object  is to determine whether one can
      perceive in the  structure of a given  chemical  (or  in  certain  of
      its properties) features  which reveal  its  biodegradability  or its
      persistence.    Put  another  way,  we  might  ask  are there ways  of
      looking  at  the  structures  or   derived  parameters  of  organic
      compounds  which  can   indicate  whether  their  biodegradation  is
      permitted or constrained  in the environment?   This  approach,  if
      successful, would  of  course be  invaluable to  the EPA  Office  of
      Toxic  Substances  in its  evaluation of the numerous  chemicals for
      which regulatory decisions are  required.

                          EVALUATION OF  ASSUMPTIONS

     Is  it  possible  that  for  kinetic  purposes  we  can   treat  the  complex
microflora of  different environments as  if  they are  pure cultures?   Or,  to
take the analogy one  step  further,  as  if  we were  dealing  with  a  purified
enzyme  system  challenged by  a  range  of  potential  substrates?   This  latter
analogy  at  least  enables us to see  that  we  can  apply this  information  to a
new, potential  substrate  in an informed way  once  we find  the  answers  to the
following questions:

    (i)   What is the mechanism of the  reaction catalyzed?
    (ii)  What is  the range of compounds  serving  as  substrates and  at  what
          relative rates for given concentrations?
    (iii) What groups in  the  substrate  are recognized  by  the active  site and
          what steric factors limit the  range of substrates?

But can  we  now predict  the rate  of  its  attack?   Or are we merely limited to
statements such as "it looks as if it ought to fit  at the active site?"

     We  should realize that  in  other areas some  degree of  success has already
been achieved  by  this  approach.    Perhaps  we  can  view  the  many different
environments of our planet in the same  way  that a pharmacologist views tissues
or  target  organs  and asks:   What is the  fate of this  drug  in these tissues
when it  is  administered?   Pharmacologists  already have  at hand the knowledge
about what methods of administration to use to  best  reach a  given tissue and
what  types  of  compounds will  enter   these  tissues   and  undergo  metabolic
transformation.   In  other  words,  the  design of  tissue-targeted drugs  has
considerable  rationale.   As  long ago  as  1690 the British philosopher,  John
Locke, wrote in  his  "Essay  Concerning Human  Understanding":  "Did we know the
mechanical affections of the particles  of rhubarb,  hemlock, opium and a man as
a  watchmaker does  a  watch  we  should  be  able  to tell  beforehand  that rhubarb
will  purge,   hemlock  kill  and  opium  make  a  man  sleep."   Note  that  Locke
recognized the need to  know both  the  "mechanical  affectations" of man as well
as  those of different  particles.  The  pharmacologists'  success  in  relating
specific  structural  features  of drugs  to  precise biological  effects  is based


                                     118

-------
on  exactly  those principles.  Their application  1s  relevant to an assessment
of  the  toxlcological  effects of chemicals  in  our environment,  but we should,
however,  realize that our level of understanding of the fate of chemicals in
our  environment  will  be  limited  by  what we  know  of  the potential  of the
environment  to  deal  with  chemicals  of different types.   Whether  that can be
treated  in  such a systematic  fashion  remains to be seen.   No  matter whether
the  assessment   is  toxicological  or  biodegradative, one  is obliged  to make
reference  to  a  data base  --  an  organized  body  of  information  to  which
reference  can  be  made for  precedents and  examples whenever  a  new  chemical
comes under  the  gaze  of a  regulatory agency.  What would our assessment now be
of  the  early  alkylbenzene  sulfonate  detergents in  use in the  late 1950's?
Hopefully we would now be able to recognize  that the  extensive branching of
the alkyl sidechains  — a  feature introduced by the  synthetic methods  employed
--  is an  environmentally  undesirable  feature in such compounds because highly
branched alkanes  show  limited rates of attack and the detergent sidechains are
the usual sites  of attack  when  biodegradation is initiated.  Unfortunately we
had  to  experience the effects of  these  slowly  degradable detergents before
research  revealed that detergents with  linear sidechains, the presently-used
LAS detergents,  are much more readily  biodegradable.

                FEASIBILITY OF SAR APPROACHES  FOR  BIOOEGRADATION
                                                                              t

    The above example raises a  question about  the general  feasibility of the  SAR
approach  --  our  knowledge  of  the  biodegradation  of  different   classes  of
organics  has not begun to keep  in  step with the range of  different  synthetic
compounds in current  use, not  to  mention  those  presently under development.
It  would  appear that  at  this  time we do  not  possess  a sufficiently  adequate
data  base  from  which  to  assess  different chemicals.   Thus we  really have at
least two  important  questions  to  deal  with.   Is the  SAR  approach a  feasible
one?   If so, then what  specific  information  is needed to make  the  approach
generally useful?   If the answer to the  first  is in the affirmative, then it
is  necessary to  evaluate  different  approaches perhaps   by determining  the
predictive value  of each  in  relation  to  subsequently performed laboratory and
field studies with a  representative number  of test chemicals.

                             FUTURE  RESEARCH NEEDS

     Whatever approach is  adopted,  a body of  information on  biodegradation
must  be  available and organized in such  a  way  as to be useful  and accessible
to  all  interested parties.   Furthermore  it should  be  based on' standardized,
widely  acceptable procedures.    Whether  we  rely  on  a  large  body  of  existing
biodegradation  data  such   as  the extensive  BOD values  in  the  literature,  on
pure   culture   studies    of  pathways   and   biotransformations    or,   more
realistically, make use of some measure of  the rate of disappearance of parent
compound, are matters  for  discussion.   Personally I  favor  the  latter  approach
coupled  wherever  possible  to  some  measure  of  the  completeness  of  the
biodegradation  process  using,   for   example,   percent   conversion  to carbon
dioxide or,  in  anaerobic  situations,  to carbon  dioxide  and methane.   I think
it should be stressed that if one adopts disappearance of parent compound as a
measure  of  biodegradation,  that  an  apparently  rapid   reaction  such  as,
hydrolysis  of the methyl  ester of  24D  or of  methyl  parathion,   will  yield
readily  degradable methanol  plus  the  bulk  of  the  molecule  which may  be
considerably  more  resistant  to attack.   In other words,  we have  to  have  an
approach that will also  provide information about products so  that these too
can be evaluated in turn.
                                     119

-------
     Additional   questions  about  concentrations of  test  compounds,  type  of
inoculum, whether sediment, should be present or  if  nutrients  should be added,
are all  reflections  of the type of  environment  we  are trying to  simulate  in
the laboratory.   Do we want  to try for  the  best  possible circumstances  for
biodegradation or the worst possible scenario?   To me there seems to be a very
real  need to assess the potential for  aerobic  biodegradation  in  both types  of
environments  and to  follow  up  with  anaerobic  studies  for  compounds  which
emerge as  persistent  or  slowly biodegradable.   We  could ask  this  question
another  way  by   saying  how  far can  we extrapolate   findings  from  one  set  of
environmental conditions,  inocula,  concentrations, etc.
                                     120

-------
                A  PRELIMINARY MODEL  TO  PREDICT BIODEGRADABILITY
                            FROM  CHEMICAL  STRUCTURE
                                  G. J. Niemi
                                Ronald  R.  Regal
                              Dinah D. Vaischnav
                                Oilman  D.  Veith

                                   ABSTRACT
           The  long-term   success   in  understanding  the  impacts  of
      potentially hazardous  chemicals  in order to  ensure environmental
      protection is dependent,  in part, on the development of predictive
      models  that   reduce   the   need   for   cost-prohibitive  tests  on
      individual  chemicals.   Under the  assumption  that chemicals  of
      similar   structure   have   similar  behavior,   knowledge   on  the
      structure of  a  chemical  allows the estimation  of  its  activity by
      means of comparisons with chemicals of known activity.   We present
      the results of a preliminary analysis on the prediction of whether
      a  chemical   is   persistent  or  not  persistent  on  the  basis  of
      structure-activity relationships.  We use a data base derived from
      a  literature  search  of  the chemicals that  have  been  tested for
      biochemical  oxygen demand (BOD) and apply multivariate analysis to
      develop a prediction model.

                        APPROACH  FOR PREDICTION OF BOD

     We  calculated  three  sets  of  molecular connectivity indices  for those
chemicals  with  BOD  data  [framework,  simple  or bond-corrected,  and valence-
corrected  indices  ( Kier and Hall, Molecular Connectivity in Chemistry and
Drug Research,  Acadtmic Press,   NY, 1976)],
     Our  approach  for  the  prediction  of   BOD   was  a  two-step  process:
(1) identification   of  chemicals   using   similar   structure   with  K-means
clustering  (' Dixon and Brown,  Biomedical Computer Programs, P Series,
 Univ. of California Press, 1979) and  (2) separate discrimination of
chemicals within a cluster of high BOD (defined as those with BOD values > 13)
and those with low BOD (those with BOD values <_ 13).  We performed the cluster
analysis with principal components that were calculated  for each chemical  from
a   separate  principal  component  analysis of  45 variables (a  combination  of
variables  calculated  from  the  3 molecular  connectivity  indices)  for  16,121
chemicals.   We assumed that  this  large  data  base was  representative  of the
multivariate  space  occupied by a substantial number  of  industrial  chemicals.
For  example,  there  are  currently  about  45,000 chemicals registered  in  the
Toxic Substance Control Act (TSCA) inventory.


                       MODEL DEVELOPMENT AND  APPLICATION

     A total  of 340 chemicals  (120  with  low  BOD  and  220  with  high  BOD) was
identified from  the data  base  assembled by  Vaischnav.    Because  the  K-means
clustering algorithm  uses   Euclidean distances  to  define  clusters,  chemicals
that are outliers  in the data base have a major influence on  the definition of
clusters.  Therefore, we identified two major parts of the principal component
space, "outer"  and  "inner"  space, and performed separate  clustering analysis
on  each.    The  outer space  chemicals were  defined  approximately as  those
chemicals  >  2  standard  deviations  from  the  mean  for  any  of  the first  8


                                     121

-------
principal  components.   This  division  resulted  in  the  identification  of  62
chemicals   (18%)  in  the  outer  space  and  278  (72%)  chemicals  in  the  inner
space.  Subsequently, we identified 4 clusters  (clusters  1-4,  Table  1)  in the
outer space group of chemicals and 6 clusters  (clusters  5-10,  Table  1)  in the
inner space group of chemicals.

     Three of the  clusters  (clusters  1,  2, and  5)  required  no discrimination
between high  and  low  BOD  groups because  two  of  the  outer  space  clusters
included  only  low  BOD  chemicals, while  one  of  the   inner  space  clusters
included 95%  (18/19) of chemicals with high  BOD.   Discrimination between high
and low BOD chemicals in the  remaining 7 clusters  was performed with stepwise
discriminant function analysis (DFA)  (Dixon and Brown, op. cit.).  We included
variables  up  to  6th order  from each connectivity index  separately  in the DFA
to identify those  variables  or  combinations  of  variables  that  best  separated
the  groups.    In  the  summary  (Table  1)  we  selected  those  variables  that
provided the best discrimination in terms  of  the number of chemicals  correctly
classified into the low BOD group.

     The   clustering/discrimination   model    resulted   in    the    correct
classification of  78%  (92/120)  of the chemicals identified  as  having low BOD
and  77%  (170/220)  of the  chemicals  identified as having  high  BOD.   However,
there  is  a  differential  probability  of correct  classification  based  on which
part  of the  subspace or  cluster the chemical  is assigned.   For example, if a
new chemical  is classified into the outer space group of chemicals, then there
is  a  high  probability   (about  94%)  that  the  chemical   can  be  correctly
classified  as  having  a  high  or  low  BOD.     Similarly,  if   a chemical  is
classified  into  cluster  3  of  the  inner  space,  then there is about  an 82%
chance  of    correct   classification.      In   contrast,   the   discrimination
probabilities in cluster 4 or 5 of the inner space for the  low BOD groups are
only  slightly above  50%.


