-------
Figure 3.
HIGH SALINITY ZONE (25-30%o)
200-
100-
•%
£•
s o-
«
Vaucluse Shores, Va.
*-.- -.
/ • Zostera
* n.
j'. \ •
/ *±& $ LW,
f ? ^^vS^ /
/ J**~'^ ^^\/
i T i i i 1 l i i i l i
'£ 1979 1980
^ MID-SALINITY ZONE (I0-20%o)
co Eastern Bay S Choptank R.
^ 100 n
«g 1WW
o
OQ
a
z
— ^ _ -.
o 50-
ar
0
1
Ui
S 0-
o •
M _, Ppectinatus Q .». -: -^ '977
/ R.maritima :'{* : \ Data
/ P pgrfo/kjfus/ ' * \ \
i* \ ' i ';
-------
1
1
1 j
3-
1
1
Io
s >-
ir
i §
^ f\
_, (n 0
1 S
s
1
1
0.5-
1
In
1
— 2 -
<
1
I/"\
o!
™ Figure 4. Seasona
species
•zones.
salinit
1981.
1
(a) HIGH SALINITY ZONE
Vaucluse Shores, Va.
j\
\
I \^R. maritima
\ A ~~Z. marina
vv'"*"KA\
' * T3
*" '
f\ /^
1 1 1 1 1 1 1 1 1 1 1
1979 1980
(b) MID-SALINITY ZONE
Choptank River „
P perfoliatus-*
m
• ^Ruppia
p'
^^y^" Myriophyllum £
^^^jj-*^
I i I 1 I 1 1 1 1 1 1 1
1977 1977
(c) LEAF -AREA INDEX, LAI
High Salinity Zone :••...-••••••
• •••" \. '-.--Mixed -: . /v
^ - >x\ / A
Zostera~~ ^ '••.. /-. ^ f
s~—J>\ '•' -s'' ' 1
/~~^^^a
j ' M M'J'S'NIJ'M'M'J'S'N'
1979 1980
L! patterns of root: shoot biomass ratios of selected
of SAV for (a) high salinity and (b) mid-salinity
(c) Shows the seasonal pattern in LAI for a high
.y area. Data from Kemp et al. 1981, and Wetzel et al
442
-------
observations confirmed that below-ground structures of SAV in the upper Bay
appeared poorly developed relative to those in other areas of the Bay and
very poorly developed relative to tropical Thalassia meadows. The low root
to shoot ratios recently observed in the upper Bay may be indicative of
stressed plants.
The leaf-area index (LAI), or the amount of photosynthetic surface per
unit of biomass, is a fundamental characteristic of SAV community
structure. LAI differences between SAV communities demonstrate the
importance of light in regulating SAV communities and their adaptability to
different light regimes. Increases in plant density can lead to potential
increases in photosynthesis to a point, but can also lead to decreases in
light availability through mutual shading. Data reported by Wetzel et al.
(1981) exhibit differences between different SAV communities. Average LAI
values are greatest in a mixed Zostera and Ruppia bed followed by
successively lower values in pure stands of Zostera and Ruppia (Figure
4c). At the study site, the Ruppia bed was located in shallow water; the
mixed and Zostera beds were at successively greater depths. The authors
attribute the pattern in LAI to differing light regimes in these areas: the
shallow Ruppia bed may have been photo-inhibited; the mixed bed near to
optimal; and the deep Zostera bed intermediate because of insufficient
light. Dennison (1979) reports a similar pattern for a Zostera bed and
underscores both the importance of light in regulating SAV communities and
the adaptability of SAV to different light regimes.
In addition to different mean LAI values, SAV communities exhibited
differences in the vertical distribution of these values. In the Ruppia
bed, values were greatest near the bottom of the canopy, presumably because
of photo-inhibition nearer the surface. In the deeper, mixed, and Zostera
communities, maximum LAI values were observed closer to the surface,
probably owing to reduced light availability at greater depths. The
maximum LAI values observed by Wetzel et al. (1981) were on the lower end
of values reported for other seagrass communities (Jacobs 1979, Aioi 1980,
Gessner 1971), suggesting that at least in these beds self-shading was not
a major factor limiting light availability.
Estimates of seasonal net production rates (Pa) for SAV communities in
the upper and lower Bay are summarized in Figures 5 and 6. In the
mid-salinity environment, values of Pa correlate well with temperature,
light, and SAV biomass. In general, rates were high during July and August
when SAV biomass, light, and temperature were high and decreased sharply to
lower values during the colder months. Figure 5 emphasizes the difference
in community net production in vegetated and non-vegetated littoral areas.
Clearly, during those periods of the year when SAV is present (May to
September), the rate at which new organic matter is created is considerably
higher in vegetated littoral areas.
Additional insights concerning the metabolic characteristics of SAV
communities can be gained by comparing the ratio of Pa (new organic matter
created during the day) to respiration (Rn: consumption of organic matter
during the night). Data indicate that ParRn is greater than 1.0 during the
early SAV growth periods and that Pa:Rn is less than 1.0 during the late
summer and fall. This observation suggests that most SAV biomass is
generated in the early growing season; during the summer and fall, high
daytime rates of Pa are observed, but the daily net production is consumed
during the hours of darkness. Essentially, the metabolic demands of
443
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
•o
cu
o"
o
13
Q
O
13
2
2
O
O
10-
5-
(d)CHOPTANK RIVER
TOODS COVE
SAV
-COMMUNITY
NON-VEGETATED
COMMUNITY
j ' F'M ' A'M'
o1 N' o1
I960
5-1
(b) EASTERN BAY
PARSON IS.
SAV
COMMUNITY
NON-VEGETATED
COMMUNITY
J ' F' M' A
A' S'
1979
-2 -1
Figure 5. Net SAV community production in gCLm d , including
estimates of non-vegetated community production for (a)
Todds Cove and (b) Parson Island. Data from Kemp et al.
1981.
444
-------
I
non-photosynthetic organisms, or heterotrophs, and plant respiration
matches or slightly exceeds the generation of new organic matter. •
Estimates of macroscopic heterotroph abundance (infauna, epifauna, and •
finfish for example) correlate with this pattern in that general abundances
are low early in the growing season, but increase rapidly (as do metabolic
demands) as the season progresses. B
Seasonal patterns of Pa for a Zostera and Ruppia community in the lower •
Bay are given in Figure 6. Distinctive patterns emerge for each
community. Rates were high in the Zostera community in the spring and fall •
with a summer minimum, but rates were highest in the Ruppia bed during the •
summer. The seasonal shifts in maximum growth may partly explain the
successful coexistence of these two species. The values given in Figure 6 •
are in hourly units derived from measurements made prior to midday. I
Afternoon values were generally lower and often indicated a heterotrophic
condition (Wetzel et al. 1981). The reason for this strong diel pattern in
Pa is not known, but nutrient or CC>2 limitation is suspected. B
ANALYSIS OF THE COMPONENTS OF SAV COMMUNITY PRODUCTION
This section places the various autotrophic components into perspective B
by comparing the relative contribution of organic matter produced by
various autotrophic components of SAV beds, including epiflora, macroscopic
algae, and benthic flora. Each component contributes a certain amount to B
the overall production of the community and provides a more or less B
desirable food source for the associated heterotrophic community.
Because of technical problems, temporal and spatial variability, and •
the time-consuming nature of the measurements, there appear to be only a B
few such studies available with which to compare results obtained in
Chesapeake Bay. Estimates of production and biomass attributable to ^
various autotrophic components of SAV communities are given in Table 1. B
However, from areas outside of Chesapeake Bay, available data suggest that »
epiphytes and macro-algae constitute a significant and, at times, a
dominant feature of SAV community production and biomass.
Data from Chesapeake Bay are preliminary, but inspection suggests that
epiphytic primary producers can constitute a substantial portion of the
total community Pa. As we have shown earlier (Figure 5), phytoplankton
production can also substantially contribute to overall SAV community
production. There is little data to suggest that epiphyte or macro—algae
constitute a substantial portion of community biomass. —
One of the problems in interpreting these data involves the high •
variability associated with measurements of benthic and epiphytic ™
production rates (Murray, pers. comm.). Apparently, short-term (day-week)
changes in bottom sediments due to wave and tidal action can radically
change benthic and epiphytic community structure and associated rates.
Thus, estimation of seasonal or annual importance is particularly
difficult. However, preliminary evaluations suggest significant, although ^|
not dominant, roles for epiphytes associated with SAV. H
SAV PRODUCTION IN THE CONTEXT OF ESTUARINE ECOSYSTEMS
I
The importance of SAV production can also be assessed in terms of its
contribution of organic matter to an estuarine system. In the shallow M
445
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
JAN
MAR
MAY JUN
MONTH
SEP
NOV
Figure 6.
-2 -1
Net SAV community production in mgO m d dominated by
(a) Zostera and (b) Ruppia in the lower Chesapeake Bay.
Dots are mean values; bars are standard deviations.
446
-------
OO CO
CO
CO
«3J
rS
o
M
S3
s~\
^
*~*
P
3
rH
1
CM
1
a
0
00
A
X-X
tfl
^
2 co
O W
M M
H H
CJ H
P 2
P P
q §
PH S
P^ O
CJ
cfl >
V— ^ ^4
cn
P-*
O Pu
o
o
^ C/3
rH H
J2J JZI
O Pd
M 2
H 0
M PL,
r, Np4
PH O
PS
3 2
O Pd
CO P4
Og
W Pi]
H pq
CJ
W /->
Pd M
CO 2
^
33 p
O Pi
Pi O
pLl ^^
•^f\|
HI -
o} B
P (30
•
i— 1
w
H
CU
CJ
B
01
01
(4-1
s
g
JJ
y{
C
cfl
rH
0.
O
4J
^
f!
PM
01
Cfl
00
cfl
1
o
rl
U
01
cfl
00
CO
1
o
M
a
•H
e
a
•H
r\
4J
C
Q)
PQ
CO
cu
4-1
pXj
J=
CU
•r4
a
u
CO
co
cfl
M
00
cfl
01
CO
ca
cu
iH
o
cu
CU
CO
C
O
•H
4-1
cfl
O
O
CO
• ON ON
S rH rH
Q
O • •
O rH rH
O • cfl cfl
r-» r- en
oo I-H vo cn cu cu cu
\D ON rH CU
CTi rH i-H >-^ H rl
rH rH 0) 01 0)
Cfl rH rH >»>•>,
CO J3 01 tfl tfl 01 Q)
oi w N ,e r< 6 6
fl H 4-1 C rl 3 3
O cfl 01 01 3 cfl cfl
•-> 53 3! CM 33 MM
0 0
1 1 rH 1 CTI . .
1 1 ON 1 O rH rH
II 1*11
o co m
• •
O 0
II 1 1 1 1
II 1 1 1 1
II 1 1 1 1
1 1 r^ i LO i i
1 1 <£> 1 0 1 1
1 1 CN I -II
0
1
O O co r>- i i
O CM P^ rH | |
CM -II
O
0
•d
o
•H
CM O M
. . a)
O 1 00 O 00 CN CO CU
O 1 CN CO -vT 1 1
O 1 CO • in rH J-l
rH O • 0)
C^ H
CO 3
CO 3 CO
3 W
4-1 cfl c!
cfl cfl iH -H
•H cfl C rH
CO CO cfl cfl -H O CO
tfl Vj CO t-i M 4-1 M-i CU
CflCUiHCUOlCJM •*->
i— 1 4_> (X 4J J-> 0) 0) Cfl
cflcQcucacacucu B
,C O 3 O O • • -H
HNPiMNCUCV. 4J
CO
cfl CU
•H cfl >,
CO rH pq rH
T3 • 0 iH
•H . IM ^i • tfl
M 0) iH cfl CO P
O CO rH CJ 01
rH Cfl Cfl • ,£« ^N
PM S O 2 CJ cfl
01
CJ
B
CU
rl
0)
O
p*
4-1
,*
C!
cfl
rH
CU
O
4-1
^
rf
PM
CU
cfl
00
rH
Cfl
1
o
o
2
01
cfl
00
rH
Cfl
1
o
M
CJ
•H
0
U
•H
&
4-1
B
CU
pa
to
01
4J
(-^
P.
•H
CX
co
CO
cfl
M
00
cfl
CU
CO
CO
0)
•H
CJ
01
P.
CO
rj
o
iH
4-1
cfl
a
o
*
&
U rH rH
01 00 00
S ON ON
S rH rH
CB • •
r-. rH rH
TJ P** Cfl CO
B O CT>
Cfl r% rH 4-1 4J
ON OI 0)
rl rH 01
CU O I-H rl r<
B vo >> cfl a) oi
CO ON o ,C > >
W rH ftf B Cfl Cfl
01 CJ 01 -U 4-1
P 33 PM co co
1 III
1 III
1 III
sr
•
m co CJN
4-1 tO
ca ca Tt
01 B rH
CJ Cfl Cfl iH O
O M )H 4J IM
*rj ^j CU O M
O 4J 4J 01 OI
3 CO CO CU CU
fx, O O ' •
cj N N PM PM
MH Cfl
O B >•>
t-i 3t CO -H cfl
Cfl Cfl Vl rH PQ
01 cfl N > OI O
CU ^ B Ctf J3 M •
O CO -H rH 4J Cfl W
I-l cfl M M O CJ CU
3 rH • ,fi
447
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
estuarine systems near Beaufort, N.C., studies were conducted on four major
primary producers, with results pointing to the potential importance of
seagrasses. Wetzel et al. (1981) summarize these studies and report annual
productivity estimates of 66, 249, 330, and 73 gCm~2y-l for
phytoplankton, salt marshes, Zostera, and SAV epiphytes, respectively.
Orth et al. (1979) report that SAV is an important autotrophic component in
certain areas of the lower Chesapeake Bay. Stevenson (personal
communication) estimates that some 40 percent of in situ production in
Chesapeake Bay could be attributed to SAV in 1963, but only six percent
could be assigned to SAV in 1975. Decreasing SAV abundance, especially in
the upper Bay, make present estimates of SAV contributions to in situ
productivity even smaller than the 1975 figures.
Estimates of relative seasonal contributions of sources of organic
matter to the upper Bay provide insight into the seasonal stability of the
food supply to food webs and form the basis for further evaluation of the
nutritional quality of the various sources (Figure 7). These trends were
developed from various kinds of information including the work of Flemer
(1970), Biggs and Flemer (1972), Heinle et al. (1977), Kemp and Boynton
(1980), Taft et al. (1980), Kemp et al. (1981), and Wetzel et al. (1981).
Figure 7 suggests that because of the diverse sources, the organic matter
supply to the upper Bay is relatively constant throughout the year and may,
in part, explain the high productivity of the estuarine system (Nixon
1980). During the late winter and spring, it appears that upland drainage
is the dominant source of most organic matter; in late spring and summer,
phytoplankton production assumes a dominant role; in early fall SAV may
have been an important source in the past; and in the winter the input of
marsh vegetation via ice scouring and transport to the Bay may be important
in some regions. Benthic microalgae are probably not significant primary
producers due to the typically short euphotic zones encountered in the Bay
(light limitation) and the high rates of sediment deposition and
resuspension that deter community development.
COMPARISON OF SAV WITH OTHER MAJOR SOURCES OF ORGANIC MATTER TO THE BAY
A simplified organic matter budget is presented in Table 2 for the
upper portion of Chesapeake Bay (upstream at the mouth of the Potomac
River) for two periods (1960 and 1978). SAV was a distinctive and
quantitatively important feature of the Bay during the early 1960's, and
severely restricted in 1978.
This budget indicates that SAV may have been an important source of
organic matter to low and mid-salinity portions of the Bay. During the
1960 period, we estimated that phytoplankton production was comparable to
SAV production, and each of these was larger than riverine input. Some
evidence indicates that between 1960 and 1978 both phytoplankton production
and riverine input increased (Heinle et al. 1980, Boynton et al. 1982),
with SAV production much lower during the late 1970's. Our estimates
indicate that in the upper portion of Chesapeake Bay, SAV contributed about
30 percent of organic matter production during the 1960's when SAV was
abundant, and on the order of four percent in 1978.
Because of high variability, these estimates are only guides as to the
relative importance of various sources of organic matter in the low and
mid-salinity portions of Chesapeake Bay. Considerable year-to-year
448
-------
HIGH-
03
tr
UJ
<
2E
o
z
IT
O
LOW-
. DRIVER INPUT ^—^^PHYTOPLANKTON
/ \-MARSHES
ALGAE
I I i I T I
WINTER SPRING
SEASON
SUMMER FALL
Figure 7. Hypothetical seasonal pattern and availability of organic
matter to the Chesapeake Bay food web.
449
I
I
I
I
I
I
I
I
1
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Table 2. ESTIMATED MAGNITUDE OF THREE SOURCES OF ORGANIC MATTER TO
CHESAPEAKE BAY FOR TWO TIME PERIODS (ALL VALUES ARE IN UNITS
OF
Source3
Phytoplanktonb production
SAV productions
Riverine inputd
Time
1960
3.8 (56)
2.2 (33)
0.8 (11)
6.8 (100)
Periods
3.
0.
0.
4.
1978
8 (79)
2 (4)
8 (17)
8 (100)
a Includes area of Bay and tributaries above the mouth of the Potomac
River (1.5 x I09ra2).
b Annual rate of production estimated at 250 g C m~2y-l (Flemer 1970).
c Annual production estimated at 360 g C m~2y-l based on rates
reported by Kaumeyer et al. (1981) and Wetzel et al. (1981) for a
180-day growing season. Areal distribution of SAV estimated at
6xlQ8m2 in 1960 (Rawls, in prep.; Stevenson, pers. coram.) and
0.7x10^2 in 1978 (Anderson and Macomber 1980).
d Riverine input of organic matter from Biggs and Flemer (1972).
variability in the absolute amounts delivered from riverine sources occurs
and probably varies by a factor of one to two. Boynton et al. (1982) have
also shown that phytoplankton productivity in the mid-salinity portion of
Chesapeake Bay can vary by as much as a factor of three. The year-to-year
variability in SAV productivity has not been evaluated, although
observations by Orth (1981) indicate that there is some degree of
fluctuation. Despite the probable errors involved in this calculation, it
seems that in the early 1960's SAV was a significant autotrophic component
in the upper Chesapeake Bay, but at the present time is a minor component.
An important ecological consequence is the possiblity of less food for
higher trophic levels such as fish.
FOOD-WEB UTILIZATION OF SAV
SAV can enter animal food webs either through direct grazing of living
plants or consumption of SAV detritus at some point in decomposition
processes. Several techniques have been used to establish the degree to
which SAV is used as a food source but, unfortunately, each has substantial
limitations. The most widely used technique is direct visual
450
-------
I
identification of material in the digestive system. This technique is
relatively simple, but analyses are often time-consuming, and the degree to •
which food items can be identified is often limited to larger items that •
are resistant to digestion. A second approach involves a relatively
expensive chemical technique in which the ratio of stable Cl2:cl3 H
isotopes is determined for both plant food items and associated predators. I
The technique is based on different plant groups having characteristically
different Cl2:cl3 ratios. Animals feeding on a particular plant will, M
in time, approximately reflect the food source ratio. This technique is of •
limited value in establishing SAV food-web relationships because there are
several primary producers associated with SAV communities, each of which
has a distinctive Cl2:cl3 ratio (Bunker et al. 1981b). M
In spite of these limitations, several substantial results have emerged •
using these techniques to link SAV production to utilization in Bay food
webs. Perhaps the most definitive linkage is between SAV and waterfowl. fl|
Direct grazing on SAV by waterfowl seems to be important both in Chesapeake •
Bay and elsewhere, and grazing in itself can impact the distribution of SAV
locally. McAtee (1917) reports that generally, SAV is excellent food for
waterfowl, that leaves, stems, roots, and rhizomes are all commonly used, I
and that P. perfoliatus is a particularly desirable species. Conversely, ™
SAV can also be significantly affected by waterfowl grazing. For instance,
Jupp and Spence (1977) reports that waterfowl grazing reduces SAV biomass •
by a factor of one to five in certain areas of Loch Leven, Scotland, and •
that overall, grazing removes about 20 percent of SAV biomass from the
Loch. In Chesapeake area, Rawls (in preparation) notes that feeding by ^
swans can transform a field of clover to "a hog wallow" overnight. H
Intensive grazing by swans during the 1980-1981 winter was probably partly ^
responsible for the poor 1981 growth of P^ perfoliatus at one of our
intensive study sites in the Choptank River (Todds Cove site). •
Studies of the dependence of Chesapeake Bay waterfowl on SAV for food V
have been conduced by Wilkins (1981), Rawls (in prep.), Perry et al. (1976)
and Stewart (1962); results have been summarized by Stevenson and Confer M
(1978) and Munro and Perry (1981). Vegetable matter is an extremely H
important food item for waterfowl in the upper Chesapeake Bay (Table 3).
Of some 2,747 birds examined by Rawls (in prep.), 78 percent of food —
material was vegetable. Several species of SAV (P. perfoliatus, R. •
maritima, M^ spicatum, and N. guadalupensis) were prominant items averaging ™
about 23 percent by volume in the diet of all waterfowl species
considered. The birds analyzed in this study were collected between 1958 •
and 1968 during fall and winter hunting seasons. That so many birds ||
contained significant quantities of SAV clearly indicates that SAV in the
upper Bay persisted far longer into the fall and winter seasons than it Ml
does now. A shortened growing season, such as we now see in the upper Bay, H
is another index of stress on SAV.
Munro and Perry (1981) also developed long-term (1972 to 1980) SAV and
waterfowl distribution data to test the hypothesis that variations in •
waterfowl populations were related to variations in SAV abundance. Though 9
they found few statistically significant relationships between these two
factors, they observed that the most important waterfowl wintering areas •
were also the most abundantly vegetated areas in recent years (lower jf
Choptank and Chester Rivers and Eastern Bay). Munro and Perry further
suggest that waterfowl have adapted to the SAV decline primarily by
45.7
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
wintering elsewhere in the Atlantic Flyway and that future increases in SAV
abundance will produce positive responses in waterfowl populations. In
Loch Leven, Allison and Newton (1974) found a very good correlation between
SAV abundance and waterfowl densities, and Hocutt and Dimmick (1971) found
P. pectinatus to be the preferred food of some waterfowl species. Along
similar lines, Wilkins (1981) reports that waterfowl use of SAV areas in
Virginia is greater than in non-vegetatod zones. Feeding studies suggest
that waterfowl grazing on SAV-associated invertebrate populations is
sufficiently intense (2 to 25 g m~2 dry material removed per year) to
influence infaunal densities.
Aside from the direct grazing pathway by which SAV can enter food webs,
the vast majority of studies, including those in the Chesapeake, indicate
that most SAV material enters food webs through detrital pathways, den
Hartog (1967), for instance, states that direct grazing is not an important
feature of SAV communities; Ott and Maurer (1977) found that only a small
fraction of Posidonia oceanica was consumed while live. Mann (1971), Day
(1967), and Harrison and Mann (1975) all reached similar conclusions that
agree with the general finding that direct grazing in most macrophyte-
dominated aquatic systems is small. Mann (1972) and others indicate that
SAV (and macrophytes in general) are a relatively poor source of food while
alive because of low nitrogen content. Mann (1972) suggests the following
scheme:
Thus, the process of decomposition of leaf litter in coastal
waters may take the following form. There is an initial period of
autolysis during which soluble materials leach out. Bacteria and
fungi then colonize the material and begin to render soluble by
enzyme action some of the previously insoluble material. The
micro-organisms absorb a proportion of the material they digest,
and some escapes. Populations of predators such as ciliates and
nematodes begin to build up. Macrobenthic organisms begin to tear
off pieces of the plant material with its attached community of
micro-organisms. They strip off the micro-organisms as the
detritus passes through their guts, the feces are recolonized and
the process is repeated by coprophagy. The cumulative result of
this process is a steady reduction in particle size, with a
consequent increase in surface-area-to-volume ratio, an increase
in microbial populations and a reduction in the Carbon:Nitrogen
ratio of the detritus.
In Chesapeake Bay, food web dependence seems to follow this pattern.
Brooks et al. (1981), in a study of a seagrass bed in Virginia, report that
seabass, pipefish, pigfish, and white perch are epibenthic feeders
utilizing amphipods and shrimp that are, in turn, detrital feeders. Data
further indicate that large predators (weakfish, bluefish, and sandbar
sharks) entered the SAV bed with little food in their stomachs and left
after feeding. Food items for these predators can generally be traced back
to a detrital source, some fraction of which is probably SAV in origin.
Again there was little evidence of direct grazing based on stomach
452
-------
TABLE 3. FOOD HABITS OF WATERFOWL IN THE UPPER CHESAPEAKE BAY,
MARYLAND a»b (FROM STEVENSON AND CONFER 1978)
Waterfowl Animal Vegetable Total
species food food (percent)
(percent) (percent)
Predominant foods
percent total volume
Canvasback
Redhead
47.76 51.85
23.40 76.59
Lesser Scaup 47.56 52.47
Bufflehead 67.42 32.59
99.61 19.65 Baltic clam
18.42 Corn
16.32 Soft-shelled clam
14.29 Redhead grass
7.44 Widgeongrass
99.99 29.29 Corn
15.19 Redhead grass
14.74 Widgeongrass
10.53 Soft-shelled, Baltic, and
Mitchell's clams
6.73 Conrad's false mussel
100.03 20.48 Widgeongrass
12.32 Soft-shelled clam .
11.59 Corn
10.85 Redhead grass
6.89 Mussel
100.01 13.52 Widgeongrass
11.85 Redhead grass
10.00 Barnacle
8.52 Fish
7.22 Mud crabs
Goldeneye
63.09 36.87
Mallard
5.00 94.80
99.96 19.44 Mud crab
17.67 Corn
14.88 Soft-shelled clam
9.22 Barnacle
9.00 Bivalves (unidentified
fragments)
99.80 24.14 Corn
10.41 Redhead grass
8.17 Widgeongrass
9.13 Other submerged
macrophytes
1.64 Conrad's false mussel
1.31 Soft-shelled clam
a Based on waterfowl gizzards collected during 1959-1968 hunting seasons
" Rawls (in press)
453
-------
I
I
I
i
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
TABLE 3. (continued)
Waterfowl Animal Vegetable Total
species food food (percent)
(percent) (percent)
Predominant foods
percent total volume
Black Duck
6.44
93.54
99.98
Canada Goose
0.00
100.00
100.00
17.52 Corn
15.50 Redhead grass
14.20 Widgeongrass
8.40 Milfoil
1.91 Conrad's false mussel
1.76 Amphipods
32.42 Grasses (Gramineae)
29.61 Corn
6.97 Milfoil
5.11 White clover
2.99 Crab grass
454
-------
analyses. In an extensive study of feeding habits in upper Bay SAV
communities, Bunker et al. (1981a) found little evidence of direct grazing
by fish, although some SAV seeds and plant leaves were found in stomachs.
Energy flow appeared to enter food webs as detritus and pass through
epifaunal and infaunal invertebrates to small and large fish. Carr and
Adams (1973) and Adams (1976) report similar results for fish communities
in Thalassia and Zostera beds. Bunker et al. (1981a) note that many
epifaunal species that are important food items are also closely associated
with SAV.
Attempts were also made to quantify food sources of the invertebrate
community using stable carbon isotope techniques to more closely relate SAV
to foodweb production. While other investigators have reported some
successes with this technique, studies in the lower Bay by van Montfrens
(1981) and Bunker et al. (1981b) in the upper Bay yielded interesting but
ambiguous results. The basic problem was that there were four or five
available sources of organic matter (SAV, phytoplankton, epiflora, benthic
microflora, and sediment detritus) each having a different ^C ratio.
Thus, unless an invertebrate had an extreme ^C ratio (either high or
low) there were an unlimited number of solution to the feeding equation.
With a few exceptions, most animals had intermediate values (-13 to -18)
suggesting that they were feeding on a mixture of detrital sources or a
single detrital source with an intermediate ^C value. In contrast to
this, Fry and Parker (1979) reported that SAV detritus was an important
feature of the organic matter supply in Texas seagrass systems. They found
that inshore animals had less negative ^C ratios (-8.3 to -14.5) than
did offshore animals (-15.0 to -19.0) and that the differences corresponded
to the less negative and more negative 13 C ratios associated with SAV (-7
to -12.2) and phytoplankton (-20 to -26), respectively. There appear to be
several possible reasons for the differences in Chesapeake Bay and the
Texas studies. First, SAV in Texas were a dominant component of the
aquatic system and thus abundant SAV detritus was probably available
through most of the year. In contrast, SAV are presently a marginal item
in Chesapeake Bay. Furthermore, visual inspection of our study sites
indicates that most of the SAV biomass is probably rapidly exported from
littoral areas prior to becoming detrital particles of appropriate sizes.
Thus, animals in Chesapeake Bay SAV beds may not have sufficient
opportunity to feed on detrital SAV such that their l^C ratios closely
reflect SAV ratios. Secondly, phytoplankton are a dominant feature of
Chesapeake Bay and we have demonstrated that SAV communities can
effectively filter plankton from the water column via their baffeling
effects on currents (Boynton et al. 1981b). Thus, there is an effective
supply of nutrient-rich organic matter with a very negative ^C ratio
(-22) available to SAV food webs. In view of the above circumstances, it
is not surprising that l^C ratios did not clearly indicate SAV detritus
to be of dominant importance in Chesapeake Bay littoral zone foodwebs.
In summary, the majority of studies suggest that SAV is available
primarily as detritus and, in some localities, is very important. Because
of the complexity of organic matter sources in Chesapeake Bay and the
current marginal distribution of SAV, a quantitative assessment of SAV
importance as a food source was not possible. However, CBP results show
that SAV in the Bay is probably used by heterotrophs of one type or
another, and that SAV's physical structure concentrates other foods
455
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
(phytoplankton, epiphytic algae, benthic microalgae) for animal consumption.
From a Bay-wide perspective, several recent studies provide additional
support for this argument. Kemp and Boynton (1981) constructed seasonal
and annual carbon budgets for three meter and six meter depth zones in the
mid-salinity portion of the estuary and found that on an annual basis
virtually all carbon inputs were used. They conclude that "despite the
considerable interactions both within the benthic community...and between
the benthos and other parts of the estuarine ecosystem... photosynthesis
ultimately limits...metabolism." If heterotrophic metabolism is organic
matter limited, it follows that SAV would also be used if, as has already
been shown, this material is a suitable food source. Boynton et al.
(I981c) have also shown that, in most portions of the Bay, the amount of
organic material being sequestered into deep sediments is a small fraction
(three to five percent) of that being produced in overlying waters, again
suggesting that if suitable organic matter is available, it will tend to be
refined. Thus, loss of SAV production may well lead to loss of animal
production.
-------
SECTION 3
THE HABITAT VALUE OF SAV SPECIES
IN CHESAPEAKE BAY
It is generally accepted that meadows of SAV serve as primary nursery
habitats for a diverse assemblage of commercially valuable biota and forage
species. Though many sampling studies have documented impressive numbers
of animals in vegetated areas, the mechanisms underlying the proposed
nursery role of SAV beds have only scarcely been elucidated. The two most
obvious explanations for the great abundance of SAV-associated organisms
are that SAV provides them with food and shelter.
These two features are supported by past research. It is clear that
SAV beds are among the most productive systems known (McRoy and McMillian
1977), and it is equally clear that most of this production is not grazed
by resident organisms (Ogden 1976). Instead, SAV detritus (Mann 1972, Klug
1980) and SAV epiphytes (Morgan 1980) provide most of the energy available
to secondary consumers in SAV beds. Recent studies, using ^C:^C
ratios to trace the source of carbon present in secondary consumers, show
that SAV-derived carbon provides a significant fraction of the energy used
by secondary consumers in Texas turtlegrass meadows (Fry and Parker 1979),
but a rather small fraction in a newly established North Carolina eelgrass
bed (Thayer et al. 1978). To date there have been no comparative studies
of the growth rate of organisms living in (versus outside) SAV meadows.
Thus, only indirect generalizations can be made regarding the relative
importance of SAV-derived carbon for the growth and survival of associated
fauna.
Growing evidence suggests, however, that SAV protects its fauna from
their predators. For example, Nelson (1979) shows that eelgrass provides
amphipods significant amounts of protection from predatory finfish, and
Stoner (1980) shows that several kinds of benthic plants provide amphipods
protection from finfish. Plant surface area affords the best estimate of a
plant's protective ability. Recently, Heck and Thoman (1981) found that
turtlegrass and several species of red algae provide significant amounts of
protection to tethered crabs in field trials, and that both artificial and
live eelgrass provide grass shrimp (Palaeomonetes pugio) significant
amounts of protection from predatory killifish (Fundulus heteroclitus).
STRATEGIES AND METHODS USED IN CBP HABITAT STUDIES
Against this background of published information, a series of studies
was designed to determine the extent to which SAV beds in Chesapeake Bay
serve as sites of densely aggregated animal species (indicating the use of
SAV for food) and as areas providing, through the physical presence of the
plants themselves, important amounts of shelter from predators for a wide
range of invertebrate and fish species.
Several field sampling studies were funded under the CBP to answer the
first question. One study compared standing stock and secondary production
of all macrofaunal ( ^ 0.5 mm) organisms that inhabited the bed with
similar estimates for nearby unvegetated bottoms (lower Bay). A second
study, with similar aims, was done at two upper-Bay eastern shore beds of
457
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
mixed species composition (Parson Island and Todds Cove, Maryland). Both
these studies used a variety of sampling techniques, including seining,
trawling, and gill netting, depending on the size and mobility of the
target species. A third sampling study compared the use of the upper Bay
eastern shore grass bed at Parson Island by commercially important fishes
and blue crabs with a lower-Bay eelgrass bed, near the mouth of the York
River, and with unvegetated habitats. Trawling and gill-net sampling were
used in this study.
Field and laboratory experiments were also conducted to investigate the
ability of SAV to provide animals with shelter and protection from
predators. Field experiments used exclusion cages to evaluate the
intensity of predation by fishes and blue crabs on infaunal populations in
vegetated versus unvegetated habitats. By excluding predators from certain
areas, we estimated what predation-free infaunal population densities would
be in both vegetated and unvegetated areas. Then, by comparing ratios of
standing crop in vegetated and unvegetated areas before and after caging,
we estimated the amount of protection provided by vegetation in natural
conditions. In the lower Bay, caging experiments were performed at the
eastern shore Vaucluse Shores site; in the upper Bay, caging experiments
were carried out in the Todds Cove portion of the Choptank River.
Laboratory microcosm experiments were conducted to estimate the amount
of protection provided by SAV for infaunal bivalves, shrimps, crabs, and
fishes. The first set of experiments was designed to test the ability of
low, medium, and high density artificial eelgrass blades and rhizome mats
to provide protection for the infaunal bivalve Mulinia lateralis and for
juvenile blue crabs (Heck and Thoman 1981). Predators were adult blue
crabs. This set of experiments was conducted in wading pools (2.43 m in
diameter x 0.45 m in height) with recirculating water. The second element
used larger tanks (3.66 md x 0.9 mh) to examine protection for spot
(Leiostomous xanthurus) and silversides (Menidia menidia) by medium and
high densities of artificial eelgrass. Predators used were summer flounder
(Paralichthys dentatus) and weakfish (Cynoscion regalis). Laboratory
studies were also performed to evaluate the protection artificial and
living eelgrass, and living widgeongrass provided grass shrimp.
Experimental tanks were 1.3 md x 0.3 mh, and recirculated water was used.
Predators were killifish (Fundulus heteroclitus). Controls in all of these
experiments involved identical treatments conducted in unvegetated tanks.
Though these laboratory studies are similar, each was designed to
investigate different aspects of predator-prey relations in vegetated
habitats. These factors included studying the importance of prey escape
behavior, predator-prey size in relation to the size of SAV patches, and
differences in the amount of protection provided by different species of
SAV.
^N SITU ANIMAL ABUNDANCES
Invertebrates
In both the upper and lower Chesapeake Bay, field study results
indicate that infaunal abundance and diversity were higher in vegetated
than unvegetated areas. In the lower Bay, polychaetes dominated eelgass
458
-------
I
and widgeongrass areas, and bivalves dominated unvegetated sites, although
a large overlap in species composition existed between the two types of •
habitats. These differences in dominant taxa are due partly to the type of •
bottom. Muddy sediments deposited in seagrass beds favor deposit feeders
such as polychaetes; the sandy sediments in non-vegetated areas favor •
suspension-feeding bivalves. Some of the other differences in species |
composition occur because epifaunal, grass-blade associated invertebrate
species inevitably occur in sediment samples from SAV areas, even though M
they are not residents of the infauna. Greater abundances in SAV areas are •
due primarily to the large numbers of polychaetes, oligochaetes, isopods, *
and grass-blade organisms collected at these sites.
In the upper Bay region, polychaetes dominated both SAV and unvegetated B
sites, although numbers were greate? in vegetated areas, as were overall |
abundances of oligochaetes and isopods. Bivalves were more abundant in
unvegetated areas in spring, but became more abundant in SAV beds during •
summer (Bunker et al. 1981c, Ejdung et al. 1981). Amphipod abundances were •
greatest at the unvegetated site and in the lower Bay. These differences
reflect the suitability of fine sediments for deposit feeders in SAV areas
versus the suitability of sandy sediments for suspension feeders in •
unvegetated areas. 9
There were also some notable differences in the abundances of organisms
as related to salinity. Diversity was much lower at the low salinity (7 to •
11 ppt) sites in the upper Bay than at the higher salinity (14 to 22 ppt) |
site in the lower Bay. For example, maximum density at the high salinity
eelgrass site was 90,000 individuals per m^. The reason for the ^
relatively low infaunal abundances at the low-salinity site is not known. •
Because unvegetated habitats support a virtually non-existent epifauna,
(defined as the animal assemblage growing on SAV and other emergent bottom
features), only vegetated habitats were sampled for epifaunal organisms. •
Epifaunal density was higher at the more densely vegetated Todds Cove bed 0
than at the Parson Island bed. However, epifaunal densities per g SAV
(excluding polychaetes) were very similar at the two sites, ranging from M
around 50 to 200 individuals per g SAV biomass. The isopod Erichsonella H
attenuata was dominant in Parson Island collections, and gastropods and
tanaids were dominant at Todds Cove. Amphipods, grass shrimp, and _
chironomids were present at both sites (Staver et al. 1981). B
Epifaunal abundances at the eelgrass sites in the lower Bay were much ™
higher than those found at the low salinity SAV sites in the upper Bay even
though polychaetes were not included in the upper Bay sampling. Numbers •
ranged from around 20 individuals per g SAV in November to more than 9,200 0
individuals per g SAV in April. Dominant species included isopods,
gastropods, polycahetes, and barnacles (Diaz and Fredette 1981).
Differences in salinity between the two intensively studied areas were
probably responsible for the large abundance of barnacles in the lower Bay
and at least partly responsible for differences in total abundance between
sites.
Finfish
Finfish sampling in the protected Todds Cove bed and the exposed Parson
Island site (Figure 8) found greater abundances and species richness in
vegetated than unvegetated bottoms, with greatest numbers occurring in the
459
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
(d)FIS
20-
7 16-
E
tr
m 12-
z
0 _
4-
0
H DENSITIES
PARSON ISLAND
n
|
-
-
-
-
r^
TODDS C
ru i
n i
OVE
|
**
V
4&I
• —
1 1
NO. OF SPECIES 01 14 6 10 6 \0 B 83
MAY JUN JUL AUG SEP
14 9 12 12 13 12 14 5 14 5
MAY JUN JUL AUG SEP
(b) FISH WEIGHT DISTRIBUTION
en
tr
o
n3*8
6-]
4-
2-
1 -
n
V
V
TODDS
S)
COVE
W
PREFERENCE
^j
i
1
\
r
i
i
i
MAY
JUN
JUL
AUG
SEP
Figure 8.
Average monthly (a) fish densities at Parson Island and
Todds Cove sites for vegetated and non-vegetated (reference)
areas and (b) fish weight distribution at Todds Cove for
vegetated and non-vegetated areas. Data from Lubbers et al.
1981.
460
-------
protected SAV area. Fish densities at Todds Cove are among the highest yet
reported in the literature (Lubbers et al. 1981).
In addition, a plot of average weight per individual (Figure 7) during
the summer period suggests that the SAV bed at Todds Cove was continually
used as a nursery area for small fish while larger sized individuals
predominated in unvegetated reference areas. However, Heck (personal
communication) often found fish abundance (as indicated from otter trawl
samples) to be as large on sandy bottoms as in SAV at the Parson Island
site. This difference occurred because of the chance catch of schools of
spot (Leiostomous xanthurus) over sandy areas, and probably because SAV
abundance dropped precipitously during the course of the study. Large fish
predators such as bluefish and cownosed rays were found in both vegetated
and unvegetated habitats and, in both studies, more fish were taken at
night than during the day. There was little indication that the low
salinity SAV beds serve as nursery areas for commercially valuable finfish,
although any conclusion concerning such values might be biased due to the
severely depressed distribution of SAV in the upper Bay.
Fish sampling programs in lower Bay eelgrass meadows on the York River
and at Vaucluse Shores found much greater abundances and species richness
in these higher salinity SAV beds than on nearby unvegetated bottoms, and
much greater night than day catches (Brooks et al. 1981). Some large fish
predators, such as weakfishes and sandbar sharks, seem to forage most often
over vegetated bottoms while others, such as bluefish, appear to forage
indiscriminately over both vegetated and unvegetated areas.
The main conclusion of these studies on fish, that fish communities are
richer in vegetated than unvegetated areas, was expected, because similar
results have been found previously in other SAV habitats in North Carolina
(Adams 1976), Florida (Livingston 1975), and Texas (Hoese and Jones 1963).
What was not expected was the finding that few commercially important
finfish use the SAV beds as significant nursery habitats. This result is
surprising, because many juveniles of commercial species such as sea bass,
snappers, and groupers use SAV as nursery areas in latitudes south of
Chesapeake Bay (Adams 1976, Livingston 1975, Weinstein and Heck 1978). The
role of SAV for commercial fishes in the Chesapeake Bay system seems to be
largely that of a rich foraging place for adults although, once again, it
is -important to emphasize that the current restricted distribution of SAV
may bias these conclusions. For instance, major spawning and juvenile
habitats for striped bass once existed in the upper Bay (Susquehanna
Flats), an area that was densely populated with SAV. More representative
patterns of commercial fish use of SAV habitat might best be evaluated
through historical correlations of SAV and juvenile fish distributions;
this is being done in the CBP's environmental characterization.
Blue Crabs
Information on blue crab abundances was collected at the same time
fishes were sampled at intensive study sites. Investigators found low
numbers of juvenile blue crabs in the upper Bay, but extremely large
numbers of blue crabs in eelgrass meadows of the lower-Bay. Up to 10,000
times as many blue crabs were found at the lower Bay than upper-Bay SAV
sites. Studies by Heck (personal communication), which used identical
461
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
sampling techniques in both low and high salinity SAV beds and adjacent
unvegetated areas, found lower Bay crab numbers ranging from a few to a
thousand times more abundant than the upper Bay beds during the spring and
summer months. In addition, most of the crabs in the high salinity
eelgrass beds were juvenile females ( <^ 100 cm CW) that constituted the
breeding stock of future generations. Blue crab densities in unvegetated
areas were found to be as large or larger than those recorded from upper
Bay SAV beds. In contrast, far fewer crabs were taken on sand than in SAV
at the lower-Bay site. This difference in sand versus SAV crab abundances
between sites is probably due to the presence of many juveniles at the
eelgrass site, most of which require SAV for protection from predators.
The adult crabs in upper Bay SAV beds apparently do not require vegetation
for protection, except when molting, and occur on both vegetated and
unvegetated bottoms. The sampling gear used to collect crabs in both
locations was not efficient for collecting molting crabs. Thus, the role
of upper Bay SAV in providing protection to molting crabs may have been
greatly underestimated.
The conclusion drawn from these studies is that there seems to be only
a very limited blue crab nursery role played by upper Bay SAV beds. Lower
Bay eelgrass beds, however, serve as primary blue crab nursery habitats and
support very large numbers of juvenile blue crabs throughout the year.
STUDIES ON SAV AS PROTECTION
Results of soft-bottom predator exclusion experiments are often
difficult to interpret because of several commonly encountered problems,
including an inability to completely exclude predators from caged areas and
accurately estimate the effects that the presence of the cage itself
produces on the physical environment (Virnstein 1978, Dayton and Oliver
1980, Peterson 1979). Caging studies conducted in Chesapeake Bay
encountered these problems, and the results of these studies, therefore,
must be interpreted with caution and circumspection.
Caging experiments in the lower Bay show that infaunal densities
increased in caged areas, and that this increase was most pronounced on
unvegetated bottoms (Orth 1981). There was little evidence that cages
altered the physical environment by changing sedimentation patterns,
although predators did periodically invade caged areas. Epifaunal
densities were higher in caged than uncaged areas shortly after the
installation of cages in eelgrass, but shading by cages subsequently
reduced eelgrass biomass and also led to declining epifaunal numbers.
Caging studies conducted in the upper Bay were less conclusive than
those conducted in the lower Bay. Some evidence suggests that predation
may be important in reducing infaunal densities; however, technical
difficulties, such as small fish passing through the cage walls, weaken the
results.
The first element of experimental predation studies shows that, in the
presence of artificial vegetation and rhizomes, juvenile blue crabs
received a significant amount of protection from predation. The infaunal
bivalve Mulinia lateralis, however, received very little protection from
either the artificial leaves or rhizome mat. The second element shows that
the predators Paralicthys dentatus (summer flounder) and Cynoscion regalis
(weakfish) captured progressively fewer spot (Leiostomous xanthurus), and
462
-------
silversides (Menidia menidia) as vegative cover increased from 0 to 22
percent (Orth 1981).
The second study (Heck and Thoman 1981) found that dense amounts of
artificial and live eelgrass, and live widgeongrass, provided grass shrimp
with significant amounts of protection from fish predators. Low and medium
densities of SAV did not provide much shelter. Furthermore, widgeongrass
provided greater protection per unit of surface area than either living or
artificial eelgrass.
The seemingly disparate results of these microcosm experiments can be
understood within the following framework. Mobile epibenthic animals, such
as fish and blue crabs in the former study, and grass shrimp in the latter
study, do derive protection from predators in the presence of SAV.
However, the amount of protection received is probably a function of plant
surface area. Thus, SAV species with finely branched leaves and high
surface areas should provide better protection for prey taxa than plants
with simple leaves, all other factors being equal. Infaunal species, such
as burrowing bivalves, should generally receive less protection from
predators than epifaunal species of SAV habitats. For shallow burrowing
species like Mulinia lateralis, SAV beds provide little or no protection
from predators. However, for species that burrow below the SAV rhizome
mats, there should be reduced predator success in vegetated habitats
compared with that in unvegetated areas.
This hypothesis explains the results of the microcosm experiments and
is amenable to further testing and verification. It is likely, however,
that a number of other unstudied variables, such as size of predator and
prey in relation to SAV dimensions and the foraging strategy of the
predators, also influence predator-prey relations in SAV beds.
463
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
SECTION 4
INFLUENCE OF SAV ON SEDIMENT DYNAMICS
Sediment processes in estuarine and coastal systems have been the focus
of numerous studies in the past several decades. Results of such studies
indicate that sediment processes strongly influence light attenuation in
the water column, produce shoaling or scouring, affect the composition of
benthic invertebrate communities, and influence the exchange of materials
between sediments and overlying waters. The sediment processes which
produce such effects can be characterized as a cycle that includes the
following components: (1) yield of "new" sediments from land erosion,
runoff, and shoreline erosion; (2) deposition of suspended sediments; (3)
resuspension of deposited sediments due to tidal and wave action and; (4)
transport of resuspended sediments to different locations.
The resuspension-deposition [(3) - (2)] portion of the cycle dominates
littoral zone sediment dynamics and can affect the health of SAV. A higher
cycling rate increases seston levels and reduces light availability to
SAV. This cycle is shown diagramatically in Figure 9 and suggests the
probable magnitude of different portions of the cycle in deep and littoral
estuarine areas. The diagram shows that wave action is the major source of
energy to resuspend unconsolidated sediments in the littoral zone, and
tidal energy provides the major force to resuspend and transport sediments
in deep water. The relative contribution of major new sources of sediment
to the Bay include material washed in from the watershed and shoreline
erosion. Most shoreline material enters the deposition and resuspension
cycle from the margins of the Bay, whereas fluvial sources follow the deep
water transport path.
REVIEW OF SEDIMENT PROCESSES
Aspects of sediment dynamics and turbidity patterns have received
considerable attention in Chesapeake Bay. Net sedimentation rates have
been repeatedly estimated for various portions of the open Bay (Biggs 1970,
Schubel and Hirschberg 1977, Brush et al. 1981) and for some tributaries
(Roberts and Pierce 1976, Yarbro et al. 1981). Increases in turbidity have
been documented for both the open Bay (Heinle et al. 1980) and for some
tributaries (Kemp 1980). The increases are apparently due to increased
algal stocks and seston levels. In a crude fashion, the decline of SAV
communities in northern Chesapeake Bay parallels increasing trends in
turbidity and nutrient loading. However, most of this work has been done
in deep areas of the Bay region. In this summary, we have focused on
vegetated and non-vegetated littoral areas less than two meters in depth.
In addition, previous measurements have largely been devoted to estimating
net sedimentation rates, and as Oviatt and Nixon (1975) have pointed out,
only a small fraction of total "sediment activity" is measured when such
estimates are made.
464
-------
Study Area
Sediment Cycling Depth,m
Rate, g m~2y-l
Reference
Naragansett Bay 7
Departure Bay, B.C. 3
Surf Zone, California 7 -
Buzzards Bay, Mass.
York River, Va.
Upper Patuxent River
Lower Patuxent River 3
Littoral Zone, Non-SAV 3
Littoral Zone, SAV 0
- 18 x
x 103
330 x
6 x 104
103
105
103
106
8 x
2 x
6 x
4 x
3 x
105
105
7
32
2
15
4 -
10-12
1 -
1 -
Oviatt & Nixon, 1975
Stephens et al. 1967
Shepard, 1963
Rhodes & Young, 1970
Haven & Morales-Alamo,
Boynton et al. 1981b
Boynton et al. 1981b
Boynton et al. 1981b
Boynton et al. 1981b
1972
WATERSHED
RUNOFF
TIDAL
ENERGY
SHORELINE
EROSION
LITTORAL.
ZONE WAVE
ACTION
RESUSPENSION-
DEPOSITION CYCLES
IN (S) LITTORAL a
(§) DEEP-WATER ZONES
Figure 9. Major physical sediment processes in Chesapeake Bay showing
sources and energy for sediment transport.
465
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
TABLE 4. SEDIMENTATION RATE IN ram y~l AT SEVERAL LOCATIONS IN CHESAPEAKE
BAY AND OTHER SELECTED ESTUARIES
Study Area
Narragansett Bay
Delaware Bay
Patuxent Estuary
Upper
Upper
Lower
Lower
Chesapeake Bay
Net Sedimentation Technique
Rate, mm y~l
Reference
0.3 - 0.4
1.5
37.0
4.0 - 7.0
4.0
5.0 - 10.0
Mass Balance
Not available
Mass Balance
Pollen Dating
Pollen Dating
Sediment Traps
Farrington 1971
Oostdam & Jordan 1972
Roberts & Pierce 1976
Brush et al. 1981
Brush et al. 1981
Boynton et al. 1981
Upper
Upper
Mid
Mid
Mid
4
6
0
.5
.0
.9
- 9.0
- 10.0
1.5
1.1
- 1.2
Pb210
Pollen
Pollen
Dating
Dating
Mass Balance
Pb210
Hirschberg &
1979
Brush et al
Brush et al
Biggs 1970
Schubel
. 1980
. 1980
Hirshberg & Schubel
1977
A considerable number of measurements of net sedimentation rates and
sediment cycling rates (summation of resuspension-deposition) were made in
estuarine environments using a variety of techniques. We summarize some of
these measurements in Table 4 with special emphasis on Chesapeake Bay. Net
sedimentation estimates for areas in the tidal Bay system ranged from 0.3 to
37 mm y~l. This broad range is not surprising in view of the strong
gradients in seston concentration and sediment input rates encountered in
estuarine systems. In the turbid upper section of Chesapeake Bay, for
example, estimates ranged from 4.5 to 10 mm y~^; in the mid-salinity region,
rates ranged from 0.9 to 1.5 mm y~l. A similar pattern was evident in the
Patuxent River. To compare the magnitude of net sedimentation with sediment
cycling rates (i.e. deposition-resuspension-deposition), accumulation rates
(in mm) were converted to a weight basis. On this basis, net sedimentation in
Chesapeake Bay ranged from about 600 to 6,000 g m~2y~l of dry sediments.
In sharp contrast to these values, sediment cycling rates were far higher,
especially in shallow water environments, and indicate that cycling dominates
sediment processes. Significantly, an estimate from an SAV community
(Choptank River) was among the lowest we encountered in estuarine systems and
illustrates the importance of these communities in stabilizing sediments at
the surface.
ROLE OF SAV IN SEDIMENT PROCESSES
Several previous investigations have led to an understanding of the
mechanisms by which SAV can modify sediment substrates (Ginsburg 1956,
Ginsburg and Lowenstam 1958, Wanless 1981). Specifically, the rhizome-root
466
-------
I
complex can stabilize sediments, and the physical structure of seagrass blades
and epiphytes can slow currents, allowing sediments to settle. This complex •
can also substantially reduce current-and-wave-induced resuspension. •
Considerable evidence suggests that SAV can play an important role through
those mechanisms in nearshore sediment dynamics. Scoffin (1970) found that •
dense beds of Thalassia protected bottom sediments from current speeds up to I
70 cm sec~l; extensive bottom-sediment erosion did not begin until current
speeds reached 150 cm sec~l. in Florida, Ball et al. (1967) found that _
bottom erosion was minimal in seagrass covered areas following the passage of I
a hurricane, but exposed sand areas were extensively modified. Also in ™
Florida, Wanless (1981) found sedimentary sequences that probably resulted
from trapping and consolidation of suspended particles by SAV. He states that •
the vertical sediment record indicates increased trapping of storm-generated B
sediments and decreased bedload transport as SAV became established. In a
Zostera bed in Denmark, Christiansen et al. (1981) infer, from inspection of im
sediment cores and historical SAV distributions, that the Zostera die-back in •
the 1930's resulted in disturbance and mobilization of nearshore sediments and
a movement of sediments into a local harbor. Moreover, Christiansen _
determined that coastal morphology was stable during periods when eelgrass was •
present, but significant changes occurred when it was absent. ™
In the Chesapeake area, Orth (1977) reports that sediment particle
diameter decreased, and organic matter content and infaunal densities •
increased in bottom sediments in areas with SAV compared with those that did •
not have such coverage. Based on these findings and observations that showed
less sediment disruption during storms in vegetated zones and less dispersion ^
of dyed sand patches, Orth concludes that SAV is effective at trapping and •
consolidating suspended sediments. It appears that substantial beds of SAV ™
can effectively modify littoral zone sediment dynamics through sediment
trapping and consolidation of sediments at the surface. Because sediment •
processes may be most active in littoral zones, sediment processes in deeper B
areas may also be affected by lateral transport and deposition (Webster et al.
1975). •
CHESAPEAKE BAY PROGRAM STUDIES
We hypothesized that SAV communities can play a significant role in
modifying littoral zone light regimes by baffling of wave and tidal currents,
thus reducing sediment resuspension. Conversely, we hypothesized that high
turbidities in some areas of the Bay have contributed to the decline of SAV
communities. Chesapeake Bay Program studies were designed to (1) document
patterns of light attenuation on several time scales (seasonal, diel, tidal
cycle) in littoral communities having SAV and in those not having SAV; (2)
relate observed light attentuation patterns to concentrations of materials in
the water column to identify the relative importance of light attenuating
factors; and (3) examine the potential of SAV communities as natural sediment
traps.
Data on suspended sediments and light attenuation from intensive study
sites are presented in Figure 10. This figure shows differences between
vegetated and non-vegetated areas plotted against tidal stage for Todds Cove
in the Choptank River and Parson Island sites in Eastern Bay. These plots
indicate that as turbid offshore waters enter SAV beds on rising tides,
sediments are effectively removed, thus increasing light transparency. It
467
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
HIGH
LOW
60-
40-
20-
in
_o
!o
o
UJ X
o _
z u
oc "o
UJ <5
t ^ 0-
0 c
o
»- Z
Z .
UJ (A
o >
S^ fiO
o. ® 60-
o
0>
0)
40-
20
TIDAL HEIGHT
(a) SESTON CONCENTRATION
(b) EXTINCTION COEFFICIENT
I
6
I
9
Low Tide
TIME , hours
High Tide
12
Low Tide
Figure 10. Percent difference between vegetated and non-vegetated habitats for
(a) suspended sediment and (b) attenuation coefficient during a tidal
cycle.
468
-------
I
appears that by high tide, turbid inflowing waters have exceeded the filtering
capacity of the bed. As tidal height decreases, the bed apparently M
effectively filters sediments because the gradient between vegetated and •
non-vegetated areas again increases to a maximum. ™
Other qualitative observations show that SAV varies in its ability to
decrease turbidity. Boynton et al. (1981b) found that areas with SAV, •
dominated by P. perfoliatus (a highly branched species), were more successful •
at decreasing turbidity than at a site where P. pectinatus (a thin-bladed
single leaf species) dominated. Even under conditions when SAV biomass was •
comparable between the two areas, it appeared that j?. perfoliatus was more I
effective in clarifying surrounding waters. On several occasions, turbid
water was observed entering a P. pectinatus bed; turbidity increased rapidly ^
in the P. pectinatus sections, but remained considerably lower in the P_^ •
perfoliatus areas. Visually, the P_._ perfoliatus beds appeared as clear areas '
against a turbid background.
Other qualitative data show that water clarity is affected by the size of B
an SAV bed. On several occasions, Boynton et al. (1981b) noted turbidity |
gradients within an SAV bed. Turbidity was greatest at the edge of the bed
and decreased with distance into the bed. It seems that for a given off-shore •
turbidity regime there is a "critical bed size" above which SAV can •
effectively modify the local environment in a fashion favorable for continued
growth (reduce seston levels; increase light penetration). Small SAV beds may
not be able to so modify local light regimes and would thus be disadvantaged •
if light is limiting growth. •
Interpretation of data concerning sediment cycling in littoral zones is
quite difficult. Boynton et al. (1981b) hypothesized that there would be •
substantial differences in the amount of material collected in both surface I
and bottom cups of sediment traps deployed in SAV beds and non-vegetated
reference areas. They anticipated that the structure of SAV would effectively ^
reduce resuspension and hence values from the bed would be markedly lower. In •
fact, while values from the SAV bed were lower, dramatic differences were not ™
consistently evident. It is possible that most resuspension - deposition
occurred during storm events and that during these events wave energies were fl
high enough to overcome the baffeling effect of SAV in these marginal B
communities, leading to substantial deposition in all areas. Further
inspection of climatic data may clarify this possibility. Another possibility M
is that material collected in cups in the SAV area was a mixture of H
resuspended materials and true sedimentation, while resuspended material made
up the bulk of the collection in the reference area. Substantial reduction in _
seston-based turbidities (Figure 9) support this suggestion. H
In spite of the lack of large differences betwen SAV and references area ™
collection, there was a reasonably consistent pattern evident with respect to
collection rate and SAV biomass, particularly for the bottom collection cups fl
(Figure 10). Both surface andvbottom cups had small collection rates when B
biomass was above 150 g m~2 and rates five to 10 times higher when biomass
was bewlow 50 g m~2. When viewed in this fashion, it appears that im
resuspension is clearly reduced in proportion to SAV biomass. H
Given the dynamic nature of the sediment-water interface in littoral
environments, estimates of net sediment retention/compaction are exceedingly _
difficult to obtain and clearly beyond anything that can be inferred from H
sediment traps. A crude estimate can, however, be obtained utilizing the •
seston data presented earlier. If we attribute the tidally related changes in
469
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
seston concentration to dposition, then seasonal estimates can be made. If we
take 90 mg L~l as an estimate of mean seston concentration in littoral
reference areas (Boynton et al. 1981b) and, as suggested in Figure 9, assume
that approximately 50 percent of the suspended material is deposited on each
tide, then daily deposition can be estimated. Taking six months as the period
when substantial SAV biomass is present allows expansion of diel estimates to
seasonal estimates. This procedure yields daily deposition rates of about 63
g m~2d-l and seasonal estimates (180 days) of 1200 g m~2. Assuming that
there is about equivalent to 0.2 cm per growing season. The potential errors
associated with such a calculation are obvious, but it is interesting that
such a reasonable value emerges. Little information is currently available to
suggest whether or not this material is sufficiently consolidated to be
considered as lost to the sediment deposition-resuspension-deposition cycles.
For example, we do not know if material deposited during the summer period
when SAV are present is subsequently lost when SAV die-off in the early fall.
Considering the important role of roots and rhizomes in this process and the
low below-ground biomass observed in Chesapeake Bay SAV communities, it seems
doubtful if this material is permanently consolidated at present, although it
may have been in the past.
Some evidence suggests that SAV can cause sediment to compact, thus
preventing resuspension of sediments. Net sediment retention-compaction can
be crudely estimated by using the suspended sediment data presented earlier.
If we attribute the tidally related changes in seston concentration to
deposition, then seasonal estimates can be made. If we take 90 mg L~l as an
estimate of mean suspended sediment concentration in littoral reference areas
(Boynton et al. 1981b) and, as suggested in Figure 10, assume that
approximately 50 percent of the suspended material is deposited on each tide,
then daily deposition can be estimated. Taking six months as the period when
substantial SAV biomass is present allows for expansion of diel to seasonal
estimates. This procedure yields daily deposition rates of about 63 g
m~2d-l and seasonal estimates (180 days) of 1,200 g m~2. Assuming that
there is about 0.6 g cm~3 of inorganic material in consolidated sediments,
this deposition is equivalent to 0.2 cm per growing season.
The potential errors associated with such a calculation are obvious, but
it is interesting that such a reasonable value emerges. Currently, little
information is available to suggest whether or not this material is
sufficiently consolidated to be lost to the sediment
deposition-resuspension-deposition cycles. Because of the important role of
roots and rhizomes in this process and the low below-ground biomass observed
in Chesapeake Bay SAV communities, it seems doubtful that this material is
permanently consolidated at present. However, SAV biomass levels
characteristic of the early 1960's may have been high enough to prevent
significant resuspension of bottom sediments. Likewise, root-rhizome
structure of these times may have more effectively consolidated bottom
sediments.
COMPARISON OF SEDIMENT SOURCES WITH DEPOSITION IN SAV BEDS IN CHESAPEAKE BAY.
To place the sediment-trapping characteristics of SAV in the context of
larger-scale sediment processes in Chesapeake Bay, we have developed a series
of calculations that compare the magnitude of two major sediment sources to
the deposition rate observed in SAV communities (Table 5).
470
-------
1000-
'E
o>
UJ
z
o
o
0.
UJ
Q
z
UJ
Q
UJ
500-
NEW SEDIMENTS
VEGETATED i BARE BOTTOM
^--BOTTOM VALUES
O
SURFACE VALUES
riii
50 100 150 200
SAV BIOMASS, gm"8 (dry weight)
250
Figure 11. Relationship between SAV biomass and sediment deposition (adapted
from Boynton et al. 1981b).
A71
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
TABLE 5. ESTIMATED ANNUAL SEDIMENT DEPOSITION IN SAV COMMUNITIES RELATIVE
TO SEVERAL SEDIMENT SOURCES IN CHESAPEAKE BAY FOR 1960 AND 1978.
(ALL VALUES IN METRIC TONS PER YEAR x 106)
Sources
Riverine Input3
Shoreline Erosionb
Total
Deposition in SAV Communitiesc
Time
1960
0.491
0.375
0.866
0.72
Periods
1978
0.491
0.375
0.866
0.08
alncludes Bay and tributaries above the mouth of the Potomac River
(1.5 x 109m2).
bAnnual estimates of riverine and erosional sediment inputs from Biggs
(1970). Assumed that inputs were relatively constant between time periods,
Deposition in SAV communities estimated to be 1200 g m~2y-l (Boynton
et al. 1981; Ward, pers. comm.)
Major sediment sources include riverine input and shoreline erosion to
the portion of Chesapeake Bay above the mouth of the Potomac River. We
assume that estimates developed by Biggs (1970) are representative of both
the early 1960's and late 1970's periods. The amount of sediment deposited
during the SAV growing season was calculated from data of Boynton et al.
(1981b), who estimate that some 1,200 grains of sediment may have been
deposited per square meter of SAV community over an estimated 180-day
growing season. Table 5 indicates that a large percentage of sediment may
have been deposited in SAV communities during the 1960 period. However, in
the late 1970's, when SAV distributions were severely reduced, the amount
of deposition was less than 10 percent of the input. Although this
calculation is preliminary, it suggests that SAV in the past may have
played an important role in sequestering sediments in Chesapeake Bay, and
that the amount of sediment presently deposited in SAV communities is small
relative to estimates of sediment input.
LIGHT LIMITATION OF PHOTOSYNTHESIS
Although there appear to be emerging patterns concerning the role of
SAV in modifying littoral zone turbidity and sediment cycling processes, it
is still necessary to establish relationships between ambient light
intensities and functions of SAV growth. Kemp et al. (1981) conducted a
number of experiments to establish SAV responses (photosynthetic rate) to a
range of light intensities. The two species investigated were P.
perfoliatus and M_._ spicatum. They found that light saturated
photosynthesis occurred at about 500-600 uEinsteins for both species, and
that about 150 uEinsteins provided enough light to reach 50 percent of the
maximum rate of photosynthesis (1/2 Pmax is similar to the Michaelis-Menten
half saturation constant).
472
-------
I
To extend these data to broad geographic regions of northern Chesapeake
Bay, we examined attentuation coefficients characteristic of several •
locations during growing seasons (May to September). In Figure 12, a •
typical summer light intensity (just below the water surface) of 1,000
uEinsteins was attenuated using a range of attentuation coefficients (1.0, a
2.0, 3.0) so as to display the light energy reaching various depths. Also •
plotted on the diagram (dashed horizontal lines) are the light intensities ™
at Pmax, 1/2 Pmax, and 1/4 Praax for P^_ perfoliatus. Thus, if an
attentuation coefficient of 1.0 was observed, sufficient light to maintain
light saturated photosynthesis reached a depth of 0.6 meters. The depth at
which sufficient light penetrates to maintain photosynthetic rates at 1/2
Pmax is also given for various locations (Table 6). m
These data suggest that in most locations light saturated I
photosynthesis does not occur in water depths greater than 0.25 to 0.5
meters. Moreover, sufficient light does not penetrate beyond 1.0 meter to
maintain photosynthetic rates at 1/2 Pmax. Thus, it appears that only in •
the most shallow or most clear environments is light not limiting to SAV •
photosynthesis. These calculations may underestimate the limiting role of
light because they are based on a subsurface light intensity of 1,000 •
uEinsteins, a value reached mainly during the middle of a typical summer I
day. On overcast days and in the early morning and late afternoon, values
are considerably lower, and the depths of Pmax would be more shallow. «
Additional work, now in progress, may allow the development of better •
relationships between photosynthesis and light, as well as between ^
photosynthesis and biomass.
TABLE 6. LITTORAL ZONE LIGHT EXTINCTION COEFFICIENTS DURING THE SUMMER
IN CHESAPEAKE BAY DEPTHS AT WHICH 1/2 Pmax OCCURS ARE SHOWN
DATA FROM TWILLEY 1981
Extinction
Location Coefficient
(1)
(2)
(3)
(4)
(5)
(6)
(7)
Upper Bay
Lower Bay
Tributaries
Upper Patuxent
Lower Patuxent
Eastern Bay
Lower Choptank
2.2
2.4
2.4
3.5
1.7
1.3
1.9
Depth 1/2 PMax
0.8
0.7
0.7
0.5
1.1
1.4
0.9
473
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1000
v>
c
'3
in
c
UJ
co
z
LU
100-
Figure 12. Relationship between surface light intensity and light
attenuation, of the water column, expressed at attenuation
coefficients (K). Data from Boynton et al. 1981.
474
-------
SECTION 5
NUTRIENT PROCESSES IN SAV COMMUNITIES
I
I
This section summarizes current knowledge of the effects of SAV •
communities on littoral zone nutrient regimes. Submerged vascular plants
have two potential sources of nutrients available for uptake and
incorporation into new biomass. Dissolved nutrients in the water can be •
taken up by leaves and stems, with some species using sediment nutrient •
reservoirs. SAV can also modify chemical conditions in sediments so that
oxidized and reduced conditions prevail and can lead to several •
transformations of nitrogen and phosphorus. •
This section addresses four categories of nutrient processes
including: (1) nutrient concentration and fluxes in SAV communities; (2)
nutrient regulation of SAV growth; (3) nitrogen transformations, including
fixation, nitrification, and denitrification; and (4) nutrient releases
associated with decomposition processes.
NUTRIENT CONCENTRATIONS AND FLUXES
Recent studies in Chesapeake Bay and global literature support the «
notion that SAV communities buffer nutrients by removing them from the •
water column, thus reducing concentrations. Pertinent examples from
studies in Chesapeake Bay include the work of Twilley et al. (1981) and
Kaumeyer et al. (1981). Twilley conducted a series of water quality •
measurements in the Choptank River estuary, an eastern shore tributary. B
Measurements were taken along the longitudinal axis of the estuary on a
monthly basis from April through September. At one point adjacent to an •
intensively studied SAV community, water quality measurements were taken in I
an SAV bed in waters of moderate depth (4 m), and along the longitudinal
axis of the estuary (deep water). Throughout this period, nutrient «
concentrations were consistently and dramatically lower in littoral, as •
opposed to deeper, sections along that sampling transect. Specificially, ™
ammonium concentrations were one to 10 times lower, nitrate two to 10 times
lower, and orthophosphate generally two to four times lower in the SAV
community than in deeper offshore waters. Similar results were obtained at
the Parson Island site, where nutrient concentrations appeared to be lower
in an SAV community than in adjacent offshore waters (Kemp et al. 1979).
An important ecological implication of these findings is that SAV may
compete with phytoplankton for nutrients, thus reducing potential excessive
algal blooms.
To elucidate mechanisms causing nutrient concentrations to be lower in
the littoral zone, Kaumeyer et al. (1981) initiated a series of studies
using a variety of sampling chambers in an SAV bed in the Choptank River
estuary. The chambers were spiked with different levels of ammonium,
nitrate, and phosphate, and concentrations of these nutrients (as well as
dissolved oxygen) were measured hourly over six to 12 hour periods during
both day and night. A typical set of results is given in Figure 13. A
Nutrient concentrations rapidly decreased from initial-spiked •
475
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
o
o»
y>
<
cr
UJ
o
z
o
u
z
UJ
cc
UJ
en
i
o
(0) NON-VEGETATED ( Depth 0.127m)
POJ
SPIKE: 80//g-ot r*NHj
!6//g-ot r'POj
90
(b)PLANKTON
60-
30 H
90-i
60-
1^ T I I I
(C)SAV (Depth 0.51m)
30-
DAY
23 00
NIGHT
02Hr&
Figure 13.
Chamber nutrient flux at Todds Cove, Choptank River,
July 1980 for ammonia-nitrogen and dissolved inorganic
phosphate. Day and night nutrient concentrations in
experimental chambers all given for (a) non-vegetated
sites, (b) plankton, and (c) vegetated sites.
476
-------
concentrations to lower levels, and generally returned to near-ambient in
less than 24 hours. Uptake rates, in most cases, depended on nutrient
concentrations in the chambers. In addition, Kaumeyer et al. found where
littoral communities were exposed to both nitrate and ammonium, both forms
of nitrogen were taken up, although ammonium was generally taken up
somewhat faster. These investigators did not attempt to partition uptake
between plankton, benthos, and SAV, but apparently uptake rates were fast
in all littoral zone communities investigated. These results suggest that
there are mechansisms, not restricted to SAV, through which nutrients,
entering the littoral zone, can be rapidly removed. Howard-Williams (1981)
reports similar results from dosing a dense bed of P. pectinatus. He found
that this community could rapidly reduce nutrient concentrations, and that
filamentous algae associated with SAV were responsible for most of the
phosphorus uptake.
In addition to these studies, substantial observational and
experimental evidence indicates that SAV removes dissolved nutrients from
the water column at a high rate. Mickle and Wetzel (1978) investigated
SAV-nutrient exchanges in laboratory systems containing Scirpus and
Myriophyllum. In these flow-through systems, nitrate and ammonium were
introduced, and output concentrations monitored. Ammonium and nitrate
concentrations decreased substantially after passing through SAV,
particularly in the Myriophyllum beds. We conclude that littoral SAV
systems are effective in damping higher concentrations entering the
littoral zone following rainfall events. McCord and Loyacano (1978)
further found that Chinese water chestnut (Eleocharis deucis) in freshwater
ponds is effective in removing nitrate and ammonium from the water column.
In their studies, ponds with water chestnut had lower concentrations of
both nutrients and phytoplankton. Net nitrogen removal rates were
estimated to be in the range of 4 mg m~^d~^. Twilley et al. (1981)
found that nutrients (ammonium, nitrate, and dissolved inorganic phosphate)
are removed at substantial rates from brackish-water ponds dominated by P.
perfoliatus and R. maritima.
Although it appears that SAV reduces nutrient concentrations, existing
evidence suggests that when loading rates and concentrations of nutrients
reach certain levels, SAV is no longer effective. In fact, SAV can be
stressed at these elevated levels of nutrients through several mechanisms.
For example, Jupp and Spence (1977) found that in Loch Leven, Scotland, the
diversity of SAV was reduced from about 23 species in 1910 to about 12 in
1975, and that this pattern of decreased diversity and abundance generally
paralleled the increase in cultural eutrophication. They found that when
phosphorus levels approached 2 ug-at I~l, algal stocks increased
(particularly blue-greens), while SAV distribution, species diversity, and
abundance decreased to very low biomass levels (0 to 20 g m~2). They
further suggest that algal blooms may decrease SAV vigor through
attenuation of light and increases in pH. Chlorophyll levels in Loch Leven
were reported to exceed 200 ug I"*-, a concentration far in excess of
those normally found in Chesapeake Bay at this time. They found that SAV
tended to recover when chlorophyll levels were decreased to the vicinity of
20 to 40 ug I"1.
In summary, Jupp and Spence (1977) constructed the following story
concerning the effects of cultural eutrophication on SAV distribution.
These events are similar to the sequence of decline in Chesapeake Bay SAV.
477
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Increased loading of nutrients to Loch Leven, in particular phosphorus,
increased algal stocks that led to a decrease in available light, both
through/attenuation in the water column and through fouling on SAV by
epiphytic species. The light restriction led to a restricted depth zone in
which SAV species could flourish, and this, of course, was in shallower
water. This restricted zone of growth was in an area where SAV was
subjected to increased stresses by both wave action, and decreased light
due to resuspension of littoral sediments and intensive grazing by
waterfowl.
In other studies, similar results were found. Mulligan et al. (1976)
found that SAV subjected to very high levels of nitrogen and phosphorus
fertilization (4,000 g-at L~l of nitrogen, 25 ug-at L-l of phosphorus)
were eliminated in pond ecosystems. They conclude that high loading rates
of nitrogen and phosphorus favor phytoplankton stocks. Similar conclusions
were reached by Sand-Jensen (1977) and Phillips et al. (1978). In the
Chesapeake Bay area, experimental work by Twilley et al. (1981) suggests
that loading rates, resulting in initial nitrogen (N) and phosphorus (P)
concentrations of 60 and 6 ug-at L~l, respectively tend to favor the
development of algal stocks and the elimination of SAV.
Thus, it appears that the role of SAV in buffering nutrient
concentrations in the nearshore zone has at least two aspects. If loading
rates are moderate, SAV (and other littoral zone components) can rapidly
decrease these concentrations to low levels. If loading rates and
resulting concentrations are sufficiently high, SAV is disadvantaged and,
in some cases, lost from the system and replaced by a phytoplankton
component.
NUTRIENT REGULATION OF SAV GROWTH
Over the past fifty years, considerable (though sporadic) research has
been directed toward understanding sources from which SAV obtains
nutrients. Various studies indicate that root uptake is the major
mechanism through which nutrient demands are met (MeRoy and Barsdate 1970,
Cole and Toetz 1975, Nichols and Keeney 1976, Twilley et al. 1977). In
contrast to this, other evidence suggests that foliar uptake, under some
conditions, is the predominant pathway (Nichols and Keeney 1976, Cole and
Toetz 1975). We suggest that nutrient uptake is facultative in that if
nutrient concentrations in the water column are very low, and adequate
nutrient reserves exist in the sediment, then root uptake will dominante.
Alternatively, if adequate nutrients are present in the water column, then
foliar uptake will predominate.
To define ammonium uptake kinetics (Marbury et al. 1981), experiments
were conducted in the upper Bay using P. perfoliatus. Foliar uptake
matched classical Michaelis-Menten kinetics for both day and night
conditions, with root uptake also partially described by these kinetics.
In several experiments, Marbury found that K^ (j^ is the substrate
concentration with the rate of nutrient uptake's one-half of the maximum
uptake) approximates 15 u moles and corresponds to an uptake rate of
approximately 0.19 to 0.31 mg N g of plant-J-hr'1. These rates are
comparable with those reported by other authors for several different
species of SAV (Nichols and Keeney 1975, McRoy and Alexander 1975, Cole and
Toetz 1975). During periods of rapid growth (May-July), total
478
-------
I
concentration of inorganic N in the water column often approaches or
exceeds values of 1^ and the interstitial concentrations of ammonium are I
in the range of one to three m moles. Because of this, it is doubtful that I
nitrogen limits submerged macrophytes in the mid-salinity and brackish
water portions of Chesapeake Bay. (Marbury, personal communication). To •
the contrary, both observational and experimental evidence indicate that I
nutrient loading, particularly of N, may be sufficiently high to favor the
replacement of SAV communities by phytoplankton (Phillips et al. 1978, _
Twilley 1981). •
In the lower Bay, SAV growth may be somewhat more regulated by nutrient ^
availability. Orth (1977) added large amounts of commercial N .and P
fertilizer to the sediment surface in Zostera beds and found significant •
increases in length, biomass, and number of stems. Sediments were sandy jj
and may have had low concentrations of interstitial nutrients as has
previously been reported for such sediments. This, coupled with .
characteristically low-nutrient concentrations in the water column, may •
produce nutrient-limited growth. ™
NITROGEN FIXATION, NITRIFICATION, AND DENITIRIFICATION •
This section discusses three important processes in the nitrogen
cycle: nitrogen fixation, nitrification, and denitrification. SAV1s •
ability to convert dissolved nitrogen gas into an organic form ("fixing") I
is important during times of inorganic nitrogen impoverishment.