                             FUTURE  RESEARCH  NEEDS

      It  is  apparent  that  chemicals  are  persistent  because of a  variety of
microbial  mechanisms  that  relate  to  degradability.    Therefore,   a  global
approach  that  puts  all   persistent  chemicals  in  one  group  and   all  non-
persistent  chemicals  in  another,  such  as  a  discriminant  analysis  of all
chemicals  considered simultaneously,  is  unsatisfactory.   Our  second approach
was  to classify  chemicals into  similar,  relatively homogeneous  groups, and
then  search  for  the common  structural  elements that separated persistent from
non-persistent  chemicals.    In  general,  this  preliminary  model  is relatively
good  for  the prediction  of persistence, but  we believe  that the model can be
improved  considerably.   The following   should be  considered  in  subsequent
applications.

     1.  We  suspect errors in the  BOD values due to  a  lack  of acclimation of
        chemicals  or due to  test  measurements.   Re-evaluation and  re-testing
        of chemicals will be  necessary to clear  potential  problems in the data
        base.

     2.  The  number of possible  combinations  of  clusters  and  variables that  can
        be  used  for 340  chemicals is  enormous.  If chemicals  within  a  cluster
        have  low  BOD  but  different structural  reasons  for being persistent,


                                      122

-------
then there is no guarantee  that  discrimination  analysis  can recognize
these structural  differences.    A  variety  of clustering models  using
combinations of  different  numbers  of clusters and  variables  needs  to
be evaluated.

3.  We used the multivariate techniques described here as exploratory
tools to elucidate structural configurations that may be associated
with biodegradation.  Therefore, we included variables with
relatively low F-ratios to enter into the discrimination equation.
Subsequent investigations, preferably with a larger sample size,
need to be attentive to potentially spurious correlations in the
process of screening a large number of variable for inclusion in
a model.  For this reason we have emphasized the potential
usefulness of this overall strategy and have not focused on
specific structural attributes that were associated with
discrimination.
                             123

-------
Table 1.   Summary of cluster analysis and discriminant  function  analysis  for
           340 industrial  chemicals.
SPACE CLUSTER ANALYSIS
N-Chemi
High
BOD
Outer 1 0
2 0
3 5
4 13
Inner 5 18
6 35
7 10
8 60
9 29
10 50
cals
Low
BOD
6
16
13
9
1
18
12
15
14
16
DISCRIMINATION ANALYSIS
% Correctly
Classified

High Low Variables Included
BOD BOD -F Ratio
-
_
60 85 SCL4
SCL6
85 78 FCL5
FP6
FP4
FCL3
_
60 78 SPC6
SPC5
80 83 VPO
VP1
VCL4
VCH6
83 53 VPO
VPS
VP4
VCL3
72 64 SCL5
SCH5
SPC4
SPC5
76 75 SP1
SP3
SP5
SCH6


- 4.95
- 2.03
- 4.92
- 5.40
- 2.84
- 1.28

- 1.18
- 2.67
- 1.16
- 2.23
- 3.83
- 2.27
- 4.78
- 7.60
- 3.04
- 1.37
- 2.63
- 3.79
- 2.55
- 1.65
- 2.63
- 4.12
- 1.42
- 5.91
Total
220
120
77
78
                                      124

-------
                                 PANEL REPORT

     To obtain  information  on the biodegradation  of  representative compounds
for a  data  base by which to  assess  the biodegradation  of  many  other similar
chemicals of  interest  to  OTS, we employed two  general  approaches.   The first
of these makes use of pure cultures,  an approach that reveals only the genetic
potential  of microorganisms  to bring  about biodegradation.  It does not supply
information about the  role  of these  organisms under environmentally realistic
conditions.  The second approach employs  mixed  cultures under conditions that
simulate environmental  circumstances and  provides  biodegradation information
such as  the kinetics of  compound  disappearance and the need for acclimation
periods.  It should be stressed that  the only test systems from which reliable
data  extrapolation  can  occur  are   those  that  employ mixed  cultures  and
environmentally realistic conditions.

     In the view  of  this  panel, it  does  not  appear  to  be possible to predict
biodegradation rates from structure-activity relationships (SAR).  For a small
category of molecules, however, it does appear possible  to make qualitative or
even  semi-quantitative  judgments.     The  limitation   is  due to  lack  of  a
sufficient data base.  Compounds which are identifiable  as the simpler members
of discrete chemical  classes,  such as aromatic  hydrocarbons,  esters, phenols,
and amides, generally fall into this  category.

     If we  are  to  explore whether predictions  of  a  qualitative  or perhaps at
best a  semi-quantitative  nature  are  possible,  it  is  necessary  to enlarge and
extend  the  base  of information  from which extrapolations might  be  made.   In
this connection it  is important  that an  acceptable  procedure is adopted by a
majority of  workers   in this  area.    Use  of  BOD  data   has  suggested  that  SAR
correlations can be  profitably pursued but  it  appears  that  a more acceptable
methodology should provide  a  reliable measure of  primary biodegradation under
conditions  simulating different  environments  and  with  regard  to  ranges  of
substrate  concentration  variations   such  as  temperature,  aerobic/anaerobic
conditions, presence  of sediment, eutrophic versus oligotrophic  water,  etc.
This panel  recommends that such  a   procedure be  adopted for the  purposes  of
generating  information  useful  for SAR considerations  but wherever practicable
should  be  accompanied by parallel  studies  of  rate measurements  of ultimate
biodegradation.

     Having  generated  a  body  of  primary  biodegradation   data for  various
classes  of  chemicals,  it   should   be  evident  what   chemical   or  physical
parameters  might  logically  be explored  for  correlations that  might  provide
indicators of  biodegradation.   Chemicals should be selected  so  that they can
be grouped  into different  clusters  based on  chemical   classes  by  functional
groups.   within  these clusters  applications  of  statistical techniques  can
establish  whether  significant  correlation exists   between  structure  and
biodegradation rates.

     We suggest  that  the following  are workable  approaches  which  should  be
implemented into research  plans:

    1.   Choose optimal test  systems  acceptable for wide-spread use.

    2.   Choose chemicals or  classes  of chemicals from a   broad cross-section of
        those representative of different classes  of environmental interest.


                                      125

-------
3.   Determine  whether  biodegradation  data  for  simple  compounds  can  be
    related  to  more complex  structures  possessing several  of the  same
    structural  entities.
                                 126

-------
Panel Members:  Gary Klecka, Dow Chemical
                Dick Raymond, Groundwater Environmental  Consultants
                Jim Spain, Georgia State University
                Harvey Holm, U.S. EPA
                John Walker, U.S. EPA
                Robert Brink, U.S EPA
                Art Kaplan, U.S. Army-Natick
         EXTRAPOLATION OF LABORATORY ftlODEGRADATION DATA TO THE FIELD

Objective:  To develop a conceptual and experimental framework for testing the
            usefulness of  kinetic  expressions of  biodegradation  processes in
            predicting  that  biotic  fate  of chemicals   in  complex  natural
            situations.

     Decay curves  are commonly obtained  from  laboratory  experiments in which
little attempt  is  made to maintain conditions similar  to those  in the field.
Resulting  biodegradation  rates  may,   therefore,   be   potentially  inaccurate
estimations  of  rates  in  the  field.    What  should  be  done   to  test  the
environmental  significance  of  these  rate determinations?  Are  field  studies
required?  How should field studies be designed?   What  is the minimum size and
scope  for a  field  test?    What  additional  environmental  factors  should be
examined  for their  effect  on  biodegradation?   How will  laboratory-derived
rates   and   their   associated  kinetic  expressions   be  used   to  predict
biodegradation  in  the field where  the  effect of  variations  in  environmental
factors  may  be additive  or integrative?   What  is  the  role  of  mathematical
models in relating laboratory data to the field?
                                     127

-------
         EXTRAPOLATION  OF LABORATORY  BIODEGRADATION  DATA  TO  THE  FIELD

                                Gary  M.  Klecka


                                   ABSTRACT

           Studies  on the biodegradation of  chemicals  in  the environment
      are extremely complex.   In addition to microbial  transformation,
      other   environmental    processes    such    as    volatilization,
      photodecomposition,  hydrolysis,   and   adsorption   are  likely  to
      influence   the  concentration  of  a  particular  chemical.    As  a
      result, model  systems are usually  established  in the  laboratory  to
      study the  decomposition  of a compound  under controlled conditions,
      and  the  rates  observed  are then  extrapolated to  describe  the
      process in the field.   Since the key  to  successful  extrapolation
      ultimately depends on the  laboratory  data,  special attention  must
      be paid when  designing  and  conducting experiments such  that  the
      information obtained will  enable  one  to make a reliable  estimate
      of the rates  at which a  process will  occur in  the  environment.

         FACTORS CONTROLLING THE FATE OF CHEMICALS IN THE ENVIRONMENT
                                                                             r
     Fundamental  to predicting  the  biodegradation  of  a  chemical  in  ttfe
environment  is  the  identification of the environmental   compartments  in  which
the  chemical  is likely  to reside.   To  address this question,  a  number  of
simple  equilibrium  models  have  been   developed,  such   as   the  one  recently
described by Neely  and  Mackay (6).  Once the partitioning  of  the  compound  is
known,  the  importance   of  biodegradation  in  controlling   the  fate  of  the
chemical  in the environment  will  require  an  understanding  of:    (a) the
mechanism(s) involved  in decomposition,  (b)  the  rates  at  which these processes
occur, and (c)  factors  which are likely to influence the  rates.

     Laboratory   studies  with   pure cultures  of  microorganisms  and  individual
compounds  have   revealed considerable  information   regarding  the  mechanisms
involved in  the  biodegradation of a  wide variety of  organic chemicals.   Many
of these  reactions  can  be classified as oxidation,  reduction,  hydrolysis,  or
conjugation.  One  of the most  important features of  microbial  degradation  is
the  ability  of aerobic  microorganisms  to catalyze   the  incorporation  of
molecular oxygen into  the  substrate  (4).   Thus,  the  bacterial  degradation  of
many  aromatic molecules  has  been  shown to   require molecular  oxygen  for both
hydroxylation  and  ring-fission   reactions.     Alternatively,   a   number  of
compounds,  including  several   of the  organochlorine  insecticides,  have been
shown  to be degraded  more rapidly under anaerobic  conditions  (7).   Thus,  a
knowledge  of the  mechanisms  involved  in   microbial  metabolism provides  an
indispensable guide  for understanding  the  fate  of  organic chemicals  in the
envi ronment.

      In  order  to   define  the  role  of  biodegradation  in  determining  the
environmental   concentration   of  a   particular  chemical,   a   quantitative
expression  describing  the rate  of the  reaction  is  required.   The microbial
degradation  of  a number of  chemicals  in  natural waters  and soils  has been
shown  to be described  by  first-order  kinetics,  and  thus   the  rates  of   these
biological  reactions  can  be  expressed in  terms of rate  constants.    Since


                                     128

-------
first-order  rate  constants  are  often subject to considerable variability, the
extrapolation   of   rate   constants   to  different  systems  will  require  an
understanding   of  the  importance   of  microbial  numbers  or   activities  in
determining  the  reaction  rate.    One  approach to  this  problem  involves
normalizing  rate  constants  by  the  size of  the  total  microbial  population as
enumerated by  conventional  techniques  (2).   However,  a considerable number of
studies  have  shown  that  biodegradation  rates  do not  consistently correlate
with  any measurement  of  the total  microbial  population  (1,  3,  10).   Thus,
further  studies are necessary  to determine  the  relationship between reaction
rate   constants  and   the   concentration   or   activity  of   the  specific
microorganisms  responsible  for  degradfng a compound.