Nitrification is the bacterial-mediated oxidation of ammonia to nitrate in
the presence of free oxygen; denitrification is the reverse process of •
bacterial-mediated reduction in the absence of free oxygen. These last two •
processes provide energy to certain bacteria depending on whether the
environment is aerobic or anaerobic. •
These three processes are of ecological as well as of water quality |
significance, because they represent sources and sinks of nitrogen and may
reflect the potential for regulating phytoplankton growth in many areas of •
Chesapeake Bay. A substantial range in N-fixation rates has been observed •
in seagrass communities. It appears that in nutrient-poor waters (low
ambient concentrations of N in both the water column and in sediments), SAV
growth can be N-limited (Patriquin 1972). Much of the nitrogen used in SAV •
growth may be supplied by N-fixation (e.g., Capone et al. 1979). Patriquin •
(1972) reports high rates of N-fixation in Thalassia beds, and Patriquin
and Knowles (1972) conclude, based on studies in a variety of Thalassia •
beds in the Caribbean, that most of the N requirements are supplied by |
fixation. Fixation rates in these studies range between two to 10 mg-at
m~ld"l, rates that are capable of supporting most, if not all, of the «
calculated N demand. I
In contrast to these results, Lipschultz et al. (1979) report low rates
of N-fixation in seagrass meadows in the Choptank River estuary. In the
areas investigated by Lipschultz, nitrogen was abundant in the water I
column, and sediment reserves were substantial. Thus, it appears that •
N-fixation is facultative in the sense that if severe N-limitation exists
in environments otherwise amenable to seagrass growth, N-fixation becomes a M
prominent feature. Conversely, in those systems, such as the mid-salinity I
and brackish zones of Chesapeake Bay, where abundant reserves of ammonium
are contained in interstitial waters, N-fixation is simply not required. _
479
I
I
-------
I
I
I
I
I
I
I
I
I
fl
I
I
I
I
I
I
I
I
I
Unfortunately, less is known about the rates of nitrification and
denitrification in seagrass ecosystems. To our knowledge, the only
published information available at this time is the work of lizumi et al.
(1980). In this study, a Zostera bed was investigated using N*5
techniques. The authors report that rates of denitrification ranged from
0.5 to 1.2 x 10~9 g-at g~^h~^, and that nitrification rates were
quite similar. When these values are converted to an areal basis,
denitrification and nitrification are important aspects of the
sediment-water nutrient cycle. lizumi et al. (1980) report that high
nitrification rates are directly coupled to denitrification, as expected,
because the entry product (nitrate) to the denitrification pathway is the
end product of nitrification. Moreover, they found that nitrification in
anoxic sediments was made possible by the transport of oxygen from the
foliar portion of SAV to the root zone. Thus, there were small microzones
of oxidized sediment in which nitrification could proceed. After nitrate
was produced, it diffused into the anoxic zone, where denitrifying bacteria
rapidly transformed nitrate to nitrogen gas.
In studies conducted in an SAV community in the Choptank River and in
brackish-water experimental ponds, Twilley et al. (1981) found that
denitrification rates in both areas ranged from 50 to 100 uM m~2Q~l;
however, rates tended to be lower in SAV than in non-vegetated littoral
zones (although such differences were not statistically significant).
Jenkins (personal communication) found much higher rates of denitrification
(about 200 to 300 ug-at N m~~2d-l) in deeper portions of Chesapeake Bay
waters in the spring when nitrate was abundant in overlying waters. Rates
were low or undetectable at other times of the year when nitrate was not
present in the water column. Evidence that nitrification rates are
substantial in SAV communities is accumulating from studies of SAV beds in
the upper Bay, and these rates appear to be substantially higher than
nitrification rates in soft-bottom communities lacking SAV.
What then are the mechanisms responsible for these observations? At
this point, it seems that oxygen produced in the foliar portions of SAV is
translocated to the roots and from the roots into the interstitial waters,
supplying the oxygen needed to support nitrification. Although the nitrate
produced could be used in denitrifiction, evidence at this point indicates
that other processes may out-compete denitrification for this nitrate.
Recent studies by Terlizzi (personal communication) of diel nitrogen
cycling in P. perfoliatus-dominated microcosms (700 liter with natural
estuarine sediments and water) showed that nitrate concentrations increased
in the roots during the daylight hours and decreased at night with a
concomitant appearance of nitrite. They suggest that the oxygen produced
in this reaction was used in support of root respiration at night, yielding
nitrite. The eventual fate of the nitrite produced in these roots is not
currently known, although some of it leaked from the roots into the
interstitial and overlying waters; the nitrite had nearly vanished by the
return of daylight. Whether or not nitrite was oxidized to nitrate or
reduced to nitrous oxide or nitrogen gas is presently not known. Thus, in
contrast to the studies of lizumi et al. (1980), these results suggest that
denitrification is important in deep waters when nitrate is abundant in the
water column. In SAV communities, measurements of denitrification have, by
and large, indicated that rates are small. Nitrification, on the other
480
-------
I
hand, appears to be enhanced by the trans location of oxygen to the root
zone, but all of the nitrate produced does not appear to enter the
denitrification pathway.
NUTRIENT RELEASE AND OXYGEN DEMAND ASSOCIATED WITH SAV DECOMPOSITION
A great deal of evidence points to the importance of submerged and
emergent macrophytes as a source of detritus available to coastal and
estuarine heterotrophs. Paralleling this, there is a considerable amount I
of scientific literature concerning the decomposition and release of B
nutrients for some higher plants, and in particular, decomposit.ion
characteristics of Spartina. Surprisingly, less is known about •
decomposition characteristics of submerged macrophytic vegetation. Several •
studies are available, however, that are pertinent to a discussion of the
decomposition process. _
In addition to the role of SAV as a detrital food-source, the relative •
impact of decomposing plants has been investigated in terms of oxygen ™
utilization. We hypothesized that submerged aquatic vegetation serves as a
temporary nutrient sink in that during the growth of SAV, N and P are taken I
up from either the water or sediment, depending on local conditions, and •
incorporated in SAV biomass. However, SAV decomposes; during this process
oxygen demand is exerted, and nutrients are presumably released back to the •
water column. •
Data from studies comparing SAV with phytoplankton and a macrophytic
alga suggest that SAV decomposition exerts a small oxygen demand, tending _
to retain nutrients to a greater extent than other plants. Some •
experiments investigated the extent of the oxygen demand exerted during the ™
decomposition process and the rapidity with which nutrients are released to
the water column. Results are summarized in Figure 14 (Twilley, personal •
communication). In these experiments, a variety of primary producers, f
characteristic of the Chesapeake Bay system including Ulva and Spartina,
were placed in small laboratory microcosms and allowed to decompose over a _
90-day period. At frequent intervals, oxygen concentration, rate of oxygen •
concentration change, and ammonium and orthophsophate concentrations were
monitored. As indicated in Figure 14, the dry-weight loss expressed as a
percent per day was highest in phytoplankton and Ulva, somewhat less in •
three SAV species, and lowest in Spartina. The mean dry-weight loss per •
day developed in these experiments was only slightly lower than those
observed in field studies. Spartina and phytoplankton species had the •
highest rates of oxygen utilization; rates for the three SAV species f
(Milfoil, Potomageton, and Ruppia) were the lowest. These results suggest
that SAV exerts only a small oxygen demand on a daily basis over the _
decomposition period. This observation is important because in some parts •
of Chesapeake Bay, bottom waters become anoxic during the summer because of
excessive deposition of labile organic material (primarily of phytoplankton
origin). I
Nutrient releases from SAV species and Spartina were low relative to •
the release observed for phytoplankton cultures and Ulva. After 70 days of
incubation in microcosms, the ammonium concentrations in experimental •
systems of Milfoil, Potomageton, Ruppia, and Spartina were on the order of •
one to two ug-at L~J-, while phytoplankton and Ulva decomposition resulted
in concentrations in excess of 10 to 14 ug-at L~J-. A similar, although _
481
I
I
-------
1
1
Ik SAV SPECIES -4
•T RATE OF
•o
Id 2.0-
(/>
W)
Q
|_i
i_: 1 0~
i
^^
_ <*
^H ^i
• o~
18-
T,
- 6-
1* 4-
ul
UJ 0
o 2-
1°
0-
I"L 10-
B
o»
I3 5-
+ » 3
X
z
1
™~*
1 1 I0~
>? 5-
Io ^
"
0J
2
g
5
'PHYTO
a
i WEIGHT LOSS
* "T~h^^ §
=! 1 ft ^^^^- 1
0 R ^ .ai
" ^ i r
i $ * i
i 1
RESPIRATION RATE
N
j^
z
l-
Q.
| r^
> ^ ^l
(3
S
3N-S* i: ^/ \
5 "^-^^^^^^ ^
° — fc
j ~
2 ^
' '
\
X
g
z
-J
o
h-
>
X
a.
f
NHj RELEASE
(DAY 70)
\ =i §
\
\
\
P^_
5
X
O ^
j 1 a
.2 1 5> |
V s I1 ^
_X Q fl Q
"T— A -m &
1 I r 1 1 1 1
i 1 i
^ POl RELEASE
^
z
-ANKTO
a.
o
»-
>-
a.
(DAY 70)
0
^
i —
c-
n i r
5 i -5 1
Si! 1
i
-------
I
not quite so radical, difference was also noted for orthophosphate. After
70 days of incubation, phosphate concentrations in the phytoplankton tanks I
were about 50 ug-at L"1, while in the SAV and Spartina microcosms I
concentrations ranged from about 9 to 15 ug-at L~l.
Harrison and Mann (1975) conducted decomposition experiments in the •
laboratory using Zostera blades exposed to 20°C (68°F) temperatures. •
They observed that Zostera lost up to 35 percent of its dry weight in 100
days in decomposition. The decomposition rates for whole leaves and
particles less than one millimeter were approximately 0.5 percent and one •
percent a day. Leaching of organic matter was responsible for a large "
fraction of organic matter loss. In terms of the nutrient content of
detrital material, the addition of bacteria markedly increased the nitrogen •
content of organic matter, but did not substantially change the decay rate |
of detrital particles. The addition of protozoa with the bacteria
increased both the nitrogen content and the decay rate of detritus; C:N _
ratios changed from about 20:1 in living blades to a minimum of 11:1 in •
detrital particles subjected to bacterial and protozoan treatments. *
Harrison and Mann further found that total organic matter, dissolved
organic C, particulate organic C, and N were also highest in new Zostera I
leaves and decreased rapidly after death. I
In studies using the same species, Thayer et al. (1977) found that
during senescence, N content decreased and subsequently increased as blades m
became detrital. They attributed this action to microbial growth and •
further speculated that most of the nitrogen increase was due to microbial
immobilization of N from surrounding waters. If bacterial immobilization
of dissolved N is a general feature of the decomposition process, then it •
-represents yet another mechanism by which SAV can reduce ambient nutrient •
concentrations in the water column.
In studies conducted in Chesapeake Bay, Staver (personal communication) •
placed above-ground portions of living P_._ perfoliatus in three-millimeter |
and one-millimeter mesh nylon bags and suspended these in the field. Bags
were retrieved at different times, and the amount of SAV material remaining «
was measured. Results of these studies indicate that, at the temperatures •
commonly encountered [25 to 30°C (77 to 86°F)], decomposition in these
bags was rapid, averaging about two percent a day (Figure 15). Although
C:N ratios of this material are not available, we expect that over time the B
N content of remaining material would increase. Although data concerning B
decomposition and the nutritive status of decomposing material are far from
complete, evidence from other areas indicates that as submerged macrophytic •
material dies, there is an initial loss in many components, including N, Hj
followed by an increase in N content, probably mediated by bacterial
incorporation of N from the surrounding medium. This material probably _
serves as an adequate food source for many heterotrophs that ingest •
detrital particles, metabolize the microorganisms, and excrete the detrital ™
fragment.
COMPARISON OF NUTRIENT BUFFERING CAPACITY OF SAV WITH IMPORTANT SOURCES •
To evaluate the potential nutrient buffering role of SAV in the context •
of Bay-wide nutrient sources, we have developed a crude budget for which •
the magnitude of nitrogen sources to the upper Chesapeake Bay are compared
with the amount of nitrogen incorporated into SAV biomass during a normal
growing season. •
483
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
JULY-AUG
1980
3mm MESH
I mm MESH
DAYS OF DEPLOYMENT
Figure 15.
Decomposition rates of _P. perfoliatus estimated using in situ
litter bags. Data were collected in the vicinity of the
Todds Cove study site in the Choptank River (Staver,
personal communication).
484
-------
I
As indicated in Table 7, something on the order of five percent of the
total nitrogen input to the upper Chesapeake Bay could have been •
immobilized by incorporation into SAV biomass during the 1960's. An •
extremely small percentage of total nitrogen input may be immobilized by
SAV uptake at the present time (0.5 percent). Estimates of sewerage input •
during the 1960's were not available, and we were not able to contrast SAV j|
uptake relative to sewerage input. However, it is interesting to note that
SAV uptake in the 1960's could account for approximately 50 percent of the _
present sewerage input. Uptake represents only one of several possible I
mechanisms used by SAV to buffer the nutrient regime in estuarine waters. *
As indicated earlier, denitrification may represent a substantial sink,
although at this point the exact magnitude of this process remains •
unclear. It should be pointed out that Table 7 provides no estimate of |
atmospheric input; however, it probably approximates 10 to 12 percent of
the total value for 1978 (Smullen et al. 1982). _
TABLE 7. ESTIMATED INPUTS OF NITROGEN TO THE UPPER CHESAPEAKE BAY FROM "
RIVERINE AND SEWAGE SOURCES, AND UPTAKE OF NITROGEN BY SAV
Sources
Riverine Input3
Sewage Inputs**
Total
SAV Uptake0 (During growing season)
Time
1960
50
d
» 50
2.4
Periods
1978
50
5.3
55.3
0.3
(ALL VALUES ARE IN UNITS OF KgNy-ixlO6) •
(MACOMBER 1980; STEVENSON, PERSONAL COMMUNICATION) »
I
I
I
_ I
aRiverine source of nitrogen calculated using regression relationships _
between Susquehanna River flow and nutrient concentrations (Guide and •
Villa 1972). "
^Sewage input data from Smullen (personal communication) _
cUptake calculated using N content of SAV of 2% and SAV Standing crop of •
200 gM~2 for both time periods. Areas of SAV coverage were estimated as
600xlo6m2 and 66xl06m2 in 1960 and 1978, respectively (Rawls, in
prep.; Anderson and Macomber 1980; Stevenson, personal communication). •
dNot available.
I
I
I
I
485 •
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
SECTION 6
SUMMARY
Four distinct processes related to the ecological role and value of SAV
were examined in the Chesapeake Bay Program: (1) estimating the magnitude
of SAV organic matter production and availability to local food webs; (2)
examining habitat value of SAV to infaunal and juvenile nekton species; (3)
estimating the role of SAV in modifying, reducing, and serving as a sink
for nearshore sediments; and (4) examining the role of SAV in modifying
nutrient dynamics of nearshore regions.
As we have shown in previous sections of this report, it appears that
SAV influences each of these processes. However, the importance of the SAV
component in the Bay community at the present time is probably small
because of the restricted distribution of this vegetation. At one time,
SAV probably played a substantial role in organic matter production,
habitat maintenance, and sediment and nutrient dynamics. The purpose of
this section is to highlight findings concerning the role of SAV in the
above processes and to place processes associated with SAV in the context
of large portions of Chesapeake Bay.
ORGANIC MATTER PRODUCTION AND UTILIZATION
The productivity rates of SAV communities in Chesapeake Bay are
comparable to with those observed in other SAV systems distributed over
large latitudinal ranges and environmental gradients. Net productivity
values associated with several types of SAV in Chesapeake Bay were as high
as those reported for other species in other areas. In sharp contrast to
the comparability of production values between SAV systems, estimates of
SAV biomass exhibited a large over all range, and substantial differences
were evident within the same type of system. In general, higher standing
stock values of SAV occurred in areas where the water is relatively clear,
deep (enough to allow for substantial vertical growth of SAV), and devoid
of extensive wave action. Moreover, average biomass (and even maximum
biomass) estimates in Chesapeake Bay were low relative to those reported
for other areas. For instance, average values of Zostera and Ruppia in the
lower Bay were generally below 200 g m~2; values for Potomageton
pectinatus and P^_ perfoliatus in the upper Bay were generally below
100 g m~z. At the present time, sufficient light to support vigorous
growth of SAV does not penetrate much beyond one meter in most littoral
regions of the upper Chesapeake Bay. Thus, growth is restricted in very
shallow regions where there is a limited water column to support the
vertical development of SAV, and the potential for wave, thermal, and
waterfowl grazing stresses is maximized. In earlier years (pre-1970) when
light penetration was not so restricted, SAV in the upper Bay may have
grown in waters of greater depth and been characterized by higher standing
stocks.
Comparison of SAV biomass for several species in the lower and upper
Chesapeake Bay has made several differences apparent: (1) peak biomass of
M. spicatum was greater than Z. marina, and the biomass of this species
was greater than EU_ maritima; TT) the peak biomass of R_._ maritima, _?._
486
-------
I
pectinatus and £._ perfoliatus approximated each other; (3) with the
exception of M. spicatum, mean biomass values were consistently higher in •
the lower Bay, often by a factor of two or more; (4) in the lower Bay, V
above-ground biomass persisted through winter months, but in the upper Bay,
above-ground material was present only during the warmer months; and (5) •
periods of peak biomass occurred earlier in the year (June) in the lower •
Bay than in the mid-salinity zone (July to August). We do not have
quantitative information concerning biomass levels or seasonal persistence _
prior to the initiation of the decline. However, anecdotal information I
suggests that biomass values were higher in the upper Bay than they are at ™
the present time and persisted through the fall months.
Submerged aquatic vegetation can enter heterotrophic food webs either B
by direct grazing of living plants or by consumption of SAV detritus. The |
majority of studies conducted suggest that SAV is an adequate food item,
that it is primarily available as detritus and, in some localities, it may •
be a dominant food source. Several substantial results link SAV production •
to use in Bay food webs, with perhaps the most definitive connection
between SAV and waterfowl. Numerous authors found that vegetable matter
was an extremely important food item for waterfowl in the upper Chesapeake I
Bay. Furthermore, the most important waterfowl wintering areas are also •
those most abundantly vegetated. It appears that direct grazing on SAV by
waterfowl is important both in the Chesapeake and elsewhere, and that *
grazing in itself can locally impact the distribution of SAV. |
Aside from this direct grazing pathway, the vast majority of studies,
including those in Chesapeake Bay, indicate that most SAV material enters _
food webs through detrital pathways. For example, in the lower Bay, sea •
bass, pipefish, pigfish, and white perch are epibenthic feeders, using *
amphipods and shrimp that are, in turn, detrital feeders. Data further
indicate that large predators enter SAV beds with little food in their I
stomachs and leave after feeding. Food items for these feeders can •
generally be traced back to detrital sources, some fraction of which is
probably SAV in origin. In an extensive study of feeding habits in the •
upper Bay, little evidence was found for direct grazing by fish on SAV, J|
although some SAV seeds and plant parts were found in stomachs. Energy
flow appears to enter food webs as detritus and to pass through epifaunal _
and infaunal invertebrates to small and large fish. Numerous epifaunal •
species, which are important food items for many consumers, were also ™
closely associated with SAV.
Because of the complexity of organic matter sources in Chesapeake Bay •
and the current marginal distribution of SAV, a quantitative assessment of |
SAV's importance as a food source is not possible. However, it is
reasonable to argue that the available SAV is used by heterotrophs, and •
that SAV's physical structure concentrates other foods (phytoplankton, •
epiphytic algae, and benthic macroalgae) for animal consumption. Studies
conducted concurrently with SAV research indicate that on an annual basis
virtually all carbon inputs in Chesapeake Bay were utilized by heterotrophs •
of one sort or another. Thus, if heterotrophic metabolism is •
organic-matter limited, it follows that SAV would also be used if, as has
already been shown, this material is a suitable food source. Furthermore, •
loss of SAV production may well lead to loss of fishery production, jf
especially if production by phytoplankton fails to compensate for the loss
of food to higher trophic levels. «
487
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
HABITAT VALUE OF SAV
Studies in the upper and lower Chesapeake Bay indicated that infaunal
abundance and diversity is higher in vegetated than in unvegetated areas.
Because unvegetated habitats support a virtually non-existent epifauna,
epifaunal densities were naturally higher at the vegetated sites and were
important food items in Chesapeake Bay food webs.
Finfish sampling at sites in the upper Bay indicate greater abundances
and species richness in vegetated than in unvegetated bottoms; fish
densities were among the highest yet reported in the literature. In
addition, average weight per individual during the summer period was low in
SAV communities as compared with unvegetated areas, suggesting that SAV
communities are continually used as nursery areas for small fish, and
larger-sized animals predominate in unvegetated areas. Fish sampling
programs in the lower Bay also found greater abundances and species
richness in eelgrass meadows than in nearby unvegetated bottoms. Some
large fish predators, such as weakfish and the sandbar shark, foraged most
often over vegetated bottom, whereas others, such as bluefish, appeared to
forage indiscriminately over both vegetated and unvegetated areas.
The main conclusion of these field studies is that fish communities are
richer in vegetated than unvegetated areas. However, few commercially
important finfish were found to use SAV beds as significant nursery
habitats. The role of SAV for commercial fishes in the Chesapeake system
seems to be largely that of a rich foraging place for adults, and not that
of a nursery habitat although, once again, it is important to emphasize
that the current restricted distribution of SAV may bias these
conclusions. (For instance, major spawning and juvenile habitats for
striped bass once existed in the upper Bay in an area that was densely
populated with SAV.) More representative patterns of commercial fish use
of SAV habitat might best be evaluated through historical correlations of
SAV and juvenile fish distributions.
Information concerning blue crab abundance was collected at the same
time fish were sampled at sites in the upper and lower Chesapeake Bay.
During comparable months, up to 10,000 times as many blue crabs were found
at the lower Bay site. In addition, most of the crabs in the high-salinity
eelgrass beds were juvenile females that constituted the breeding stock for
future generations. The conclusion drawn from these studies is that SAV in
the upper Bay serves as a very limited blue crab nursery. Lower Bay
eelgrass beds, however, serve as primary blue crab nurseries, supporting
very large numbers of juvenile blue crabs throughout the year. It should
be noted, however, that upper Bay SAV beds may well provide a protective
habitat for molting adult blue crabs.
Experimental studies involving exclusion of predators from certain
areas of SAV beds indicate that predation rates on some infaunal taxa were
lower in vegetated than unvegetated areas, and predation rates on epifaunal
species seemed to be lower in SAV habitats than elsewhere. Laboratory
microcosm experiments supported the notion that SAV provides less
protection for infauna than it provides for epifauna of SAV beds. It also
seems that artificial eelgrass provides protection roughly equivalent to
that of live eelgrass, and that SAV species with finely divided leaves
provide (other factors being equal) more protection than do SAV with
simple, unbranched leaves. It is clear that SAV-associated animals do feed
488
-------
in the beds, and that the food supply is considerably greater in SAV
communities than in other available habitats.
I
I
SEDIMENT PROCESSES
Results of recent studies indicate that SAV can substantially influence •
sediment dynamics in littoral zones. Specifically, SAV stabilizes
sediments; slows currents, allowing sediments to settle (which increases _
light penetration into the water column); and substantially reduces •
current-and-wave-induced resuspension. In Chesapeake Bay, Orth (1977) •
reports that sediment particle diameter decreased, and organic matter
content and infaunal densities increased in sediments in areas with SAV, as •
compared with those that did not have such coverage. Other findings and fj
observations showed less sediment disruption during storms in vegetated
zones, and less dispersion of dyed sand patches. Based on this information «
it was concluded that SAV is effective at trapping and consolidating •
suspended sediments.
Data developed at intensive study sites in the upper Chesapeake Bay
indicated that as turbid water entered SAV beds on rising tides, sediments •
were effectively removed, increasing light transparency. It appears that •
by high tide, turbid inflowing waters normally exceed the filtering
capacity of SAV beds. As tidal height decreases, the bed effectively m
filters sediments, and the turbidity gradient between vegetated and •
non-vegetated areas again increases to a maximum. In addition,
resuspension was reduced in SAV communities with the reduction proportional
to SAV biomass. I
Because of the dynamic nature of the sediment-water interface in ^
littoral environments, estimates of net sedimentation are exceedingly
difficult to obtain. A crude estimate was made from observations based on •
differences in seston concentrations inside and outside SAV beds. These £
calculations indicated that daily deposition rates of sediment were about
63 g m~2d~l, yielding seasonal estimates on the order of 1,200 g tm
m~2. if we assume that there is about 0.6 g ctn-^ of inorganic material •
in consolidated sediments, this deposition is equivalent to about two
millimeters per growing season. At the present time, we do not know if
material deposited when SAV was present is subsequently lost when SAV dies •
in the fall. Considering the important role of roots and rhizomes in the •
process of sediment consolidation and the low below-ground biomass observed
in Chesapeake Bay SAV communities, we doubt that this material is •
permanently consolidated at present, although it may have been in the past. •
To place the sediment-trapping characteristics of SAV in the context of
larger-scale sediment processes in Chesapeake Bay, we have developed a _
series of calculations that compare the magnitude of two major sediment •
sources with the deposition rate observed in SAV communities. Major *
sediment sources include riverine input and shoreline erosion to the
portion of Chesapeake Bay above the mouth of the Potomac River. The amount V
of sediment deposited during the SAV growing season was based on the data •
of Boynton et al. (1981b). They estimated that some 1,200 grams of
sediment were deposited per square meter of SAV community over an estimated •
180-day growing season. A large percentage of sediment input could have •
been deposited in SAV communities during the 1960 period. However, in the
late 1970's, when SAV distributions were severely reduced, the amount that
489
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
could have been deposited was reduced to something less than 10 percent of
the input. Although this calculation should be considered preliminary, it
suggests that SAV in the past may have played an important role in
sequestering sediments in Chesapeake Bay, and that the amount of sediment
deposited in SAV communities at the present time is small relative to
estimates of sediment input. Presumably, the difference in sediment
trapped in 1960 versus 1978 is spread over the bottom of the Bay, with some
part available to resuspension.
NUTRIENT PROCESSES IN SAV COMMUNITIES
In an earlier section of this report, we argued that SAV communities
are capable of buffering nutrients between littoral and pelagic zones of
the estuary. Recent studies in Chesapeake Bay tend to support this
notion. Measurement of nutrient concentrations in offshore areas and in
SAV communities indicates that nutrient concentrations are consistently
lower in the SAV communities. In addition, experimental studies involving
the addition of nutrients to SAV communities indicate that nutrients are
rapidly removed from the water column; ambient nutrient levels are
reestablished 12 to 24 hours after additions. It has not been determined,
however, which autotrophic component is most responsible for the uptake of
these nutrients.
Several experimental studies were also conducted to examine rates of
nitrification and denitrification in SAV communities. These experiments
attempted to quantify the potential of SAV as a nutrient sink. In studies
conducted in an SAV community in the Choptank River and in brackish water
experimental ponds, we found that both areas exhibited substantial
denitrif ication rates (50 to 100 ug-at m~2cj~l) and rates tended to be
lower in SAV than in nonvegetated littoral zones. Studies in the upper Bay
also showed that nitrification rates are substantial in SAV communities.
These rates were higher in SAV areas than in soft-bottom communities not
having SAV. Although nitrate produced from this reaction could be used in
denitrification, evidence at this point indicates that other processes may
out-compete denitrification for this nitrate.
In addition to the role of SAV as a detrital food source, the relative
impact of decomposing plants on oxygen utilization and nutrient release
rates was investigated. We found that Spartina alterniflora and
phytoplankton species had the highest rates of oxygen utilization, and
three SAV species had the lowest rates. These results suggest that SAV
exerts only a small oxygen demand on a daily basis over decomposition
periods. This observation is important in that bottom waters in some parts
of Chesapeake Bay become anoxic during the summer because of excessive
deposition of labile organic material.
Nutrient release rates from SAV species and Spartina were low relative
to release rates observed for phytoplankton cultures and Ulva. After 70
days of incubation, ammonium concentration in experimental systems of SAV
was one to two ug—at L~l, while phytoplankton and Ulva decomposition
resulted in concentrations in excess of 10 to 15 ug-at L~l,
respectively. These data suggest that SAV exerts a small oxygen demand
while decomposing and tends to retain nutrients relative to phytoplankton
and to the one macrophytic alga tested.
490
-------
I
To evaluate the potential nutrient buffering role of SAV in the context
of Bay-wide nutrient sources, we developed a crude budget in which the m
magnitude of nitrogen sources to the upper Chesapeake Bay were compared •
with the amount of nitrogen incorporated into SAV biomass during a normal
growing season. About five percent of the total nitrogen input to the _
upper Chesapeake Bay could be immobilized via incorporation into SAV •
biomass during the 1960's. In contrast, an extremely small percentage of ™
total nitrogen input (0.5 percent) could be immobilized via uptake by SAV
at the present time.
491
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Literature Cited
Adams, S.M. 1976. Feeding Ecology of Eelgrass Fish Communities. Trans.
Amer. Fish. Soc. 105:514-519.
Aioi, K. 1980. Seasonal Change in the Standing Crop of Eelgrass (Zostera
marina L.) in Odawa Bay, Central Japan. Aquat. Bot. 8:343-354.
Allison, A., and I. Newton. 1974. Waterfowl in Loch Leven, Kinross. Proc.
R. Soc. Edinb. B. 74:365-81.
Anderson, R.R., and R. T. Macomber. 1980. Distribution of Submerged
Vascular Plants, Chesapeake Bay, Maryland. Report to USEPA by Aero
Eco, Reston, VA.
Ball, M., E. A. Shinn, and K. Stockman. 1967. The Geologic Effects of
Hurricane Donna in South Florida. J. Geol. 75:583-597.
Biggs, R. B. 1970. Sources and Distribution of Suspended Sediment in
Northern Chesapeake Bay. Mar. Geol. 9:187-201.
Biggs, R. B., and D. A. Flemer. 1972. The Flux of Particulate Carbon on
an Estuary. Mar. Biol. 12:11-17.
Boynton, W. R., C. A. Hall, P. G. Falkowski, C. W. Keefe, and W. M. Kemp.
1981a. Phytoplankton Productivity in Aquatic Ecosystems. Encyclopedia
of Plant Physiology, New Series, Vol. B. Part 18.6. 0. L. Lange, ed.
Springer-Verlag, NY. (In review).
Boynton, W. R., W. M. Kemp, L. Lubbers, III, K. Kaumeyer, S. Bunker, K.
Staver, and J. Means. 1981b. Influence of Submerged Macrophyte
Communities on Turbidity and Sedimentation in Littoral Zones of
Northern Chesapeake Bay. In: Submerged Aquatic Vegetation in
Chesapeake Bay: Its Ecological Role in Bay Ecosystems and Factors
Leading to Its Decline. W. M. Kemp, J. C. Stevenson, W. R. Boynton, J.
C. Means, eds. Horn Point Environmental Laboratories, Cambridge, MD.
pp. 842-878.
Boynton, W. R., W. M. Kemp, K. V. Wood, C. W. Keefe, and C. F. D'Elia.
1981c. The Significance of Nutrient and Carbon Fluxes Across the
Sediment-Water Interface Along Major Estuarine Gradients. Unpublished
manuscript. Chesapeake Biological Laboratory, Solomons, MD.
Boynton, W. R., W. M. Kemp, C. G. Osborne, and E. Spaulding. 1981d.
Sediment Dynamics in the Patuxent Estuary. Unpublished data.
Chesapeake Biological Laboratory, Solomons, MD.
Boynton, W. R., W. M. Kemp, and C. W. Keefe. 1982. A Comparative Analysis
of Nutrients and Other Factors Influencing Estuarine Phytoplankton
Production. Chesapeake Biological Laboratory, Solomons, MD. In review.
492
-------
I
Brooks, H. A., J. V. Merriner, C. E. Meyers, J. E. Olney, G. W. Boehlert,
J. V. Lascara, A. D. Estes, and T. A. Munroe. 1981. Higher Level «
Consumer Interactions. Chapter 4. Virginia Institute of Marine •
Sciences, Gloucester Point, VA.
Brush, G. S., F. W. Davis, and S. Rumer. 1980. Biostratigraphy of •
Chesapeake Bay and Its Tributaries: A Feasibility Study. USEPA No. ™
60018-80-040.
Brush, G. S., F. W. Davis, and S. A. Stenger. 1981. Sediment |
Accumulation and the History of Submerged Aquatic Vegetation in the
Patuxent and Ware Rivers: A Stratigraphic Study. The Johns Hopkins «
University. Baltimore, MD. M
Bunker, S., L. Lubbers, K. Kaumeyer, K. Staver, W. Boynton, M. Kemp, and
C. Spaulding. 1981a. Comparative Food Habits of Littoral Fish •
Communities Associated with Vegetated and Non-Vegetated Sites in ||
Chesapeake Bay. In: Submerged Aquatic Vegetation in Chesapeake Bay:
Its Ecological Role in Bay Ecosystems and Factors Leading to Its M
Decline. W. M. Kemp, J. C. Stevenson, W. R. Boynton, J. C. Means, •
eds. Horn Point Environmental Laboratories, Cambridge, MD. pp.
609-633.
1
Bunker, S., W. Boynton, M. Kemp, L. Lubbers, and P. Parker. 1981b. •
Relationships Among Lower Trophic Levels in an Estuarine Macrophyte
Community as Indicated by Stable Carbon Isotopic Analysis. In: •
Submerged Aquatic Vegetation in Chesapeake Bay: Its Ecological Role in £
Bay Ecosystems and Factors Leading to Its Decline. W. M. Kemp, J. D.
Stevenson, W. R. Boynton, J. C. Means, eds. Horn Point Environmental ^
Laboratories, Cambridge, MD. pp. 634-662. •
Bunker, S., K. Kaumeyer, L. Lubbers, K. Staver, C. Spaulding, W. Kemp, and
W. Boynton. 1981c. The Infaunal Macroinvertebrate Community •
Associated with Submerged Macrophytes in the Chesapeake Bay. In: •
Submerged Aquatic Vegetation in Chesapeake Bay: Its Ecological Role in
Bay Ecosystems and Factors Leading to Its Decline. W. M. Kemp, J. C. m
Stevenson, W. R. Boynton, J. C. Means, eds. Horn Point Environmental •
Laboratories, Univ. of Md., Cambridge, MD. pp. 423-446.
Carr, W. E. S., and C. A. Adams. 1973. Food Habits of Juvenile Marine •
Fishes Occupying Seagrass Beds in the Estuarine Zone near Crystal •
River, Florida. Trans. Am. Fish. Soc. 102(3):511-540.
Capone, D. G., P. A. Penhale, R. S. Oremland, and B. F. Taylor. 1979. |
Relationship Between Productivity and ^^2^) Fixation in a
Thalassia testudinum Community. Limnol. Oceanogr. 24:117-125. ^
Christiansen, C., H. Christoffersen, J. Dalsgaard, and P. Nornberg. 1981. *
Coastal and Near-Shore Changes Correlated with Die-Back in Eelgrass
(Zostera marina, L.). Sedimentary Geology. 28:163-173. •
493
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1
I
Cole, B. S., and D. W. Toetz. 1975. Utilization of Sedimentary Ammonia by
Potamogeton nodosus and Scirpus. Verh. Internat. Verein. Limnol.
19:2765-2772'
Day, J. H. 1967. The Biology of the Knysna Estuary, South Africa. In:
Estuaries Amer. Assoc. Adv. Sci. Publ., 83. G. H. Lauff, ed.
Washington, DC. pp. 397-407.
Dayton, P. K., and J. S. Oliver. 1980. An Evaluation of Experimental
Analyses of Population and Community Patterns in Benthic Marine
Experiments. In: Marine Benthic Dynamics. K. R. Tenore and B. C.
Coull, eds. University of South Carolina Press, Columbia, SC.
pp. 93-120.
den Hartog, C. 1967. The Structural Aspect in the Ecology of Seagrass
Communities. Helgolander Wissen. Meersunter. 15:648-653.
Dennison, W. 1979. Light Adaptation of Plants: A Model Based on the
Seagrass Zostera marina L. M.S. Thesis, Univ. of Alaska, Fairbanks. 69
pp.
Diaz, R., and T. Fredette. 1981. Secondary Production of Some Dominant
Macroinvertebrate Species Inhabiting a Bed of Submerged Vegetation in
the Lower Chesapeake Bay. Chapter 3, Section IV. Virginia Institute
of Marine Sciences, Gloucester Point, VA.
Ejdung, G., G. Romare, K. Staver, K. Kaumeyer, S. Bunker, L. Lubbers,
W. Boynton, and W. M. Kemp. 1981. The Influence of Predation on
Macroinvertebrate Infauna as Modulated by Submerged Macrophytes in the
Northern Chesapeake Bay. In: Submerged Aquatic Vegetation in
Chesapeake Bay: Its Ecological Role in Bay Ecosystems and Factors
Leading to Its Decline. Kemp, J. C. Stevenson, W. R. Boynton, J. C.
Means, eds. Horn Point Environmental Laboratories, Univ. of Md.,
Cambridge, MD. pp. 575-607.
Farrington, J. W. 1971. Benthic Lipids of Narragansett Bay - Fatty Acids
and Hydrocarbons. Ph.D. Thesis. Univ. R.I., Kingston, RI.
Flemer, D. A. 1970. Primary Production in the Chesapeake Bay. Ches.
Sci. 11:117-129.
Fry, B., and P. L. Parker. 1979. Animal Diet in Texas Seagrass Meadows:
13c Evidence for the Importance of Benthic Plants. Est. Coastal
Marine Science. 8:499-509.
Gessner, F. 1971. The Water Economy of the Sea Grass Thalassia
testudinum. Marine Biol. 10(3):258-260.
Gessner, F., and L. Hammer. 1960. Die Primarproduktion in Mediterranean
Caulerpa-Cymodocea Wiesen. Bot. Marina. 2:157-163.
494
-------
I
Ginsburg, R. N. 1956. Environmental Relationships of Grain Size and
Constituent Particles in Some South Florida Carbonate Sediments. Am. M
Assoc. Petroleum Geologists Bull. 40:2384-2427. •
Ginsburg, R. N., and H. A. Lowenstam. 1958. The Influence of Marine Bottom
Communities on the Depositional Environment of Sediments. J. Geol. •
66:310-318. •
Guide, V., and 0. Villa, Jr. 1972. Chesapeake Bay Nutrient Input Study. •
Tech. Rept. 47. Region III, EPA, Annapolis Field Office, Annapolis, MD. |
Harrison, P. G., and K. H. Mann. 1975. Chemical Changes During the
Seasonal Cycle of Growth and Decay in Eelgrass (Zostera marina) on
Atlantic Coast of Canada. J. Fish Res. Board Can. 32(5):615-621.
Haven, D. S., and R. Morales-Alamo. 1972. Biodeposition as a Factor in M
Sedimentation of Fine Suspended Solids in Estuaries. In: •
Environmental Framework of Coastal Plain Estuaries. B. W. Nelson, ed.
Geol. Soc. of America, Memoirs 133. pp. 121-130. •
Heck, K. L., Jr., and T. A. Thoman. 1981. Experiments on Predator-Prey
Interactions in Vegetated Aquatic Habitats. J. exp. Marine Biol.
Ecol. 53:125-134. •
Heinle, D. R., D. A. Flemer, and J. F. Ustach. 1977. Contributions of
Tidal Marshlands to Mid-Atlantic Estuarine Food Chains. In: Estuarine •
Processes, Vol. II. M. L. Wiley, ed. Academic Press, NY. pp. 309-320. Q
Heinle, D. R., C. F. D'Elia, J. L. Taft, J. S. Wilson, M. Cole-Jones, A. B. ^
Caplins, and L. E. Cronin. 1980. Historical Review of Water Quality •
and Climatic Data From Chesapeake Bay with Emphasis on Effects of
Enrichment. Final Rept. Grant R806189010, EPA Ches. Bay Program.
Hirschberg, D. J., and J. R. Schubel. 1979. Recent Geochemical History of •
Flood Deposits in the Northern Chesapeake Bay. Estuarine Coastal Mar.
Sci. 9:771-784. •
Hocutt, G. E., and R. W. Dimmick. 1971. Summer Food Habits of Juvenile
Wood Ducks in East Tennessee. J. Wildl. Mgmt. 35:286-292. ^
Hoese, H. D., and R. S. Jones. 1963. Seasonality of Larger Animals in a *
Texas Grass Community. Publ. Inst. Mar. Sci. Univ. Tex. 8:212-215.
Howard-Williams, C. 1981. Studies on the Ability of a Potomageton |
pectinatus Community to Remove Dissolved Nitrogen and Phosphorus
Compounds from Lake Water. J. Appl. Ecol. 18:619-637. am
lizumi, H., A. Hattori, and C.P. McRoy. 1980. Nitrate and Nitrite in
Interstitial Waters of Eelgrass Beds in Relation to the Rhizosphere.
J. Exp. Mar. Biol. Ecol. 47:191-201. •
Jacobs, R.P.W.M. 1979. Distribution and Aspects of the Production and •
Biomass of Eelgrass, Zostera marina L., at Roscoff, France. Aquat.
Bot. 7:151-172. tt
I
495 .
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Jones, J.A. 1968. Primary Productivity by the Tropical Marine
Turtlegrass, Thalassia testudinum Konig, and Its Epiphytes. Ph.D.
Dissertation, Univ. Miami, Coral Gables, FL. 196 pp.
Jupp, B.P., and D.H.N. Spence. 1977. Limitations of Macrophytes in a
Eutrophic Lake, Lock Leven. II. Wave Action, Sediments and Waterfowl
Grazing. J. Ecol. 65:431-446.
Kaumeyer, K., W.R. Boynton, L. Lubbers, K. Staver, S. Bunker, W.M. Kemp,
and J.C. Means. 1981. Metabolism and Biomass of Submerged Macrophyte
Communities in Northern Chesapeake Bay. In: Submerged Aquatic
Vegetation in Chesapeake Bay: Its Ecological Role in Bay Ecosystems
and Factors Leading to Its Decline. W.M. Kemp, J. C. Stevenson, W. R.
Boynton, J. C. Means, eds. Horn Point Environmental Laboratories,
Cambridge, MD. pp. 353-400.
Kemp, W.M. 1980. Physical and Geological Processes in the Patuxent
Estuary. In: Reviews of Patuxent Estuary Data Base. J.A. Mihursky
and W.R. Boynton, eds. Chesapeake Biological Laboratory Ref. No UMCEES
78-157 CBL. Solomons, MD. pp. 52-70.
Kemp, W.M., and W.R. Boynton. 1980. Influence of Biological and Physical
Processes on Dissolved Oxygen Dynamics in an Estuarine System:
Implications for Measurement of Community Metabolism. Estuarine and
Coastal Mar. Sci. 11:404-431.
Kemp, W.M., and W.R. Boynton. 1981. External and Internal Factors
Regulating Metabolic Rates of an Estuarine Benthic Community.
Oecologia. (in press).
Kemp, W.M., J.C. Stevenson, W.R. Boynton, and J.C. Means, eds. 1981.
Submerged Aquatic Vegetation in Chesapeake Bay:. Its Ecological Role in
Bay Ecosystems and Factors Leading to Its Decline. Univ. of Md. Horn
Point Environmental Laboratories, Cambridge, MD.
Klug, M.J. 1980. Detritus-Decomposition Relationships. In: Handbook of
Seagrass Biology. R.C. Phillips and C.P. McRoy, eds. Garland STPM
Press, NY. pp. 225-246.
Lipschultz, F., J.J. Cunningham, and J.C. Stevenson. 1979. Standing Crop
and Root:Shoot Ratio of Myriophyllum spicatum, Potamogeton perfoliatus,
and Ruppia maritima in the Upper Chesapeake Bay. Unpublished
manuscript. Horn Point Environmental Laboratories, University of
Maryland, Cambridge, MD.
Livingston, R.J. 1975. Impact of Kraft Pulp-Mill Effluents on Estaurine
and Coastal Fishes in Apalachee Bay, Florida. Marine Biol. 32:19-48.
496
-------
I
Lubbers, L., S. Bunker, K. Staver, W. Boynton, N. Burger, M. Meteyer, W.M.
Kemp. 1981. Comparative Abundance and Structure of Littoral Nekton •
Communities at Vegetated and Non-Vegetated Sites in the Chesapeake |
Bay: Its Ecological Role in Bay Ecosystems and Factors Leading to Its
Decline. Kemp, J. C. Stevenson, W. R. Boynton, J. C, Means, eds. Horn A
Point Environmental Laboratories, Univ. of Md., Cambridge, MD. pp. •
461-574. •
Mann, K.H. 1971. Ecological Energetics of the Seaweed Zone in a Marine I
Bay on the Atlantic Coast of Canada. I. Zonation and Biomass of m
Seaweeds. Marine Biol. 12(1):1-10.
Marbury, D., J. Metz, L. Lane, J.C. Stevenson, W.M. Kemp, and R.R. •
Twilley. 1981. Nitrogen Uptake Kinetics for the Submerged Estuarine
Macrophyte Potamogeton perfoliatus. In: Submerged Aquatic Vegetation
in Chesapeake Bay: Its Ecological Role in Bay Ecosystems and Factors •
Leading to Its Decline. W.M. Kemp, J. C. Stevenson, W. R. Boynton, J. »
C. Means, eds. Horn Point Environmental Laboratories, Cambridge, MD.
pp. 804-841. •
Marshall, N. 1970. Food Transfers Through the Lower Trophic Levels of the
Benthic Environment. In: Marine Food Chains. J.H. Steele, ed. ^
Oliver and Boyd, Edinburgh. I
McAtee, W.L. 1917. Propagation of Wild-Duck Foods. Bull. 465. U.S.
Dept. of Agriculture. •
McCord, C.L., Jr., and H.A. Loyacano, Jr. 1978. Removal and Utilization
of Nutrients by Chinese Water Chestnut in Catfish Ponds. Aquaculture. •
13:143-155. |
McRoy, CiP. 1970. Seagrass Productivity: Carbon Uptake Experiments in
Eelgrass, Zostera marina. Aquaculture. 4:131-137. •
McRoy, C.P., and R.J. Barsdate. 1970. Phosphate Absorption in Eelgrass.
Limnol. Oceanogr. 15:6-13. •
McRoy, C.P., and C. McMillan. 1973. Production Ecology and Physiology of
Seagrasses. International Seagrass Workshop, October 22-26, 1973. fj|
Teiden, The Netherlands. 29 pp. •
McRoy, C.P., and V. Alexander. 1975. Nitrogen Kinetics in Aquatic Plants
in Arctic Alaska. Aquatic Botany. 1:3-10.
McRoy, C.P., and C. McMillan. 1977. Production Ecology and Physiology of
Seagrass. In: Seagrass Ecosystems. E.P. McRoy and C. Helfferich,
eds. Marcel Dekker, Inc., NY. pp. 53-88.
I
Mickle, A.M., and R.G. Wetzel. 1978. Effectiveness of Submerged _
Angiosperm-Epiphyte Complexes on Exchange of Nutrients and Organic •
Carbon in Littoral Systems. I. Inorganic Nutrients. Aquatic Botany. *
4:303-316.
497
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1
I
I
I
Morgan, M.D. 1980. Grazing and Predation of the Grass Shrimp
Palaeomonetes pugio. Limnol. Oceanogr. 25:896-902.
Mulligan, H.F., A. Baranowski, and R. Johnson. 1976. Nitrogen and
Phosphorus Fertilization of Aquatic Vascular Plants and Algal in
Replicated Ponds. I. Initial Response to Fertilization.
Hydrobiologia. 48:109-116.
Munro, R.E., and M.C. Perry. 1981. Distribution and Abundance of
Waterfowl and Submerged Aquatic Vegetation in Chesapeake Bay. U.S.
Fish and Wildlife Service, Migratory Bird and Habitat Research
Laboratory, Laurel, MD.
Nelson, W.G. 1979. Experimental Studies of Selective Predation on
Amphipods: Consequences for Amphipod Distribution and Abundance. J.
Exp. Marine Biol. Ecol. 38:225-45.
Nichols, D.S., and D.R. Keeney. 1976. Nitrogen Nutrition of Myriophyllum
spicatum: Uptake and Translocation of 15N by Shoots and Roots.
Freshwater Biol. 6:145-154.
Nixon, S.W. 1980. Between Coastal Marshes and Coastal Waters: A Review
of Twenty Years of Speculation and Research on the Role of Salt Marshes
in Estuarine Productivity and Water Chemistry. In: Estuarine and
Wetland Processes. P. Hamilton and K.B. MacDonald, eds. Plenum Publ.
Co., NY. pp. 437-525.
Odum, E.P. 1971. Fundamentals of Ecology. W.B. Saunders Co.,
Philadelphia, PA. 574 pp.
Ogden, J.C. 1976. Some Aspects of Herbivore-Plant Relationships on
Caribbean Reefs and Seagrass Beds. Aquat. Bot. 2:103-116.
Orth, R.J. 1977. The Importance of Sediment Stability in Seagrass
Communities. In: Ecology of Marine Benthos. B.C. Coull ed. Univ. of
South Carolina Press. pp. 281-300.
Orth, R.J. 1981. Resident Consumer. In: Functional Ecology of Submerged
Aquatic Vegetation in the Lower Chesapeake Bay. R. L. Wetzel, ed.
Final Report USEPA Chesapeake Bay Program, Annapolis, Md.
Orth, R.J., K.A. Moore, and H.H. Gordon. 1979. Distribution and Abundance
of Submerged Aquatic Vegetaion in the Lower Chesapeake Bay. EPA Rept.
No. 600/8-79-029/SAV1.
Oostam, B.L., and R.R. Jordan. 1972. Suspended Sediment Transport in
Delaware Bay. In: Environmental Framework of Coastal Plain
Estuaries. B.W. Nelson, ed. Geol. Soc. Am., Memoirs 133.
Ott, J., and L. Maurer. 1977. Strategies of Energy Transfer from Marine
Macrophytes to Consumer Levels: The Posidonia oceanica Example. In:
Biology of Benthic Organisms. B.F. Keegan, P.O. Ceidigh, and P.J.S.
Boaden, eds. Pergamon, NY. pp. 493-502.
498
-------
I
Oviatt, C.A., and S.W. Nixon. 1975. Sediment Resuspension and Deposition
in Narragansett Bay. Estuarine Coastal Mar. Sci. 3:201-217. •
Patriquin, D.G. 1972. The Origin of Nitrogen and Phosphorus for Growth of
the Marine Angiosperm Thalassia testudinum. M
Patriquin, D., and R. Knowles. 1972. Nitrogen Fixation in the Rhizoshpere
of Marine Angiosperms. Mar. Biol. 16:49-58.
Penfound, W.T. 1956. Primary Production of Vascular Aquatic Plants. •
Limnol. Oceanog. 1:92-101.
Penhale, P.A. 1977. Macrophyte-epiphyte Biomass and Productivity in an •
Eelgrass (Zostera marina L.) Community. J. Exp. Mar. Ecol. 26:211-224.
Penhale, P.A., and G.W. Thayer. 1980. Uptake and Transfer of Carbon and •
Phosphorus by Eelgrass (Zostera marina L.) and Its Epiphytes. J. Exp. ™
Mar. Biol. Ecol. 42:113-123.
I
Perry, M.C., R. Andrews, and P.P. Beaman. 1976. Distribution and
Abundance of Canvasbacks in Chesapeake Bay in Relation to Food
Organisms. Presentation, Atlantic Estuarine Research Society, Cape «
May, NJ. 11 pp. I
Peterson, C.H. 1979. Predation, Competive Exclusion, and Diversity in the
Soft-Sediment Benthic Communities of Estuaries and Lagoons. In: •
Ecological Processes in Coastal and Marine Systems. R.J. Livingston, V
eds. Plenum, NY. pp. 233-264.
Phillips, G.L., D. Eminson, and B. Moss. 1978. A Mechanism to Account for •
Macrophyte Decline in Progressively Eutrophicated Freshwater. Aquatic
Botany. 4:103-126. —
Phillips, R.C. 1974. Temperate Grass Flats. In: Coastal Ecological ™
Systems of the United States. H.T. Odum, B.J. Copeland, and E.A.
McMahan, eds. Conservation Foundation, Washington, DC. H
Rawls, C.K. 1965. Field Tests of Herbicide Toxicity to Certain Estuarine
Animals. Ches. Sci. 6(3):150-161. M
Rawls, C.K. (In prep.) Food Habits of Waterfowl in the Upper Chesapeake
Bay, Maryland. Univ. of Maryland Center for Environmental and
Estuarine Studies. 140 pp. •
Rhodes, D.C., and O.K. Young. 1970. The Influence of Deposit-Feeding
Organisms on Sediment Stability and Community Trophic Structure. J. of •
Mar. Res. 28:150-178. |
Roberts, W.P., and J.W. Pierce. 1976. Deposition in Upper Patuxent •
Estuary, Maryland, 1968-1969. Est. Coast. Mar. Sci. 4:267-280. •
Sand-Jensen, K. 1977. Effect of Epiphytes on Eelgrass Photosynthesis.
Aquatic Botany. 3:55-63. V
499
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Schubel, J.R., and D.J. Hirschberg. 1977. Pb210-Determined Sedimentation
Rate, and Accumulation of Metals in Sediments at a Station in
Chesapeake Bay. Ches. Sci. 18(4):379-382.
Schulthorpe, C.D. 1967. The Biology of Aquatic Vascular Plants. Edward
Arnold Ltd., London. 610 pp.
Scoffin, T.P. 1970. The Trapping and Binding of Subtidal Carbonate
Sediments by Marine Vegetation in Bimini Lagoon, Bahamas. Jour.
Sedimentary Petrology. 40:249-273.
Smullen J., J.L. Taft, and J. Macknis. 1982. Nutrient and Sediment Loads
to the Tidal Chesapeake Bay System. In: Chesapeake Bay Program
Technical Studies: A Synthesis, (in Print).
Staver, K.W., W.R. Boynton, and W.M. Kemp. 1981. The Epifauna Communities
Associated with Two Different Submerged Grass Beds in the Euryhaline
Region of the Chesapeake Bay. In: Submerged Aquatic Vegetation in
Chesapeake Bay: Its Ecological Role in Bay Ecosystems and Factors
Leading to Its Decline. Kemp, J. C. Stevenson, W. R. Boynton, J. C.
Means, eds. Horn Point Environmental Laboratories, Univ. of Md.,
Cambridge, MD. pp. 401-422.
Shepard, F.P. 1963. Submarine Geology, 2nd Edition. Harper and Row, New
York.
Stephens, K., R.W. Sheldon, and T.R. Parsons. 1967. Seasonal Variations
in the Availability of Food for Benthos in a Coastal Environment.
Ecology. 48(5):852-855.
Stevenson, J.C. , and N.M. Confer. 1978. Summary of Available Information
on Chesapeake Bay Submerged Vegetation. Final Draft of Report to U.S.
Fish and Wildlife (Office of Biological Services, Washington, D.C.)
14-16-008-1255
Stewart, R.E. 1962. Waterfowl populations in the Upper Chesapeake
Region. U.S. Fish Wildl. Serv. Spec. Sci. Rept. Wildl. No. 65. 207 pp.
Stoner, A.W. 1980. Perception and Choice of Substratum by Epifaunal
Amphipods Associated with Seagrasses. Marine Ecol. 3:105-111.
Taft, J.L., W.R. Taylor, E.O. Hartwig, and R. Loftus. 1980. Seasonal
Oxygen Depletion in Chesapeake Bay. Estuaries. 3(4):242-247.
Thayer, G.W., D.W. Engel, and M.W. LaCroix. 1977. Seasonal Distribution
and Changes in the Nutritive Quality of Living, Dead and Detrital
Fractions of Zostera marina L. J. Exp. Mar. Biol. Ecol. 30:109-127.
Twilley, R., M. Meteyer, N. Kaumeyer, J. Means, W. Boynton, W. Kemp, K.
Kaumeyer, K. Staver, and A. Hermann. 1981. Nutrients, Sediments, and
Herbicides in Agricultural Runoff and the Distribution of These Water
Quality Variables in the Middle and Upper Regions of Chesapeake Bay.
500
-------
In: Submerged Aquatic Vegetation in Chesapeake Bay: Its Ecological
W.M. Kemp. 1981. Seston Budget of the Choptank River Estuary. Horn
Point Environmental Laboratories, Cambridge, MD.
501
I
in: auoinergeci Hquacit- vegetation in ^nesapease aay. its acoiogicai ^
Role in Bay Ecosystems and Factors Leading to Its Decline. W.M. Kemp, •
J. C. Stevenson, W. R. Boynton, J. C. Means, eds. Horn Point
Environmental Laboratories, Cambridge, MD. pp. 45-80.
Virnstein, R.W. 1978. Predator Caging Experiments in Soft Sediments: B
Caution Advised. In: Estuarine Interactions. M.L. Wiley, ed. pp.
261-273. •
Wanless, H.R. 1981. Fining-Upwards Sedimentary Sequences Generated in
Seagrass Beds. J. of Sedimentary Petrology. 51(2):445-454. _
Webster, T.J.M., M.A. Paranjape, and K.H. Mann. 1975. Sedimentation of ™
Organic Matter in St. Margaret's Bay, Nova Scotia. J. Fish Res. Bd.
Can. 32(8)-.1399-1407. •
Weinstein, M.P., and K.L. Heck, Jr. 1978. Ichthyofauna of Seagrass
Meadows Along the Caribbean Coast of Panama and In the Gulf of Mexico:
Composition, Structure and Community Ecology. Marine Bio. 50:97-107. I
Wetzel, R.F. 1964. A Comparative Study of the Primary Productivity of
Higher Aquatic Plants, Periphyton and Phytoplankton
Lake. Int. Rev. Gesamten. Hydrobiologia. 49:1-64.
Higher Aquatic Plants, Periphyton and Phytoplankton in a Large Shallow I
Wetzel, R.L., P.A. Penhale, and K.L. Webb. 1981. Plant Community •
Structure and Physical-Chemical Regimes at the Vancluse Shores Study M
Site. Chapter 2, Section A. Virginia Institute of Marine Sciences,
Gloucester Point, VA. —
Wilkins, E. 1981. Aspects of Waterfowl Utilization of a Mixed Bed of '
Submerged Vegetation in the Lower Chesapeake Bay. Chapter 3, Section
VI. Virginia Institute of Marine Sciences, Gloucester Point, VA. B
Yarbro, L. , P. Carlson, T. Fisher, J. Chanton, R. Grump, N. Burger, and
I
I
I
I
I
I
I
I
-------
I
I
HERBICIDES IN CHESAPEAKE BAY AND THEIR EFFECTS
_ ON SUBMERGED AQUATIC VEGETATION:
A Synthesis of Research Supported by U.S. EPA
• Chesapeake Bay Program
I
I
_
•
I
I
I
I
W. M. Kempl
J. C. Means2
T. W. Jonesl>3
J. C. Stevenson 1-
December 1981
M ^The University of Maryland Center for Environmental and
Estuarine Studies (UMCEES) , Horn Point Environmental
^ Laboratories, Cambridge, MD.
™
, Chesapeake Biological Laboratory, Solomons, MD.
fl -^Biology Dept., Salisbury State College, Salisbury, MD.
I
I
I
502
-------
CONTENTS
I
I
Page
Figures 504 •
Tables 506 *
Section
1. Introduction 507 fl
2. Rationale for Selection of Compounds Studied during CBP .... 511 V
Herbicide Chemistry and Use 511
Rationale for Selection 516 •
3. Distribution of Herbicides in the Bay 517 •
Open-Bay concentrations 517
Tributary concentrations 517
Runoff concentrations 520 •
Other runoff studies in the Bay region 522 9*
Major factors affecting runoff 522
4. Environmental Behavior of Herbicides 525 •
Sorption reactions 525 9
Herbicide degradation 528
5. Toxicity of Herbicides in the Estuary 534 ^
Toxic mechanisms 534 •
Toxicity to animals 535
Mutagenicity , 535
SAV phytotoxicity 536 V
Effects on Photosynthesis and Respiration 537 W
Effects on Population, Biomass, and
Physiomorphology 545 •
Other Factors 550 £
Acute versus chronic exposure 550
Mode of uptake 550 —
Combined stresses 551 •
Metabolites 552 ~
6. Summary and Implications ' . . 555
Summary of research findings 555 •
Did herbicides cause the SAV decline? 556 9
Are herbicides a problem? 556
Literature Cited 559 •
503
1
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
FIGURES
Number Page
1. Trends in herbicide use in the United States between 1949 and
1976 508
2. Schematic representation of the fate, transport, and effects of
herbicides in the Chesapeake Bay region 509
3. Estimated use of major herbicides in (a) Maryland and Virginia
and (b) the Choptank watershed from 1975 512
4. Concentrations of atrazine and linuron along salinity gradients
in (a,b) Chesapeake Bay and (c,d,e) the Choptank River for summer
months in 1977 and 1980 518
5. Atrazine concentrations: (a) spatial distribution along the
Rappahannock River and' (b) temporal concentrations in headwaters
of Severn River for 1979 and 1980 519
6. (a) spatial distributions of atrazine and linuron in the Choptank
River-Estuary and (b) temporal patterns of atrazine concentra-
tions in runoff from Choptank and Horn Point watersheds for spring
1981 521
7. Effects of (a) basin slope and (b) time-interval to first runoff
event on the loss of atrazine from agricultural fields, where loss
is taken as percent total applied that runs-off to watercourse . . . 523
8. (a) Freundlich adsorption isotherms for atrazine on estuarine
colloids, sediments and agricultural soils. (b) Effects of
salinity on adsorption coefficients (KQC) for estuarine sediments
and colloids 527
9. Approximate persistence in soil (i.e., time until >90 percent of
initial application disappears from site) for nine herbicides used
in Chesapeake Bay region 529
10. Loss of l^C-ring labeled atrazine from experimental systems, and
percent of total residuals as parent compound and as two major
metabolites 531
11. Atrazine mass-balance after one, three, and 30 days in experimental
microcosms containing estuarine water and sediments along with the
submerged macrophyte, Potamogeton perfoliatus 533
12. Typical results showing apparent photosynthesis over time for
experimental microcosms containing Potamogeton perfoliatus treated
with atrazine (0-1 ppm) 538
13. Ratio of apparent photosynthesis to night respiration for
Potamogeton perfoliatus Created with two levels of atrazine (and
control) 540
14. Typical patterns of diel 02 under in s itu domes covering Zostera
marina communities tricated with atrazine and shading in Guinea
Marsh, VA 541
15. Regression of "loss in apparent photosynthesis" versus herbicide
(atrazine and linuron) concentrations for three species of
submerged estuarine macrophytes 546
16. Summary of measurements of plant biomass in duplicate microcosms
containing (a) Potamogeton perfoliatus and (b) Myriophyllum
spicatum treated with linuron (0-1 ppm) 547
17. Effects of atrazine on mortality and average height of Zostera
marina in estuarine microcosms after 27-day exposure 549
504
-------
I
FIGURES (Cont'd) V
18. Effects of plant vigor (as indicated by peak experimental values
of apparent photosynthesis) on the response of Potamogeton
perfoliatus at 25 ppb atrazine 553
19. Correlation between an index of potential diffuse loadings
(watershed area/estuarine volume) and percent occurrence of
I
submerged macrophytes at randomly chosen stations visited in 1974 . 557 •
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
505
-------
I
•
TABLES
Number Page
• 1. Chemical Properties of Major Herbicides in the Chesapeake Bay
Region ............................... 513
^ 2. Uses of Major Herbicides in the Chesapeake Bay Region ....... 515
• 3. Summary of Atrazine Degradation Rates in Agricultural and
* Estuarine Environments ..................... 532
4. Mutagenicity of Major Herbicides in the Chesapeake Bay Region . . . 536
15. Summary of Selected Structural Characteristics of Potamogeton
perfoliatus Populations in Microcosm Communities Treated with the
Herbicide, Atrazine (Cunningham et al. 1981a) ........... 548
I
I
I
I
I
I
I
I
I
I
I
I
I
• 506
-------
SECTION 1
INTRODUCTION
I
I
I
The widespread use of herbicides for weed control in the last several
decades has contributed substantially to expanding agricultural production ^
in North America. Annual applications of herbicides in the United States •
currently amount to some 175,000 metric tons active ingredients. The ™
dramatic increase in herbicide use since 1950 has followed a general
pattern of exponential growth as seen in Figure 1. Also depicted in this I
figure is the ever increasing importance of the ^-triazine herbicides (and J|
in particular atrazine) between 1960 and 1975.
Inevitably, a fraction of the herbicides applied to agricultural fields «
is transported to nearby watercourses by runoff and subsurface interflow. I
Significant concentrations of these compounds have been observed in
streams, lakes, and estuaries throughout North America (Richard et al.
1975, Truhlar and Reed 1976, Newby et al. 1978, Frank and Sirons 1979, •
Hermann et al. 1979). Since many of these compounds are also registered 9
for aquatic weed control (individually or as part of a formulation), there
appears to be considerable potential for inadvertent damage to non-target M
plant species in the hydrosphere. •
Submerged aquatic vegetation (SAV) in Chesapeake Bay has undergone a
marked decline throughout the estuary since the mid-1960s (Stevenson and _
Confer 1978). Both the piedmont, and coastal-plain portions of the Bay's •
watershed are actively farmed, and herbicide use in this region has ™
generally followed trends in the rest of the United States. The general
coincidence in timing of events (that is, introduction of ^-triazines I
versus the initial decline in SAV) led to a serious concern among •
scientists, resource managers, and other citizens of the region as to the
potential role that these herbicides may have played in the loss of SAV in «
Chesapeake Bay. •
The U.S. Environmental Protection Agency's (EPA) Chesapeake Bay Program
(GBP) established SAV as one of three major themes of a multi-year research
effort. Causes of the SAV decline, with considerable emphasis placed on •
investigating the interactions between herbicides and SAV in the estuary, •
were among the issues addressed in this program. Numerous aspects of
herbicide fate, transport, and effects were examined in this research. The
interrelationships among various processes and the potential linkage
between herbicide application and effects on SAV are depicted in Figure 2.
Some of the herbicides placed on agricultural fields percolate into _
subsurface waters where they reach a sorption equilibrium with soil I
particles and are taken up by weeds. The herbicide compound usually kills
the weeds. A portion of the compound degrades to various metabolites, and
a portion enters the estuary through runoff, leaching, and streamflow. •
Some of the herbicide may be volatilized and/or transported with dust, B
thereupon entering the estuary through fallout. The herbicide may then be
taken up by SAV, causing them phytotoxic stress. In the estuary, the •
herbicide partitions to sediment and water in response to the physical M
factors of salinity, pH, and temperature as well as to the specific
chemistry of the sediment and herbicide. Again, some of the herbicide is _
lost to degradation. I
I
507
I
I
-------
NONTRIAZINE
HERBICIDES
1950
1955
I960
1965
YEAR
Figure 1. Herbicide use in the United States. (Data are
from Eichers et al. 1978 as adapted in Stevenson
et al. 1981.)
508
-------
d
o
•H
oo
0)
CO
pa
cd
0)
a,
nj
en
cu
•H
U5
CD
T3
•H
O
•H
J3
V)
OJ
CM
0)
509
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
In the following pages, the results of herbicide-related research from
the CBP are synthesized. First the nature of these herbicides and the
rationale for selection of two compounds for intensive study is discussed,
then these results are examined in the context of the conceptual framework
of Figure 2 and in relation to pertinent research done elsewhere. Finally,
the overall implications of these research findings are evaluated in terms
of the role of herbicides in the SAV decline.
510
-------
I
I
I
SECTION 2
RATIONALE FOR SELECTION OF COMPOUNDS STUDIED IN THE CHESAPEAKE BAY PROGRAM
A wide assortment of herbicides is used within the Chesapeake Bay
watershed, and it would be impossible to study all of them in detail. It
was decided that initial research should, therefore, focus on the two major I
compounds. All pertinent criteria, in terms of potential impact on SAV, ™
were considered, and atrazine and linuron were chosen. In this section we
discuss the general chemistry of the important herbicides in the Chesapeake •
Bay region, as well as patterns of their use. We then present the ||
rationale for selecting these two particular compounds for intensive study.
HERBICIDE CHEMISTRY AND USE •
There are over 140 herbicidal compounds listed in the current Herbicide
Handbook (Weed Science Society of America, 1980), which probably represents I
the majority of those weed-control substances registered with EPA. Eight W
compounds from six chemical groups were chosen for discussion here, based
on amounts of each used in the Bay region. The annual use-rates for major •
herbicides in 1975 are summarized in Figure 3(a) for Maryland and Virginia, £
and Figure 3(b) for the Choptank River watershed. Clearly, the four most
heavily used compounds are atrazine, alachlor, linuron, and simazine. .
Application rates are also shown for six additional compounds, of which •
four have been chosen for further discussion. *
Many of the important herbicides are produced by chlorination of
aromatic compounds, including 2,4-D and dicamba; other compounds include B
chlorinated aliphatic acids, heterocyclic derivatives, and organometals •
(Mrak 1974). In Table 1, some chemical properties of herbicides, grouped
in terms of their ionic and acidic nature, are summarized. Water •
solubility, molecular weight, and vapor pressure are presented. In •
addition, octanol-water partition coefficients (KQW) are listed to
provide a relative index of the compound's hydrophobicity. The Kow is a
highly correlated with the ability of an herbicide to bioaccumulate, or be •
biologically incorporated across a membrane lipid bilayer. The uses of
these compounds depend largely on their chemical characteristics. Several
key aspects of herbicide use in the Chesapeake Bay region are provided in •
Table 2. Various information is compiled here, including the year that the M
herbicide was introduced for public use, the main crops (in Maryland and
Virginia) with which it is used, and the associated planting and tillage •
practices, as well as the timing and rate of application. •
The cationic and acidic herbicides are generally more water soluble
than the others (Table 2). Paraquat, as a salt, is highly soluble in _
water, but virtually insoluble in organic solvents; 2,4-D is more generally •
soluble. The £-triazine compounds (atrazine and simazine) are among the *
least water soluble with moderate organic solubility; trifluralin dissolves
readily in octanol, but not so readily in water. Compounds with high vapor •
pressure, such as dicamba, are more likely to volatilize under wet £
conditions and enter the hydrosphere with precipitation; the ^-triazines,
with their low vapor pressure, are less likely to follow that route of _
transport. •
511
I
I
-------
1
1
1
m 150
I P
E
"•*
IUI
%
ui 100
|Q
S£
CD
OC
UI
1 »
N 50
1
1
1
1 S
— - 4A
™TW
UI
1 I
9 30
Ia
m
cr
LU
• X2°
1 "
• on
1
1
r (a)
-
-
-
™*
.
^
/
'
/
'/
/
/
/
/
/
/
/
UI
IM
OC
h-
p*
•:•
•:•
•:•
:|:
E3 MARYLAND
7] E3VIRGINIA
/
/ z
^s s § « i i «
/= — o uj^noo
/5NtE wojs
<2Z °a-<
/-
'<
y.
i
y/.
y/ z
// 1
/ / — j
/ / UI 1— UI <
'Z, z < IT ^ o:
'O' - u < Q 3
(E^ N or -J , _i
'-j < rf > ' u.
/»' 3E S *- * —
'=., - < => -IT
'_J OT d. GO CM H-
'/,
y,v,
y/ '//
2345678
RANK
1
Figure 3. Estimated herbicide use in the Chesapeake Bay region for
11975: (a) Maryland and Virginia and (b) Choptank watershed
(Stevenson and Confer 1978).
• 512
-------
00
s g
o> 1-1
O -U
co vi
cfl
PU
cj
(-1
CO 4J
•H 42
3 oo
a TH
at a)
r-l 13
co
01 /->
M U OOO
o 3 sci
CX CO O
tO to g rH
> 0) E
t-i -H /-N
0) rH 6
•u -H a,
n3 42 CX
13 3 ^
rH
O
CO
ts
U 0)
•H 6
6 co
0) 3
U
I
CJ
CO
efl
s
z
o
CO
CO
3
CJ
co
vO
oo •
m
CM
§
§
C
o
H
I
CJ
CM
CM
oo
oo
en
CM
en
•
o
CM
o
CM
vO vO
o cn
o o
o d
i
CM
CM
CM
CM
«a-
•
o
O
r~~
u-|
en
en
cn
u-i
o
m
o
o
hy 1-4,4-
ium-dichloride
4-1 d
0) -H
S T3
13 r-l
rH CX
« iH
rH 43
'o
5 ,
S co
CO 1
li'o'
4: c
W 1-1
oi g
^ co
1 rH
1 CX
O O CD
to to G
O CX i-t
rH O N
45 CO CO
CJ -H i-l
1 Sv/ M
CM vO 4-1
4-bis)ethyl-
triazine
i i
o co
to 1
O r*
rH O
43 C
a -H
| g3
CM CO
lorophenoxy)
id
oro-0-
id
X U rH CJ
CJ CO 43 Cfl
•H 0
T3 CJ i-l CJ
1 -H T3 -H
-a- w i co
« 0) SO -H
CM a « c
^ a en co
i ^
ra oo
CO 00
rQ 1-1
•H
O
IT)
en
en
o •*
• CM
O I
I
•H I
"O P.
I I
CM (X
I O
o to
to CX
3 -H
^H ^3
£ z o)
M •> d
4J Z 1-1
I I T3
cfl O -H
I to 3
CO 4J rH
I -H O
CO C 4J
CJ
1
JJ
g
o
o
CJ
§
01
g
cfl
Z
£
0
g
0
CJ
,*~\
01
s
cfl
01
TJ
cfl
J«j
H
**~s
CO
w
Q
rH
CJ
M
CQ
Pi
W
sc
0
IH
z
o
rH
4-1
CO
a
-H
c
o
•H
4->
cfl
O
•
3
CT
cfl
to
cfl
PH
r-i
4-1
Cfl
3
a-
Cfl
}-i
cfl 01 *-*•
CX C X
O -HO)
43 N r-l
4J Cfl 4J
to O
0 -H
••— • CO
CO
PQ
t
rH
M
r-l CO
<3 v— '
0)
c
•H
N
cfl
* Q
•H
CO
s~*
CX
0)
CJ
a
•H
(-1
a,
VwX
CO
43
o e
0
•H
•a
•H
CJ
<£
•
rH
M
M
1 CO
^3" CJ
« -H
CM ca
co
O
M
CJ
EC
O
§
rH
rL
co "e
r-l CO
3 rH
rH IW
14-1 01
•H to
H£
513
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
x-s
TJ
01
3
C
•H
•U
C
O
u
\-s
•
rH
W
hJ
59
2
H
T)
00
C
•H
•u d
c o
n ) * »>
m *ri ,>
> -U O
rH -H txj
O -U
CO J-l
CO
PM
U
1 .
tfl 4J
rH 43
3 00
O -H
01 0)
1 f^
o 3
a
CO
01
V-l rl O£XD
0 3 331
cx co o
CO CO g rH
> oi i
H ^ X
PH
U
>%
4J
\-t -rl s-*
01 rH B
•u -H ex
cfl 42 ex
S 3 ^
^j
^^
0
CO
43
i-H
Cfl
U 01
•H e
B cfl
01 &
CJ
^"
1 CO
-d-
CT> O
-a- r-
CM CM
m CM
rH CM
m oo CM
r» sr «*•
rH CM
1
I
I cd 5S
rH 0) 1
>-. rl rH
d 3 5s
01 rH 43
43 >, 4J --^
CX 43 01 rH
O JJ 1-t fS
rl 0) t) 43
OS 1 W
rH | VO 01
43 rH « S Q)
01 CM J^T3
•H ts 1 X i-t
T3 X 0 O rH
1 O rl rH -H
-3- 43 O f^ C
•> 4J rH 43 CO
en 0) 43 W 4J
N-' B o 01 o)
II I B u
en rH CM "-^ CO
*o
CJ
CO rl
14H 01
3 M
d 3
cfl 4J
a
o
4-1
w d
d co
o co
cx d
3 0
Q a
co
01
S3
O
M
H
<
U
M <-N
fe Q) 0)
M B S
CO Cfl Cfl
co s S3
<
iJ d 01
U O T3
i co
nJ 1 rl
. *J o -~*
rl 0 X 3 rH O
3 )H O U 43 CO
rH 3 rl -H O CO
>, (3 O U cfl tfl
d -H (J 00 rH ,J
01 rJ ^^ & *$ ^*>
43 3
P-K CO
•
• M
> >
y— N
00
f^^
ON
rH
)-l
0)
MH x^.
d rH
O 00
<— > U CTi
•* rH
P* T3
ON d •
rH Cfl rH
cd
d
rH O 4J
Cd CO 01
d
4J 01 O
0) > CO
01 Oi
4»! u s^-
cd • co
M ?
a • o
O S4
oo
O CTi •
OO rH 4J
O\ r»
rH <3 0)
CO iH
< CO O
CO S -H
CO v^ iw
13 ^-N >4H
^^ OO U 01
P^ O O
CJ ON p* O
O rH CM
m ^ i d
CM en o
1 0) CM -H
O CX 4-1
CM O -U -H
43 cfl 4J
4-> O M
cd 3 >> CO
cfl 4J CX
0) 3 -H
tH rH M
3 •> -H 0)
CO ^-v _Q 4J
CO O 3 cfl
0) 00 rH S
!-i ON O 1
CX rH CO rH
v-/ O
rl !H d
O S S 0
cfl 43 O T3
514
-------
O
M
O
u
PC
W
w
en
w
CO
w
Q
M
O
M
w
PC
pi
o
en
w
CO
;=>
CM
W
PQ
H
O
en
CJ
o
M
§
33
O
U-l
o
ft
C 'cfl
0 JS
U-< -H ^.
O 4J •
cfl -H
01 0 •
4-> -H CO
Pi 0. 60
0. US
< >-'
T3
A
CJ
a
u-i O
O -H
4-1
00 CO
c a
•H -H
e rH
•H d.
H a
^0
•
O
1
CO
o
4_|
c
cfl
rH
Cu
1
01
PU
01
•a
o
01
60
cfl
rH
rH
•H
H
2
«t
Q
-a
a
jz
•n
01 CO
4J £L
Cfl O
•H M
0 O
O
cn
CO
O
CO
M
C
o
u
cfl
13
Ol
o
M 3
CO T3
01 O
>H 1-1
4J
c
M
\D
^O
Ol
rH
0)
e
cfl
2
c
o
£
g
o
CJ
CO
U 4J
Q CO
M O 3
CJ -H CT
M C CO
pq O l-i CJ
K -H cfl -H
r^J *^ f\ | C/}
BJ CO cfl
CJ pq
O
M
2
O • M
MM M
CM -3-
• •
CM rH
1 1
r~- CM
rH CM
-,
,£3 .C 03 fi 0
33 CO CO
B CO U
. 3 « 01
c 60 a & • c
J-l r-l M >> >% 1-1
0 0 O O 0 O
O en u co cn cj
00 CM CO CM
in v^ ^o vo I
ON ON ON ON
rH rH rH rH
CO CO
01 01
C TJ
•H -H
rH rH
CO -H -H
we c
Q < C ,C 4Jcfl
•H OCOMC-H COrH
CM P MJ3HO)|J -Q • M
2 M > >
•
s-x
rH
CO
rH
rH
CO 60
c
4J -H
oi cu
a.
CO O
C M
01 CJ
> 1
01 0)
4-1 rH
cn ,a -
•a 3
o
-T3
in II
rx.
CN Q
rH
iH
rH rH
CO lH
4-1
4-> 1
Oi B
V^ £ oO
Oi -H r-»
K* C CT\
CO -H rH
01 g
3 r-l
U M 01
O M-t
C
/-s rH O
rH rH O
r» -H
ON 4J T3
rH 1 C
•— ' O Cfl
C
rH C
CO . II 0
CO
4J 2 C
Ol 0)
E 01 01
CO OO 4-1
jr co co
00 rH ^-"
C 4-* Q)
rH C 4J
•~-s 01 CO
> cfl
rH C W
Cfl 0
'J rJ
4J O
0) II U-l
4J O CO
(-1 0)
Cfl 4J
g •• Cfl
01 CO M
4-> C
co o -a
cfl -H 01
4J T3
cfl C
.. -H 01
co > i
o) oi B
O r-l O
M rf U
3 2 01
O
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Paraquat is highly sorbed, tending to adhere essentially irreversibly
to surfaces of soil particles. Hence, it is used as a "contact" herbicide,
sprayed directly on the weed foliage. It is used at low application rates
before planting of the crop, particularly with no-till fanning and double
cropping (Table 2). The postemergent herbicides, such as 2,4-D and
dicamba, must >»~ highly specific to broad-leaf weeds, and are, thus, used
primarily with corn and small grains at low application rates. The
^-triazines are also effective in control of broad-leaf weeds and are very
versatile, being applied at relatively high rates both pre- and post-
emergence of corn, under either conventional or no-till conditions.