     The extrapolation of laboratory-derived rate constants to the environment
will  also  require an  understanding  of the  various  biological,  chemical, and
physical  factors  that  influence biodegradation  rates,  and the  integration of
these   effects   into  mathematical   models.     Biological  variables   include
concentration,  spatial   distribution,   species   composition,   activity,  and
previous history of  the microbial population.  In addition, biodegradation may
be  affected  by  interactions among members  of the community such as mutualism,
commensalism,  competition and  predation,  although  little is  known  about the
influence  of   microbial    interactions  on  degradation   rates.    Substrate
concentration may also be a significant factor affecting  the susceptibility of
an  organic  compound to microbial attack.    There  is  some evidence suggesting
the existence  of  threshold  concentrations  below  which biodegradation does not
occur;  however, the data  are  inconclusive.   The  biodegradation  of   organic
chemicals  is known  to be  affected  by  a  number of  environmental variables,
including   sorption,  mixing,   temperature,   oxygen  concentration,   redox
potential,  pH, ionic  strength, and the  presence  of  organic  and  inorganic
nutrients.   Unfortunately,  the significance  of many  of these  variables  is
poorly  understood,  and few quantitative relationships are  available.   These
effects  will  need to  be  evaluated  if  laboratory data  are to 'be  successfully
extrapolated to predict the fate of  chemicals in the environment.


              USE  OF LABORATORY  DATA  TO PREDICT  FATE  IN THE FIELD

     In  spite  of  the many  uncertainties,  several   investigators  have  had
considerable  success in   using  laboratory  biodegradation   data to  predict  the
fate  of  a chemical  in the  environment.   The  key  to  these  initial  attempts
involved the  application  of site-specific rate  constants.   This approach was
used by Games  (5) for modeling  the fate of linear alkylbenzene sulfonate (LAS)
in Rapid Creek, South Dakota.   LAS is introduced into the  stream from a single
point-source  input   of domestic  sewage.    The   choice   of  LAS  as  the  test
substrate  simplified  the  modeling  process,  since  biodegradation  has  been
identified as  the  principal  transformation mechanism.    Rate  constants  were
measured  using   water  and  water-sediment  mixtures  obtained  from  sites
downstream  from  the  Rapid  City  wastewater  treatment   plant.    The  U.S.
Environmental  Protection  Agency Exposure  Analysis Modeling System (EXAMS)  was
then used for predicting  the  fate of LAS in  Rapid Creek,  and the results were
compared  to  actual  LAS  concentrations  in  the  stream  below  the  treatment
plant.  Steady-state concentrations  of  LAS in the water and sediment predicted
using  the  EXAMS model  agreed fairly well  with  the measured  concentrations.
However, a sensitivity analysis of  the model to  errors  in flow  rate,  aqueous
and  sediment   biodegradation   rate   constants,   dispersion   coefficient  and


                                     129

-------
adsorption  coefficient   revealed  that  the  model  calculations  were  highly
dependent  on   the   least  understood   parameters,   namely  the   dispersion
coefficient  and the  sediment  biodegradation  rate  constant.   Variation  in
either  of  these  parameters   resulted  in  a  significant  deviation  of  the
predicted concentrations from the measured values.   It  is  clear that a better
understanding of sediment biodegradation and  sediment-water interface dynamics
is  required  to improve  predictions  of  the fate  of  chemicals  in  the  aquatic
envi ronment.

     Walker  (8) demonstrated  the  use  of  site-specific  rate  constants  for
predicting  the  fate  of the   herbicide  propyzamide  in  experimental  field
plots.  Laboratory studies conducted using soil  samples  obtained from the test
site indicated that  biodegradation was first-order with  respect  to propyzamide
concentration,  and  the  effects  of  temperature  and   soil  moisture  on  the
reaction rate  could be  described  using  simple equations.   Field  studies were
conducted on  two  separate  occasions,  and test sites were  treated to simulate
both surface applied and incorporated herbicide  formulations.  Measurements of
propyzamide  concentration,  soil   temperature,   and moisture   content  were
performed  throughout  the  field  experiment.   For  the   computer  simulations,
biodegradation  rate  constants  were  corrected  using   data   obtained  for
variations  in  soil  temperature  and  moisture,  and   changes   in  herbicide
concentrations  were  predicted  by   integration   of  the  decomposition  rates
calculated  for  the  entire   experimental  period.     The  decomposition  of
propyzamide  predicted  using the  computer model   was found to  correlate well
with  the measured  concentrations  of  both surface  applied  and  incorporated
forrnul ations.

     Walker  (8)  subsequently  modified  the  basic simulation  model  to include
methods  for  estimating  soil  temperatures and moisture  contents from standard
meteorological  data.  When  used in  conjunction  with  site-specific degradation
rate constants, the model has  been  shown  to  predict  the environmental fate of
the   pesticides  simazine,   atrazine,    propyzamide,    linuron,   metamitron,
trifluralin,   metribuzin,   chlorthal-dimethyl,    prometryne,   asulam,   and
napropamide   (9).    With  certain  compounds,  such  as  simazine,  atrazine,
propyzamide,  and metribuzin,  there  was  a tendency to underestimate the rates
of  disappearance.   With these  pesticides,  it   appears  that mobility  of the
compound in the soil was responsible for the inconsistencies between predicted
and  measured  concentrations.    Deviations were  also noted  in  an  attempt to
model the  fate  of  surface applied  formulations  of  napropamide, where loss of
the compound was affected by photodecomposition.   These observations emphasize
the  importance  of accounting  for all fate processes when  attempting to model
the fate of chemicals in the environment.
                                      130

-------
                             FUTURE  RESEARCH  NEEDS

     In summary, predicting the role of biodegradation in determining the fate
of chemicals in  the  environment in many cases  will  involve the extrapolation
of  laboratory  data  to  the  field.    It  is  clear  that  the key  to successful
predictions will require a fundamental understanding of:

    1.   mechanisms involved in decomposition
    2.   kinetics of biological reactions
    3.   relationship of microbial activity to reaction rates
    4.   environmental variables likely to influence the rates.


     Although  a  considerable  amount  of   research   remains  before  reliable
extrapolations  of  laboratory  data  can  be  made on  a  routine  basis, several
investigators  have  had  considerable  success  by  applying  simple  models  in
conjunction  with site-specific  rate  constants.   It  is   likely  that further
studies of  this type will  be important for  improving  our ability to predict
the fate of chemicals in the environment.
                               LITERATURE CITED
    Bartholomew, G.  W.  and  F. K. Pfaender.  1983.
      temporal  variations  of  organic  pollutant
      estuarine environment.  Appl. Environ.
                                    Influence  of spatial and
                                biodegradation  rates  in  an
                           Microbiol.  45:103.
    Baughman,  G.   L.
      expression  of
      Dickson,  and
      Chemicals  in
   ,  D.   F.  Paris,  and  W.  C.  Steen.   1980.   Quantitative
   biotransformation  rate,  p.  105.   In  A.  W. Maki,  K.  L.
   J.   Cairns,   Jr.,  Eds.,   Biotransformation   and   Fate  of
   the  Aquatic  Environment.     Washington,  D.C.:  American
      Society for Microbiology.

    Bourquin, A.  W.,  J.  C. Spain,  and  P.  H.  Prichard.  1982.   Biodegradation
      activity  correlations  with  biological  and  environmental   variables.
      Abstr. Annu. Meet. Am. Soc. Microbiol.  N91.
    Da g 1 ey,  S.
      biosphere.
1977.    Microbial  degradation  of  organic  compounds  in  the
Survey Prog. Chem.  8:121.
    Games, L. M.  1982.  Field validation of exposure analysis modeling system
      (EXAMS) in a  flowing  stream,  p.  325.   In K.  L. Dickson, A. W. Maki, and
      J.  Cairns,  Jr.,  Eds.,  Modeling  the  Fate  of Chemicals  in  the Aquatic
      Environment.  Ann Arbor, MI: Ann Arbor Science Publishers.

    Neel y ,  W.  B.  and D.  Mackay.    1982.    Evaluative model  for  estimating
      environmental   fate,   p.  127.    In  K.  L.  Dickson, A.  W.  Maki,  and  J.
      Cairns,  Jr.,   Eds.,   Modeling  the  Fate  of  Chemicals  in  the  Aquatic
      Environment.  Ann Arbor, MI: Ann Arbor Science Publishers.

    Sethunathan, N.   1973.   Microbial degradation  of  insecticides  in flooded
      soil and in anaerobic cultures.  Residue Rev.  47:143.
                                      131

-------
8.  Walker,  A.    1974.    A  simulation  model   for  prediction  of  herbicide
      persistence.  J. Environ.  Qual.   2/396.

9.  Walker, A.    1978.   Simulation of  the persistence of  eight  soil-applied
      herbicides.  Weed Res.  18:305.

10. Wright, R. T.  1978.  Measurement  and significance of specific activity in
      the   heterotrophic   bacteria  of   natural   waters.     Appl.   Environ.
      Microbiol.  36:297.
                                      132

-------
                               EXTRAPOLATION OF
                 LABORATORY BIODEGRADATION DATA TO THE FIELD

                              Richard L.  Raymond
                                   ABSTRACT

           Field experience utilizing  biodegradation as a  cleanup tool
      for hydrocarbon contaminated soils and groundwaters has shown that
      many factors, presently  beyond the control  of the operator, affect
      the kinetics  of  degradation.   It  seems  highly unlikely  that  any
      laboratory test can be devised that  will come  close  to describing
      the kinetics  to  be expected in the  field.   A few  examples will
      illustrate the problem.


           USE OF LABORATORY DATA  FOR FATE PREDICTIONS IN THE FIELD

     Numerous laboratory  and  pilot  plant  studies have  shown that  many  pure
hydrocarbons   and   mixtures    thereof    are   degradable    at   rates   above
0.5 Ibs/gal/day.   The  conditions  for  this high  rate  of  turnover  optimize
oxygen demand,  substrate  and  inorganic  nutrient  concentrations, temperature,
and physical  mixing.   Optimization is  rarely  achievable  in  the  field.   In
Table 1,  Table 2,   and   Figure 1,  the   effect   of   substrate   type   and
concentration,  soil  composition,   and  probably  adsorption  on the rate  of
biodegradation is illustrated  for two biofarming projects.

     Equally  dramatic  effects  are  shown  in field situations  for  groundwater
cleanup  of  gasoline.   Two  examples are shown in  Figures  2 and 3.   Rates in
gallons  of  aquifer  storage   capacity  are  0.003  and  0.005  Ib/gal/month,
respectively.   In  groundwater cleanup,  oxygen  transfer is  the  most limiting
factor in the rate of degradation.


                               LITERATURE  CITED

1.  Raymond,  R.  L., J.  0.  Hudson,  and  V. W.  Jamison.    Land  application of
    oil.  AIChE Symposium Series.  Water,  1978.  pp. 340-356.

2.  Raymond,  R.  L.,  V.  W. Jamison,  and  J.  0.  Hudson.   Beneficial  stimulation
    of  bacterial  activity  in   groundwaters   containing   petroleum  products.
    AIChE Symposium Series.  Water,  1976.  pp. 390-404.
                                      133

-------
Table 1.   Biodegradation rates observed in landfarming.