Linuron and alachlor are also versatile compounds with a wide range of
uses, although linuron is associated most closely with soybeans.
RATIONALE FOR SELECTION
At the outset of this research program in the spring of 1978, six
criteria were used for selecting the two herbicidal compounds for focus
during the CBP. These criteria are related to the fate-and
effects-pathways described in Figure 2. Starting with total application to
agricultural lands in the Bay region, we considered how long the herbicide
persists on the field (that is, available for runoff to the estuary). The
solubility and actual percentage of each compound transported into
surrounding waterways suggest something about its relative mobility. In
1978 there was a distinct paucity of information concerning either the
actual concentrations of these compounds occurring in the Bay, or their
toxicity to SAV, but we used what scant data were available. Necessarily,
the weighting of these factors was relatively subjective, representing our
perception of importance and reliability of information. A ranking among
the six most important herbicides led to the following: atrazine,
alachlor, linuron, paraquat, trifluralin, and 2,4-D. It might be noted
that though its current use is substantially reduced, 2,4-D was included
here because it was one of the most common compounds used in the 1960's.
Simazine was not considered in this ranking because of its close similarity
to atrazine.
By these criteria, atrazine ranked clearly as the major compound with
trifluralin falling to the bottom of the list. The relative importance
among the other four herbicides was virtually indistinguishable, and each
probably deserves further scrutiny in its own right. Nevertheless, we
selected linuron as the other substance for CBP focus, primarily because of
its relative longevity reported for agricultural soils, and because it is
associated so closely with soybean production. Over recent years, corn and
soybeans have become the most important crops in the region, and the two
selected herbicides, atrazine and linuron, are, respectively, most
significant for those two crops.
516
-------
SECTION 3
DISTRIBUTION OF HERBICIDES IN THE BAY
I
I
I
This section summarizes observed concentrations of atrazine and linuron
in Chesapeake Bay and its tributaries, and attempts to relate these •
concentrations to runoff rates. We will initially examine concentrations |
along the axis of the main Bay and then move into successively higher-order
tributaries toward the source-waters that drain agricultural lands. «
OPEN-BAY CONCENTRATIONS •
The maximum concentrations of atrazine and linuron reported in either B
the open waters of the main-stem Bay or a first-order tributary, such as B
the Choptank and Patuxent Estuaries, between 1976 and 1980 were about 3.5
ppb (surface water). In Figures 4a and 4b, we present data for the main •
Bay for June and July of 1977 and 1980 (from Austin et al. 1978, Newby et I
al. 1978, Means et al. 1981b). Concentrations of atrazine and linuron
never exceeded about 1.3 ppb, and were generally highest at lower
salinities. Patterns of concentration-versus-salinity exhibited I
nonconservative behaviors, probably reflecting either non-steady-state B
input conditions or significant sources other than the Susquehanna River
(Stevenson et al. 1981). General trends for the two years were quite •
similar. |
TRIBUTARY CONCENTRATIONS —
Herbicides were also monitored in two major estuarine tributaries of
the Bay. Mixing diagrams of herbicide concentration-versus-salinity are
also provided for 1980 data from the Choptank River (Figure 4c, 4d, 4e). B
The absence of a relationship in the June data was probably owed to the B
meager runoff that occurred in late May through June of that year, and the
small runoff experienced during July generated a weak relationship for that •
month. Linuron concentrations were relatively high at the head of the B
estuary as well as at about 13 ppt salinity, suggesting runoff sources both
up-river and down-estuary. Linuron concentrations in June and July were _
virtually undetectable, and the higher August values correspond to the July B
planting of double-cropped soybeans. Zahnow and Riggleman (1980) reported *
no detectable aqueous concentrations of linuron in the Choptank and other
tributaries in 1977 to 1978, although some herbicide was found in up-river B
sediments. Atrazine was measured at numerous stations throughout B
Virginia's Bay waters, and two longitudinal profiles along the Rappahannock
River are presented in Figure 5a for June and August of 1979. Highest •
values were 3.5 ppb in the freshwater reaches; estuarine concentrations fi
never exceeded 1.0 ppb (Hershner et al. 1981).
Samples were obtained for analysis of estuarine sediments and suspended
particulate matter at most stations in the Bay and tributary surveys. B
Atrazine was detected periodically in estuarine sediments at low B
concentrations (about 5.0 ppb) in Maryland waters (Means et al. 1981b) .
Similarly, sediment concentrations were rarely detectable in Virginia,
although one value in excess of 30 ppb was reported for a sample from the
517
I
I
I
-------
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
•1
1
1
*
E~
0.
a.
z
0
cc
z
Lul
O
O
o
IU
0
o
cr
LJ
X
Figure 4
1.5-
1.0-
0.5-
1.5-
I.On
0.5 -^
(a) CHESAPEAKE BAY
, JUNE 1977 8 1980
yv^W
*\ ........ __
^•••••^-^^^
^
i
Ab) CHESAPEAKE BAY
" JULY 1977 a 1980
M
"Ty\
^A.
b — N_n_
o-i •= 1
i.o-
0.5
1.0-
0.5-
1.5-
i.o-
0.5-
(C) CHOPTANK RIVER
JUNE 1980
0 • 0 ATRAZINE (I960)
•• — -«UNURON (1980)
1 a 1
— Q.
~-^^ ^
(d) CHOPTANK RIVER
CL JULY 1980
^\^^
(e) CHOPTANK RIVER
^AUGUST 1980
""•"^^
X
o^^.
^ — ^_
0 5
SALINI
Concentrations of atrazine and
June and July of 1977 and 1980
June through August of 1980.
i 1
^
^ fr^^Sn !
10 15
TY (ppt)
linuron in Chesapeake Bay for
and in the Chop tank River for
(Data for 1980 are from Means
et al. 1981b; and for 1977 from Austin 1978, and Newby et al
1978.)
518
-------
(a)
$
.d"
<
tr
UJ
o
z
o
o
2-
8 JUNE 1979
0-
10 AUGUST 1979
I
60
T I I 1
40 20
DISTANCE FROM MOUTH (km)
P '-"-v
(b)
u
c
o-1
15-
IL.
RAINFALL
-
a.
a.
<
a:
UJ
o
o
10-
5-
o-
Figure 5.
ATRAZINE
APRIL
MAY
980
Concentration of atrazine (a) in the Rappahannock River,
Virginia; and (b) in runoff from the Severn River, Virginia
(Hershner et al. 1981). Inches of rainfall for April and
May are also shown in this figure.
519
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
head of the tidal creek portion of Severn River (Figure 5b) (Hershner et
al. 1981). Atrazine concentrations were never detected in suspended
estuarine solids sampled in the field during 1980 (Means et al. 1981b).
It seems surprising, at first, that herbicide concentrations in the
Choptank were no greater than those observed in the main Bay during 1980
surveys, in view of the far shorter transit time between field and estuary
in the tributary. The spring of 1980, however, was a period of
extraordinarily low runoff, particularly in the eastern shore region; total
rainfall for May and June was 8.8 cm (3.5 in). Precipitation in the
Choptank watershed for May and June of 1981 was 26.2 cm (10.3 in), much
greater than the previous year and even slightly greater than normal, with
a 20.4 cm (8 in) average for 1971-1980. Thus, 1981 represents a year when
relatively high herbicide concentrations would be expected in the estuary.
RUNOFF CONCENTRATIONS
Means et al. (1981b) monitored atrazine and linuron concentrations
during base flow and after all runoff events in spring and summer 1980-1981
at the creek and small embayment draining a 94 ha (232 acres) experimental
watershed at Horn Point Environmental Laboratories (HPEL). These data (for
1981) are summarized in Figure 6a. Herbicide concentrations were also
measured in the Choptank River headwaters and estuary after a major storm
in mid-May 1981 (Figure 6b) . Concentrations of atrazine in the river
reached 9.0 ppb and exceeded 2.0 ppb well into the estuary. Linuron
concentrations of 2.0 to 3.0 ppb were found in both fresh and brackish
waters, with no apparent relation to salinity. Such high values of linuron
were unexpected, since this event preceded soybean planting, and they
probably represent localized runoff from treated fields of small grains.
Atrazine concentrations in 1981 runoff from the HPEL watershed (draining
primarily corn fields) reached peak levels of about 20, 45, 10, and 13 ppb
during the four spring runoff events described in Figure 6a.
Concentrations at the drainage creek during the same period in 1980
exceeded 3.0 ppb for only one short event (May 1), when peak values were
18.3 ppb. The flow in the Choptank headwaters at Beaver Dam exhibited a
marked maximum (9.0 ppb) only during the first two closely spaced events in
1981. Concentrations as high as 20 ppb were observed once in the small
estuarine embayment (Lakes Cove) receiving direct runoff from the HPEL
watershed.
Rainfall during the spring of 1980 was considerably greater on the
western shore of Virginia, where almost 10 cm (3.9 in) of rain fell during
the eight-day period April 25 to May 2. Hershner et al. (1981) monitored
for atrazine in the headwaters of the Severn River (draining extensive
agricultural land), and reported maximum concentrations of about 16 ppb
during the runoff generated by two successive downpours of 3.0 cm (1.2
in). These data are provided in Figure 5b. Somewhat higher concentrations
were reported by Hershner et al. (1981) for 1979, with four values above 10
ppb measured at a small tidal creek in the upper Severn River during an
April runoff event. One extreme value (108 ppb) was observed during this
episode in a drainage creek. The wet spring of 1981 was not studied in
Virginia, but it appears that general spatial and temporal distributions of
herbicides are similar in upper and lower Bay regions, both being highly
responsive to hydrologic conditions.
520
-------
.a
a.
•--• 20-
z
o
UJ
o
LU
Z
N
CC
H-
.0-
10-
(a)
INITIAL
HERBICIDE
APPLICATION
Watershed
HPEL
-RUNOFF FROM
AGRICULTURAL
HELD HPEL
FLOW OVER
BEAVER
\ DAM
MAY
JUNE
*. IO~
.a"
&
z ft _
O
H
'
u IS
1
[^ ATRAZINE
fH LI NURON
^
v, n
I ^
$» t rvr-n *
75 65 60 v 35 25 20
DISTANCE FROM MOUTH (km)
10
Figure 6.
(a) Temporal patterns of atrazine concentrations in runoff
from Choptank watershed (spring), and (b) spatial distribution
of herbicides in Choptank River and estuary (May 10-13, 1981)
(Means et al. 1981b).
521
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
OTHER RUNOFF STUDIES IN BAY REGIONS
Herbicide concentrations in the Rhode River on Maryland's western shore
have been intensively studied. Correll et al. (1978), Wu et al. (1977),
and Wu (1980) reported 1976 concentrations and runoff rates for atrazine
and alachlor in the Rhode River basin and estuary. In general, their
results are consistent with those of Means et al. (I981b) and Hershner et
al. (1981), where peak runoff concentrations of dissolved herbicide were
about 35 ppb and 3.0 ppb for atrazine and alachlor, respectively. They
also reported concentrations of herbicides sorbed to suspended solids,
which were periodically on par with the dissolved form but were in excess
of values observed by Means et al. (1981b). Herbicides found in Rhode
Estuary never exceeded 1.0 ppb (dissolved) for either compound. Atrazine
has also been measured by the U.S. Geological Survey in the Susquehanna
River (at Harrisburg and Conowingo) and several small tributary creeks for
1978-1980 (Ward 1980 quoted in Stevenson et al. 1981). Concentrations were
generally in the range of 1.0 to 5.0 ppb, though one exceptionally high
value (68 ppb) was found at Goods Run in May 1980.
Wu (1980) estimated that about one percent of the atrazine and 0.2
percent of the alachlor applied to agricultural fields in the Rhode River
basin entered the watercourse. These runoff rates are within the range,
but on the low side, of values reported in the literature (Wauchope 1978).
Of almost 50 estimates of atrazine runoff compiled by Schueler (1979) from
various North American fields, we calculate a mean of 2.6 percent from a
range of 0 to 17 percent. Much less information is available for alachlor
and linuron; however, reported values range from 0.02 to 14 percent and
appear to be near (slightly less ) atrazine values. Data from the HPEL
flume have not yet been analyzed in terms of percent losses, but these
forthcoming values may add some insights to this issue.
MAJOR FACTORS AFFECTING RUNOFF
Numerous factors influence the rate and concentration of herbicide
runoff and .should be considered when interpreting results from the
Chesapeake Bay region. Among these factors are: chemical nature of the
compound; slope of the land; rainfall intensity, duration, and timing; soil
type; plant cover; and drainage density. Slope and precipitation are
particularly important factors that can profoundly influence runoff. We
have plotted overall percent loss of atrazine applied to the field, versus
the topographic slope of the field, for six different sites in the Eastern
and Central United States in Figure 7a. There is considerable scatter in
these data, because variables other than slope are also operative.
Nonetheless, there appears to be a positive relationship that follows a
hyperbolic, or logistic shape, with greatest effects found in the region of
five to 10 percent slope. Although comparison of data from different
watersheds must be viewed with caution, one might infer that runoff data
from the coastal-plain portions of the eastern and western shores of the
Bay (generally less than about six percent slope) may be comparable to one
another.
Another important consideration is the time interval between
application of herbicide to the field and a given rainfall-runoff event.
Herbicide concentrations are highest in the first runoff and generally
522
-------
T3
CD
"a.
Q.
O
-I
Ld
O
OC.
CO
CO
o
"O
CD
o.
Q.
UJ
H.
O
(XL
CO
CO
Q
8
6-
4 -
2-
(Q)
4°.
I WHITE etal. 1967
2 HALL 1974
3 RITTER etal. 1974
4 SMITH etal. 1974
5 WU etal. 1977
6 LONGDALE et al. 1978
0
I
5
SLOPE
10
15
OF LAND
(b)
8-
6-
4-
2-
.ATRAZINE CONCENTRATION
(Smith etal. 1978, quoted in
Wauchope 3 Leonard 1980)
\
\
\o2
\
\
\
\
\
\
\
3o
ATRAZINE
LEAKAGE
I WHITE et al.1967
2 HALL 1974
3 LANGOALE etal. 1978
4 TRIPLETT etal. 1978 °4
04
30
30
I
10
\
20
I
30
I
40
^-M-
u.
u.
o
z
- 10
o
i-
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1
I
I
I
decrease exponentially in subsequent events, as indicated for atrazine by
the solid line in Figure 7b. This effect is the result of several factors,
including degradation, plant uptake, leaching, and depletion of initial
mass; it emphasizes the fact that highest herbicide concentrations occur in
the period shortly after field application. Wauchope and Leonard (1980)
have used literature data to develop an empirical function that generalizes
this relation:
Ct = AR(1 + 0.44t)~1-6 (1)
where Ct is the runoff concentration at event time t, R is the
application rate, and A is the availability index (a function of the
chemistry of the particular compound). Moreover, the total amount of
herbicide transported into surrounding watercourses over the whole season
is also a function of the timing of the first runoff event after
application (Figure 7b, dashed line). A similar, first-order decay
function generally describes this relation. The data compiled in this
figure suggest that if no runoff occurs within the first 10 days after
application, total atrazine loss to the watercourse will probably be less
than one percent of that applied, while runoff within three days following
application can lead to seven percent loss. Thus, both the concentration
and total amount of herbicide entering the estuary may depend largely on
the time interval between herbicide application and rainfall-runoff events.
524
-------
SECTION 4
ENVIRONMENTAL BEHAVIOR OF HERBICIDES
I
I
I
Two key processes that determine the fate of herbicides in the
environment are adsorption and degradation. As suggested in Figure 2, •
these processes occur in both terrestrial and aquatic (estuarine) |
environments. Before the initiation of CBP research in 1978, very little
was known about the nature of these processes as they occur in estuaries. —
Hence, parallel experiments were designed to examine adsorption and •
degradation in simulated estuarine and soil systems, representative of •
typical conditions in the Bay and its watershed. Emphasis was placed on
atrazine, although experiments were also performed with linuron and other •
compounds. Atrazine degradation appears to proceed more rapidly in systems |
with sediments and/or soil, than with water alone, and Jones et al. (I981b)
have postulated that most of the degradation may be preceeded by sorption, •
followed by desorption. Hence, the two processes are intimately coupled. •
SORPTION REACTIONS
The adsorption of dissolved herbicides to solid surfaces proceeds as a •
function of aqueous concentration (C) until an equilibrium is achieved
between C and the adsorbed concentration [x/m, where x is the weight of •
herbicide adsorbed to solid (micrograms), and m is the weight of solids •
(grams)]. The equilibrium relation is often described using the Freundlich
equation, _
I
x = (Kd)(C)1/n (2)
m
where n is a constant describing the shape of the equilibrium relation, and |
K^ is the sorption coefficient (for example: Giles et al. 1960, Bailey
and White 1970, Kempson-Jones and Hance 1979, Travis and Etnier 1981). •
High Kjj values indicate a stronger tendency for adsorption. Several •
studies suggest that organic matter in the substrate tends to be the
controlling factor for adsorption of many non-polar organic compounds such
as herbicides (Bailey and White 1964; Karickhoff et al. 1979; Means et al. I
1979, 1980). Therefore, it is convenient to normalize Kj values to the *
organic matter of the sorbant,
KQC = K(j/(decimal fraction organic carbon) (3) |
Values of Koc for atrazine and linuron have been reported for a wide •
variety of soils, ranging from 47 to 394 (atrazine) and 124 to 2678 •
(linuron), with typical values being 170 and 670, respectively (Rao et al.
1981).
Sorption isotherms for atrazine with agricultural soils, estuarine m
sediments, and estuarine colloids, and for linuron with estuarine sediments •
and colloids, were determined from the Bay region (Means et al. 1981a;
Means and Wijayaratne 1981). The atrazine data are summarized in Figure
8a. All isotherms were linear over the range of concentrations tested (n =
525
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1.0). The Koc were about 10 times greater for colloids (2,000-14,000)
than those for sediments (200-400), which were in turn generally greater
than those obtained for soils (100-200). Values of KQC for linuron with
sediment and colloidal matter from the Patuxent River were 3.4 times
greater than for atrazine with the same substrates. This relationship is
almost identical to the relationship between the two compounds for soils,
where linuron KQCs were 3.9 times greater. These values of Koc for
colloids appear to be consistent with the findings of Wu et al. (1980), who
reported that estuarine surface microlayers at Rhode River were typically
enriched with atrazine by a factor of about 10 to 30 over bulk water
concentrations. If it is assumed that this enrichment is due to sorption
by the hydrophobic colloidal matter concentrated at this air-water boundary
with associated organic carbon of 5.0 ppm, then the Koc values would be
about 2,000 for the samples of Wu et al. (1980).
Correll and Wu (1981) have reported Kd values ranging from 5.0 to 260
(depending on C) for atrazine dissolved in distilled water and adsorbed to
Rhode River estuarine sediment. These values are two to 100 times greater
than those of Means et al. (1981a) and other investigators. The highest
Kjj's of Correll and Wu would correspond to Kocs of Means et al., only
if the Rhode River sediment were 50 percent organic carbon. They do report
extremely high organic carbon percentages; however, even these are too low
(five to 27 percent) to explain the differences (Correll et al. 1978).
Moreover, their data imply that Freundlich isotherms would be non-linear in
the same general concentration range as that given by Means et al. (1980)
and others and, using their data, we calculate n = 2.3 with Kj = 126
(Equation 2). It is difficult at this point to resolve these discrepancies.
The adsorption of atrazine and linuron has been extensively studied on
a wide variety of soils (Talbert and Fletchall 1965, McGlamery and Slife
1966, Green and Obien 1969, Harris and Warren 1967, Weber et al. 1969,
Bailey and White 1970, Grover and Hance 1970, Hurle and Freed 1972, Colbert
et al. 1975, Hiltbold and Buchanan 1977, Dao and Lavy 1978). A number of
factors have been identified that influence the adsorption of these
compounds to soils, including pH, temperature, moisture, electrolytes, and
organic matter. Of these factors pH and salinity were examined to
determine their potential effects on herbicide adsorption under estuarine
conditions. Salinity exerted a small (five percent) negative effect on
adsorption with sediments between 5.0 and 15 ppt, but the overall effect
from 0 to 15 ppt was erratic and probably nonsignificant (Figure 8b). For
colloids, on the other hand, salinity between nine and 19 ppt appeared to
have a substantial negative effect (29 percent). This pattern was
consistent between experiments (where salinity was manipulated) and field
observations (where salinity varied along the estuarine axis), but was
opposite to that which would be predicted as a result of "salting-out" of
the hydrophobic herbicide. This finding suggests that salinity affects
more the nature of the colloidal material than the solubility of the
herbicide (Means and Wijayaratne 1981). It was found that pH also
influenced Koc for colloids with atrazine and linuron. Both herbicides
exhibited maximum KOC at a pH approximating that of the estuarine
environment from which water and colloids were taken. Increasing or
decreasing pH by one unit (that is, between pH 7.0 and 9.0) caused a 2.0 to
20 percent decrease in KQC for the two herbicides, although at pH 5.0 to
6.0 KQC dropped by 25 to 35 percent.
526
-------
(a) FREUNDLICH ADSORPTION
ISOTHERMS FOR ATRAZINE
ESTUARINE
SEDIMENTS
ESTUARINE
COLLOIDS
AGRICULTURAL
SOILS
in
•a
o
o
X
in
-O
-------
I
I
I
I
I
I
I
I
I
I
1
I
I
I
I
I
I
I
I
The implications of these sorption data are that atrazine and linuron
are readily susceptible to runoff and leaching from agricultural fields,
even without particulate soil erosion. Control of soil erosion alone will
not control atrazine transport to the estuary. Means et al. (1981) suggest
that once in the estuary, dissolved atrazine will adsorb readily to
suspended sediments and colloidal material. Sediment-sorbed atrazine will
move only through resuspension and sediment transport; colloidal-bound
herbicide may travel great distances and concentrate in organic film at the
air-water interface, typical of coastal waters, particularly during the
fall. The expected fate of linuron would be analogous, except less would
leave the field in dissolved form, but more would adsorb to estuarine
particles.
HERBICIDE DEGRADATION
An important factor contributing to the potential toxicity of
herbicides, such as atrazine and linuron, is the longevity of these
substances in the field or estuary. Numerous studies have described the
kinetics of atrazine degradation in various soil environments, and a wide
range of physical factors (such as pH, temperature, moisture, clay, and
organic content of soils) has been shown to affect this process (for
example, Swanson and Dutt 1973; Best and Weber 1974; Hiltbold and Buchanan
1977; Hance 1979; Kempson-Jones and Hance 1979; Kells et al. 1980). Direct
experiments were conducted for atrazine degradation in flasks with
estuarine water and sediments maintained in natural light under field
temperatures (Jones et al. 1981b). We also monitored atrazine and linuron
concentrations in laboratory microcosms [25 L (6.6 gal) and 700 L (184.9
gal)] over eight weeks and thus, indirect estimates of degradation were
obtained (Cunningham et al. 1981a, 1981b).
Much of the information on herbicide longevity developed by agronomists
and soil scientists refers to persistence in agricultural soils (that is,
the time required for 90 percent or more of the compound to disappear from
the site of application). Reported values of "field persistence" result
both from degradation and mobility of the compound. Nonetheless, to the
extent that mobility may not vary excessively among the compounds,
persistence provides a rough estimate of relative degradation rates in the
field. A summary of persistence data (defined as above) for nine
herbicides important in the Chesapeake Bay region is presented in Figure 9
(Stewart et al. 1975). Except for paraquat (the persistence of which is
largely related to its highly sorptive nature), the ji-triazines, atrazine
and simazine, are the most persistent of these compounds. Thus, it appears
that atrazine is among the more persistent herbicides in common use, and
understanding its degradation should provide a conservative perspective on
herbicides in general.
Herbicide degradation is often described as a first-order decay process
Ct/C0 = e-kt (4)
where ct is the amount of the compound remaining at time t; Co is the
initial amount and; k is a decay rate coefficient. Others have suggested
that higher order processes better describe the herbicide degradation (for
528
-------
PARAQUAT
1
!
^
-///////////////////////////////////////////*
rr^
ATRAZINE
SIMAZINE
1
TRIFLURALIN
/////////////////A
LINURON
V////////////X
\ DICAMBA
V////////A
UJ
CO
ALACHLOR
-
-
CQ
Ul
O
Z
Q
1
\ FV!
00
-------
I
I
I
I
I
I
I
I
I
I
I
1
I
I
I
I
I
I
I
example, Hamaker 1972; Kempson- Jones and Hance 1979) so that the more
general relation applies
Ct " [Co (1~n) + (n-Dktld/d-11) (5)
where n is the apparent order of reaction. The overall rate of reaction is
often described by the half-life (T^/2^> °r time required for
disappearance of 50 percent of the original substance. This
equal to (0.693/k) for first-order reactions.
Atrazine can degrade through chemical and biological processes into
metabolites, some of which may be toxic. The degradation of atrazine to
its metabolites can occur through: chemical hydrolysis to hydroxyatrazine;
dealkylation to either the de-ethylated, deisopropylated, or deaminated
atrazine forms followed by hydrolysis; or conjugation. The dealkylations
and ring cleavage are generally considered to be biologically
(enzymatically) mediated, but the hydrolysis to hydroxyatrazine is
controlled by physical parameters, most notably pH (Armstrong et al.
1967). Ring cleavage of atrazine is a very slow process, typically causing
losses of only a few percentage points over several years; however, prior
hydrolysis to hydroxyatrazine does increase the rate of ring cleavage
(Armstrong et al. 1967). The biological degradation is performed by soil
fungi and bacteria, with the organisms using mainly the side-chains as
carbon sources, nitrogen sources, or both (Kaufman and Blake 1970).
The degradation of l^C ring-labeled atrazine in two estuarine
water-sediment microcosms (from Choptank and Tangier) and two soil systems
(well-drained Sassafras and poorly drained Mattapex) was compared over an
80-day period under high- and low-oxygen tensions (Jones et al. 1981b) . In
the estuarine systems, total residues moved from water to sediments over
the course of the experiment, and the relative percentage of parent and
daughter compounds changed rapidly during the first several weeks (Figure
10). The initial degradation products generated in the estuarine systems,
as revealed by thin-layer chromatography and autoradiography , appeared to
be the same as for the soil systems, with hydroxyatrazine being the major
short-term metabolite. By the 21st day of the experiment, the percentage
of total extracted residues corresponding to atrazine, monodealkyalted
atrazine, and hydroxyatrazine were 65, 10, and 25 for the Choptank, and 15,
eight, and 77 for the Tangier (the estuarine systems); 66, 5.0, and 29 for
the Sassafras, and 93, 2.0, and 5.0 for the Mattapex (the soil systems).
Atrazine degradation was far more rapid in the estuarine systems than
in the soils. Half-lives for the herbicide ranged from three to nine days
in overlying estuarine water, and 15 to 20 days for estuarine sediments, as
compared with 330-385 days for agricultural soils. A portion of the
residues adsorbed to sediments and soils was nonextractable, and these
half-life estimates employed the conservative assumption that extractable
and nonextractable residues were similarly distributed among the three
major metabolites. Oxygen tension appeared to have negligible effect on
atrazine degradation; however, the low oxygen systems were not completely
anoxic. The relative degradation rates of atrazine in the three
environments, as observed in this experiment and other studies from the
literature, are summarized in Table 3, where mean half-lives are 14, 45,
and 180 days, respectively, for estuarine water, aquatic sediments, and
agricultural soils. Thus, atrazine is less likely to be a problem in the
530
-------
en
LU
Z)
Q
V)
LU
CC
LU
_l
CD
<
O
<
cc
I-
X
LU
LU
O
tr
LU
a.
WATER
(a) CHOPTANK
M ATRAZINE PARENT
p| HYOROXYATRAZINE
I OEALKYLATED ATRAZINE
I4C ACTIVITY IN WATER
(b) CHOPTANK -02
SEDIMENTS
(a) CHOPTANK + Oj (b) CHOPTANK - 02
(a) SASSAFRAS +
SOILS
(b) SASSAFRAS - 0.,
100-
80-
60-
40-
20-
1
7
•
1
14
•V
1
1
21
*w
\
1
28
•«
1
45
4*
1
59
B
1
1
80
I
f3
7 14 21 28 45 59 80
DAYS FROM START OF EXPERIMENT
14
Figure 10. Loss of C-ring labeled atrazine from experimental systems
and percent of total residuals as parent compound and as two
major metabolites (adapted from Jones et al. 1981).
531
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
1
I
I
I
I
I
1
t
I
I
I
estuary, because it rapidly degrades; however, because the herbicide's
half-life on farmland soils is longer than in the estuary, it remains
potentially available for runoff long after application.
TABLE 3. SUMMARY OF ATRAZINE DEGRADATION RATES IN AGRICULTURAL AND
ESTUARINE ENVIRONMENTSa
Environment Half-life (days)b
Mean Range
Agricultural Soils 180 10-1200
Aquatic Sediments 45 6-145
Estuarine Water 14 3-30
aSummarized after Jones et al. (1981).
"Time required for 50 percent degradation of original compound.
Herbicide data from our estuarine microcosm phytotoxicity experiments
(Cunningham et al. 1981a, 1981b) indicated that overall disappearance of
atrazine from the water-sediment environment occurred somewhat more slowly
than in the estuarine flask systems of Jones et al. (I981b), but still more
rapidly than for soils (Figure 11). Half-lives of atrazine were on the
order of 60-80 days in the microcosms. The slower decomposition in these
systems may be a function of the fact that the experiments were performed
in artificial lights, which would contribute less to photodecomposition
(Jordan et al. 1964), or perhaps higher pH and/or reduced organic
substrate. A value of t]/2 from similar microcosm data of Correll and Wu
(1981) was calculated to be about 30 days, which is closer to the combined
sediment-water t\/2 of 10 to 20 days from Jones et al. (I981b).
Microcosm experiments involving linuron indicated much faster degradation
of this herbicide, with ?i/2 = 10 days (Cunningham et al. 1981b), a
result consistent with the relative persistence of the two compounds in
soil (Figure 9) .
It appears that atrazine is a relatively good agricultural weed-control
compound, in that it persists in soils (where it can perform its designed
function) 10 times longer than in the estuary (where nontarget species
might be exposed to its toxic effects). Apparently, atrazine is one of the
more persistent herbicides in use, and its half-life in the estuary is at
least six times greater than linuron. This 6:1 relationship is similar to
the 3.5:1 relationship reported for the persistence (time for 90 percent
disappearance from field application) of these two compounds in the field.
Therefore, the field persistence data (such as Figure 9) may, in some
cases, provide a crude index of potential estuarine longevity.
532
-------
co S
w i
o
o
o
o:
o
2
z
LU
O
rO
_| Q
<
CD
CO
CO
S
UJ
•z.
N
ct:
H <
< Q
•
:
E
H O
I10
o>
E
(E 10
^ 9 o
1
55
rO
^^
^0
^> O
0:0°
g 6£
1
CO
a
CO
a
0
PLANTS
PLANTS
*
CO
PLANT
£££
o» .0
E «>
N rH
cd oo
O > CTl
01 rH
0)
Z C i-l i-H
O cd t-> cd
H « § w
tf CU W 0)
0 0 2 S S
«f w SB
TS i ,C co
Q? C j_i hf\ r!
O C 60
uj " tj ti
•r cu c -H
Q c -H c
o cd c
^ M C O
0) O ^
•U O
cd co
-------
I
I
I
I
I
I
I
I
I
I
1
I
I
I
I
I
I
I
I
SECTION 5
TOXICITY OF HERBICIDES IN THE ESTUARY
In this section the toxicity of herbicides used in the Chesapeake Bay
region is considered. The major concern is herbicide phytotoxicity to SAV,
particularly in Chesapeake Bay. We also review other aspects of herbicide
toxicity for animals, algae, and emergent aquatic plants, as well as
mutagenic action of these compounds. We begin with a general review of the
known mechanisms of toxicity, again emphasizing atrazine and linuron.
TOXIC MECHANISMS
Herbicides can kill plants by interfering with photosynthesis,
respiration, and other aspects of plant metabolism. The major herbicides
in use today can be categorized as to their site of action. Four sites are
recognized: the chloroplast, the mitochondria, protein synthesis, and
membrane permeability. Of these, the chloroplast-related group of
herbicides are in the widest use in the Chesapeake Bay watershed. Two
herbicides in this group are atrazine and linuron. Both of these compounds
appear to inhibit the Hill reaction of photosynthesis at the same location
within the chloroplast, stopping electron transport leading to the
production of the reduced cofactor (NADP^) for the fixation of C02-
There is disagreement among investigators as to the exact location of this
attack (Ebert and Dumford 1976), but most concede that it is between the
initial electron acceptor Q in photosystem II and plastoquinone (Gysin and
Knusli 1960, Moreland and Hilton 1976). More specifically, recent
information indicates that both atrazine and linuron compete for the same
protein-binding site on the thylakoid membrane, possibly causing a
conformation change, blocking electron transport (Brewer et al. 1979).
The fact that the site of inhibition is within the chloroplast itself
can dictate the relative toxicity of a compound or its organic solubility.
The lipoid-rich membrane environment of the chloroplast makes penetration
of more polar compounds difficult, thus reducing their access to the
binding site. In the case of atrazine, the daughter products show a
decreasing phytotoxicity in the order of de-ethylated atrazine >
de-isopropylated atrazine > hydroxyatrazine. This toxicity inversely
correlates with their relative polarities (Lamoureux et al. 1970) and the
order relates to how soluble herbicides are. Therefore, solubility data
can provide some insight into relative toxicity.
Resistance to photosynthetic inhibitors in plants is manifested either
in the ability to degrade the parent compound to nontoxic metabolite(s), to
complex the compound through conjugation, or to acquire altered binding
sites on the chloroplast membrane through genetic selection. Degradation
of the parent compound may be enzymatically or nonenzymatically
controlled. Corn contains the compound benzoxazinone that nonenzymatically
hydrolyses atrazine to hydroxyatrazine. Corn also contains enzymes that
degrade atrazine to its dealkylated products. Sorghum can conjugate
atrazine by a glutathione s-transferase enzyme, which removes the chlorine
from the molecule, allowing a bond to form between the triazine ring and
the sulfur of glutathione (Shimabukuro 1968). Pfister et al. (1979)
534
-------
I
suggest that plant species may develop resistance to herbicides by the •
evolutionary selection of altered binding sites on the chloroplast *
membrane. Theoretically, SAV might develop resistance to herbicides
through similar genetic mechanisms which provide a means of increasing tt
degradation of the herbicide within the plant cells, or tying it up. |
TOXICITY TO ANIMALS _
Since the toxic mechanisms for many of the important herbicides act
directly on the chloroplast, it would be anticipated that effects of these
compounds on heterotrophic organisms would be substantially less (Stevenson •
et al. 1981). Toxicity data for various estuarine animals tend to support •
this hypothesis. For example, the fiddler crab, Uca pugnax, was observed
to withstand concentrations up to 100 ppm atrazine with no demonstrable fl|
effects in bioassays (Davis et al. 1979). Only at 1,000 ppm was the |
escape-response ability of fiddler crabs damaged, so that normal activities
in the saltmarsh ecosystem were impaired. Even when fed cordgrass —
(Spartina alterniflora) containing atrazine, box crabs showed little M
behavioral response (Pillai et al. 1979). Similarly high levels of ™
atrazine resistance have been reported for mud crabs, Neopanope texana
(Newby et al. 1978). Shrimp and oysters have been shown to be somewhat B
more sensitive to atrazine in bioassay experiments, with shrimp exhibiting 0
30 percent mortality in 96 hours at 1.0 ppm atrazine, and oysters showing
no effects at this concentration (Butler 1965). *
MUTAGENCITY
An increasing concern in recent years has been the discovery of the •
mutagencity of pesticides and/or pesticide metabolites. The issue has been •
complicated further by the instances in which a nonmutagenic parent
compound can be activated by either plant or animal metabolism into a •
mutagenic substance. Herbicides, like most pesticidal compounds, often f
contain chlorine or bromine substituents, aromatic rings, and amine
groups. These functional moieties have often been associated with ^
mutagenic activity in organic molecules. It should be noted that, in some •
cases, mutagenicity has been associated with a trace byproduct from *
commercial production of the pesticide (that is, dioxins in 2,4-D and
nitrosamines in trifluralin). A summary of available information on the •
mutagencity of the major herbicides used in the Chesapeake Bay region is m
presented in Table 4. The issue of genetic toxicity of the herbicides was
not addressed in any of the research funded by the CBP-SAV program. It
remains, however, an important question that needs to be considered in the
future.
535
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
i
i
i
i
i
i
i
i
i
i
i
TABLE 4. MUTAGENICITY OF MAJOR HERBICIDES IN CHESAPEAKE BAY REGIONS
Compound
Relative Mutagenicity
Comment s
Sodium Azide
Atrazine
Simazine
Cyanazine
Diquat
Paraquat
2,4-D
Dicamba
Trifluralin
Linuron
Alachlor
Propachlor
NT
NT
plant activated
plant activated
plant activated
dioxin contaminant,
by-product +++
nitrosoamine,
by-product ++
plant activated
plant activated,
synergistic enhancement
with triazines
aSource
"Symbols: +++ extremely large effect; ++ large effect; + some effect;
- no effect; NT not tested
PHYTOTOXICITY FOR ALGAE AND EMERGENT PLANTS
Phytoplankton vary widely in their susceptibilities to atrazine, with
toxic concentrations ranging from 20 to 1,000 ppb (Stevenson et al. 1981).
Davis et al. (1979) reported that 100 ppb atrazine caused some effects on
mixed algal assemblages from coastal waters. For the diatoms,
Thallassiosira and Nitzschia, the LD5Q was 1,000 ppb. Reductions in cell
density of 10, 90, and 100 percent were obtained at concentrations of 20,
200, and 500 ppb, respectively, for the chlorophyte, Chlamydampnas spp.
(Loeppky and Tweedy 1979, Hess 1980). Pruss and Higgins (1974) reported no
lasting effects on algal populations in a lake treated with 100 ppb
simazine. Chlorella pyrenoidosa possesses a high degree of resistance to
atrazine, where 1,000 ppb were required for 50 percent reduction in
chlorophyll-^ (Kratky and Warren 1971). Metz et al. (1979) found similar
resistance for Chlorella strains, which was attributed to the ability of
this species to exist heterotrophically. Bryfogle and McDiffett (1979)
showed that productivity in algal cultures treated with 400 ppb simazine
was actually enhanced, although species diversity was reduced with
536
-------
I
resistant Chlorella strains dominating the experimental systems. •
Atrazine effects on Spartina alterniflora have been studied extensively •
by Davis et al. (1979). Cordgrass biomass was unaffected at 10 ppb; at 100
ppb, a 34 percent reduction in biomass was observed; at 1,000 ppb, •
approximately 46 percent of biomass was lost. Apparently, Spartina has |
some detoxification capability for atrazine, by a mechanism similar to that
of sorghum (Pillai et al. 1977). Thus, laboratory studies suggest that _
neither algal nor marsh grass populations would be seriously damaged at •
atrazine concentrations in the range observed in Chesapeake Bay; however,
herbicide treatment could conceivably contribute to phytoplankton species
shifts that allow monospecific bloom conditions. 1|
SAV PHYTOTOXICITY
The crucial relationship in this,discussion is the potential phytotoxic I
effect of herbicides on SAV. Two herbicides, atrazine and linuron, were
tested against SAV species Potamogeton perfoliatus (a dominant native) and
Myriophyllum spicatum (an exotic that was extremely abundant just before V
initial SAV decline in 1964). Though historically important in Chesapeake w
Bay, both of these species are of freshwater origin. The effects of
atrazine on the marine and estuarine seagrasses, Zostera marina and Ruppia •
maritima, have also been tested (Hershner et al. 1981). Experimental ^
exposures ranged from incubations of six to 24 hours both in situ and in
vitro, to five weeks in laboratory microcosms of three sizes and designs. M
Similar microcosm studies were performed by Correll and his colleagues, I
using atrazine with three additional Bay species: Potamogeton pectinatus,
Zannichellia palustris, and Vallisneria americana. Thus, a broad data base
now exists on this topic, with seven species, two herbicides, and six •
experimental designs. V
Effects on Photosynthesis and Respiration Wt
The general response of P. perfoliatus to atrazine treatment in
laboratory microcosms is shown in Figure 12 (Cunningham et al. 1981b) , ^
where mean values for apparent photosynthesis, Pa (net Q£ production •
during the day), are given for microcosms under control, and under six *
herbicide dosages over a nine-week experimental period. The shaded portion
of each graph represents the departure of actual metabolic rates from M
expected values (based on both pretreatment and control data). Similar 0
data have also been reported for atrazine effects on M. spicatum, and for
linuron toxicity to both SAV species (Cunningham et al. 1981b). In M-
addition, P. perfoliatus response to low concentrations (1.0 to 25 ppb) of •
the atrazine was retested.
In general, the response patterns of SAV to herbicides were similar to
that shown in Figure 12, where marked decreases in photosynthesis were •
observed at concentrations greater than 50 ppb, with some less pronounced 9
effects at lower concentrations. Myriophyllum, however, exhibited slightly
I
was actually enhanced over controls. Recent short-term experiments
indicated that the Pa response of R. maritima to atrazine was similar to «
that of P. perfoliatus (Jones, unpublished data). Simple two-way analysis I
greater resistance to atrazine, but virtually identical response to linuron
as compared with P. perfoliatus. At 5.0 ppb atrazine, Pa for M. spicatum
537
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
•o
r
>*
O
2
o
QC
o.
Z
Ul
e
3E
Q.
6
4
2'
6
4
i
2,
6
4
2
POTAMOGETON / ATRAZINE - I
CONTROL
SOppb
100 ppb
500 ppb
1000 ppb
34567
WEEK OF EXPERIMENT
8
Figure 12. Summary of measurements of apparent photosynthesis for
microcosms containing Potamogeton perfoliatus treated with
atrazine (0 to 1 ppm). Data points are means of duplicate
measurements on duplicate systems (n = 4) (Cunningham et al.
1981b).
538
-------
I
of variance suggests that effects were always significant (p < 0.05) for •
concentrations greater than 50 ppb, and sometimes significant at 5.0 ppb; W
however, further statistical analysis is still in progress.
At all concentrations less than 50 ppb, Potamogeton Pa exhibited a M
trend of recovery toward control levels after the second post-treatment •
week. The same pattern was found for M. spicatum with atrazine, and for
both species with linuron. For both species, however, incipient recovery ^
from linuron treatment occurred after the first post-treatment week. Rates •
of recovery were similar in all cases, being about 0.5-1.5 mg *
"compensation point"; Pa:R > 1 indicates net growth, and Pa:R< 1
suggests net loss of plant material. It was found that Pa:R offered a
The ratio of apparent photosynthesis to dark respiration (Pa:R) ft
provides a measure of the energy balance for plants and has been used as an ft
index of stress. The point where Pa just equals R is termed the
a:R
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
CO
10
UJ
2
UJ
Q.
X
UJ
UJ
UJ
CNJ
q
cvi
q
d
ouva
BlAHdOdOVW
C cfl
O A
•H 60
4-1 C
CO -H
K c
•H c
a, 3
co o
c cu
CO
o o
O 4-1
CL.JS
4J
4J -H
c s
CU
CO CU s~*
O. 4J Cfl
O. Cfl rH
Cfl CU OO
)-l cfl
cfl fli 4J
« v^ CU
00
•H
540
-------
O
CO
O>
w
uj cr
z<
CO
CM
m 3
CM o
N39AXO Q3A10SSia
00
CU N rH
0 cd
O M •
T) -W r-l
CO CO
0)
M
C
•H "O
0) X
n -u to
01 cd i-i
•o u -"
U
CM <;
o co >
•H -H CO
T) C M
*i*
O CO
co u cu
CU C
4-> -H
^J il
CO CO
B
CO rl
U 0)
a. co
60
CO
CO
00
541
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
a
o
X
0)
I
4-1
g
O
00
u
-------
3
O
0)
C
•H
j_l
C
O
u
3
00
N39AXO Q3A10SSIQ
543
I
I
I
I
1
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
1
I
I
I
I
I
O
00 X
ff> to
- cc
<
H S
CO
o
•8
o
.8
n
O
o
.O
• 10
(VI
O
. o
o
•s
3
o
I
UJ
2
•d
0)
C
o
o
-------
I
for 34 varied experiments. Some differences were found between atrazine •
and linuron effects, and among the three SAV species. Of all species, m
Zostera exhibited the greatest effect at low herbicide levels, with an
apparent threshold concentration (intercept of x-axis) of about 1.0 ppb; M. A
spicatum was the most resistant with a threshold of about 6 ppb. This I
model predicts that at 10 ppb atrazine, the resulting reductions of SAV
Pa for Myriophyllum, Potamogeton, and Zostera would be approximately 0,
17, and 33 percent, respectively. Thus, in the lower-Bay waters small •
concentrations may have greater effect on SAV (that is, Zostera) W
photosynthesis.
Correll and his colleagues (Correll et al. 1978, 1978; Correll and Wu M
1981) have reported other herbicide-SAV experiments for Chesapeake Bay £
plants. They have investigated atrazine effects on a second species of
Potamogeton (P. pectinatus) and on Z. marina, as well as on two additional ^
freshwater genera (Zannichellia palustris and Vallisneria americana). They I
have also reported some results of linuron effects on Z. palustris.
Time-course experiments (21-48 days) have been performed for a range of
herbicide concentrations. In general, the patterns of responses reported B
are similar to those in Figure 15, where, for example, linuron effects seem V
to be greater than those of atrazine. Correll's results, however, appear
to suggest considerably greater resistance to atrazine for all test ip
species. For example, the maximum effect found for any species at 75 ppb •
was for Z. palustris, which exhibited about a 40 percent reduction in Pa,
and the minimum effect reported by Cunningham et al. (I981b) was 42 percent
for M. spicatum at 100 ppb atrazine. Moreover, two of four species tested •
exhibited significant enhancement of Pa by 75 ppb (Correll and Wu 1981). ™
One of those enhanced species was Z. marina, the same plant that Hershner
et al. (1981) reported never exhibited less than 47 percent reduction in •
Pa at 100 ppb for five experiments. 9
Few other comparable data are available in the literature. Walker
(1964) reported that Potamogeton sp. was effectively controlled in fish »
ponds (that is, removed from ponds for at least two months) with treatments •
of 0.5-1.0 ppm simazine. Herbicide bioassay experiments with the'submerged
vascular plants, Myriophy11um brasiliense and Elodea canadensis, showed
that oxygen evolution was suppressed (25 and 40 percent less 02, •
respectively) by simazine at concentrations of 120 ppb (Button et al. •
1969). Fowler (1977) has shown, more recently, that effects of a related
£-triazine herbicide (DPX 3674) on Myriophyllum verticillatum and P. m
pectinatus could be detected at 125 ppb. Stevenson et al. (1981) have M
reported some unpublished data of J. Forney for E. canadensis, where a one
percent growth inhibition was found at 3.2 ppb atrazine, and a 50 percent _
inhibition occurred at 80 ppb. •
Effects on SAV Population, Biomass, and Physiomorphology
Total SAV biomass responded to herbicide treatment in a manner •
analogous to plant photosynthesis. Typical biomass data from microcosm
experiments are provided in Figure 16, where j?L perfoliatus exhibited M
significant reduction in plant matter at concentrations greater than 50 ppb M
linuron, and M. spicatum followed a secular decrease in biomass when
exposed to linuron concentrations from 5.0 to 1,000 ppb, although the loss
was significant (compared with control) only at concentrations greater than V
545
i
I
-------
I
I
I
I
I
I
I
I
I
1
I
I
I
1
I
I
i
I
I
w
UJ
X
100
80
60
o
Q- 40
z
z
o
O
Ul
IT
20-
O KTRAZ\HE./Potomogaton perfoliatus
(NUMBERS REFER TO DIFFERENT
EXPERIMENTS)
D LINU RON / /? perfoliatus
• ATRAZIN E /Zostera marina
® ATRAZINE/A
-------
in
o
a
o
(T
o
CO
CO
S
O
CD
O
X
a.
o
cr
a
10
10
m
(a)
Potamogeton perfo/iafus
(* i S)
(b)
Myriophyllum sp/cafum
(* t S)
CONTROL 5 50 100
LINURON DOSE, ppb
500
1000
Figure 16. Sunnnary of measurements of plant biomass in microcosms
containing (ai Potamogeton perfoliatus, and (b) Myriophyllum
spicatum, treated with linuron (0 to 1 ppm) (Cunningham et al.
1981b).
547
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
1
I
I
I
I
I
I
I
I
100 ppb. A slight, but insignificant increase in biomass was found at 5.0
ppb for M. spicatum. These data are from the final week of the experiment,
and time-course effects of herbicide dosage on biomass lagged metabolic
responses by one to two weeks. Ratios of above:below-ground biomass were
generally unaffected by herbicide treatment, but shoot density was enhanced
at moderate dosage (100 ppb) and sharply reduced at high doses (Table 5).
Although above:below-ground biomass ratios can be an index of stress, none
was found in this case. Hershner et al. (1981) reported only small,
insignificant effects on Zostera shoot height, density, leaves-per- shoot,
and mortality at atrazine concentrations less than 100 ppb after 27 days,
whereas marked effects were apparent at 1,000 ppb (Figure 17). However, P.
perfoliatus exhibited significant etiolation of stems and increases in
chlorophyll £ content of leaves at 100 ppb atrazine (Table 5), both of
which are typical responses to light stress.
TABLE 5. SUMMARY OF SELECTED STRUCTURAL CHARACTERISTICS OF POTAMOGETON
PERFOLIATUS POPULATIONS IN MICROCOSM COMMUNITIES TREATED WITH THE
HERBICIDE, ATRAZINE
(CUNNINGHAM ET AL.
1981a)a
Structural15
Characteristic
Chlorophyll-a
(mg m~2)
Foliar Biomass (Ba)
(g d.w. m~2)
Rhizobial Biomass (Bb)
(g d.w. m~2)
Ratio, Bb:Ba
Unit Length of shoots
(cm g~I)
Shoot density
(no. m~2)
Control
28 + 8
44.3 + 17.1
40.0 + 12.9
0.93 + 0.22
24
468
Treatment
Low
(0.1 ppm)
158 + 16
24.3 + 8.7
20.0 + 8.6
•
0.94 + 0.54
53
495
High
(1.0 ppm)
114 + 5
0
0
-
63
134
aData are from samples taken in the final (6th) week of the experiment.
"Given are mean values +_ standard deviation, where n = 12 for chlorophyll
and n = 6 for biomass. Values for shoot length and shoot density are
measurements from harvest of entire plant population for duplicate micro-
cosms at each treatment.
Correll and Wu (1981) examined the effects of atrazine on Vj_ americana
in some detail, after initial screening experiments indicated that it was
the most sensitive of the four species they examined. They monitored
mortality, vegetative reproduction, and leaf growth as indices of herbicide
stress; they observed about a 35 to 40 percent increase in mortality above
control at 12 ppb atrazine, with similar losses of reproduction and
growth. No significant effects were seen at 3.2 ppb atrazine. This effect
548
-------
- 80
QC
O
2 40
Q[ 20
UJ
X
UJ
O
z
<
X
u
+ 20
0
-20
-40
-60
-80
-100
a) ATRAZINE-INDUCED MORTALITY
OF ZOSTERA MARINA
b) ATRAZINE-INDUCED
CHANGES IN HEIGHT OF
Z. MARINA
CONTROL O.I I 10 100 1000
ATRAZINE CONCENTRATION (ppb)
Figure 17. Effects of atrazine on (a) percent mortality of Zostera
and (b) height of Zostera (Hershner et al. 1981).
549
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
of 12 ppb is reasonably consistent with the decreases in biomass that were
found for P. perfoliatus at 50 ppb (either linuron or atrazine), where up
to 60 percent reduction (for example, Figure 16) occurred (Cunningham et
al. 1981b). Correll and Wu (1981), however, found almost a 40 percent
increase in mortality of Vallisneria at 12 ppb atrazine, but only a 20
percent decrease in Pa at 75 ppb. It would seem that effects on Pa
should be greater than those on survivorship, since reduction in
photosynthesis does not necessarily lead to death, whereas the inverse is
true. We have no explanation for this apparent inconsistency.
Other Factors Affecting Phytotoxicity
To place these bioassay experimental results into proper perspective,
it is necessary to consider several factors that influence the ultimate
effect of herbicides on SAV.
Acute Versus Chronic Exposures—
Somewhat surprisingly, the relative toxicities of herbicides to SAV
appeared to be independent of exposure-time. Experiments involving
incubations of six to 24 hours (Jones et al. 198lb, Hershner et al. 1981)
yielded results virtually identical to those obtained from exposures of
four to five weeks (Cunningham et al. 1981a, 1981b). In the linuron
experiments, dose-response patterns closely followed those for atrazine,
even though only 10 percent of the parent linuron remained after a
four-week period. On the other hand, Pa of Potamogeton treated with 50
ppb atrazine dropped to 35 percent of controls two weeks after treatment,
but recovered to 70 percent of control levels two weeks later, even though
herbicide concentration remained at 75 percent of initial levels (Figure
12). Thus, it appears that the initial short-term exposure to herbicides,
at a given concentration, largely determines the subsequent pattern of
stress and recovery.
Based on recently-conducted, time-series measurements of ^C-labeled
atrazine uptake by Pj_ perfoliatus, it appears that most uptake occurs
within one to two hours, and no additional incorporation can be measured
after two days of exposure to constant concentrations (T. Jones,
unpublished data). This finding suggests that, in nature, even ephemeral
exposure of SAV to herbicides, such as that following a runoff event, can
induce the same general phytotoxic response and recovery as observed in the
batch microcosms experiments (Figure 12). The rate at which P. perfoliatus
loses previously-incorporated herbicide is currently being examined at
University of Maryland Center for Environmental and Estuarine Studies
(UMCEES) with continuous, subsequent exposure to herbicide-free water.
Apparently, the metabolic recovery shown in Figure 12 involves some sort of
enzymatic detoxification rather than depuration (active or passive removal
via excretion or some other means).
Mode of Uptake—
Jones et al. (1981b) reported that £._ perfoliatus uptake of
•^C-labeled atrazine could occur either through shoot or root pathway,
although shoot uptake appeared to dominate. This finding is generally
consistent with findings of previous investigators. Aldrich and Otto
(1959), for example, reported that ^ pectinatus was equally capable of
550
-------
I
either root or shoot incorporation of 2,4-D-l-C, and Funderburk and •
Lawrence (1963) found that simazine was also taken up through both routes w
by the SAV, Heteranthera dubia, but that other herbicides showed no
root-to-shoot translocation through the stem. Frank and Hodgson (1964) A
also reported uptake of fenac by both roots and shoots of P_._ pectinatus, f
but little or no translocation in either direction. Hence, while some
herbicide uptake by SAV roots is possible, limited ability to translocate —
up the stem reduces the importance of this mechanism for chloroplast-active •
compounds such as atrazine and linuron. In addition, the relatively high '
adsorptive tendency of these herbicides to sediment surfaces may reduce the
herbicide exposure of roots. B
Atrazine uptake by P. perfoliatus occurs hyperbolically, as a function £
of external concentration (Ce) . At Ce less than about 450 ppb internal
plant concentrations of atrazine, G£, were less than Ce, while G£ M
approaches Ce at about 500 ppb. Moreover, as G£ approaches Ce, Pa I
approaches zero. Thus, atrazine incorporation does not follow a strict
Fickian diffusion at low herbicide concentration, although the first-order
process is approximated at Ce<500 ppb (Jones et al. 1981b) . •
It had been postulated that herbicides bound to suspended sediments 9
and/or colloids, which subsequently settled on SAV leaf surfaces,
represented a potential mechanism for magnifying the concentrations to •
which plants are exposed. However, recent experiments have indicated that I
P. perfoliatus shows little or no uptake of atrazine from herbicide-bound
sediments placed on SAV leaves (T. Jones, unpublished data). In addition, _
Correll and Wu (1981) found that atrazine concentrated in surface M
microlayer films exhibited no greater phytotoxicity than the same quantity •
mixed in a large water-volume bathing SAV in microcosm experiments.
Combined Stresses— (
Although it is important to understand the individual effects of
herbicide stress on SAV, many environmental factors act simultaneously on •
the plants in nature. Therefore, the significance of combined effects of •
herbicide-herbicide, herbicide-light, and herbicide-nutrient interactions
must be considered.
Some preliminary investigations of atrazine-linuron combinations with •
P. perfoliatus suggest that 25 ppb of each herbicide reduced Pa more than B
50 ppb of either herbicide combination for the last two post-treatment
weeks. However, there was no difference estimated by the Colby (1967) •
formulation for the first two post-treatment weeks (Cunningham et al. •,
1981b). Additional experiments with herbicide combinations are in
progress, but these results are not yet available. The agricultural _
weed-control literature contains some information on combined herbicide •
effects, but these reports are also inconclusive. For example, Horowitz *
and Herzlinger (1973) found that only one out of seven combination
experiments with diruon, simazine, trifluralin, and fluometurn at 0.1 and W
0.5 ppm exhibited significant synergism. Akobundu et al. (1975) observed 9
synergistic action between atrazine and alachlor, but this nonlinear effect
was small. Appleby and Somabhi (1978) investigated the reported £
antagonisms between atrazine (or simazine) and glyphosate and found that •
the interaction was due more to the physical binding in spray solution than
to any biochemical mechanism. At this point, the importance of
herbicide-herbicide synergisms for SAV toxicity in the Chesapeake Bay •
551
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
region is unclear; however, the effect is probably small.
A consistent pattern of interaction between light and herbicides has
been observed for SAV and other plants. When subjected to 30 percent
shading, the relative response of Z. marina to atrazine treatment (Figure
14d) was markedly reduced as compared to 100 percent light (Hershner et al.
1981). When the effects of atrazine (50 ppb) on P. perfoliatus
photosynthesis were tested over a full range of light intensities, it was
found that the relative toxicity (reduction in Pa) is greater at
less-than-saturating light intensities (T. Jones et al., unpublished
data). Hodgson and Otto (1963) showed that two species of Potamogeton were
more sensitive to contact herbicides at high, rather than low light
intensities, and similar results have been reported for algae (McFarlane et
al. 1972) and weeds (for example, Hammerton 1967). This relationship
probably exists, because the more active chloroplasts operating at high
light levels are more susceptible to herbicidal damage. In a related
experiment, we observed that epiphytic sediments significantly reduced
atrazine uptake by E^ perfoliatus leaves (T. Jones et al., unpublished
data). This effect may be attributable to a combination of physical
buffering and sorption by epiphytic sediments, as well as the
light-herbicide relationship mentioned above.
Under conditions of nutrient sufficiency, which occur for SAV
throughout most of the upper and middle Bay, nutrient additions would be
expected to show little effect on herbicide phytotoxicity. This conclusion
was reached after a series of recent experiments. However, SAV grown for
months in microcosms can experience nutrient (or 002) limitation (Kemp et
al. 1980) and thus provide a simple way of addressing the question of
nutrient limitation. In Figure 18, the results of four different atrazine-
Potamogeton experiments are summarized in a manner that may reveal such a
relationship. A negative trend was found between herbicide toxicity at 25
ppb and maximum Pa at full incubator light intensity (150 uE
m~2s~l). It might be inferred that nutrient limitation (or some other
environmental stress reducing peak Pa) increases herbicidal action on
SAV. The potential effect of nutrient-induced epifloral growth attached to
SAV on herbicide stress has not been examined, but it might be expected to
act in much the same fashion as did epiphytic sediments.
Metabolites of Herbicides—
One herbicide-SAV issue that has yet to be addressed in this research
is the potential toxicity of herbicide metabolites, or degradation
products. In a previous Section, it was shown that the dealkylated
daughter products of atrazine degradation occur at persistently low levels
« 10 percent of original atrazine) in soils, sediments, and water.
Various investigators have shown that the dealkylated degradation products
of atrazine retain some toxicity (although considerably less than the
parent compound) to terrestrial plants (Shimabukuro 1968, Kaufman and Blake
1970). The carry-over toxicity, which has been reported for atrazine-
treated fields from one year to the next, has been attributed by some to
the persistence of the N-de-ethylated metabolites. Both Sirons et al.
(1973) and Dao et al. (1979), for example, have reported carry-over
toxicity after atrazine application to croplands for as long as two years.
At present, the toxicity to SAV from metabolites of atrazine and other
herbicides is not known, nor are the levels of these compounds in the
552
-------
UJ
z
N
cr
H
<
Q.
O.
CM
a
a.
u.
O
O
O
a
UJ
-------
I
•environment. Although the potential for metabolite build-up in estuarine
sediments may appear remote, the issue remains the one major gap in our
understanding of the overall herbicide-SAV issue.
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
554
-------
SECTION 6
SUMMARY AND IMPLICATIONS
SUMMARY OF RESEARCH FINDINGS
I
I
I
In this paper we have highlighted the results of extensive research B
supported by EPA-CBP to investigate the behavior of agricultural herbicides •
in an estuarine environment, particularly in relation to Chesapeake Bay's
dwindling communities of submerged aquatic vegetation. The relative, •
potential importance of various herbicidal compounds in relation to SAV was I
considered, and atrazine and linuron were selected for primary focus. The
watercourses in this coastal region have been systematically sampled for _
herbicide concentration from the mainstern Chesapeake Bay, to primary I
tributaries, to secondary bays and coves, to creeks that drain agricultural V
fields. Maximum observed concentrations of these two major herbicides in
the four levels of tributaries were about four ppb, 7.0 ppb, 20 ppb, and •
100 ppb, respectively. High herbicide concentrations of about 10 to 20 ppb £
were observed to occur in estuarine waters for ephemeral periods of two to
eight hours. The length of time between herbicide application to the —
cropland and the first rainfall-runoff event, and the extent and intensity •
of rainfall, are key factors governing the transport of herbicides from the *
field to the estuary.
The degradation of atrazine in estuarine environments appears to occur •
far more rapidly than in agricultural soils, with half-lives of two to 26 (|
weeks, respectively. The longevity of linuron is less than that of
atrazine and, in fact the latter compound appears to be one of the most m
persistent herbicides used in the watershed. Atrazine exhibits moderate I
tendency for adsorption to soils and estuarine sediments. Most of the
herbicide running off from field to watercourse does so in the dissolved
form, rather than bound to soil particles, and most of the atrazine in the •
estuary is similarly found in the dissolved state. Estuarine colloids have V
about 10 times greater ability to bind atrazine than do sediments and
soils. Salinity and circum-neutral pH appear to exert little influence on
herbicide—sediment sorption; however, increased salinity does tend to
decrease the proportion of herbicide bound to colloidal matter.
Atrazine brings about a dramatic stress response for several species of ^
SAV at concentrations of 50 to 100 ppb. At these concentrations, •
reductions of photosynthesis are always significant, and full recovery of
photosynthetic rates may not be attained. The relation between percent
loss of photosynthesis and herbicide concentration generally follows a •
semilogarithmic function for all species and both compounds tested. This •
model predicts threshold toxicities at herbicide concentrations ranging
from 1.0 to 7.0 ppb. Combining all experimental data for three species and A
two herbicides yielded a highly significant regression (r^ = 0.89), which £
predicts about 10 to 20 percent loss of SAV photosynthesis at 5.0 to 10.0
ppb herbicide. Similar herbicidal effects were observed for plant _
structural characteristics. •
Experiments with P. perfoliatus and atrazine indicate that herbicide "
uptake, which is a function of external concentration, proceeds to
equilibrium within one hour, and that depuration (loss of herbicide) upon •
I
555 g
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
exposure to clean water occurs very slowly, with toxic effects still
apparent after days of cleasing. Reduced light level (above compensation
light) and/or presence of epiphytic sediments appear to decrease the
relative stress effect of herbicide on SAV, but nutrient deficiency and
plant senescence may tend to increase herbicide effects. There are
currently few data to support the hypothesis that combinations of two or
more herbicides act in any other than an additive fashion. The potential
toxicity of herbicide metabolites (degradation products) to SAV is a matter
about which very little is known.
DID HERBICIDES CAUSE THE SAV DECLINE?
From the evidence that has been compiled here, the answer to this
question is most likely no. Herbicide concentrations in excess of 20 ppb
were not found in estuarine waters in various surveys since 1977, under a
range of situations, including those which approach worst-case runoff
conditions. Under such extreme conditions, concentrations of 10 to 20 ppb
were observed, but rarely lasted more than four to eight hours. Although
exposures to 20 ppb (of one hour or more) will cause significant loss of
productivity, full metabolic recovery would be expected within one to four
weeks following initial contact. Moreover, herbicides degrade rapidly in
the estuarine environment, with half-lives measured in days and weeks;
residual concentrations do not appear to build up in sediments. The
hypothesized mechanisms of increasing SAV exposure to herbicides via
concentration of the compounds in epiphytic sediments or surface-layer
films do not appear to represent significant factors. One of the caveates
that remains unresolved is the fact that very little is known about
estuarine concentrations and SAV toxicities of major herbicide
metabolites. The de-ethylated daughter products of atrazine degradation do
tend to persist for months under estuarine conditions, and weed-control
literature attributes carryover toxicity (after atrazine application) to
this metabolite.
ARE HERBICIDES A PROBLEM?
Ephemeral herbicide concentrations in excess of 5.0 ppb do occur
periodically in some estuarine water that once contained extensive SAV
beds. In general, such concentrations appear to cause losses in SAV
productivity of 10 to 20 percent, even when exposures are brief (about an
hour), and recovery may take days to weeks. The effects of repeated,
brief exposures to such concentrations are not known. A reasonable
assumption, however, would be that if the time interval between runoff
events (which might yield such deleterious concentrations) is greater than
SAV recovery time, then the partial loss of photosynthesis may persist.
Such reductions in SAV productivity will definitely add to the generally
stressed conditions that these plants currently experience in the estuary.
The sources of most of these stresses include such factors as salinity
extremes, waterfowl grazing, uprooting by cownose rays, and turbulent
waters or violent wave action caused by major storm events, as well as
water-column turbidity and the accumulation of epiphytic materials.
Herbicide-induced loss of productivity (though minor) could act in concert
with many of these stressors to create intolerable conditions for SAV
existence.
556
-------
1
1
1
1
**•»
1
CO
UJ
>-
I30
1
2
0
UJ
o
a:
UJ
2 20
3
CO
fc
B
2
UJ
cc
3 10
8
i-
UJ
u
o:
Stf
• EASTERN BAY
\
\
\
\
\
\
\
Nx »CHOPTANK RIVER
\
\
CHESTER *x
RIVER x
\
\
^ Y* 23.7- 26. 4X
\ (ra-0.91)
\
\
\
V
^
\
\
\
\
PATUXENT \
RIVER \
• V
\
WICOMICO Nv NANTICOKE
RIVER • x RIVER •
i i
1
1
1
1
1
I
1
1
1
1 10 "•
POTENTIAL DIFFUSE LOADING, m2 m3 •
1
Figure 19. Correlation of potential diffuse loadings and percent occurrence
of SAV in 1974 (Stevenson and Confer 1978). •
1
1
557 m
1
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
The source of herbicidal compounds to Chesapeake Bay is agricultural
runoff. There appears to be a relationship between potential loadings
(that is, watershed area divided by estuarine volume) of nonpoint, or
diffuse-source materials (including herbicides, nutrients, and sediments),
and SAV abundance in six major tributaries (Figure 19). This correlation
suggests that the greater the loadings from runoff, the more extensive the
decline has been.
Although the development of recommended farming practices is well
beyond the scope of this research endeavor, it is hoped that these research
results and their environmental implications will be considered by the
agricultural community in the evolution of improved farming approaches.
The importance of agriculture in the socio-economic milieu of the
Chesapeake region is unquestionable. Our recommendation is simply that the
estuarine resource values be considered in concert with the land-based
values to develop balanced patterns of human enterprise.
558
-------
LITERATURE CITED
Report. UMCEES Ref. No. 78-136. Horn Point Environmental Labs,
Cambridge, MD. 41 pp.
Colby, S.R. 1967. Calculating Synergistic and Antagonistic Responses of
Herbicide Combinations. Weeds. 15:20-22.
559
I
I
Akobundu, 1.0., R.D. Sweet, W.B. Duke, P.L. Minotti. 1975. Weed Response
to Atrazine and Alachlor Combinations at Low Rates. Weed Sci. 23:67-70. •
Aldrich, F.D., and N.E. Otto. 1959. The Translocation of 2,4-D-l-C in
Potamogeton pectinatus, a Submerged Aquatic Vegetation. Weeds. _
7:295-299. •
Appleby, A.P., and M. Somabhi. 1978. Antagonistic Effect of Atrazine and
Simazine on Glyphosate Activity. Weed Sci. 26:135-139. •
Armstrong, D.E., C. Chesters, and R.F. Harris. 1967. Atrazine Hydrolysis
in Soil. Soil Sci. Amer., Proc. 31:61-66. •
Austin, J.J. , R.C. Bubeck, and T. Munson. 1978. Monitoring of the Upper
Chesapeake Bay for Symmetrical Triazine Herbicides and Simazine, U.S.
Environmental Protection Agency, Region III Central Laboratory, •
Annapolis, MD. (Unpubl. Ms.). V
Bailey, G.W., and J.L. White. 1970. Review of Adsorption and Desorption •
of Organic Pesticides by Soil Colloids with Implications Concerning •
Pesticide Bioactivity. J. Agric. Food Chem. 12:324-332.
Best, J.A., and J.B. Weber. 1974. Disappearance of S-triazines as •
Affected by Soil pH Using a Balance-Sheet Approach. Weed Sci. ™
22:364-373.
Breisch, L.L., and W.M. Kemp. 1978. Nitrogen and Phosphorus Sources and 0
Water Quality Characteristics of the Choptank River Estuary. Student
I
Brewer, P.E., C.J. Arntzen, and F.W. Slife. 1979. Effects of Atrazine,
Cyanizine, and Procyazine on the Photochemical Reactions of Isolated •
Chloroplasts. Weed Sci. 27:300-308. •
Bryfogle, B.M, and W.F. McDiffett. 1979. Algal Succession in Laboratory •
Microcosm as Affected by an Herbicide Stress. Amer. Midi. Nat. |
101:344-354.
Butler, P. 1965. Effects of Herbicides on Estuarine Fauna. So. Weed •
Cont. Conf. 18:576-580.
Colbert, P.O., V.V. Volk, and A. P. Appleby. 1975. Sorption of Atrazine, B
Terbutryn and GS-14254 on Natural and Lime-Amended Soils. Weed Sci. •
23:390-394.
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Correll, D.L., J.W. Pierce, and T.L. Wu. 1978. Herbicides and Submerged
Plants in Chesapeake Bay. IN: Amer. Soc. Civil Eng. (ed.) Coastal
Zone-78. pp. 858-877.
Correll, D.L. and T.L. Wu. 1981. Atrazine Toxicity to Submersed Vascular
Plants in Simulated Estuarine Microcosms. Aquatic Botany (In review).
Cunningham, J.J. , W.M. Kemp, J.C. Stevenson, and M.R. Lewis. 1981a.
Effects of Herbicide Stress on the Structure and Metabolism of the
Submerged Macrophyte, Potamogeton perfoliatus, in Estuarine
Microcosms. Aquatic Botany.(Unpublished manuscript).
Cunningham, J.J. , W.M. Kemp, J.C. Stevenson, W.R. Boynton, and J.C. Means.
1981b. Stress Effects of Agricultural Herbicides on Submerged
Macrophytes in Estuarine Microcosms, In: Submerged Aquatic Vegetation
in Chesapeake Bay. Annual Kept, to U.S. EPA. W.M. Kemp et al. eds.
UMCEES, Horn Point Environ. Labs., Cambridge, MD. pp. 147-182.
Dao, T.H., and T.L. Lavy. 1978. Atrazine Adsorption on Soil as Influenced
by Temperature, Moisture Content and Electrolyte Concentration. Weed
Sci. 26:303-308.
Dao, T.H. , T.L. Lavy, and R.C. Sorensen. 1979. Atrazine Degradation and
Residue Distribution in Soil. Soil Sci. Soc. Am. J. 43:1129-1134.
Davis, D.E., J.D. Weete, C.G.P. Pillai, F.G. Plumley, J.T. McEnerney, J.W.
Everest, B. Truelove, A.M. Diner. 1979. Atrazine Fate and Effects in
a Salt Marsh. U.S. Environmental Protection Agency Research and
Development Report. (EPA-600/3-79-111). National Tech. Inform. Serv.,
Springfield, VA. 84 pp.
Ebert, E., and S.W. Dumford. 1976. Effects of Triazine Herbicides on the
Physiology of Plants. Residue Rev. 65:1-103.
Eichers, T.R., P.A. Andrilenas, and T.W. Anderson. 1978. Farmers' Use of
Pesticides in 1976. Agric. Econ. Rep. No. 418, U.S. Dept. Agric.,
Washington, DC. 58 pp.
Fowler, M.C. 1977. Laboratory Trials of a New Triazine Herbicide
(DPX3674) on Various Aquatic Species of Macrophytes and Algae. Weed
Res. 17:191-195.
Frank, P.A. , and R.H. Hodgson. 1964. A Technique for Studying Adsorption
and Translocation in Submersed Plants. Weeds. 12:80-82.
Frank, R., and G.J. Sirons. 1979. Atrazine: Its Use in Corn Production
and its Loss to Stream Waters in Southern Ontario, 1975-1977. The
Science of the Total Environment. 12:223-239.
Funderburk, H.H., and J.M. Lawrence. 1963. Adsorption and Translocation
of Radioactive Herbicides in Submersed and Emerged Aquatic Weeds. Weed
Res. 3:304-311.
560
-------
I
Giles, C.H. , T.H. MacEwan, S.N. Nakhwa, D. Smith. 1960. Studies in •
Adsorption. Part XI. A System of Classification of Solution £
Adsorption Isotherms. J. Chem. Soc. :3973.
Green, R.E., and S.R. Obien. 1969. Herbicide Equilibrium in Soils in •
Relation to Soil Water Content. Weed Sci. 17:514-519-
Grover, R., and R.J. Hance. 1970. Effect of Ratio of Soil to Water on I
Adsorption of Linuron and Atrazine. Soil Sci. 109:136-138.
Gysin, H., and E. Knusli. 1960. Chemistry and Herbicidal Properties of
Triazine Derivatives. Adv. Pest Control Res. 3:289-358.
Hall, J.K. 1974. Erosional Losses of s-Triazine Herbicides. J. Environ. _
Qual. 3:174-180. •
Hamaker, J.W. 1972. Organic Chemicals in the Soil Environment. Vol. I
C.A. Goring and J.W. Hamaker, eds. Marcel Dekker, NY. 1:253 pp. •
Hammerton, J.L. 1967. Environmental Factors and Susceptibility to
Herbicides. Weeds. 15:330-336. _
Hance, R.J. 1979. Effect of pH on the Degradation of Atrazine, ™
Dichlorprop, Linuron and Propyzamide in Soil. Pestic. Sci. 10:83-86.
Harris, C.I., and G.F. Warren. 1967. Adsorption and Desorption of |
Herbicides by Soils. Weed Sci. 15:120-126.
Heinle, D.R., J.L. Taft, C.F. D'Elia, J.S. Wilson, M.Cole-Jones, and A.B. I
Vivian. 1980. Historical Review of Water Quality and Climatic Data in
Chesapeake Bay. Publ. Mo. 84, Chesapeake Res. Consortium, Annapolis, _
MD. •
Hershner, C., K. Ward, and J. Illowsky. 1981. The Effects of Atrazine on
Zostera marina in the Chesapeake Bay, Virginia. (Rept. to US EPA). •
Virginia Instit. Marine Sci., Gloucester Pt., VA. |
Hess, F.D. 1980. A Chlamydomonas Algal Bioassay for Detecting Growth p
Inhibitor Herbicides. Jour, for Weed Sci. 28:515-520. •
Hiltbold, A.E., and G.A. Buchanan. 1977. Influence of Soil pH on
Persistance of Atrazine in the Field. Weed Sci. 25:515-520. B
Hodgson, R.H., and N.E. Otto. 1963. Pondweed Growth and Response to
Herbicides under Controlled Light and Temperature. Weeds. 11:232-237. •
Hermann, W.D., J.C. Tournayre, and H. Egli. 1979. Triazine Herbicides
Residues in Central European Streams. Pesticides Monitoring Journal. _
13:128-131. •
Horowitz, M., and G. Herzlinger. 1973. Interactions Between Residual
Herbicides at Low Concentrations. Weed Res. 13:367-372. •
561
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Hurle, K.B., and V.H. Freed. 1972. Effect of Electrolytes on the
Solubility of Some 1, 3, 9-Trazines and Substituted Ureas and Their
Adsorption on Soil. Weed Res. 12:1-10.
Jones, T.W. , J.C. Means, J.C. Stevenson, and W.M. Kemp. 1981a. Uptake and
Phytotoxicity of Atrazine in Potamogeton perfoliatus, In: Submerged
Aquatic Vegetation in Chesapeake Bay. Kept, to US EPA. W.M. Kemp et
al. eds. UMCEES, Horn Point Environ. Labs. Cambridge, pp. 208-230
Jones, T.W., W.M. Kemp, J.C. Stevenson, and J.C. Means. 1981b.
Degradation of Atrazine in Estuarine Water-Sediment Systems and
Selected Soils. J. Environ. Qual. (In review).
Jordan, L.S., B.E. Day, and W.A. Clery. 1964. Photodecomposition of
Triazines. Weeds. 12:5-6.
Karickhoff, S.W., D.S. Brown, and T.A. Scott. 1979. Sorption of
Hydrophobic Pollutants on Natural Sediments. Water Res. 12:241-249.
Kaufman, D.D., and J. Blake. 1970. Degradation of Atrazine by Soil
Fungi. Soil. Biol. Biochem. 2:73-80.
Kells, J.J. , C.E. Rieck, R.L. Blevins, and W.M. Muir. 1980. Atrazine
Dissipation as Affected by Surface pH and Tillage. Weed Sci.
28:101-104.
Kemp, W.M., M.L. Lewis, J.J. Cunningham, J.C. Stevenson, and W.R. Boynton.
1980. Microcosms, Macrophytes, and Hierarchies: Environmental
Research in Chesapeake Bay, In: Microcosm Research in Ecology. J.
Giesy, ed. ERDA Conf. 781101. pp. 911-936 .
Kempson-Jones, G.F., and R.J. Hance. 1979. Kinetics of Linuron and
Metribuzin Degradation in Soil. Pestic. Sci. 10:449-454.
Kratky, B.A., and G.F. Warren. 1971. The Use of Three Simple Rapid
Bioassays on Forty-Two Herbicides. Weed Res. 11:257-262.
Langdale, G.W., A.P. Barnett, R.A. Leonard, and W.G. Fleming. 1979.
Reduction in Soil Erosion by the No-Till System in the Southern
Piedmont. Trans. M. Soc. Acric. Eng. 22:223-278.