     Substrate                        Loading            	Rate	
                                    Ibs/cu.  ft.          Ibs/cu.  ft./month

     Decane                            0.012                0.0061

     Octadecane                        0.01                 0.0070

     Tetramethylpentadecane            0.01                 0.0020

     Biphenyl                          0.01                 0.0070

     Phenanthrene                      0.02                 0.0100

     Phenylundecane                    0.01                 0.0200

     #2 Fuel Oil                       1.6                  0.1100

     #6 Fuel Oil                       1.6                  0.0750

     Crude Oil                         1.6                  0.0810


These rates are orders of magnitude less than predicted from laboratory  data.
                                      134

-------
Table 2.  Percentage reduction in oil concentration
                                     Location  of  Field Sites
Type of Oil

After one year:
Used crankcase oils
Crude oils
Home heating oils (#2)
Residual oil (#6)

After three years:
Used crankcase oils
Crude oils
Home heating oil  (#2)
Residual oil (#6)
Marcus Hook,
Pennsylvania
69.2
54.2
86.0
48.5
90.6
87.3
98.3
89.2
Tulsa,
Oklahoma
73.8
77.5
90.0
65.5
99+
99+
99+
94.0
Corpus Christi,
Texas
60.8
54.2
86.0
59.4
79.8
78.0
94.0
85.5
Avera
67.9
61.9
87.3
57.8
89.8
88.1
97.1
89.5
                                      135

-------
5
*
I2O


100


80


60


40


20
                            o>
                            E
120


100


80


60


40


20
     6/1  7/1  8/1  9/1  IO/I

         CH -(CH )  -CH
            3283
                             6/1   7/1  8/1  9/1  IO/I
  6/1   7/1  8/1  9/1  IO/I
                              120

                              IOO


                              80


                              60


                              40


                              20
                                6/1  7/1  8/1  9/1  IO/I
     Figure 1.  Examples of biodegradation  under field  conditions.
                           136

-------
      120

      100
   z
   -  80

   1 60

      40

      20
  160

  140

_, 120

z loo

\  80
E
   60

   40

   20
                                      204
                                      T
        6/1  7/1  8/1  9/1  (0/1
     6/1  7/1   8/1  9/1  IO/I
Figure 1 (continued).   Examples  of  biodegradation under field
         conditions.
                      137

-------
Figure 2.   Millville,  NJ,  biostimulation


     Aeration capacity

     Nutrient distribution

     Nutrient composition and quantity
       Ammonium sulfate
       Sodium phosphates
       Trace inorganics - Mg, Mn,  Fe, etc.

     Water control
     Duration
     Gasoline degraded
20 SCFM

Batch and continuous


1200 Ibs.
 600 Ibs.
  25 Ibs.

Injection and discharge to waste

6 months

1000 gals.
Figure 3.  Whitemarsh Township PA, biostimulation
     Aeration capacity

     Nutrient distribution

     Nutrient compostion and quantity
       Ammonium sulfate
       Sodium phosphates

     Water control
     Duration
     Gasoline degraded
25 SCFM

Batch


58 tons
29 tons

Discharged to waste

One year

45,360 gals.
                                      138

-------
                  COMPARISON OF £-NITROPHENOL BIODEGRADATION
                     IN FIELD AND LABORATORY TEST SYSTEMS

                                  J.  C.  Spain
                                 INTRODUCTION

           Laboratory  test  systems  provide  the  most  practical  means to
      obtain  data  that can  be used  to  predict the  biodegradation and
      fate  of  organic pollutants.    More  accurate   predictions   could
      probably be made if fate tests could be conducted in the field for
      each chemical  under consideration.  Such  studies  have often been
      conducted  for   the  application  of   pesticides   to  soil,  but
      constraints of  time and expense do  not  permit  such studies  to be
      carried  out   routinely  in  aquatic  habitats.    The  second best
      approach  is to  conduct a few field  tests  with  selected chemicals
      and  to  compare  the results with  those from laboratory  tests to
      assess the relevance and utility of  the laboratory data.  When the
      strengths  and  weaknesses of the test  systems  are understood, the
      degree   of  confidence  with   which   laboratory  data   can  be
      extrapolated can be evaluated.
     We  have used  several  types  of  biodegradation  test  systems,  including
shake  flasks,  eco-cores,  and  microcosms, to  study  the  biodegradation  of _p_-
nitrophenol  (PNP) in the laboratory.   In most  instances,  microbial communities
degraded nitrophenol  after a  lag  period  of several  days.   The length of the
lag  period was  variable,  however;  and  in  samples  from estuarine  or marine
sites  and  some  freshwater  sites, there was  no biodegradation for weeks.  The
inclusion of sediment also seemed to affect the biodegradation, but it was not
clear whether the effect made the results more or less  realistic.

     The purpose of this study  was  to  compare biodegradation of j}_-nitrophenol
and  concomitant  responses  of  microbial  communities  in  laboratory~test systems
with those in the  field.   We  prepared laboratory test  systems  with samples
from a freshwater pond  and then treated all  of the laboratory  systems and the
pond simultaneously with the test compound so  that direct comparisons could be
made.    Two  questions  were  of primary  importance:    (1)  Is  the adaptation
process  in the  field similar  to that   in  the laboratory?   (2) Which  type of
laboratory  test  system  (what   level of complexity) best predicts  the  field
results?
                           RESULTS FROM FIELD STUDY

     In  the  pond,  PNP  disappeared  slowly  for  the  first   120-150  hours
(Figure 1), then much more rapidly.  Because of fluctuations in  the pond water
level  due  to  rain and  runoff,  it was not  clear  from the  shape of the curve
between 100  and  200  hours  whether  the  biodegradation rate  had increased or
whether the test  compound  was being washed  out of  the pond.   Therefore, when
the PNP concentration fell  below 50 ug/1  in the pond, an additional aliquot of


                                     139

-------
PNP was added to bring the concentration to 134 yg/1.    After 23 hours with no
measurable rain or runoff, the nitro compound  was  no  longer detectable, which
Indicated that the biodegradation rate  had increased dramatically.

     The effects of prior exposure on  the  microbial  community  were determined
by measuring  the  biodegradation  rate  of PNP in the pond  exposed  to the nitro
compound and  in  the control  which had no prior exposure.   The  test compound
disappeared from the exposed pond within  one day  (Figure  2).  During the same
period, there was no detectable  degradation in  the control  pond.   The rate of
biodegradation in the first 24 hours  was two orders of magnitude faster in the
exposed pond  than  in the control  (-9.84 compared to  0.0998 ug/1).   After an
acclimation period  of 3-4 days,  the   test  compound  was also  degraded  in the
control pond  (data not shown).

     Bacteria able to mineralize PNP were enumerated in samples taken from the
pond  during   the  initial  exposure  to   determine  whether  changes   that  would
account for the Increased degradation rates could be detected in the microbial
community.  There was  an  increase of  three orders  of  magnitude in the number
of  PNP degraders in  the  exposed pond  between  100 and 240  hours  (Figure 3),
which  corresponded to  the time period  during which  PNP degradation increased
dramatically.  There was  no commensurate  increase  in total  heterotrophs,  and
there was no  proliferation of PNP degraders in  the control pond.


                        RESULTS  FROM LABORATORY TESTS

     Results  in  laboratory test  systems  were   similar  to those  in  the  pond.
For example,  data from the small  microcosms and  from  shake flasks with sediment
are shown in  Figures 4 and 5, respectively.  The shake flasks without sediment
(Figure 6) were the only exception to the close correlation  between  laboratory
and  field data.    The lag  period extended  for  almost  500  hours,  and the
subsequent degradation rate was slower than in  the other systems.

     Increases  in  biodegradation rates   correlated  well  with increases   in
populations  of  PNP degrading  bacteria.    We  are  currently  investigating the
relationships  among   microbial   growth,   acclimation,   and  degradation   of
xenobiotic compounds.

     The  results  of this study  show  that there  can  be dramatic  increases  in
microbial  degradation  rates   in  the  field after  exposure  of the  resident
microbial community  to the test compound.   Such  increases  must be  considered
in predicting the fate of pollutants that could serve as growth substrates for
bacteria.  Our earlier work has shown  that there can be wide variations in the
abilities of  communities  to adapt to degrading a test compound.  We  now have  a
better  understanding of how  laboratory test systems can be  used to  assess the
biodegradation  responses.   For  example,  the role  of  the sediment  in the
acclimation   and  the  degradation is  readily  apparent.     We  are  currently
conducting experiments to determine the factors that control such  responses  so
that they can be predicted.
                                      140

-------
                                  CONCLUSIONS

     Several  conclusions  can  be  drawn  regarding extrapolation of  laboratory
data to predict results in the field:

    1.   Laboratory  systems  must  include  the  components  that   affect  the
        biodegradation  rates  and  the  complexity  necessary  to  simulate  the
        environment under consideration.   In this  case,  inclusion  of sediment
        from the test site provided an acceptable level  of agreement.

    2.   Measurement  of   biomass   or   any  other   surrogate   for   predicting
        biodegradation  probably  will   not  provide  predictions  that  are  as
        accurate as measuring the biodegradation rate in the environment.

    3.   If a  simple biodegradation test  system can  be shown to  predict  the
        biodegradation  process   in   the   environment,   the  most   reasonable
        approach would  be to use the test  system  with an inoculum  from  the
        site to predict biodegradation  at  that  site.   That is to  say, if some
        parameter  must be  measured  to  predict  biodegradation   rates  at  a
        particular site,  then measuring the  biodegradation rate  itself with a
        simple  laboratory  system  would be the most  effective means currently
        available.

                                LITERATURE  CITED

     1.  Spain,  J.C.,  P.H.  Pritchard, and A.W. Bourquin.   1980.  Effects of
        adaptation  on hiodegradation rates  in sediment/water cores  from
        estuarine  and freshwater environments.  Appl. Environ. Microbiol.
        40:726-734.

     2.  Spain,  J.C.,  and P.A.  Van  Veld.  1983.  Adaptation  of  natural
        microbial  communities  to degradation  of xenobiotic  compounds:
        effects of  concentration,  exposure  time,  inoculum,  and
        chemical structure.  Appl.  Environ. Microbiol.   45:428-435.
     3-  Spain,  J.C.,  P.A.  Van  Veld, C.A. Monti, P.H.  Pritchard, and
        C.R. Cripe.   1984.  Comparison of j>-nitrophenol  degradation in
        field and  laboratory test  systems.  Appl. Environ.  Microbiol.
        (Accepted).
     4.  Van Veld,  P.A., and J.C. Spain.  1983.  Degradation of selected
        xenobiotic  compounds in  three  types of aquatic  test systems.
        Chemosphere.   12:1291-1305.
                                     141

-------
     350
     250

   OL
   Z
   Q.
      150
       50
                      •


50      100     150

        HOURS
                                      200
                      250
Figure   1.  Concentration of PNP in the test pond,  ihe test
           pond received an additional treatment of PNP after
           220 hours.
  250
Q_

Q_
   50
                               CONTROL
                  TREATED
              20
      60

HOURS
                               100
Figure  2.  Concentration of PNP  in treated and control ponds
           The treated pond was  exposed to PNP for 10 days;
           the control pond received no pretreatment.
                          142

-------
      I06
           HETEROTROPHS
  £L
                       TREATED
                       CONTROL
                  PNP DEGRADERS
               48
                 144

             HOURS
240
Figure  3.
Total  heterotrophs  and PNP  degraders enumerated
by MPN technique in  previously treated and control
ponds.  Samples were diluted appropriately and
incubated in PNP nutrient media for 2 weeks.
         200
       o»
         100
         50
                       .    A!    .
                      100
                     200
   300
                         HOURS
  Figure  4.  Degradation of PNP in small  (3  1) microcosms.
            The microcosms contained sediment and water
            (removed prior to treatment  from the pond).
            The microcosms received an additional treatment
            of PNP after 180 hours.
                          143

-------
   200
en
I  100
      0
                   100        200

                       HOURS

 Figure  5.  Degradation of PNP in  flasks containing water
            and 500 mg sediment/1.  The sediment was  removed
            from the pond prior to  treatment.  The flasks
            received an additional  treatment of PNP after
            170 hours.
    200
  - 150
  CL
  Z
  CL
     100
      50
 Figure  6.
    200    400     600

         HOURS

Degradation of PNP  1n flasks containing
water from the field site. The flasks received
an additional treatment of PNP after 600 hours,
                          144

-------
                                 PANEL REPORT

     The  importance   of  field  studies  has  become   increasingly  evident  to
participants  at  this  workshop.     At  this  point,  there  is  considerable
uncertainty   among   workshop   participants   regarding  the   validity   for
extrapolating laboratory data  to  predict the biodegradation of a  chemical  in
the environment.   Consequently,  the  validity of  laboratory  studies  must  be
tested in the field.