Lamoureux, G.L., R.H. Shimabukuro, P.H. Swanson, and H.R. Frear. 1970.
Metabolism of Atrazine in Excised Sorghum Leaf Sections. J. Agric.
Food. Chem. 18:81-86.
Loeppky, C., and B.C. Tweedy. 1979. Effects of Selected Herbicides Upon
Growth of Soil Algae. Weed Sci. 17:110-113.
MacFarlane, R.B., W.A. Glooschenko, and R.C. Harriss. 1972. The
Interaction of Light Intensity and DDT Concentration Upon the Marine
Diatom, Nitzschia delicatissima. Hydrogiologia. 39:373-382.
562
-------
I
McGlammery, M.D., and F.W. Slife. 1966. The Adsorption and Desorption of •
Atrazine as Affected by pH, Temperature and Concentration. Weeds. |
14:237-239.
Means, J.C. , J.J. Hassett, S.G. Wood, and W.L. Banwart. 1979. Sorption •
Properties of Energy-Related Pollutants and Sediments. In:
Polynuclear Aromatic Hydrocarbons. P.W. Jones and P. Leber, eds. Ann
Arbor Sci. Pub. Ann Arbor, MI. pp. 327-340.. I
Means, J.C. , S.G. Wood, J.J. Hassett, and W.L. Banwart. 1980. Sorption of
Polynuclear Aromatic Hydrocarbons on Sediments. Envir. Sci. Techn. •
14:1524-1528. g
Means, J.C. , T.W. Jones, T.S. Pait, and R.D. Wijayaratne. 1981a.
Adsorption of Atrazine on Chesapeake Bay Sediments and Selected Soils, I
In: Submerged Aquatic Vegetation in Chesapeake Bay. Annual Rept. to *
U.S. EPA. W.M. Kemp et al., eds. UMCEES Horn Point Environ. Labs,
Cambridge, MD. pp. 284-296. •
Means, J.C., J.C. Stevenson, W.R. Boynton, and W.M. Kemp. 1981b.
Herbicides in Maryland Chesapeake Bay. A Listing of Concentrations «
Measured in 1980-81. Unpublished data. Ches. Biol. Lab. Solomons, MD. •
Means, J.C., and R. Wijayaratne. 1981. Role of Natural Colloids in
Transport of Hydrophobic Pollutants. Science. (In Press). •
Metz, J.J., M.R. Lewis, and R. Galloway. 1979. Effect of Atrazine on Two
Algal Isolates. In: Submerged Aquatic Vegetation in Chesapeake Bay. •
Annual Rept. to U.S. EPA. W.M. Kemp, J.C. Stevenson, and W.R. Boynton |
eds. UMCEES Horn Point Environ. Labs, Cambridge, MD.
Moreland, D.E., and J.L. Hilton. 1976. Actions on Photosynthetic Systems, I
In: Herbicides: Physiology, Biochemistry, Ecology. L.J. Audus, ed. ™
Academic Press, NY. 1:493-523.
Mrak, E.M., ed. 1974. Herbicide Report. Hazardous Materials Advis. •
Comm. EPA-SAB-74-001. 196 pp.
Newby, L.C. , R.A. Kahars, K. Adams, and M. Szolics. 1978. Atrazine I
Residues in the Chesapeake Bay. Unpubl. MS, Ciba-Geigy Corp.,
Greensboro, NC. _
Pfister, K., S.R. Radosevich, and C.J. Arntzen. 1979. Modification of *
Herbicide Binding to Photosystem II in Two Biotypes of Seneci vulgaris
L. Plant Physiol. 64:995-999. •
Pillai, C.G.P. , J.D. Weete, and D.E. Davis. 1977. Metabolism of Atrazine
by Spartina alterniflora. I. Chloroform-Soluble Metabolites. J. •
Agric. Food Chem. 25:852-856 I
Pillai, P., J.D. Weete, A.M. Diner, and D.E. Davis. 1979. Atrazine
Metabolism in Box Crabs. J. Environ. Qual. 8:277-280. •
I
563 _
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Pruss, S.W., and E.R. Higgins. 1974. Effects of Low Levels of Simazine on
Plankton Algae in a Non-Stratified Lake. Proc. N.E. Weed Contr. Counc.
28:124-131.
Rao, P.S.C. et al. 1981. EPA Rept. (in press).
Richard, J.J., G.A. Junk, M.J. Avery, N.L. Nehring, J.S. Fritz, and H.J.
Svec. 1975. Analysis of Various Iowa Waters for Selected Pesticides:
Atrazine, DDE and Dieldrin-1974. Pesticides Monitor J. 9:117-123.
Ritter, W.S., H.P. Johnson, W.G. Lovely, and M. Molnau. 1974. Atrazine,
Propachlor, and Diazinon Residues on Small Agricultural Watersheds.
Environ. Sci. Technol. 8:38-42.
Schueler, T.C. 1979. Summary of Runoff and Loading Coefficients of
Diffuse Source Pollutants: Herbicides, Appendix Fl-14, In: Submerged
Aquatic Vegetation in Chesapeake Bay. Annual Rept. to U.S. EPA. W.M.
Kemp, J.C. Stevenson, and W.R. Boynton, eds. UMCEES Horn Point
Environ. Labs, Cambridge, MD. .
Shimabukuro, R.H. 1968. Atrazine Metabolism in Resistant Corn and
Sorghum. Plant Physiol. 43:1925-1930.
Sirons, G.R., R. Frank, and T. Sawyer. 1973. Residues of Atrazine
Cyanazine and Their Phytotoxic Metabolites in a Clay Loam Soil. J.
Agric. Food Chem. 21:1016-1020.
Smith, C.N., R.A. Leonard, G.W. Langdale, and G.W. Bailey. 1978.
Transport of Agricultural Chemicals from Small Upland Piedmont
Watersheds. U.S. EPA, Res. Rep. Ser. EPA-600/3-78-056. 364 pp.
Smith, G.E., F.D. Whitaker, and H.G. Heineman. 1974. Losses of Fertilizer
and Pesticides from Claypan Soils. Environmental Protection Technology
Series EPA-600/2-74-068. Washington DC.
Stevens, G.A., J.W. Wysong, and B .V. Lessley. 1981. Farm Data Manual.
Cooper. Exten. Service Publ., Dept. Agric. Res. Econ. Univ. Maryland,
College Park. AREIS 20. 117 pp.
Stevenson, J.C. , and N.M. Confer. 1978. Summary of Available Information
of Chesapeake Bay Submerged Aquatic Vegetation. U.S. Dept. Inter.
FWS/OBS-78/66. NTIS, Springfield, VA. 333 pp.
Stevenson, J.C., T.W. Jones, W.M. Kemp, W.R. Boynton, and J.C. Means.
1981. An Overview of Atrazine Dynamics in Estuarine Ecosystems. In:
Agrichemicals and Estuarine Productivity. J.D. Costlow, L.E. Cronin,
T.B. Duke, and W. McClellan, eds. John Wiley Publ. (In press).
Stewart, B.A., D.A. Woolhiser, W.H. Wischmeier, J.H. Cars, M.H. Frere, J.R.
Schaub, L.M. Boone, K.F. Alt, S.L. Horner, and H.R. Cosper. 1975.
Control of Water Pollution from Croplands. USDA Rept. ARS-H-5-1, U.S.
EPA Rept. EPA-600/2-75-026a. NTIS, Springfield, VA. 1:111 pp.
564
-------
I
Sutton, D.L., D.A. Durham, S.W. Bingham, C.L. Foy. 1969. Influence of tt
Simazine on Apparent Photosynthesis of Aquatic Plants and Herbicide •
Residue Removal from Water. Weed Sci. 17:56-59.
Swanson, R.A., and G.R. Dutt. 1973. Chemical and Physical Processes that I
Affect Atrazine and Distribution in Soil Systems. Soil Sci. Soc. Amer.
Proc. 37:46-52.
Talbert, R.E., and O.H. Fletchall. 1965. The Adsorption of Some •
s-triazines in Soils. Weeds. 13:46-52.
Travis, C.C. , and E.L. Etnier. 1981. A Survey of Sorption Relationships p
for Reactive Solutes in Soil. J. Environ. Qual. 10:8-17.
Triplett, G.B., B.J. Conner, and W.M. Edwards. 1978. Transport of I
Atrazine and Simazine in Runoff from Conventional and No-Tillage corn. *
J. Environ. Qual. 7:77-84.
Truhlar, J.F., and L.A. Reed. 1976. Occurence of Pesticide Residues in m
Four Streams Draining Different Land-Use Areas in Pennsylvania,
1969-71. Pesticide Monitor. J. 10:101-110. g
Walker, C.R 1964. Simazine and Other s-triazines Compounds as Aquatic
Herbicides in Fish Habitats. Weeds. 12:134-139.
Ward, J. 1980. Atrazine in the Susquehanna River and its Tributaries. •
U.S. Geological Survey, Washington, DC. Unpublished data.
Wauchope, R.D. 1978. The Pesticide Content of Surface Water Draining from p
Agricultural Fields - A Review. J. Environ. Qual. 7:459-472.
Wauchope, R.D., and R.A. Leonard. 1980. Maximum Pesticide Concentrations •
in Agricultural Runoff: A Semi-Empirical Formula. J. Environ. Qual. ™
9:665-672.
Weaver, L.O., O.D. Morgan, and J.G. Kantzer. 1975. Pest Control - 1975 |
Recommendations. Field Crops. Cooperative Exten. Serv. Univ. Md.,
College Park, Bull. 237, 32 pp. •
Weber, J.B., S.D. Weed, and T.M. Ward. 1969. Adsorption of s-triazine by
Soil Organic Matter. Weed Sci. 17:417-421.
Weed Science Society of America. 1980. Herbicide Handbook. 35th •
ed. WSSA. Champaign, IL. 450 pp.
White, A.W., A.D. Barnett, B.C. Wright, and J.H. Holladay. 1967. Atrazine |
Losses from Fallow Land Caused by Runoff and Erosion. Environ. Sci.
Technol. 1:740-744. —
I
I
565 -
-------
Wu, T.L., N.J. Mick, and B.M. Fox. 1977. Runoff Studies of the
Agricultural Herbicides Alachlor and Atrazine from the Rhode River
Watershed during the 1976 Growing Season. In: Watershed Research in
Eastern North America. D.L. Correll, ed. Smithsonian Institute Press,
Washington, DC. pp. 707-724.
Wu, T.L., L. Lambert, D. Hastings, and D. Banning. 1980. Enrichment of
the Agricultural Herbicide Atrazine in the Microsurface Water of an
Estuary. Bull. Environ. Contam. Toxicol. 24:411-414.
Zahnow, E.W., and J.D. Riggleman. 1980. Search for Linuron Residues in
Tributaries of the Chesapeake Bay. J. Agric. Food Chem. 28:974-978.
566
-------
LIGHT AND SUBMERGED MACROPHYTE COMMUNITIES IN
CHESAPEAKE BAY: A SCIENTIFIC SUMMARY
I
I
I
I
I
I
I
by |
I
Richard L. Wetzel, Robin F. van Tine, and Polly A. Penhale
Virginia Institute of Marine Science M
and School of Marine Science College of William and Mary •
Gloucester Point, Virginia 23062 *
I
I
I
I
I
I
I
I
567 •
-------
I
I
I
I
I
I
I
3. Light and Photosynthesis in Chesapeake Bay SAV Communities. 599
I General Review of Photosynthesis 599
Photosynthesis of Submerged Vascular Plants in Relation
to Light and Temperature 600
_ Photosynthesis-Light Studies in Chesapeake Bay 603
• P-I Relationship of Major Species 603
* Microcosm Studies 608
In Situ Studies of Community Response to Light .... 608
I
CONTENTS
Page
Figures 567
Tables 571
1. Introduction 572
Background 572
The Research Program on Light and SAV: An Overview . . . 574
2. Light in Chesapeake Bay 577
General Characteristics of Estuarine Optical Properties . 577
Light Attenuation in Chesapeake Bay 582
Comparison of Light Attenuation in Vegetated and
Unvegetated Sites in the Bay 584
Historical Data Bases and Optical Properties of the
Chesapeake Bay 587
4. Summary 620
• Literature Cited 623
Summary and Conclusions 632
I
I
I
I
I
I
I
_ 568
-------
FIGURES
3 Diffuse downwelling spectral attenuation coefficients for
Chesapeake Bay 583
4 Lower Chesapeake Bay light study stations 585
5 Diffuse downwelling spectral attenuation coefficients for
vegetated and unvegetated sites 535
I
I
Number Page
1 Path of light from the atmosphere to benthic estuarine •
macrophytes 579
2 Downwelling spectral quanta irradiance in a Zostera bed 580 •
I
6 Downwelling PAR attenuation coefficients for vegetated and •
unvegetated sites 588 •
7 Historical Chesapeake Bay Secchi disk values 589 m
8 Summary of the historical chlorophyll a data for the upper and
lower Chesapeake Bay 591
9 Historical chlorophyll a data for three regions of •
Chesapeake Bay 592
10 Enriched areas of Chesapeake Bay 597 |
11 Diagramatic photosynthesis-light curves 601 M
12 Photosynthesis-light curves for two upper Chesapeake Bay
species 604
13 Photosynthesis-light curves for two lower Chesapeake Bay •
species 605
14 Vertical distribution of leaf area index for Ruppia and |
Zostera 609
15 Total chlorophyll in Ruppia and Zostera leaves 610 •
16 Suspended solids, light availability, and Potamogeton
photosynthesis 611 I
17 The effect of light flux on upper Chesapeake Bay SAV
photosynthesis 612 •
18 Diagramatic representation of light flux and calculated
photosynthetic parameters for an upper Chesapeake Bay site . . 614
569
I
I
-------
I
FIGURES (Continued)
Number Page
* 19 Apparent productivity and light flux at a Ruppia site ...... 615
I 20 Apparent productivity and light flux at a Zostera site ..... 616
21 Apparent productivity versus light flux for three sites in the
• lower Chesapeake Bay ..................... 617
I
I
I
I
I
I
I
I
I
I
I
I
I
570
-------
TABLES
Number Page
1. Comparison of PAR Attenuation Coefficients inside and
outside an SAV bed ...................... 590
I
I
2. Secchi Disc Data, upper Chesapeake Bay ............ 593 I
3. Chlorophyll a Concentrations in the lower Potomac River .... 594
4. Freshwater Flows and Hurricanes in Chesapeake Bay ....... 595 m
5. Suspended Sediment Transport in the Susquehanna River ..... 596 •
6. Photosynthetic Parameters for Ruppia and Zostera ....... 606
7. Literature Review of Photosynthesis-Light Experiments ..... 607 •
8. In situ Oxygen Productivity and Light Experiments ... ..... 618
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
SECTION I
INTRODUCTION
The initial focus of submerged aquatic vegetation (SAV) research in the
U.S. Environmental Protection Agency (EPA), Chesapeake Bay Program (CBP)
was evaluation of the structural and functional ecology of these
communities. In the upper Bay, Myriophyllum spicatum and Potamogeton
perfoliatus are the dominant species; the dominant species in the lower Bay
are Zostera marina and Ruppia maritima. Studies centered on various
aspects of productivity (both primary and secondary), trophic structure,
and resource utilization by both ecologically and economically important
species. Much of the initial research was descriptively oriented because
of a general lack of information on Chesapeake Bay submerged plant
communities. These investigations created the data base necessary for the
development of ecologically realistic simulation models of the ecosystem.
Following these initial studies, the research programs in both Maryland and
Virginia evolved toward more- detailed analyses of specific factors that
potentially limit or control plant growth and productivity. Previous
results indicated certain environmental parameters and biological processes
that possibly limited and controlled SAV distribution and abundance.
Specifically, these included light, nutrients, herbicides and fouling
(epibiotic growth). Laboratory and field studies were devoted in the later
phases of the CBP-SAV program toward investigating these interactions.
This work is among the first studies in North America to investigate light
quality as a major environmental factor affecting the survival of sea
grasses.
The overall objectives of this later work were to evaluate more
precisely environmental and biological factors in relation to submerged
aquatic plant community structure and function. Both the published
literature and the results of CBP-SAV program studies indicate that the
interaction of these environmental parameters, together with other physical
and biological characteristics of the ecosystem, determine the longer term
success or failure of SAV communities (den Hartog 1970, den Hartog and
Polderman 1975, Williams 1977, Wetzel et al. 1982).
BACKGROUND
A major goal of CBP-SAV research was to investigate the response of Bay
grasses to various environmental variables. Studies centered on the four
dominant submerged aquatics in the Bay. Understanding the relationship
between environmental factors and the productivity and growth of SAV was
determined to be the first step necessary in attaining the overall goals of
the management program. Natural and man-made changes in environmental
quality may favor one species or another, or result in alteration of the
entire community. The basic responses of the grasses, as well as the
entire community, must be determined before environmental change can be
evaluated in terms of specific management criteria.
Studies in the various CBP-SAV research programs that address
environmental regulation and control of SAV communities focused on nutrient
regulation [primarily nitrogen as ammonium (NHj) and nitrate
572
-------
I
light and photosynthesis, and other biological and •
physical-chemical factors influencing light energy distribution. •
The results of studies in the lower Bay communities suggest a net
positive response to short-term nutrient additions and support the •
observation by others that these communities are nutrient limited (Orth I
1977). The most consistent positive response is associated with Ruppia
dominated communities, and the most variable is associated with the deeper —
Zostera community (Wetzel et al. 1979). In contrast, Kemp et al. (1981b) •
observed that upper-Bay SAV communities did not appear nutrient limited, ™
but were perhaps limited by suboptimal light conditions. These results,
together with community metabolism studies, suggest that light and the H
environmental factors controlling available light are key factors governing |
plant community growth and productivity. Light-temperature-turbidity
regimes and their interaction may explain, in large part, observed •
variability in distribution and abundance. Changes in these parameters, •
governed by either natural or man-induced events and, perhaps, determined
over longer time scales, influence variation in distribution and abundance
in Chesapeake Bay ecosystem as a whole. •
Throughout Chesapeake Bay, submerged aquatic plant communities exhibit •
a distinct zonation pattern from the shallower inshore high-light area to
the deeper, low-light area of the beds. These characteristic distribution •
patterns also suggest different physiological responses to and control by I
local environmental conditions, principally light.
Studies were initiated in August, 1979, on lower-Bay Ruppia-Zostera _
communities and continued for an annual cycle to investigate the effects of •
light and temperature on specific rates of seagrass photosynthesis. The ^
experiments were l^C uptake studies in which plants were removed from the
sediment, placed in a set of screened jars, and incubated in a running B
seawater system using ambient sunlight. The plants were exposed to 100, •
50, 30, 15, 5, and 1 percent of ambient light to determine the effect of
light quantity on photosynthesis. Experimental designs comparable with •
these were also conducted for upper-Bay species. Results are discussed •
later in this paper in Section 3.
In conjunction with these studies, measures of leaf area index (LAI)
were also conducted. Physiologically, the photosynthesis-light •
relationship determines the light levels at which SAV can grow and •
reproduce, that is, succeed. A greater leaf area exposed to light results
in greater productivity; however, light reaching the plants is not only •
determined by physical factors controlling light penetration through the |
water column, but by plant self-shading. Maximum plant biomass can in part
be related to leaf area. The leaf area index (plant area per sediment •
surface area) estimates maximum leaf density and thus potential area •
available to intercept light (Evans 1972, cited in McRoy and McMillan 1979).
Leaf surface area also provides a substrate for epiphytic growth. Leaf
area samples were collected to characterize the three main vegetation zones I
typical of lower-Bay communities. These data were used to provide a more •
accurate description of light penetration through the plant canopy as well
as to evaluate potential morphological adaptation of the plants to various •
light environments. To complement these specific l^C studies and LAI |
measures, field studies were completed to determine the effect of in situ
light reduction through artificial shading. Light reductions of 70 to 20 _
percent of ambient were used. The results of these studies support the I
573
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
hypothesis that total community metabolism is governed by, and is very
sensitive to, available light. During the course of these investigations,
light data collected in the field for various environmental (climatic)
conditions indicated that natural light reductions of these magnitudes were
common. To determine the overall effects of light reduction, specific
factors were investigated more thoroughly using both laboratory and in situ
experimental approaches for light-photosynthesis relationships, as well as
studies that determined those environmental variables controlling light
energy distribution and availability to the plant communities.
Studies initiated during the later phases of the CBP-SAV research
program investigated the effects of epiphytic growth and metabolism, and
the interactive effects of light and acute exposure to the herbicide
atrazine. Studies on epiphyte colonization were along two lines: the
epiphytic community as a primary producer and food source, and as a
competitor with the vascular plant community for available light.
Experiments completed suggest that the epiphyte community at times
dominates metabolism of the community and limits light available for
vascular plant photosynthesis. What remained to be determined was what
environmental conditions favor colonization, and at what point does the
resulting colonization stress the vascular plant.
These various research activities provide a data and information base
that serve management needs and identify specific research areas where
additional information is required for integration and synthesis. The work
proposed in the later part of the CBP-SAV program centered on filling what
were considered major gaps in information and the data base. The synthesis
report that follows is directed to our current state of understanding of
light energy properties and distribution in Chesapeake Bay and to the
relation of this information to past and current knowledge about SAV
community growth and survival.
THE RESEARCH PROGRAM ON LIGHT AND SAV: AN OVERVIEW
It has been the working hypothesis of the Chesapeake Bay Program-SAV
group that changes in such water quality variables as suspended
particulates (both living and non-living), dissolved substances, and
nutrients alter, directly or indirectly, underwater light regimes in such a
way as to limit benthic macrophyte primary production. Plants absorb light
energy for the process of photosynthesis, converting water and carbon
dioxide into organic compounds. White light (visible sunlight) is composed
of a spectrum of colors that are used selectively by green leaves based on
the plant's specific pigment complexes. Chlorophyll requires mainly red
and blue light for photosynthesis; these wavelengths are absorbed, and the
green and yellow bands are reflected. The accessory pigments also absorb
in the blue region.
As light penetrates the water column, the energy content and spectral
quality are changed by absorption and scattering. Water itself, dissolved
substances, and particulate materials are responsible for both the
absorption (conversion into heat energy) and the scattering of light.
Selective absorption and scattering by these factors result in attenuation
of specific light wavelengths causing a "color shift" (Kalle 1966, Jerlov
1976). Scattering, the change in direction of light propagation, returns
some of the incident radiation toward the surface and thus further reduces
the total light energy available to support photosynthesis. Phytoplankton
574
-------
I
act as both scattering and selectively absorptive and reflective particles p
and are in direct competition with other primary producers for the same
wavelengths of light—the red and blue bands. _
The temporal and spatial distribution of particulate materials and m
dissolved substances are largely determined by climatic variables and ™
biological processes. Wind velocity and direction, tidal amplitude and
frequency, current velocity, rain, and land runoff all interact to induce I
variations in water quality parameters and subsequently the spectral |
composition of light in the water column (Dubinsky and Herman 1979, Kranck
1980, Anderson 1980, Thompson et al. 1979, Scott 1978, Riaux and Douville •
1980). •
Based on these general premises, the light research program encompassed
four basic facets: (1) description of the submarine light environment
together with measures of various water quality parameters; (2) description •
of climatic and oceanic forcing functions; (3) detailed studies of •
photosynthesis-light relations by individual species and for entire SAV
communities; and (4) analysis of the relationships and correlations among •
the above data and other available information. The measurement and f
collection of light, water quality parameters, climatic and oceanic forcing
functions were made simultaneously with the light-photosynthesis _
investigations. Studies on both shores of the upper and lower Chesapeake •
Bay in vegetated and non-vegetated regions were undertaken. •
Characterization of the light environment was accomplished using a
Biospherical Instruments Model MER-1000 Spectroradiometer (Booth and U
Dunstan 1979). Specific attenuation in 12 biologically important |
wavelengths and integrated photosynthetically active radiation (PAR) values
were calculated from these data. The spectral irradiance measurements were •
made in quantum units as suggested for biological studies by the Special I
Committee on Oceanographic Research (SCOR) of the International Association
of Physical Oceanographers (IAPO).
There is a paucity of data on spectral irradiance in marine •
environments (Jerlov 1976). There are even fewer studies reporting data ™
for estuarine waters, Chesapeake Bay being no exception. Burt (1953,
1955a, b), using a shipboard spectrophotometer, analyzed filtered seawater •
samples from Chesapeake Bay and concluded that the primary factor in light |
extinction was the filterable, particulate matter. Seliger and Loftus
(1974) studied the spectral distribution of light in shallow water in a •
subestuary in the upper Bay in July and found a marked reduction of light I
in the 400-500 nm region of the spectrum. Champ et al. (1980) report an *
observed "orange-shift" for measurements made in the upper Bay during
August, 1977, using a submersible solar illuminance meter equipped with I
optical filters. They suggest that there is a continuum of spectral shifts •
toward the penetration of longer wavelengths from oceanic to coastal to
estuarine waters. This corroborates and extends Kalle's "yellow shift" •
theory (Kalle 1966). Kalle contends that the shift to longer wavelengths J
is more pronounced as the concentrations of suspended particles increases.
These investigations make up, in large part, the only complementary data _
base and, to our knowledge, no data exists in and around SAV habitats. I
Broad band (PAR) transmittance was determined with a Montedoro-Whitney ™
i? situ combination beam transmissometer and nephelometer. The
transmittance data were used to calculate the attenuation coefficient
"defined as the absorption coefficient plus the total scattering
575
I
I
I
-------
I
I
I
I
I
I
I
1
I
I
I
I
I
I
1
I
I
I
I
coefficient" (Jerlov 1976, Kiefer and Austin 1974). van Tine (1981) found
significant correlations between absence of submerged aquatic vegetation
and low transmittance values in an estuary in the Gulf of Mexico.
Total particulate matter (TPM), particulate organic matter (POM),
particulate-ATP, particulate chlorophyll _a, particulate inorganic matter
(PIM) were monitored in light spectral studies. These various measures
were used to estimate phytoplankton, zooplankton, detritus, and inorganic
fractions of the TPM.
Wind velocity and direction, water current velocity, tidal stage and
depth were determined concurrently with the other measures. Kiley (1980)
suggests a close relationship between wind and current for the York River.
In an effort to explain turbidity values, Williams (1980) calculated
significant positive correlations between wind and turbidity for upper-Bay
subestuaries. Ginsburg and Lowenstam (1957) and Scoffin (1970) showed a
baffling effect of SAV on currents that caused particulate matter to settle
out, generally improving the local light environment. Collection and
analyses of these data formed the basis for characterization of the natural
light environment and of the factors that are principal controls.
Various lines of evidence, as discussed earlier, suggest light in
general as a major factor controlling the distribution and productivity of
seagrasses. Preliminary studies demonstrated both potential nutrient and
light quantity effects on plant community metabolism. In the later phases
of CBP-SAV research, both field and laboratory studies were designed and
carried out in a more quantitative sense on photosynthesis-light relations
in Chesapeake Bay SAV communities.
For the field approaches, the entire SAV community and its interactions
were included in experimental designs. Short-term shading experiments
reflected the community response to daily variations in light quantity due
to such natural phenomena as cloud cover, tidal stage, and storm events.
Long-term shading studies reflected community response to possible
situations where water quality deteriorates to the point where light
penetration is reduced. The purpose of these studies is to estimate at
what point, relative to light quantity, the SAV communities would die out.
For the latter effort, sets of neutral density mesh canopies were placed in
selected SAV areas for long term studies. Shaded and control areas were
studied at regular intervals over the course of these experiments (1-2
months). With this design, community metabolism and various plant
community parameters (e.g., leaf area index, chlorophyll £ and b_, biomass,
and other plant meristic characters) were measured. Studies were carried
out in spring, summer, and early fall, 1981, to include the major growth
and die-back periods.
Past research programs in the CBP-SAV program resulted in several
hypotheses that might explain both the short and longer term survival of
Bay grasses. Among these, the potential for light, including those
variables influencing light, or more specifically light-energy
distribution, as a major environmental variable controlling SAV
distribution, growth, and survival was postulated. The intent of the
remaining sections of this report is to provide the general characteristics
of light in natural aquatic systems with emphasis on Chesapeake Bay; to
summarize the research results throughout the Bay relative to light and Bay
grasses; and to discuss the potential for light or light-related casualty
of Bay grass declines.
576
-------
SECTION 2
LIGHT IN CHESAPEAKE BAY
and indirectly proportional to the wavelength
•t.
where TV is Planck s universal constant, and C- is the speed of light in a
vacuum. This means that quanta of shorter wavelengths contain more energy
than quanta of longer wavelengths.
The complete spectrum of downward irradiance for incoming solar
radiation at the top of the atmosphere, at sea level, and at several water
depths is illustrated in Figure la. Most of the energy reaching the
earth's surface is contained within the shorter wavelengths (0.4 to 1 u or
400 to 1,000 nml). Not surprisingly, this region includes the
wavelengths of greatest biological importance, that is, 400 to 700 nm, the
photosynthetically active region of the spectrum termed PAR or PHAR. There
is almost no energy outside the PAR region at a depth of 1 m. Most of the
"missing" energy has been converted to heat by absorption. Only four to 11
percent of incident irradiance between 300-700 nm is reflected from the
surface or backscattered out of the water column (called albedo) (Clark and
Ewing 1974).
The properties and concepts in optical oceanography are usually divided
into two mutually exclusive classes, inherent and apparent. Inherent
properties, such as absorption and scattering, are independent of changes
in insolation (incoming light), whereas apparent properties, such as
underwater irradiance, vary with changing solar and atmospheric conditions.
As light passes through the water column, its energy content and
spectral quality are changed by absorption and scattering due to water
itself, dissolved substances, and suspended particles. The combined effect
of these processes is termed attenuation. The spectral distribution of the
total attenuation coefficient (o0 , measured with the beam transmissometer ,
generally shows high attenuance at both ends of the PAR. Since o\ is an
aggregated coefficient, it is informative to consider the component
parameters that cause the observed attenuance.
1 1 nm = 10-3 um = 10-9
m
I
1
I
GENERAL CHARACTERISTICS OF ESTUARINE OPTICAL PROPERTIES
The study of the interaction of solar energy with estuarine waters |
necessitates not only an understanding of the properties of light and
water, but also of the myriad living and non-living entities, both rm
dissolved and suspended, which affect the propagation of light in aquatic •
environment s .
The sun emits electromagnetic radiation in discrete packs or quanta (Q)
of energy called photons. The energy content (£) of each quantum is •
directly proportional to the frequency (^), S
577
-------
I
I
I
I
I
I
1
1
I
I
I
I
I
I
I
i
i
i
i
Scattering is the change in direction of light propagation caused by
diffraction, refraction, and reflection due to particles, water molecules,
and dissolved substances. Scattering is wavelength dependent, but in an
irregular and complex manner. Absorption is a thermodynamically
irreversible process wherein protons are converted to thermal, kinetic, or
chemical energy; photosynthesis is an example. Much of the attenuance in
the long wavelengths is due to the water itself, as shown by James and
Birge (1938) for pure water and by Clarke and James (1939) for filtered
seawater (see Figure 1). The effect of sea salts on attenuance is
insignificant. Pure water or pure seawater show a constant light
attenuation. Of course, natural water bodies (particularly estuaries) are
not pure, but contain constantly varying particulate and dissolved
substances. Burt (1958), using uncontaminated filtered seawater samples,
was able to determine the attenuance due to dissolved substances. By
subtracting this from the total attenuation coefficient of non-filtered
seawater, he was able to calculate the light attenuance due to particulate
matter. The energy of blue and red wavelengths is selectively absorbed by
particles, as shown in the example given by Prieur and Sathyendranath
(1981) (Figure Ib). The shorter wavelengths are also attenuated by yellow
substance or Gelbstoff (see Figure Ib), the collective name given to a
complex mixture of organic compounds by Kalle (1966). Gelbstoff is formed
from carbohydrates resulting from organic matter decomposition. Sources
are both allocthonous (swamps, marshes, land runoff) and autocthonous
(planktonic and benthic organisms). Flocculation of fine suspended and
colloidal materials in estuaries probably promotes the reaction, as does
the presence of amino acids (Kalle 1966).
The apparent optical properties of a body of water result from the
measurement of natural light fields underwater, that is, the measurement of
in situ radiant flux. Irradiance (E) (the flux of light per unit area) is
usually collected with a flat circular opal glass (or plastic) diffuser
(2 fV collector). The diffuser is designed so that light received from all
angles is transmitted to the sensor according to Lambert's cosine law. In
other words, the irradiance transmitted is proportional to the incident
radiant intensity multiplied by the cosine of the angle of incidence.
Jerlov (1976) reports that the ratio of cosine collection of downwelling
irradiance (E^) to equal hemispherical collection (E0) is generally in
the range of 0.75 to 0.85 downwelling. 2 If irradiance is the apparent
property of water bodies most commonly measured for biological purposes,
and was the measure used in CBP-SAV research. Of course, irradiance can be
expressed as either energy or quanta and measured in broad spectral
regions, such as the PAR, or at discrete wavelengths (spectral
irradiance). A family of downwelling spectral irradiance curves, in
quanta, is shown in Figure 2 for a Zostera marina bed on the eastern shore
of Chesapeake Bay. This figure shows that both total light energy and that
of specific wavelengths are lost with depth. At 0.1 meter, for example, a
lot of surface insolation, particularly in the photosynthetically important
400-500 range, has been lost.
Primary producers or autotrophs contain light-capturing pigments to
carry out photosynthesis. Most phytoplankton possess a pigment complex
similar to that of seagrasses and other higher plants. These pigment
systems absorb strongly in the blue and red regions (chlorophyllous
pigments). Figure Ib illustrates how combinations of water column
578
-------
NON-CHLOROPHYLLOUS
PARTICLES >^
ATTENUAT10N
COEFFICIENT
0.5
400
I
500 600
WAVELENGTH (nm)
700
Figure 1. Theoretical path of light from top of atmosphere to benthic
estuarine macrophytes. (a) Spectral energy distribution of
light at top of atmosphere, at the surface of the earth, and
at two depths in the ocean on a clear day (redrawn from Jerlov
1976 and Gates 1971). (b) Relative spectral absorption of
various constituents of estuarine waters (redrawn from Prieur
and Sathyendranath 1981). (c) Typical spectral irradiance
and attenuation in a Chesapeake Bay seagrass bed (Wetzel
et al. 1981). (d) Mean quantum action spectrum for higher
plants. 1.0 represents the highest photosynthetic response
observed by Inada in an individual species (redrawn from
Inada 1976).
579
I
1
I
I
I
1
1
1
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
1
I
I
I
I
I
I
I
I
I
I
9
I
I
I
I
SURFACE INSOLATION
400
500 600
WAVE LENGTH (nm)
700
Figure 2. Downwelling spectral quanta irradiance at the surface and at
several depths (Z) above the canopy of a Zostera marina bed on
the eastern shore of lower Chesapeake Bay (Vaucluse Shores)
at 1230 E.S.T. on a cloudy April day. The scale for the
insolation is on the right (Wetzel et al. 1982).
580
-------
(22
Often incorrectly termed extinction coefficient.
I
constituents cause specific spectral attenuation patterns. As these I
constituents change both temporally and spatially, the resultant spectral •
absorption pattern changes. Prieur and Sathyendranath (1981) have
attempted to classify water bodies based on combinations of these factors. •
The diffuse downwelling (or vertical) attenuation coefficient^ (K^) m
expresses the decay of irradiance as an exponential function,
I
I
where E2 is the irradiance at depth 2.2', E^ is the irradiance at depth «
Zi; and (Z2 - Z^) is the distance between the two measurement depths B
in meters. The units of Kd are m~l.
If (Z2 - Z\) brackets the air-water interface, it will include the
effects of reflection and inflate the estimate of K^. K
-------
I
I
1
f
I
I
1
I
I
I
I
I
1
t
i
I
I
I
I
portions to which higher plants such as seagrasses respond the most
efficiently. The mean quantum action spectrum for 50 species of higher
plants is presented in Figure Id (Inada 1976). A photosynthetic action
spectrum is produced by exposing a plant to controlled amounts of energy
(or quanta) at discrete wavelengths and by measuring its photosynthetic
response. The action spectrum in this figure is normalized to the highest
observed photosynthetic rates for red light. The curve presented here is
an approximation of the likely action spectrum for seagrasses. A major
peak falls in the 400-500 nm (blue) range, a region in estuarine waters
where very little light is available because of absorption by inorganic
particles, phytoplankton, and Gelbstoff.
Temporal variations in light distribution, both in the atmosphere and
underwater, are due directly and indirectly to the relative motions of the
earth, moon, and sun. The distance between the earth and sun and between
the earth and moon determines not only the amount of energy received by the
earth, but also the depth of water through which it must travel to reach
the seagrasses. The seasonal distribution of nutrients and the resultant
plankton blooms and runoff (with particulate and dissolved loads and
changed salinity regimes) also cause temporal variations in estuarine
underwater optical properties. Storms and wind increase land runoff,
currents, and waves. In shallow areas, this action increases
resuspension. Scott (1978) found that it took 11 days for the submarine
irradiance to return to pre-storm levels in an estuary in Australia. In
littoral regions, average submarine light conditions may be partly
controlled by the interaction of the local coastal morphology with
prevailing wind patterns.
Diurnal variations have two components: solar elevation and tidal
variation (amplitude and frequency). Since the interface between water and
air is a boundary between media of different optical densities, an
electromagnetic wave striking it splits into a reflected and a refracted
wave. Reflection of combined sun and skylight from a horizontal, flat
surface varies asymptotically with solar elevation between three to six
percent at angles greater than 30° from the horizon. Below 30°, the
reflectance increases dramatically up to 40 percent at 5°. Reflection
below 30° is wavelength dependent. The longer waves are reflected more
because the changing quantity of diffuse atmospheric light at low sun
angles (Sauberer and Ruttner 1941). Wave action, on the other hand,
reduces reflection at low angles.
Tidal cycles in estuaries not only change water bodies and their
associated seston and dissolved components, but also cause resuspension of
sediments and differences in depth. These are, of course, highly
idiosyncratic for specific systems (Burt 1955b, Scott 1978).
LIGHT ATTENUATION IN CHESAPEAKE BAY
A comparison of diffuse downwelling spectral attenuation coefficients
reported for Chesapeake Bay and its tributaries is presented in Figure 3
along with Jerlov's (1976) most turbid coastal water classification curve
(Type 9). For Chesapeake Bay, the earliest measurements of kd(X) were
made by Hurlburt (1945) (Figure 3a). His values fall in the lower range of
more recent in situ measurements. The shaded areas in Figure 3a represent
the range of values measured by Wetzel et al. (1982) from March through
582
-------
E
c
X
o
Ul
1
* 34 AX 1V1SV09
AVI 3)l»a*»S3HO
-I
(i-uj)j.N3ioidd3Co Nou.vnN3i.Lv
CO
4-1
C
rl
m oo
m ON w
r-l O
r-l W 4J
O rd (U (U
r-l CO >
r-l 4J M ft,
d)
C r-l
-------
I
I
I
I
1
I
I
I
I
I
I
I
1
1
I
I
I
I
1
July, 1981, in shallow regions of the lower Bay «3 m). Jerlov's curve
falls in these observed ranges, showing that the data fall within the range
of the most turbid coastal waters. Champ et al. (1980) conducted a light
characterization survey of Chesapeake Bay during August, 1977. Their mean
values are shown in Figure 3a along with their specific site measurements
in and near the mouths of the Sassafrass, Patuxent, Potomac and Chester
Rivers in Figure 3c. Their mean values fall within the upper ranges
measured in the lower Bay (Wetzel et al. 1982).
Pierce et al. (1981) intensively monitored the Rhode River during 1980
and 1981. Their annual mean attenuation values for an upriver station and
one at the mouth are plotted in Figure 3b. The upriver station was found
to be consistently more turbid, presumably because of its proximity to
autocthonous sources. Attenuation at both stations was higher for green,
yellow, and red wavelengths than observed in the lower Bay; however,
attenuations in the shorter wavelengths were in the same range. Maximum
penetration was at 575 nm and minima at 775 and 425 nm. Lower Bay maxima
were similar, and minimum measured was at 410 (775 was not measured).