     Fundamental to this process  is  the selection of  laboratory  test systems
that  are  similar  to  the   various  environmental  compartments  (soil,  water,
groundwater, etc.).   Each  of these laboratory systems  should  incorporate the
important components  of  the compartment under consideration, such as a field
source  of  inoculum,  the  appropriate  solid  surfaces  (e.g.,   sediment),  and
environmental conditions.  Once a  laboratory system is developed,  it  will then
be important to verify the  ability of the system to predict  the biodegradation
of  a  chemical   in  the  corresponding  field  site.    The  design  of  these
experiments  should  incorporate a  statistical  analysis, in  order  to minimize
the  variability likely  to  be  encountered  in  the  environment.   Should  the
results of initial  trials indicate that the laboratory system does not predict
the field observations, the system should be modified as appropriate.

     Upon establishment  of  a reliable laboratory model, the system  should be
validated  by  using several  chemicals of different  classes and  enough field
sites  to  establish  confidence  in the system.    Following  validation,  the
laboratory system can then  be  used as a tool  to predict the biodegradation of
other chemicals at a  given  site.   Furthermore,  the laboratory  system can also
be used with confidence  to  address a  variety  of  additional  questions, such as
the relationship of biodegradation rates among different sites  and the effects
of a controlling environmental  factor on the biodegradation  rate.
                                     145

-------
             SUMMARY AND RESEARCH PLAN FOR  BIODEGRADATION STUDIES

                       BASED ON OUTPUTS AND DISCUSSIONS
                    AT THE BIODEGRADATION KINETICS WORKSHOP

                                 J. M.  Suflita
                                P.  H.  Prichard
                                 INTRODUCTION

     The microbial ecology effort of the EPA has evolved to the point where it
is no longer sufficient  for  researchers  to  enumerate  the  numbers and types of
microorganisms  in  a particular  environment.    Past  experience has  taught us
that this  approach  provides  little  insight  into the  cycling  of materials and
energy  in  the  ecosystem to  which the  organisms belong.   More importantly we
have become  interested  in and  involved  in  the activity  of  microorganisms in
nature  and  wish   to   describe  that   activity  in   a   quantitative  manner.
Therefore, we feel  that  future  research  should  describe kinetics of microbial
activity,  predominant variables  controlling  the reaction  rates, and potential
techniques to extrapolate laboratory biodegradation information to the field.

     The  greater degree  of  complexity  and  sophistication  inherent  in  this
research endeavor presents numerous  problems in approach and methodology.  The
biodegradation  kinetics  work  has pointed up a  number of  equivocal approaches
and  interpretational difficulties in biodegradation  studies.   However, during
the   course   of  this   meeting,  certain   underlying   generalizations  were
established; these can serve as a basis for the formulating of future research
directives.

     We  have  attempted  to organize  these directives  into a  specific research
plan.    Clearly,  the various  avenues   of  kinetic  research  are  not mutually
exclusive.   Progress  in  one  area  will  assist  the  development  of another.
However,  we  feel strongly  that certain  research  areas  deserve  an immediate
effort, whereas others will evolve in the future.

     Kinetic models of  biodegradation   processes  were generally considered  by
the  participants   at  the  workshop  as  necessary  elements   of   a  research
program.   Models  are based on assumptions that  can form the basis  for specific
testing  of  hypothesis   and   related  experimentation.    They  force  us,  for
pragmatic  reasons,  to describe the complexity of biodegradation with a minimum
number   of  measurable   parameters.     The   parameters    idealy  represent   an
integration  of  the  most  important  biotic  and  abiotic factors affecting the
biodegradation  of xenobiotic chemicals.  Although  the time and  research needed
to  search for  and  test  those parameters will  be  extensive,  we feel that the
potential  outcome will  justify the effort.   Kinetic  models will  also  provide
predictive capabilities  that  will   be   invaluable  for  our  making  regulatory
decisions.   Such models  will  require  us  to examine  the factors  controlling
adaptation events   in  natural  ecosystems,  since  these  phenomena  must  be
included  in  our  predictions.   We believe  that the  testing of kinetic models
represents a major  research  directive.

      Equally  important  are  research  endeavors  focusing  on  the  problem  of
extrapolating   laboratory  biodegradation  data  to  the  field  and extending


                                     146

-------
biodegradation  kinetic  predictions  from  one   ecosystem  to  another.    The
development of  a  conceptual  framework  for  extrapolation will  be  the impetus
and direction for  gathering  data from  both  field  and laboratory experiments.
It will  also force us to make decisions about how to determine efficiently the
spatial  and  temporal  variability  in  biodegradation  rates.    In  addition,  it
will facilitate the regulatory decision-making process and base it on observed
variability.

     These two areas, model testing and extrapolation, will therefore form the
basis for our research  plan.   Discussions and panel reports from the workshop
helped define the goals of this plan.


                                 MODEL TESTING

KINETIC MODELS  FOR BIODEGRADATION

     Several  types  of  kinetic  models  of  varying  degrees of  complexity were
discussed at  the  workshop.   However, two  simple  models,  a  first-order decay
model and  the so-called  "second-order  approach" or  Athens  model,  as we will
subsequently refer to it,  seem to be  the most reasonable to  incorporate  into  a
research plan.

     A first-order decay model contends that  the rate of substrate  (xenobiotic
chemical) disappearance over time in any environmental  sample will depend on
the  substrate  concentration  by a   proportionality  constant  (k-^)  if other
conditions  in  the  test  system  (cell  numbers,   cell  activity,   nutrients,
competing   substrates,   etc.)  remain  the  same   over   the   course  of  an
experiment.   The  proportionality constant  is   a  reflection  of  the actively
degradating microbial biomass in  the  sample  being  tested and the  inherent rate
at  which the  enzymatic  machinery  in  the   cells  will   attack  the  xenobiotic
chemical in question.   It  obviously will  vary, depending on  when  and where the
environmental   sample   is  taken.     The  model's  dependency  on   substrate
concentration will  only be  true  when substrate  concentrations are  relatively
low.

     The Athens model   reasonably contends  that  the rate of biodegradation is
proportional  to both substrate  concentration and  actively degrading biomass.
Typically,  a  first-order  rate  constant  (kj)  is  normalized  by  a   biomass
measurement  to  produce a  proportionality constant (kp).   In  this case k-, is
actually a  reflection  of  the  Michaelis kinetic parameters  Vmax and Km  (R-,  =
V   /Km).   The  proportionality constant, k~,  is specific  for each  xenobiotic
cnemical,  and  its  value  is  assumed  to  be  the  same   for  all   microbial
communities,   regardless   of  their  size,  activity,  or   source   (site-
independent).    This  suggests   that   the   biochemical  kinetic  constants  for
degrading any one xenobiotic chemical  are essentially  "universal."

     The Athens model,  however,  suffers from several  practical constraints and
inadequately  tested  assumptions.   Most obvious in this  respect  is  the  term in
the  expression  which requires  information  on the  actively  degrading biomass.
Several  biomass measurements have  been used in attempts  to test  this  model.
Measurements  of  CPU,   AODC  counts,  LPS measurement,   ATP,  etc.  have  found
limited  utility   with   relatively   few  substrate  classes   in  only   a  few
habitats.   To  be  sure, there  is general  disagreement  on the "best"  biomass
measurement currently  in  use.
                                      147

-------
     In  addition,  the  derivation  of  the   model   assumes   that  substrate
concentration is  much  less than  K^.    Since  many laboratory  experiments  use
starting substrate  concentrations  in  the range of 20-200 yg/1,  it  is  assumed
that Km must  therefore  be  relatively  high.   To ensure  that  degradation rates
are measured  in the first-order region,  substrate  concentrations should be at
least one-tenth the Kffl value.   Practically, this  means  Km  values should be in
the  low yg/1  range.   Thus, it  is quite possible that a biodegradation  rate
determination might be  unknowingly performed  in the  zero-order  region rather
than the first-order region.   Before  substrate concentrations can  be  assumed
to  be  well  below  K ,  an   indication  of  the  magnitude in  K   for  different
xenobiotic  chemicals must be obtained.   The implicit  assumption that K  values
are  constant  relative  to   other  parameters  such  as  Vm;)V  also  needs to  be
tested.                                                 max

ACTIVITY MEASUREMENTS

     It  is  apparent  in  the  development  and  use of  biodegradation  kinetic
models  that  (a)  the  activity  of  a   microbial   community  (which  includes
different  bacterial populations),  (b)  the  relativeconcentrations  of  the
active bacteria,  and (c) the  environmental  factors controlling their activity
need to be  assessed using  a minimum number of  parameter estimates.   The more
encompassing  and  integrative  the  parameter estimates,  the  less  are required.
We  believe  that EPA should support research to develop activity measurements
that  are  consistent  with   known   biological  theory,  easily  performed  by  a
variety  of   laboratories,  and  applicable   to  a  wide  variety  of  xenobiotic
chemicals and ecosystems.   The initial  use of  total  viable  cell  numbers  in a
microbial   community was,  in  essence,  an   attempt  to  provide  an  "activity"
measurement.   It  was  assumed  that all  bacteria,  or  some  constant proportion
thereof, were  active  in degrading a  chemical  and that they  were all  equally
active.  Evidence presented at  this workshop  and  from other sources indicates
that this "activity" measurement  is apparently  applicable to a limited number
of  chemicals  and  ecosystems.    This  is  due in  large  part  to  problems in  the
original assumptions.    Similarly, measurements  of   the  numbers  of  specific
chemical degraders  within  a population are limited  because  of methodological
problems and  the  realization  that the mere  presence of a  degrader  does  not
necessarily mean it is  active.  Thus,  new approaches  for relating the activity
of a microbial community to its degradative potential  are required.

     Two promising  approaches  are the  measurement   of the  maximum  velocity
(Vm,J of degradation and the first-order decay rates (k,).
  ma x                                                   i

     In  the   saturating  or   zero-order   region  of  substrate  decay,   the
biodegradation rate (Vmax)  is limited by the active catalyst concentration and
not  by  the  substrate  concentration.    Therefore,  the  rate  observed  at
saturating  substrate levels  is a   direct reflection of  the  actively degrading
biomass.   This  is exactly  the quantity  required  in  the Athens model   (that is
k2  = k,/V   )> but  it  uses a  rate of  biodegradation  instead  of cell  number.
At  first  glance,  it  would  appear that  the., units  for this  proportionality
constant  are  unusual   (i.e.,  liters.umoles    or  reciprocal   of  K   units).
However, when this kinetic constant  is properly  employed, -dS/dt  will  have
recognizable units  (i.e., iimoles.l~.h~ ).

     The  first-order  region   of   substrate  decay  is  also  a  reflection  of
activity degrading  biomass when   the substrate  is present  at concentrations


                                     148

-------
well below  the  K  of  the  reaction.    In  this  region, a  constant  fraction  of
substrate  undergoes   conversion  to  product  at  any   given  moment,  and  the
absolute velocity decreases as the substrate is  consumed.   Thus, the rate for
a given reaction is substrate dependent.

     It is important to note that first-order kinetics will often be displayed
if some other factor mitigates the consumption  of a substrate by the requisite
microorganisms.      For   instance,   this   behavior   will   be  observed  if
biodegradation  is  limited  by   diffusion  of  the  molecule to the  site  of
microbial  activity.     Such  phase   transfer   resistance   is  common  for  the
consumption  of  gaseous  substrates   or possible  for   chemicals  that tend  to
partition strongly to solids.

     Both approaches  (V    ,  k^  have minimal  interpretation difficulties, and
each is  a  direct measure  of  activity.   Secondly,  they  are not  limited to a
particular  type of  activity  (e.g.,   hydrolysis,  hydroxylation,  deamination,
etc.),  and  they are  not  restricted to a given  chemical  class or  series  of
structurally  related  compounds.   Both  approaches are  also independent  of the
nature  of  the catalytic  unit.   It matters  little if  the biotransformation is
actually catalyzed  by an extracellular enzyme,  a bacterium,  a fungus,  or an
anaerobic consortium  of  organisms.   The  approach  can  be  used  in a variety of
environments, including single phase systems (i.e., water alone) or multiphase
systems   (i.e.,  sediment/water  slurries).      In   addition,   the  activity
measurement  will  already  reflect an integration  of  the  environmental  factors
influencing  biodegradation.    In  general, V    determinations  do  not  require
the analytical  sophistication associated witn the measurement of low substrate
concentrations;  as  a  corollary  then,  the  requirements  for  radiolabeled
material  may  largely  be   circumvented.    Biodegradation  rates  obtained  at
saturating concentrations of substrate  (V a ) permit subsequent predictions of
rates   at   any   concentration   (assuming  the   Athens  model   is  essentially
correct).   Biodegradation  rates  obtained in the  first-order decay region are
only  predictive at  lower  substrate  concentrations.   Finally,  the suggested
activity measurements  can  be  made by all laboratories using relatively simple
incubation systems.

METHODOLOGY

     Ideally  we would like to  see  research designed  to  give  all  the salient
kinetic  parameters  (K[r), V    , kp  and  kp) in  a single experiment  or  set of
experiments.    It  is  desirable  to  get  the  maximum   amount of  pertinent
information  from  a limited  amount  of  experimentation,  particularly when the
inoculum tends  to  be  difficult  to obtain.  Based on the information presented
at  this workshop,  it would  appear  that there  are  two  promising  methods  to
accomplish  this goal:   (a)  progress   curve experiments   and  (b)  experiments
generating   typical   substrate  concentration-velocity   data   pairs.     Both
approaches have advantages and disadvantages, and neither  is applicable to all
chemicals  in  all  environments.   The techniques  should be chosen with special
regard  to practical constraints  such as  the ease  of obtaining the measurements
and  of  gaining  the maximum amount of  information with a minimum of inoculum,
time, effort, and cost.

     1.   Progress  curves.   Progress  curves are experiments designed to  follow
substrate depletion or product formation  through  the zero-, mixed-, and  first-
order   regions.    By  analyzing  these  experiments   using  either  nonlinear


                                      149

-------
parameter  estimation   techniques  (as  recommended  by  the  statisticians)  or
linearizing transformations of the data, the  various Michaelis  constants V
and K  can be obtained, and thus k,  and k?  can also be  calculated.   The rear x
advantage  to  this  technique   over  others  js_  that  the  kinetic  parameter
estimates can be derived from a single  experiment.  Progress curve  experiments
also   lend   themselves   to    manipulating   environmental   factors   (i.e.,
partitioning, temperature,  etc.)  which will   help  describe  quantitatively the
influence of these factors on biodegradation  processes.

     A  limitation  of   this  technique  may  be   the  difficulty  of  obtaining
saturation concentrations  for  some  substrates.   However,  this remains  to be
seen.   If  the degradation  of  xenobiotic  compounds is  enzymatically catalyzed,
saturation kinetics  should be observed.

     Obviously, we realize that progress curve experiments are not  designed as
simulations  of  complex  environments.    Rather,  they  are  tools  to  help  us
understand   the   intricacies   of  nature  with   an   eye   toward   predicting
biodegradation kinetics.

     Unlike  other types  of  experiments designed to  estimate  biodegradation
rate   parameters,  the   assumptions   in  this   experimental  approach  are
straightforward.  Simply  stated,  decreases in reaction velocity with time are
due only to  decreasing  catalyst  saturation,  and  the  biodegradative process is
unlinked to  significant  changes  in  catalyst   concentration  over the  course of
the experiment.

     The latter assumption presumes  that the   particular biodegradation process
is unrelated to significant microbial  growth.  Of course, this assumption will
not always be the case, and microorganisms  will  proliferate as a result of the
metabolism of  a  xenobiotic carbon and energy source.    However, we  feel that
the   recommended   approach  and  ensuing   rate   estimations   are  basically
conservative.   There  will  be  a  certain amount  of error associated with any
rate estimation,  and  from an environmental policy viewpoint,  we  feel  that it
is better  to err  on the  conservative  or low  side rather than to overestimate
the natural  biodegradation rate.   If  microbial  growth does occur  as a  result
of  xenobiotic  compound  utilization,  this  ultimate  expression  of metabolism
will  be   reflected  in   an   adaptation  phenomenon   (see  below),  and   rate
estimations can be periodically updated.

     2.    Substrate concent rat i on -velpc i ty   experiments.    The same  kinetic
parameters can also be obtained by performing typical  sTfbstrate concent ration -
velocity  experiments   and  analyzing the  results  by   nonlinear  regression or
linearizing  techniques.   These  experiments  involve   determining  the  initial
rate    of    substrate    disappearance   at    several    different    substrate
concentrations.   Either labeled or unlabeled  substrates can be used, depending
on  analytical  sensitivity required.    Typically,  such  experiments  assume  that
the  reaction proceeds  to  a  negligible extent during  the  course  of  the  assay
(i.e.,  less  than  5%   substrate  depletion).    Under  some  and perhaps  most
experimental  conditions,  it may  not  be  possible  to  restrict the  reaction to
this  extent.   At best,  these  experiments  usually employ 5 or 6 data points.
This  limitation  is  not  inherent   in  progress  curve  analysis,   since the
Michael is-Menten  expression  is  a differential velocity equation that  is  valid
over the entire concentration  range  and  as many data  points as desired  can be
incorporated into the analysis.
                                      150

-------
     Several  important  factors may limit the  widespread  applicability of the
substrate  concentration-velocity  method.    First,  radiolabeled  substrates are
often required to detect very  low concentrations or small changes in substrate
concentrations  against  a  large  background.    Considering  the  myriad  of
chemicals  and their  potential  availability  and  cost,   it  is  unlikely  that
techniques  relying on radiolabeled materials will gain widespread acceptance.

     Secondly,  parameter  estimations  from  substrate  concentration-velocity
data  pairs  often  rely  on  linearized  plotting  techniques.    Good   examples
include  heterotrophic activity experiments, which  usually use plots analogous
to the double reciprocal plot  of  Lineweaver-Burk.  However, to obtain  accurate
parameter  estimates using  this approach,  some a priori  knowledge is required,
since the  chosen substrate concentration  range is  critical  and must be in the
neighborhood  of  K  .   If  the  chosen  substrate  range  is  too  high (relative to
Km), the observed velocity is  insensitive to changes in substrate.  Thus V
can be accurately  estimated,  but K  will  be  inordinately low.   If the chosen
range  is  much  lower  than  Km, the  curve  will  intercept the graph  axis too
closely to  the origin to allow accurate Km  or Vmax estimates  (Km and V    will
be very large).  The substrate increments chosen can also critically influence
parameter  estimation.   Other  techniques  of  plotting substrate-velocity data
have  similar  ideal   concentration  ranges.    As  a  corollary  to  the   above
cautions,  saturation  is almost  never  checked  in these  experiments to verify
that one  is truly in the appropriate concentration  range.   Nevertheless, if
appropriate concentration  ranges  are chosen, accurate parameter estimation can
be made using this technique.

     Lastly,  from  a  practical  standpoint,   many  substrate  concentration-
velocity experiments  rely  on  CO,, formation and detection.   The assumption is
that primary  degradation  is   the  rate  limiting process  and  that  subsequent
conversion  to CO.? is  relatively  rapid.   However, even a cursory review of the
literature  reveals  that COo may  not  represent the  ultimate  fate  of  the test
substrate.   Primary  degradation reactions may yield intermediates  that can
bind, sorb, or are generally  more persistent  than,  starting chemicals.   Thus,
before experiments are  designed, C02 formation must  first  be correlated with
substrate  disappearance.

                                 RESEARCH  PLAN

MODEL VALIDATION

     Research effort  should continue  to  test  further  the usefulness  of the
Athens  model.  We need  to  determine  if  it  is desirable and feasible to obtain
a  k^  proportionality  constant  for  every  xenobiotic chemical  (or  class  of
chemical  if the  benchmark  approach is utilized).   Research  to date has  shown
that the  use of  total  numbers  of bacteria  as the  normalizer  for different
degradative  activities   among  microbial  populations   has   been  partially
successful.    Some   compounds   that   are  apparently  degraded   by  hydrolytic
mechanisms  give consistent k2  values when  the  first-order  rate  is divided by
total   cell  numbers.     However,  further  work  along  these  lines  is  not
recommended for two  reasons.   First, there does not  seem to be a  consistency
trend  indicating  which  chemicals  are  amendable  to  the  total  cell  number
approach.   For example,  some chemicals that are hydrolytically attacked do not
give consistent  ko  values while  others  do.   Secondly,  since  the  activity of
microbial communities is unpredictably related to cell numbers, the conceptual
basis  behind using total cell  numbers is obscure.
                                     151

-------
     We believe,  as we  have  indicated above, that  a  more direct  measure  of
microbial  activity  is  needed  to obtain  and  compare k^ values.   To determine
which activity measure to use, we recommended that a decision be made relative
to the  magnitude  and variability in  values  of the  Michaelis  half saturation
constant,   K  ,  for  a  variety  of  xenobiotic  chemicals.    In  other  words,
assumptions ^kl>out  Km used  in the  formulation of  the Athens  model must  be
carefully   and  thoroughly tested.   There  is currently  a paucity  of  kinetic
information upon which to make such decisions.   We  feel  strongly that this is
where the  research effort in EPA should be focused.

     The  conceptual  framework for a  research effort  involving  the Michaelis
half-saturation constant  is shown  in  Figure  1.   The Km  values for xenobiotic
chemicals   can  be  relatively  high or  low.   Enzymes capable  of metabolizing
pollutant   chemicals  that  closely  resemble  native  substrate(s)  would  be
expected  to exhibit  high affinities  for the  xenobiotic  compound and thus low
K 's.  Chemicals totally foreign to the environment would probably have enzyme
systems with  relatively  high  K  values.   If  K   is  consistently high for most
xenobiotic  chemicals,  then  an   adequate  measure   of  microbial  degradative
activity  for any environmental sample  would  be  the  first-order decay constant
(ki).  This would  be  an  integrative measure  representing the inherent rate at
which the catalytic unit(s) would attack the chemical, the number of catalytic
units involved, and the  factors that control  catalysis.  Knowledge that 
-------
                                      nr
                           HIGH
                            v
                        RELATIVELY
                         CONSTANT
                 LOW
              RELATIVELY
               CONSTANT
           4T—
          YES
      "T"
       NO
        MEASURE
       k1  OR Vmax
     (I^POSSlM)
      YES
                                MEASURE
                                   max
    INTERMEDIATE CASE

    (CAN USUALLY MAKE
    PREDICTIONS BASED
    ON KINETIC MODEL)
       v
   WORSE  CASE

(NO GENERALIZING
 KINETIC MODEL
  APPLICABLE)
   BEST CASE

(CAN  USE  KINETIC
   MODELS TO
 PREDICT  RATES)
Figure 1.  Flow diagragm  for  research  strategy in  testing  kinetic models  of
           biodegradation.
                                     153

-------
     If Km values for xenobiotic chemicals  are reasonably constant  relative to
V    values, then three hypotheses follow.   First,  the  average affinity of the
catalytic units for the substrate is the same  or  similar from one  environment
to another.   Similarity may only  be  apparent after certain  categorizations,
i.e., oligotrophic  versus  eutrophic  water bodies, polluted  versus  unpolluted
areas, naturally occurring  versus  exotic  chemicals,  etc.   It  may  be  that the
biochemical mechanisms associated with substrate transport  into  the  cells and
initial   catabolic  attack  are  quite  similar.   Biodegradation  processes  are
therefore  controlled  more  by  environmental   factors  and substrate  chemistry
rather than by biochemical  factors,  i.e.,  a large  array of metabolic  machinery
has  apparently  not  evolved   to  transform  xenobiotic  chemicals.    Second,
assuming  biodegradation  can   occur,   the   rate  of   transformation   will  be
controlled  largely  by  the number  of  actively  degrading  microorganisms  or
catalytic units.  This  activity will  encompass and  reflect  the  effect of the
immediate  environment.   This  activity,  as  we have  suggested above,  can  be
equated to the V    value.   Knowledge  about the relative constancy  in Km would
be a  major step in determining  if  the complexity of  biodegradation  rates in
natural  ecosystems  can  be  generalized.   If K  's are  also relatively constant
for a general  chemical class,  then perhaps all  chemicals in a given  class can
be expected to have invariant  K 's in widely  different  environments.   Third,
since ki = V   /K , a relatively  constant  K   will  mean that k, varies as does
V   .  Knowledge of either  kj  or  Vmax  can  therefore  be used in biodegradation
rate predictions.

     Information on K   will facilitate derivation of  k,,  values  (k]/V   ) and
the  testing  of  the Athens model.   Experiments  to  determine  k~  neecT to be
conducted with  a  variety  of chemicals and environmental  samples from diverse
areas.    If  k-  can  be  shown  to  be  the  same or similar  for chemicals  by
normalizing differences in degradation rates  with V    , then V    would appear
to be an  excellent  parameter for  assessing actively  degrading biomass.  Thus,
V     or  k,  could  be used  as   the  basic  parameter  estimate  for  determining
cWeria for extrapolation.

ESTABLISHING KINETIC PARAMETERS

     To be sure, there is a limited amount of information on the basic kinetic
parameters of  Km,  V   , k^, and  k£  governing  pollutant  biodegradation.   Our
current thoughts  on biodegradation kinetics  are  an  integration and  a direct
outgrowth of the information presented at the workshop.  We feel that  in order
to test models for biodegradation kinetics, efforts using judiciously  selected
model  substrates   in   a   number  of  habitats  should   be  expanded  toward
establishing the parameters mentioned above.

     The  criteria  for selection of the  model substrates include:   (a) their
potential  for analysis,  (b) their  high water  solubility,   (c)  their  lack of
significant sorption,  (d)  the  availability of radiolabeled compounds,  (e) the
amount of background information on their biodegradation, (f) their ability to
be  metabolized   aerobically   and  anaerobically,   (g) representativeness  of
different   chemical   classes,   and  (h) their  potential   for  exhibiting   a
relatively wide  range  of biodegradation  rates.  A potential  list  of  chemical
types that could  be tested  is  presented  in the workshop report.   In addition,
we   suggest   that   the   initial   efforts   start  with   three  substrates:
4-nitrobenzoate,   4-cresol,  and  2,4-dichlorophenoxyacetic  acid.      These
compounds  should be tested  by establishing their kinetic parameters  (Km,  Vmax»


                                     154

-------
ki, and  k~)  in habitats  ranging  from oligotrophic environments  to eutrophic
ones.   For example, biodegradation  in  a  polluted  portion  of an estuary may be
compared with a pristine area.  Another example would include freshwater lakes
of differing  (but well  defined)  trophic status.   These  habitats  as  well  as
others (sewage, sludge, streams, rivers, open ocean, groundwater, etc.) should
be used  to  help  determine if predictive kinetic  models  of biodegradation are
truly feasible.

     The techniques used  in  this  effort  should include  the shake flask, since
it is this  method which is the most  likely  to  be used  on a widespread basis.
The design  and analysis of these experiments  will  be  those described earlier
in this document.  Degradation studies with suspended sediments should also be
included.

ENVIRONMENTAL FACTORS AFFECTING BIOOEGRADATION

     As pointed out by  the  workshop panel  report, there are many factors that
can   serve   to   accelerate   or   retard  the   biodegradation   of  xenobiotic
susbstances.   Considering  the  multiplicity  of  these  factors,  one questions
whether or not it is expeditious for EPA to support research to quantitate the
influence of these factors.

     Our  approach  to  assessing  metabolic  activity  by  measuring  laboratory
biodegradation rates seems  to be a way  of  accounting  for these environmental
factors  without  actually knowing what  they are.    For  instance,  such results
will  tell us if a biodegradation rate is relatively slow or fast.   Even though
such experiments  will  not allow us to conclude why this  is so.   We feel that
by  measuring  rates we  can  successfully  integrate  the  many  environmental
factors  influencing biodegradation  and  can  limit  the  number of false positive
indications of biodegradation.

     It  is  fully  recognized  that  it is desirable  to learn as much  as possible
about  the  factors  controlling  biodegradation  rates  in  the  environment.
However, we  believe that pragmatically a detailed  knowledge  of these factors
is not   necessary  for   rate  predictions.   As  more and  more  factors  become
elucidated,   quantitative   relationships   describing   their   influence   on
biodegradation will  be established.   The   basic  biodegradation  kinetic model
will    thus   grow  in   degree   of  sophistication   to   accommodate   these
relationships.   However,  for each new factor  described,  field information on
when  and where that factor  will  have  its effect must be obtained to make more
sophisticated  predictions.    The  subsequent  monitoring  effort  will  increase
exponentially with  each new factor or concern.   Our  philosophy suggests that
research  efforts must  initially  be  oriented in  the  direction  of  finding
parameter estimates  that integrate the effect  of these  environmental  factors
into  a  small  number   of  measurements.    We  feel  that  either  V     or  k,
measurements  lend themselves to  this goal   and  therefore  should  be  strongly
emphasized.  If this concept does not eventually prove useful, we will perhaps
have  to  investigate more intensively the factors that  control biodegradation
rates  in  order  to  find  a  better  quantitative  basis  for  environmental
predictions of biodegradation rates.

     Work  should also  continue  on the influence  of  environmental  factors,
because  the  resulting  information  will  eventually help extrapolation research
(see  next   section).    It  may  also provide  the   basis  for  experiments that


                                      155

-------
attempt to  manipulate  the environment  to  improve biodegradation  capacity  or
rates  for  the   purposes  of  eliminating  certain   in-place   environmental
pollutants.   Dr.  Raymond's paper  describing the use of  nutrient  additions  to
accelerate hydrocarbon  degradation in  ground water is  an  excellent example.

ADAPTATION

     Numerous  published  investigations and  extensive  discussions  at  this
workshop  have  indicated  that  many   xenobiotic  chemicals  do  not  degrade
immediately when added  to an  environmental  sample.  A  lag or adaptation period
of  varying  lengths is  observed  before a chemical  is  rapidly degraded.   The
extent of the adaptation  period directly affects  the  exposure of the biota  to
the pollutant.   The  incidence of adaptation and  the  mechanism(s) controlling
adaptation  should  therefore  be   as   extensively  studied  as  biodegradation
rates.    Therefore,  lag  phases   must  be  considered   in  possible  regulatory
decisions.

     At this point,  the length of time before  the  onset of rapid degradation
is impossible to predict since the phenomena is  poorly understood.  Therefore,
it is  difficult  to model  this  portion  of  a  decay curve.   Probable reasons for
adaptation  include the  requirement  for bacterial  growth  before significant
parent chemical  is transformed,   specific substrate concentrations before the
requisite  enzymes are  inducted,  the   need  to  deplete  competing substrates
before  the  compound  of  interest  degrades,  or  the need  to   exchange  the
essential  genetic  material.   Other mechanisms of adaptation  can  certainly  be
postulated.

     At present,  a generalizing descriptive  model  for  adaptation has not been
proposed.    This  is  largely  because   of  an absence  of  information  on  the
mechanisms  controlling  adaptation.   However, the adaptation  event itself can
be readily  observed experimentally.  Adaptation,  or the  length of time before
accelerated biodegradation occurs, is an integrating parameter which, like the
V     value,  does not  require a knowledge of  the exact adaptive mechanisms  or
the factors controlling it.  We therefore believe that research efforts should
investigate  the  length  of   the   lag   or   adaptation  period.     Three  major
approaches   should  be   considered.      First,  the   effect   of  substrate
concentrations on  the  length  of  the  lag period  should be cleary  defined for a
variety   of  chemical   types   and  environmental   samples.     Arguments  that
adaptation     will     not     occur     at     environmentally     realistic
concentrations (ug/1) must  be  experimentally  substantiated.     Second,  the
relationship of  adaptation to the growth of organisms  capable of  degrading the
substrate must  be investigated.    This effort is needed to establish whether
lag  periods are simply  a  function  of  the initial number of  degraders in the
environmental  sample  and  their   eventual  growth rate in  that  sample.   This
assumes   that   if the   degraders  are  present  they  will  commence  growth
immediately following exposure to the  substrate.  Examples where  degraders are
present   but  never  grow  are  assumed  for  now to   be infrequent   and  may
consequently  form the  basis  for research  in  the  future.    If  growth can  be
quantitatively  linked  to the lag period prior  to adaptation, the development
of descriptive models becomes feasible.

      Third,  the  natural  variability in adaptation  time  should be examined to
determine   if    regulatory   decisions   can   possibly   be   based  on   certain
consistencies  in the adaptation  response.   For  example, will  the  adaptation


                                      156

-------
time for  a  specific xenobiotic  chemical  always  be the  same  if environmental
samples are  always taken  from  the same  general  area  or does  the  adaptation
time always vary widely  in  time  and  space?   If consistencies can be observed,
it  will  be  important  subsequently to  relate  them to  adaptation  mechanisms,
chemical structure trophic conditions, topographic and hydrogeologic features,
and previous exposure to pollutants.
                                 EXTRAPOLATION

     Extrapolation of laboratory data to the field is a simple concept that is
poorly   defined.      Extrapolation   means   different   things   to   different
investigators,  and a  clear  plan  that  outlines  the  research  questions  and
suggests practical, experimental  approaches  does  not exist.   In an effort to
correct  this  situation, we  propose  a  research  strategy that  would involve:
(a) establishing the "environmental  relevance" of laboratory-derived data, and
(b) categorizing the variation in biodegradation rates in natural environments
in such a manner to permit predictions of exposure concentrations.

ENVIRONMENTAL RELEVANCE OF LABORATORY DATA

     It  is  frequently  argued that  xenobiotic  chemical  biodegradation  rates
obtained  from simple  laboratory tests  (e.g.,  river die-away,  shake flasks,
etc.) are environmentally  insignificant  since,  by  design, these tests replace
the   complexity   associated  with   natural   environments   with  expediency,
standardization,  and   reproducibility.    Laboratory  tests  are  important  for
regulation and chemical registration,  but  it  is  not  clear how accurately this
information reflects what actually happens in the field.  To be  sure, industry
cannot examine every chemical in a laboratory system that simulates  real  world
conditions,   but   information  can   be  supplied  that   will   delineate  the
qualitative  and  quantitative  degree  to  which  a  simple   laboratory  method
represents field events.  Simple awareness of the deficiencies of a  laboratory
test  may  be  very useful   from a  regulatory  or  extrapolation   viewpoint.
Modifications of  existing tests  or  the addition  of supplemental  tests will
likely be the most productive outcome of  this  type  of  relevancy study.   Such
modifications will  greatly improve  our  confidence in  making fate predictions
from laboratory information.

     We propose that  the  environmental  relevance of  laboratory  test data can
be  most  efficiently  determined  using  microcosm  studies and  field studies.
Based on recent information,  microcosms  appear  to  provide good  simulations of
the  field.    They  can  be dosed  indiscriminately,  and they  are  isolated from
uncontrollable weather conditions.  They certainly provide a means of studying
biodegradation  processes   under  complex  conditions.     Microcosm  studies
therefore,  in our  opinion,  are the  methods  of  choice  for  such  relevancy
studies.  Whenever possible,  the fate  of a xenobiotic chemical  should also be
examined under  field  conditions.   Only experience  in both  types  of studies
will determine our confidence in and limits to the the use of microcosms.

     A research plan  involving  microcosms  and field  studies  would  consist of
using biodegradation information from simple laboratory tests to see if it can
adequately explain  the  biotic fate  of a chemical in  more  complex  ecosystems
such as microcosms or field  studies.   If this  hypothesis proves correct, then
we believe that the  environmental  relevance of the  laboratory test data will
have been established.
                                     157

-------
     If it cannot, then manipulation of the complex system or the isolation of
component parts  thereof  may indicate  how  to  modify appropriately  the  simple
laboratory test  so  that  it  will  accurately describe natural  events.   In some
cases, complexity  may control  biodegradation  rates to  such an extent  as to
prohibit  reduction  to simpler  terms.    If this  is the  case,  a  routine  and
standardizable complex laboratory test  system must eventually be developed and
tested.

     The  use  of laboratory data   to  explain  events  in  complex   systems  is
complicated by the  fact  that biodegradation will  not  be  the only fate process
affecting  a   chemical.      Dilution,    sorption   to  sediments,   diffusion,
bioturbation,  bioaccumulation,  photolysis, hydrolysis, volatility,  etc.,  may
also influence the  fate  of  the chemical.   Recent  investigations indicate that
the  use  of   appropriate controls, physical   and  chemical  tracers,  unique
analytical   techniques,   and  simple   mathematical   models  will   permit  a
determination  of the  biotic fate of a chemical in a  complex system and will
allow a comparison  with  information generated  in  simple  laboratory tests.  We
recommend  that  these research  efforts  be  continued  and  expanded.    Test
chemicals have to be  selected  carefully,  and  ancillary  experiments will have
to  be carried  out  for   additional  information.    Microcosms  that  simulate
specific portions of  ecosystems  need to be developed and tested.  Appropriate
field  sites  will have to be  located.    Dosing, analytical,  and data analysis
procedures need  to  be developed.

     We feel  strongly that microcosms  and  field studies have relatively  little
predictive value.   They  are  generally  too  small to be truly representative of
any  large  ecosystems.   They are instead excellent  validation tools.  Kinetic
models  for  biodegradation,  sites of  active biodegradation,  and  factors that
control  biodegradation  rates   can   be  tested in  these   systems   and  their
environmental  significance determined.

VARIATIONS IN  BIODEGRADATION RATES

     Estimating  biodegradation rates  using laboratory  studies requires, as we
have  proposed  above,  the determination  of V    value (assuming a need for a
site-independent kinetic  constant   k2)  or  tne   first-order  rate constant.
Environmental  samples are usually collected from a  field site for  this purpose
and  then used  in  a  laboratory  biodegradation test  system.    The resulting
estimation of  biodegradation,  k-,,  or  the  activity  of  the microbial community
(V    )  is representative of  that   site  only.   Its  representativeness  of the
rest  of   the  ecosystem  is  unknown.   The  values  are  likely to  vary widely,
depending on when and where the  sample is  taken.   Little appreciation for  this
natural  variation  currently exists.    Is  it  so  extensive  as  to  force us to
analyze biodegradation rates over very small spans  of  space  and time?   Perhaps
biodegradation  rates  or  the  activities of  microbial  communities  responsible
for  the  biodegradation  are  not  as  widely varying  as  we  suspect.    Or  perhaps
variability  can  be categorized  relative to  certain general charactristics of
the   ecosystem.      Will   eutrophication   tend   to  decrease   or  increase
variability?     Will  different   degrees  of  pollution  or  prior  exposure to
xenobiotic  chemicals  affect variability, or  will the  absolute  rates  just
increase  or  decrease?    Thus,  we   believe that an  extensive  research  effort
should be centered  on measuring  the variability in  biodegradation  rates.   This
appears  to  be  the  only way  that  we  will   be   able  to  truly   extrapolate
laboratory data  to the  field, to extrapolate biodegradation information  from


                                      158

-------
one  environmental   site   to  another,  and   to   extrapolate  biodegradation
information about one chemical  to another in the same chemical class.

     We suggest that the temporal and spatial  variability of k1 or V    in one
typical ecosystem  (estuary,  river,  lake,  etc.) be  established.   Statistical
designs, which  are  readily available  and  adequately tested,  will  have  to be
employed  for  the  most efficient  estimation   of  this  variability.    In  that
ecosystem, the  largest area  that will consistently  give biodegradation  rates
within a  pre-specified range must  eventually  be  determined.   In  an estuary,
for example, the area considered might include the water (at an average depth)
and sediment for the main embayment.  Determination of the range will be based
on the actual  observed  variability,  on the  toxicological  properties  of the
chemical, and on certain ecosystem  characteristics.   The objective will  be to
determine the  frequency  and  size of "hot spots"  or  "low spots,"  if any.   The
range  in   biodegradation   rates  tolerated   will   essentially  control   the
significance of hot  spots.   For  example,  one  portion of the main embayment in
an estuary  may consistently  give degradation  rates  of  V     values that  fall
outside the observed variability  in  the  rest  of the  embayment.  This hot spot
may not,  however,  be  significant if  one is willing  to  accept the additional
variability.   Acceptance  or  nonacceptance  will  again depend  on  the observed
data,  the  chemical   being  tested,  and  certain  regulatory  considerations.
Efforts  must  then  continue  to  identify  hot   spots  in  other corners of the
ecosystems  and  during  other   seasons  of  the  year.    The   maximal  use  of
statistical design will be crucial for optimized determination of variability.

     With  an   estimate of  the  variability  for  the  biodegradation  of  one
chemical  in  one ecosystem  at  hand,  cluster  analysis  of  the  data  should be
performed to determine the minimum  number  of  measurements necessary to detect
similar variability in any new ecosystem.   If  the number is too large then the
variability in degradation rates in aquatic environments is too extreme for us
to make  generalizations  over  large areas  within a ecosystem.   Degradation
rates  become very  site specific, and  extrapolation  of  laboratory  data to the
field  is  unfeasible.   However,  if the  number  of  required  measurements is
reasonable, then examination of  the  temporal  and  spatial variability in other
ecosystems should be carried out.

     We  propose  that  two  or  three  geographically  different field  sites be
chosen for each major ecosystem type (e.g., rivers, estuaries, streams, lakes,
etc.)  to  measure the  variability in  k^,  Vmax>  or  Km«   For  example,  two or
three  different  lakes  could  be  selected  in   different  geographic  areas  to
represent  this  ecosystem  type.   The  selection would  be  based,   in  part, on
accessibility  for  experimental  purposes  and,  in   part,  on  environmental
characteristics  of  the  field  site.   We believe  that  initially  field  sites
representing  extremes  (i.e.,   polluted  versus  pristine,  eutrophic  versus
oligotrophic)   should  be examined for  their  variability.   If  variability is
"similar" for each of  these  extremes,  then  it  can be argued that site-to-site
extrapolation  is  feasible within certain   limits  of error.    A chemical, for
example, may show an average  half-life of X days plus or minus some error term
for one particular site of bounded environment.  If variability can be assumed
to be  the same  from  one  site to another, the  average half life for a chemical
determined at a new  site  can be  assigned the  same variability as the original
site.
                                     159

-------
     The "large number of chemicals examined by  OTS  precludes  an assessment of
variability  in  biodegradation  rates  for  each  new  chemical.   Some  form  of
structure activity relationship  is  obviously  necessary.  We  propose  that the
biodegradation rates  of structural  analogues within  a given  chemical  class can
be ranked from the fastest to  the  slowest.   The consistency of the ranking(s)
should  then   be  examined  using samples  taken  from several   field   sites  at
different times.   If  one  were to examine  the  chemical class  of chlorinated
biphenyls, for example, would  monochlorinated biphenyls always degrade faster
than dichlorinated biphenyls, regardless of the  source of the inoculum used in
the  biodegradation  experiments?   Would  4,4-dichlorobiphenyl  always degrade
faster  than   3,4-dichlorbiphenyl,  etc.?    A  consistent  ranking  would  prove
invaluable.   It  would  mean  that the  variability in biodegradation  rates in
time and space for one representative of a chemical  class (i.e., a "benchmark"
chemical) could be linked to other chemicals of  that class through the ranking
analysis.  Consistent  rankings may only be possible  if different subsets are
established.   Ranking  using  samples  from  polluted areas may be different from
those using  samples  from nonpolluted  areas.  Extensive eutrophication may also
result  in  different  rankings.   Likewise,  rankings  using  samples  from  river
water should  not necessarily  be expected  to agreee  with those using  estuarine
samples.

     We  suggest that  the  major chemical classes  of  concern to OTS and EPA be
identified and a representative member of each class be selected.  The partial
list  of chemicals discussed  at this workshop  is an excellent  start in this
direction.   Work can  then  begin to establish  consistencies  in the rankings of
their biodegradation rates with other chemicals within  those  classes.
                                      160

-------
Appendix:  Proposed Benchmark Chemicals for Biodegradation Research

     The following  chemicals  were  recommended by an  ad  hoc  committee who met
briefly on the last day of the workshop:

    1.  pentachlorphenol
    2.  p-nitrobenzoate
    3.  2-mercaptobenzothizole (MBT)
    4.  pthalate esters
    5.  phenylcarbamate or thiocarbonate
    6.  decalin or phenanthrene
    7.  phenoxybenzoate
    8.  ABS
    9.  carboxyazobenzene (or similar)
   10.  polyethylene glycol
   11.  Dalapon
   12.  tertiary or quadamine (EDTA or  NT)
   13.  cyclohexane carboxylic ac.
   14.  substituted pyridine
   15.  parachlorobiphenyl


     The criteria for choosing a benchmark chemical were:

    1.  Range of physical properties  (e.g., anionic,  cationic, hydphobicity)
    2.  Range  of  chemical  structure  (e.g.,  heterocyclic  -  aromatic  and
        aliphatic branched, polymeric,  straight chain)
    3.  First-order fate process -- biodegradation            *
    4.  Variety of initial reaction mechanisms
    5.  Safety in handling
    6.  Availability and cost
    7.  Environmental relevance
     The  purpose  of using benchmark chemicals was  to compare laboratory test
systems and to test the integrity of those systems.
                                      161

-------