Seliger and Loftus (1974) derived curves from 41V irradiance measurements
in the Rhode River that generally agree with the measurements of Pierce et
al. (1981), except in region 500 to 700 nm. Their measures fall within the
observations made for the lower Bay (Wetzel et al. 1982). The differences
noted in the 500 to 700 nm range may be due to upwelling irradiance
measured by the spherical collector.
Results of the August, 1977, survey by Champ et al. (1980) are shown in
Figure 3c. Their attenuation measurements in the turbidity maximum zone at
the mouth of the Sassafras River are the highest reported for the Bay. As
noted, there is nearly no available light below 500 to 600 nm. Wetzel et
al. (1982) observed similar, very high attenuations in the blue region (400
to 500 nm) at lower-Bay sites during a spring runoff event following a
major rain storm. The attenuation of green wavelengths (~500 to 550 nm) in
the summer was much higher at the mouths of the Patuxent and Potomac Rivers
(upper Bay) than at the mouths of the York, Severn, and Ware Rivers (lower
Bay). Figure 4 illustrates the lower Bay sampling stations.
A summary of the recent Chesapeake Bay data on diffuse downwelling 2TT
irradiance attenuation coefficients indicates a severe attenuation of light
energy in the photosynthetically important (400 to 500 nm blue, and 700 to
775 nm near infrared) regions of the spectrum. Attenuation in the short
wavelengths was particularly marked in the turbidity maximum region of the
Bay at the mouth of the Sassafras River, and at the mouth of the Patuxent
River during August (Champ et al. 1980) and at lower-Bay sites during
spring runoffs (Figure 5). The mean Bay attenuation coefficients
calculated by Champ et al. (1980) are about 1.0 ra~l higher than Jerlov's
(1976) most turbid coastal water classification.
Comparison of Light Attenuation in Vegetated and Unvegetated Sites of the
Bay
An analysis of the spectral attenuation coefficients at shallow sites
in the lower Chesapeake was undertaken to determine if correlations existed
between the presence or absence of benthic macrophytes (Zostera marina and
Ruppia maritima) and specific spectral patterns (Wetzel et al. 1982). The
specific question, what are the light quality differences between vegetated
584
-------
Figure 4. Locations of lower Bay stations (Wetzel et al. 1982).
(1) Mumfort Is., York R. (2) Allen's Is., York R.
(3) Guinea Marshes (4) Mouth of Severn R., Mobjack Bay
(5) Four Point Marsh, Ware R. Mobjack3Bay.
585
-------
1
1
1
IV
1
I
1
1
1
I
1
1
1
1
t
I
1
1
f
1
4-
3-
j->
'E
H 2-
UJ
0
u.
u.
g '-
o
z
o
^ 0 -
— J
z • «-,
UJ
1-
*~L
^
o
!J 3-
«J
«l
UJ
5
z
1 2-
UJ
en
u.
5 i-
0-
UNVEGETATED SITES VEGETATED SITES
ii 11
MUMFORT IS.
V
\
. \
\\
^
\ \JULY
\ \
\ \
MAY\ *N
\ X
N *X\ /~^"
\ ^^^.-^'*
^vAPRIL "-"-•-••^
\9 ^'
•^•-^.^.--"
\ MAY. SEVERN R.
\ '
\ IRUNOFF
\ 1
\ i
JULY\ 1
\ 1
\l
\»
*\
\
\ ,.--<-
MARCHN/^*^--. -^'^^*~~*
^~^^>~^
|
1 GUINEA MARSH
IRUNOFF
i
MAY>
i
•^vJULY 1
^^-^ 1
\ j
\ i
\ ;
\
\
\ i
V^MARCH \ |
XN,^^^
FOUR POINT MARSH
t
\
\ JULY tJl A P^* LJ
\ ** ** w • iTi w n ^ n
\ APRIL
. \ MAY
\ \
'v -. - II 1 1 V
\ \ JULY
\ \
\ \ •
RUNOFF\ \
\ \
\ \
\ v^ ^,.—.-'
•
^•^ \ MAY
APRIL^^.^ X, ^^-''^
400 500 600 700 400 500 600 7C
WAVELENGTH IN NANOMETERS
Figure 5. Mean monthly diffuse downwelling spectral attenuation
coefficients. for vegetated and unvegetated sites in the
lower Chesapeake Bay. All coefficients calculated for the
depth interval 0.1 to 0.5m. Mumfort Island (York River) and
Severn River sites: unvegetated. Guinea Marsh and Four Point
Marsh (Ware River) • sites : vegetated (from Wetzel et al. 1982).
586
-------
and unvegetated sites, was addressed. The sites (Figure 4) were chosen •
because of their varied vegetational histories (Orth et al. 1981). The <|
Mumfort Island (York River: Station 1) and Severn River (Station 4) sites
are presently unvegetated. The Guinea Marsh (Station 3) and Four Point £
Marsh (Ware River: Station 5) sites have seagrass beds. Both the Severn •
River and Four Point Marsh sites are affected by agricultural runoff (C.
Hershner, personal communication) . The Allen's Island site, Station 2, is
presently unvegetated, but has recently been replanted by Orth and I
associates. Twelve wavelengths (410, 441, 488, 507, 520, 540, 570, 589, W
625, 656, 671, 694 nm +_ 5 nm) and total PAR were analyzed at depths of 0.1
and 0.5 m. Downwelling irradiance (£4) was measured as Quanta nm~l fe
cm~2 sec~l, each reading representing the mean of 250 scans. Diffuse •
downwelling spectral attenuation was calculated between 0.1 and 0.5 m.
The mean spectral attenuation values ranged from about 0.2 to 9.0
m~l. Integrated PAR attenuation varied from about 0.5 to 1.6 m~l A
(Figure 6). A .clear seasonal pattern of extreme attenuation of blue *
wavelengths was evident at all sites beginning in May. This was probably
due to a combination of increased particulates associated with runoff V
events and seasonal plankton blooms. j|
Mean PAR attenuation coefficients were found to be significantly lower
(mean difference of 0.47 m~l) in vegetated than in unvegetated sites •
during May, 1981 (Figure 6). This was due to a lower attenuation in the M
500 to 700 nm region of the spectrum at vegetated sites (Figure 5), despite
the effects of high blue attenuation due to runoff. A significant
difference among sites based on PAR attenuation coefficients was also •
observed in July; however, one vegetated site (Four Point Marsh) was w
grouped with the unvegetated sites having higher attenuation (Figure 6). '
This was due to the increased attenuation of wavelengths above 500 nm at •
the Four Point Marsh site during July. The only general light quality £
differences between vegetated and unvegetated sites that was evident from
these analyses were the reduced attenuation in the 500 to 700 nm region at «
vegetated sites during May. 3 M
Kaumeyer et al. (1981) measured a significant difference in PAR
attenuation coefficient inside and outside SAV beds at Todds Cove, Md.
during July, August, and September, 1980. The vegetated areas were from B
0.4 m~l to approximately 2.0 m~l lower. Significant differences were 9
not found in attenuation inside and outside grassbeds at the Parson Island
study site. Table 1 summarizes the results of their studies. ft
Historical Data Bases and Optical Properties of Chesapeake Bay Waters
Most of the historical light data for Chesapeake Bay has been collected 9
by Secchi disc. This method is not ideal, but can be used to indicate ™
trends. Heinle et al. (1980) reviewed Secchi disc light data for both
mid-Bay and the Patuxent River, which was chosen because of the extensive M
data base (Figure 7). Transparency has decreased since the 1930' s, V
Subsequent measurements and analyses extend and corroborate this conclu-
sion. Not only is the mean violet and blue attenuation lower in
vegetated sites but the variation is also less (see Wetzel et al 1982).
587
-------
I
I
I
1
1
I
1
I
t
I
1
I
I
f
I
I
I
I
I
2.0
IT
o
-------
a.
UJ
a
4—
3-
2-
I —
O 1974-77
• 1936-40
A I960, 1961, 1964
A MEDIAN 8/22 pts.
£r-& RANGE 7/23 pts.
I I
I
T
T
•X 2.0-
1.0-
^ 0.5 H
o
o
UJ
-------
I
I
I
I
I
I
I
I
I
I
I
I
1
t
I
I
I
1
I
Table 1. COMPARISON OF MEAN PAR ATTENUATION COEFFICIENTS INSIDE AND OUTSIDE
OF VEGETATED AREAS AT TODDS COVE. MD.. 1980 (KAUMEYER ET AL. 1981)
Month Location
June SAV 2.6 + 0.20
Reference Site 2.5 + 0.75
July SAV 2.5 + 0.30
Reference Site 2.9 ^ 0.70
August SAV 1.8 +_ 0.56
Reference Site 3.1 + 0.33
September SAV 1.9 _+ 0.34
Reference Site 3.8 + 0.96
especially during the winter in the mid-Bay region (Figure 7a). An
increase in turbidity, as estimated by Secchi disc measures, has been quite
dramatic in the Patuxent (Figures 7b, 7c). Mid-1970's Secchi disc data for
rivers in the upper Chesapeake Bay are reported in Table 2 from Stevenson
and Confer (1978). The values are generally low (<.1.0 m) and are similar
to those reported for the Patuxent during the 1960's and 1970's (Figures
7b, 7c).
Increases in chlorophyllous pigments, due to phytoplankton blooms
caused by increased nutrients, can have a severe effect on light
attenuation in the photosynthetically critical blue and red spectral
regions (Figures Ib, Id). Historical chlorophyll data for Chesapeake Bay
and Patuxent River are summarized in Figures 8 and 9. Chlorophyll
concentrations have increased dramatically in the upper and mid-Bay since
the early 1950's. Concentrations as high as 100 to 200 ug IT1 were not
unusual. In contrast, lower-Bay concentrations have not significantly
changed (Figure 8b). Concentrations in the Patuxent River have increased
590
-------
80-
70-
60 —
\J\s
5 50-
oi
_j 40-
o
30-
20 -
10-
o-
30-
t
20-
o>
Ol
Z
o
10-
o-
i
0 CBI 1949-1951
A CBI 1964-1966
D CBI 1969-1971
• EPA 1969-1971
I
1
fj &
126 I
(1970)
c
A
1
l'
CD 2fj A (Sf^a-g *& °
i i i i
i
2
1
>c
i
1240
1
1
A
i
If
6
1
1
Tff
i
c
p
(1963
A
1
' It
s
• 125
I
r
a
^
1
/
c
I
1
O FLEISCHER etal, (1976) 1973 DATA
O CBI 1949-1951 -Potomac to Roppahonnock (744)
• CBI 1949-1951 Lower Boy Below RappahannocK (724,707)
• PATTEN etol, (1963)
A CBI 1964-1967 (746)
Q CBI 1969-1971 (744, 744S)
• CBI 1969-1971 (7^,707) •
*
T T
a
^ T* T 1 1
a i n i
Sli ' A *
J 1 F ' M ' A ' M
>
•
1
^
d
1
c
i
i
)
I
J '
A
i
q
4
•1
Ii[
>
i
M «
»
J ' A
1
1
1
1
A
L
\
\
119651 UPPER 0
BAY
•
]
A
D
C
A
1
3C
A
A
.0 ^rt
O ii HI (m
~B" "QS
i i i
LOWER b
BAY
A
t
a
0
4^
' I* ?1
'3
S ' 0 ' N ' D
I
1
1
I
1
t
1
i
i
i
i
i
i
MONTHS
Figure 8. Summary of historical chlorophyll
(a) upper Bay. (b) lower
1
a data for the Chesapeake Bay.
Bay (Redrawn from
591
Heinle et al. 1980). ^
I
1
-------
I
I
I
I
i
i
t
i
i
i
i
i
i
t
i
i
i
i
i
100
- 80
o>
a.
o'l 60
a.
o
o
.c
O
40
20
o>
a
100
80
60
40
20
o>
100
80
60
40
20
a
Jan , Feb, Mar
May, June, July
Aug ,Sept, Oct
• Lower Marlboro
O Benedict Bridge
D Queen Tree Landing
1962 64 66 68 70 72 74 76 78
YEAR
Figure 9. Summary of historical chlorophyll £ data for three regions of
the surface waters of the Patuxent R., Md. (a) January-March
(b) May-July (c) August-October. (Cross hatching is to clarify
general trends for each site.) (Redrawn from Heinle et al. 1980)
592
-------
Table 2. AVERAGE SECCHI DISC DATA (cm) BY RIVER SYSTEM, MARYLAND
CHESAPEAKE BAY, 1972-19 76a (AS REPORTED IN STEVENSON & CONFER
1978)
River System
Elk and Bohemia
Rivers
Sassafras River
Howe 11 and Swan
Points
Eastern Bay
Choptank River
Little Choptank
River
James Island and
Honga River
Honga River
Bloodsworth Island
Susquehanna Flats
Fishing Bay
Nanticoke and
Wicomico Rivers
Manokin River
Patapsco River
Big and Little
Annemessex Rivers
Gunpowder and Bush
River Headwaters
Pocomoke Sound,
Maryland
Magothy River
Severn River
Patuxent River
1972
33.0
34.3
33.8
67.3
60.7
64.5
70.1
78.2
73.7
64.5
49.5
55.4
94.2
73.7
109.7
42.9
101.6
83.8
97.3
80.3
1973 1974 1975
35.1 - 25.7
52.3 - 29.2
75.4 - 61.2
62.5 76.5 54.6
62.5 84.3 61.5
59.4 66.8 63.8
64.0 74.2 67.1
67.3 72.6 68.8
87.6 94.7 177.0
65.5 82.6 33.8
77.0 85.6 75.7
58.9 65.8 61.0
94.7 101.3 107.4
80.0 67.8
92.7 96.3 88.1
38.3 46.7
82.0 - 96.8
97.3 73.4
70.4 79.5
80.8 61.5 66.8
Continued
593
1976
36.3
51.1
57.7
75.9
64.3
78.5
73.4
67.8
83.3
76.5
54.1
58.9
81.0
70.1
85.1
53.8
85.9
74.4
86.4
62.7
1
1
1
1
1
1
1
1
1
1
f
~*
1
I
I
-------
I
I
I
I
I
I
I
1
I
I
I
I
I
I
1
I
I
I
1
Table 2. AVERAGE SECCHI DISC DATA (cm) BY RIVER SYSTEM, MARYLAND
CHESAPEAKE BAY, 1972-1976a (AS REPORTED IN STEVENSON & CONFER
1978) (CONTINUED)
River System 1972 1973 1974 1975
Back, Middle and
Gunpowder Rivers 79.5 75.7 73.2 75.4
Curtis and
Cove Point 45.2 77.0 81.8 58.9
South, West and
Rhode Rivers 74.7 66.0 61.2 48.5
Chester River 76.2 73.4 100.1 87.9
Love and Kent
Points 89.7 74.7 117.6 72.1
Smith Island,
Maryland 78.5 76.2 89.7 139.4
Average 70.1 71.1 79.5 76.2
1976
61.2
73.7
67.1
85.1
89.9
87.6
71.4
significantly in both the upper and lower portions (Figure 9), especially
during late spring and early summer (Figure 9b). Levels in excess of 100
UgL~l were common in the summer throughout the 1970's — this is twice
the concentration measured during the previous decade.
In addition to the thoroughly documented increased chlorophyll £
concentration in the Patuxent, there have also been increases in most of
the other tributaries of the Bay. Chlorophyll a_ concentrations in the
Choptank, Chester, and Miles Rivers of the middle eastern shore are 1.5 to
Table 3. RANGES OF CONCENTRATIONS OF CHLOROPHYLL a (ug 1~1) AT SURFACE
AND BOTTOM DEPTHS IN THE LOWER POTOMAC RIVER DURING 1949-1951,
AND 1965-1966 (HEINLE ET AL. 1980)
Month
January
March- April
May
July
Oc tober-November
1949-1951
Surface
1-2
10-21
3-6
3-5
1-9+
Bottom
1-2
12-27+
9-24+
1-2 +
1-7
1965-
Surf ace
3.2-4.6
1.1-20.0
5.8-13.2
9.0-13.8
9.3-24.0
1966
Bottom
3.1-5.0
1.1-9.5
4.3-9.8
1.0-1.8
3.6-11.0
594
-------
I
2.0 times higher presently than earliest data show. There have been •
upstream increases in the Magothy, Severn (Md.), and South Rivers. •
Concentrations up to 100 uL'^were measured in the upper Potomac in the
mid-19601s. Concentrations in the lower Potomac were generally higher in —
the 1960's than 1950, except in March and April (Heinle et al. 1980). I
Increased chlorophyll £ concentrations have also been measured in the ™
Rappahannock and York Rivers during the last few years. The upper James
has had high concentrations similar to the upper Potomac since the I
Table 4. ANNUAL MEAN FRESHWATER FLOWS AND OCCURRENCE OF HURRICANES TO ALL
OF CHESAPEAKE BAY (CUBIC FEET PER SECOND) FOR 1951-1979 (HEINLE g
ET AL. 1980). I
Year
1951
1952
1953
1954 Hurricane
1955 (2) Hurricanes
1956
1957
1958
1959
1960
1961
1962
1963
1964
1965
1966
1967
1968
1969
1970
1971
1972 Hurricane
1973
1974
1975
1976
1977
1978
1979 Hurricane
Bay Annual
Average
82,100
94.300
72,800
58,700
73,400
76,000
64,400
81,400
66,400
77,300
78,000
64,800
52,400
61,900
49,000
53,300
77,200
60,100
54,900
77,200
79,000
131,800
95,200
76,900
103,100
84,400
80,100
91,300
113,800
5-Year
Average
76,260
73,100
61,220
64,540
97,180
92,400
1
t
I
I
I
1
I
I
I
t
mid-19601s, but the lower River still does not. Dense algal blooms have ^
been noted in the Elizabeth, Back, and Poquoson Rivers of the lower Bay. •
Heinle et al. (1980) summarized the state of the Bay graphically in ™
terms of enrichment that they defined as deviations in concentrations of
chlorophyll a_ from historic, natural periods of stability or steady state •
595
I
t
-------
I
I
I
I
I
I
I
1
I
I
t
1
1
t
i
I
i
i
I
concentrations. Figure 10 shows the regions of the Bay that are
categorized as moderately or heavily enriched. Many of these areas have
experienced declines in Bay grasses on a time scale overlapping the
enrichment.
Changes in dissolved organic materials, inorganic particulate matter,
and allochthonous organic particulate matter in the Bay are mainly
determined by inputs (runoff) of freshwater to the tributaries and by
additional input due to storm events. Table 4 summarizes annual mean
freshwater flow to the entire Bay and major storms during the period
1951-1979. In addition to adding large amounts of sediment to the water
column, major storm events increase nutrient loads that favor phytoplankton
blooms.
Suspended sediment transport and discharge of the Susquehanna River,
the major source of freshwater to the Bay, are given in Table 5.
Table 5. SUSPENDED SEDIMENT TRANSPORT AND DISCHARGES OF SUSQUEHANNA RIVER
(GROSS ET AL. 1978)
Calendar Year
1966
1967
1968
1969
1970
1971
1972
Agnes, 24-30 June 1972
1973
1974
.1975
Eloise, 26-30 Sept. 1975
1976
Annual suspended sediment
(millions of metric tons
Above Dam
1.5
1.7
1.7**
nd
2.0
1.4**
11.3
7.6
3.2
1.7
3.8
1.6
nd
discharged
per year)
Below Dam
0.7 (60%)*
0 . 3**
nd
0.32 (60%)*
1 . 1**
1.0
33
30
1.2 (54%)*
0.8 (53%)*
11
9.9
1.2
nd = no data
* Percent discharged during annual spring flood
** Records incomplete for the year
Gross et al. (1978) suggest that one-half to two-thirds of the suspended
sediment discharge of the Susquehanna is deposited behind the dams or in
the lower reaches of the river during years of low flow and no major
flooding. During major floods, however, these deposits are eroded and
transported into the Bay. Thus, dams effectively increase the amount and
variability of sediment discharged under flood conditions.
It is evident that major storms, such as hurricanes, significantly
increase freshwater input, but there is also an apparent wet-year, dry-year
cycle imposed on the data. The five-year-flow averages (Table 4) suggest a
mid-19601s depression followed by an increase through the 1970's. Although
these data have not been rigorously analyzed, it is apparent that long-term
changes and/or cycles in climatic conditions (rainfall, temperature, and
596
-------
Figure 10. Portions of the Chesapeake Bay considered enriched by Heinle
et al. 1980. Enrichment is defined as increase in chlorophyll a
levels from historic, natural periods of stability.
597
-------
I
I
I
I
1
I
I
1
I
I
I
1
1
t
t
I
I
I
I
major storms) influence water quality and optical properties of Bay
waters. However, cause and effect relations are still poorly understood
and resultant optical properties of Bay water are determined and controlled
by multiple influences: runoff; nutrients; suspended particulates (both
living and dead); and, as the principal driving forces, the general
climatic regime.
598
-------
SECTION 3
LIGHT AND PHOTOSYNTHESIS IN CHESAPEAKE BAY SAV COMMUNITIES
GENERAL REVIEW OF PHOTOSYNTHESIS
I
I
I
Photosynthesis is the process in which light is used as the energy •
source for the synthesis of organic compounds. Three basic steps are •
involved in the process: (1) absorption of light energy by photosynthetic
pigments; (2) processing the captured light energy to produce the compounds •
ATP and NADPH; and (3) the reduction of C02 using ATP and NADPH and the £
production of carbohydrates. The first two steps are light-dependent and
are collectively referred to as the "light reaction". The third step is
light-independent and termed the "dark reaction". •
The photosynthetic pigments have characteristic light energy absorption -»
spectra in the photosynthetically active region, 400 to 720 nm.
Chlorophyll £ absorbs light more effectively at higher wavelengths (>600 tt
nm); accessory pigments such as chlorophyll b^, carotenoids, and others are |
more effective at shorter wavelengths (<600 nm) . Chlorophyll £ and the
accessory pigments absorb and transfer light energy at varying efficiencies K,
to specialized chlorophyll a_ molecules (P700) where they are used directly I
for biochemical reactions.
The photochemical reactions are driven by units of light energy called
photons (quantum energy). The quantum energy is a function of wavelength; •
quanta of shorter wavelengths contain more energy than quanta of longer 9
wavelengths. Light energy transferred to P700 is most efficient as it is
used directly in the photosynthetic system; light energy transfer by •
chlorophyll a and accessory pigments is less efficient. The quantum yield, •
the moles of 02 produced or C02 fixed per photon of light absorbed, is
used to estimate the transfer efficiency. ^
The light utilization spectra of a particular species is called the I
action spectra, a characteristic curve obtained by combining the light ' "
absorption spectra and the quantum yield of intact plant cells. The action
spectra is an important feature because it reflects the ability of a M
species to adapt to various light spectral regimes (Figure Id). This is of jf
particular importance when considering photosynthesis of submerged plants.
In aquatic environments, spectral shifts in light energy result from the m
water itself, suspended organic and inorganic material, dissolved organic I
compounds, and other water column constituents (discussed in Section 2).
A general approach to the investigation of photosynthesis is to
construct light saturation curves for various species (Figure Ha). An •
examination of photosynthesis-light curves (P-I curves) shows that •
photosynthesis (P) increases with increasing light to a point of optimal
irradiance (!Opt) where, over a range of irradiance, the photosynthetic A
system is saturated and maximum photosynthesis (Pmax) occurs. At higher jj^
irradiance, there may be a depression in the photosynthetic rate, termed
photoinhibition. The initial slope of the curve (dP/AI or<3f) and Pmax are ^
the two major parameters used in describing P-I curves (Jassby and Platt •
1976). Alpha (
-------
I
I
I
i
i
i
i
i
i
i
i
i
i
i
t
i
i
i
i
state of the plants (Parsons et al. 1977). The term 1^, proposed by
Tailing (1957), is the irradiance at which a linear extension of the
initial slope intercepts Pmax. 1^ is regarded as indicative of the
plant's adaptation to its light regime (Steeman-Nielsen 1975). 1^ is
irradiance where P = 0.5 Pmax and is similar to the Michalis-Menten
half-saturation constant. Ic is the irradiance at the compensation
point, where photosynthesis equals respiration (P = R).
Characteristic P-I curves are shown in Figure lib. Plants adapted to
high and low light environments, termed sun and shade species, exhibit
different P-I curves. Sun species (curve 3) generally exhibit higher
^max values than shade species, which exhibit greater and lower Ic
values (curves 1 and 2). In the aquatic environment, with reduced
availability of light, species exhibiting shade-type photosynthesis
(greater photosynthetic rates at low light intensities) are at an advantage.
PHOTOSYNTHESIS OF SUBMERGED VASCULAR PLANTS IN RELATION TO LIGHT AND
TEMPERATURE
In situ studies of submerged angiosperms point to the important role of
light in seagrass production and distribution (Jacobs 1979, Mukai et al.
1980). In a study of Zostera in Denmark, Sand-Jensen (1977) showed a
positive correlation between leaf production and insolation over a nine
month period. Biomass and photosynthesis rates of Posidonia declined with
depth near Malta (Drew and Jupp 1976); this was probably due to decreased
light penetration with depth. In before and after studies of an estuary
that was closed to the sea, Neinhuis and DeBree (1977) report that the
Zostera population increases in density and extends to a greater depth;
they suggest that this is probably due to an increase in water transparency.
In situ light manipulation experiments provided evidence of the
importance of light to seagrass production. For example, at the end of a
nine-month study during which ambient light was reduced by 63 percent, in
situ Zostera densities were only five percent of that of the control
(Backman and Barilotti 1976). In similar studies, Congdon and McComb
(1979) report that lower than ambient light levels result in lower Ruppia
biomass; as shading duration increases, higher light levels are required to
sustain a high biomass.
Studies involving the epiphytic community, those organisms directly
attached to submerged angiosperm blades, suggest that epiphytes have a
detrimental effect because they shade the macrophytes. Both Kiorbe (1980)
and Phillips et al. (1978) provide data to indicate that epiphytic
development suppresses macrophyte growth. Sand-Jensen (1977) reports that
Zostera photosynthesis is reduced by up to 31 percent due to a decreased
penetration of light and inorganic carbon through the epiphytic community
to the seagrass blades. Johnstone (1979) hypothesizes that the rapid
linear growth of Enhalus leaves (up to two cm day~^) is related to a
shading effect from epiphytes. In contrast, the data of Penhale and Smith
(1977) suggest that an epiphytic community may be beneficial in certain
environments. For Zostera exposed at low tide, epiphytes prevent
desiccation damage by trapping a film of water, and probably reduce the
photoinhibitory effect of high light.
In addition to light, temperature also influences submerged macrophyte
distribution and productivity rates (Biebl and McRoy 1971, Drew 1978). The
600
-------
Pmax
a>
.c
M
O
^
.C
a.
a
Compensation
Point (P=R)
"opt
P max 3
u>
Q>
C
10
O
O
Q.
0
'I and 2
Figure 11. Diagramatic photosynthesis-light relationships. See text
for description of parameters.
601
I
I
I
I
I
I
I
I
I
1
I
I
I
I
i
i
i
i
I
-------
I
I
I
I
I
I
I
1
I
I
I
1
I
I
t
I
I
I
I
biogeography of marine and brackish water plants points to a temperature
effect on worldwide distribution; for example, genera such as Zostera,
Ruppia, Phyllospadix, and Posidonia occur mainly in temperate zones while
genera such as Thalassia, Syringodium, and Halophila occur mainly in
subtropical and tropical zones. Drew (1979) reports that the Pmax of
four seagrass species collected near Malta increases in direct proportion
to temperature, up to temperatures [30 to 35°C (86 to 95°F)] where
tissue damage occurred; decreases are not observed at environmental
temperatures. In contrast, Penhale (1977) observed a decline in Pmax
from 22 to 29°C (71.6 to 84.4°F) for Zostera in North Carolina where
environmental temperatures reach 34°C (93.2°F). The co-existence of
species such as Ruppia and Zostera in the lower Chesapeake Bay may be a
result of differential responses to both temperature and light, as
apparently is the case in a Myriophyllum-Vallisneria association described
by Titus and Adams (1979). They report that a greater for temperature
tolerance Vallisneria, in conjunction with the temperature dependence of
photosynthesis, results in a temporal partitioning of production.
Vallisneria apparently favored in midsummer conditions; Myriophyllum spring
and fall conditions.
Sun and shade species have been described for submerged macrophytes
(Spence and Crystal 1970a, 1970b; Titus and Adams 1979). Sun species
generally exhibit higher Pmax values than shade species that exhibit
lower Ic values, and lower dark respiration rates. Certain species can
adapt to a wide range of light conditions. Bowes et al. (1977) cultured
Hydrilla under high and low irradiances; subjecting the plants to high
light increased the Iopt value four-fold. Plants grown under low light
achieved Ic and 1^ at lower intensities.
In seagrass systems, pigment relationships generally vary with light
quantity or with position within the leaf canopy. The adaptive capability
of seagrass pigment systems to the light environment has been shown in
various studies. For example, Wiginton and McMillan (1979) report that the
total chlorophyll content is inversely related to light for several
Caribbean seagrasses collected at various depths. For seagrasses cultured
at several light levels, the total chlorophyll content increased with
decreasing quantum flux (McMillan and Phillips 1979, Wiginton and McMillan
1979). Within individual meter-long Zostera leaves, the chlorophyll a^ to
chlorophyll b_ ratio varied significantly, with the lowest ratio at the
basal portion of the plant (Stirban 1968). In a detailed study of
chlorophyll relationships in a Zostera system, Dennison (1979) observed no
substantial variation in total chlorophyll content within the leaves as a
function of depth of the leaf canopy in integrated samples along a depth
gradient within the bed. The chlorophyll £ to chlorophyll b_ ratio,
however, decreased from the apical to basal portion of the leaves.
Although the physiological photosynthesis-light relationship ultimately
determines the light levels at which plants grow, the morphology of
individual plants and the community canopy structure may play an important
role in production and species distribution. In a study of Myriophyllum
and Vallisneria, Titus and Adams (1979) observed that the former had 68
percent of its foliage within 30 cm (11.7 inches) of the surface, and the
latter had 62 percent of its foliage within 30 cm of the bottom.
Myriophyllum, an introduced species, has often displaced the native
Vallisneria; a contributing factor is probably the ability of Myriophylium
to shade Vallisneria. In a detailed community structure analysis of a
602
-------
I
monospecific Zostera community across a depth gradient, Dennison (1979) I
concludes that changing leaf area is a major adaptive mechanism to (B
decreasing light regimes.
PHOTOSYNTHESIS-LIGHT STUDIES IN CHESAPEAKE BAY I
Investigations of photosynthesis-light relationships carried out .
through the Chesapeake Bay Program can be categorized into three general •
experimental designs. In the first, P-I curves were constructed for the •»
four dominant species in Chesapeake Bay system: Myriophyllum spicatum and
Potamogeton perfoliatus in the upper Bay, and Zostera marina and Ruppia •
maritima in the lower Bay. These experiments used whole plants or leaves •
subjected to various light intensities (created through the use of neutral
density screens) and various temperatures. ^
The second approach used microcosms in which the effects of various •
concentrations of phytoplankton and suspended solids on light penetration -"
and on Potamogeton photosynthesis were determined.
The third experimental design involved in situ community metabolism •
measurements under a wide range of natural light regimes. In certain 9
experiments, neutral density screens were used to shade the community on a
short-term basis. The experimental design and methods for each of these ^
studies are detailed in Kemp et al. (1981b) and Wetzel et al. (1982). •
P-I Relationship of Major Species
P-I curves were constructed for whole plants of M. spicatum and P. B
perfoliatus at 21°C (69.8°F) (Kemp et al. 1981b) (Figure 12). Both9
species exhibited the characteristic photosynthetic response to light with
light saturation occurring between 600 and 800 uE m~2 sec~l. •
Myriophyllum exhibited a greater Pmax and a greater 1^ than £
Potamogeton; however, the two species exhibited similar 0(. Although these
species occur in the same general locale, they do not form dense, mixed bed _
stands where they would be in direct competition for light. •
The photosynthetic response to light and temperature was determined for
isolated Z._ marina and R. maritima leaves (Wetzel et al. 1982). Since
these species co-exist in the lower Chesapeake Bay, an evaluation of •
photosynthetic parameters of each species might suggest competitive •
strategies. Experiments carried out at six temperatures and under natural
light indicate that light saturation of Zostera occurs about 300 uE m~2 ^
sec~l while that of Ruppia occurs about 700 uE m~2 Sec~^-. •
Differences in Pmax between Zostera and Ruppia were observed and appear
related to temperature. At warmer temperatures, Ruppia exhibits a higher
^max than Zostera; the situation is reversed at colder temperatures •
(Figure 13). A summary of the data shows that Ruppia exhibits the greater ™
pmax at temperatures greater than 8°C (46.4°F) (Table 6). A
comparison between the two species shows that Zostera generally exhibits a A
greatercfj this suggests a competitive advantage for Zostera at lower light 9
levels.
603
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
c
0
a.
(M
O
c/
05
LU
X
(O
o
H
O
I
CL
LU
tr
<
Q.
Q.
- P max
MYRIOPHYLLUM SPICATUM
Pmax
200
600
LIGHT INTENSITY, M EINSTEINS m s
1000
.-2-1
Figure 12. Photosynthesis-light curves for two species of upper Chesapeake
Bay submerged vascular plants (from Kemp et al. 1981c).
604
-------
50-1
40-
2.5-
T 20-
AUGUST 29, 1979
10
20
30 40 50 60
LIGHT LEVEL (Percent Ambient)
JANUARY 29, 1980
70
80 90
100
o
i-
o
I
a.
10 20
LIGHT LEVEL (Percent Ambient)
Figure 13. Photosynthesis-light curves for Ruppia and Zostera from a
mixed bed site on the Eastern Shore, Virginia. Light is total
light flux during 4 h C incubations (from Wetzel et al.
1982).
605
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
1
I
I
I
1
1
I
I
I
I
1
I
Table 6. PHOTOSYNTHETIC PARAMETERS FOR RUPPIA MARITIMA AND ZOSTERA MARINA
LEAVES AT VARIOUS TEMPERATURES. THE LIGHT IS THE TOTAL LIGHT
FLUX DURING THE 4h 14C INCUBATIONS (FROM WETZEL ET AL. 1982)
TEMP
°C
1
8
12
18
21
28
LIGHT
E m"2
5.0
22.1
15.1
21.8
14.5
12.0
P (mg
max
Ruppia
2.15
3.12
3.91
2.60
3.82
2.39
c g-i h"i)
Zostera
2.66
3.25
2.15
2.15
3.55
1.31
INITIAL
Ruppia
0.18
0.41
0.16
0.35
0.27
0.52
SLOPE
Zostera
0.70
1.41
0.55
0.34
0.27
0.69
The data from these experiments relate to how plants capture light and
process it, and suggest mechanisms for the species distribution of Ruppia
and Zostera in the lower Chesapeake Bay. The results also show that
temperature largely influences the distribution of these plants. Ruppia
forms single species stands in shallow intertidal to shallow subtidal areas
where high light and high temperatures are prevalent during the summer.
Ruppia is generally more efficient at the higher light and temperature
regimes in these habitats. Zostera, which has the greater depth range, is
adapted to much lower light conditions as indicated by the lower light
saturation point and greater which is a function of the dark reaction under
optimal environmental conditions or a function of the inhibitor under
supoptimal conditions, ranged from 0.9 to 3.7 mg C g~l hr"1. 1'^
ranged from 110 to 225 uE m~2 sec"1 and Ik from 70 to 350 uE m~2
sec"1.
606
-------
Table 7. SUMMARY OF PHOTOSYNTHESIS-LIGHT EXPERIMENTS FOR SELECTED
SUBMERGED AQUATIC ANGIOSPERMSa (FROM KEMP ET AL. 1981c)
Plant Species
Zostera marina
n M
ii n
n n
Thalassia testudenum
n n
Cymodocca nodosa
n n
Halodule uninervis
Syringodium filiforme
Ruppia maritima
Vallisneria americana
Ceratophyllum demersum
n n
Ranunculus pseudof luitas
Myriophyllum spicatum
n n
n n
Potamogeton pectinatus
P. perfoliatus
pmaxb
1.5
2.2
1.2
1.3
1.7
2.5
2.6
1.5
1.6
3.7
1.9
2.2
3.2
2.2
3.3
2.8
1.9
1.3
0.9
1.1
Light
I'K
140
170
167
184
225
170
140
130
140
225
123
130
135
130
115
215
110
200
195
140
Parameters0
IK IG d
230
220
280
345
320
210
220
175
220
290
236
100
80
230
150
180
70
290
350
230
28
145
50
40
50
120
30
30
20
25
30
60
25
Reference
Drew 1979
Penhale 1977
McRoy 1974
Sand- Jensen 1977
Buesa 1975
Capone et al. 1979
Beer and Waisel 1979
Drew 1978
Beer and Waisel 1979
Buesa 1975
Nixon and Oviatt 1973
Titus and Adams 1979
Van et al. 1976
Guilizzoni 1977
Westlake 1967
Titus and Adams 1979
Van et al. 1976
Kemp et al. 1981c
Westlake 1967
Kemp et al. 1981c
a Most of these data were interpolated from graphical relations provided
by respective authors.
b Pmax is light-saturated photosynthetic rate in mg C g~l h~l, where
02 production data were converted to C assuming PQ = 1.2.
c Light variables: I'K = half-saturation constant; IK = intersection
of initial slope and Pmax; IQ = light compensation point where
apparent production approaches zero. Light data converted to PAR units
( uE m~2 sec"1) assuming 1 raW cm"2 = 2360 Lux = 0.86 cal cm"2
h"1 = 46 uE m~2 sec"1.
d Values for Ig are not available for experiments using the 14C method
which cannot measure negative net photosynthesis.
That submerged angiosperms have similar photosynthetic patterns is
useful from the management point of view where decisions often must be
based on information from only one or two species. However, to answer
detailed questions concerning species competition or species adaptations,
it is necessary to determine the interrelationship of photosynthetic
patterns, pigment complement, plant morphology, and community canopy
structure.
607
I
I
I
I
I
1
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
1
1
1
I
I
I
I
I
Thus, features in addition to photosynthetic parameters help determine
plant community photosynthesis. Canopy structure and chlorophyll content
were determined for a Ruppia-Zostera bed in the lower Chesapeake Bay
(Wetzel et al. 1982). Both Ruppia and Zostera showed a concentration of
leaf area (surface available for light absorption) at the lower portion of
the canopy where less light penetrates (Figure 14). The wider the bar, the
more concentrated the leaf material. This probably allows for a greater
overall net community photosynthesis than if there were a uniform vertical
distribution of leaf area. Highly significant differences were observed
between the vertical stratification of leaf area of Ruppia and Zostera.
Ruppia exhibits much greater leaf area than Zostera at the lower canopy (0
to 10 cm above substrate); this probably contributes to its success in the
mixed bed areas where it is shaded by Zostera.
Preliminary estimates of pigment content of Ruppia and Zostera suggest
differences between species (Figure 15). The highest concentrations of
chlorophyll are at mid-canopy for Zostera and at top-canopy for Ruppia
(Wetzel et al. 1982). Ruppia also showed a higher total chlorophyll
concentration than Zostera. This higher chlorophyll concentration in
combination with its canopy structure are adaptations that contribute to
Ruppia's success in mixed bed areas. These estimates give us information
on how changes in light quantity (from water quality changes) will affect
the success of mixed SAV beds.
Microcosm Studies
The microcosm studies of Kemp et al. (1981b) show a negative effect of
suspended sediments on Potamogeton photosynthesis (Figure 16). Two
concentrations of fine sediment particles (^64 m in diameter,
representative of particle size in nature), kept in suspension with
recirculating pumps, reduced light availability in the two treatments and
resulted in significantly lower photosynthesis of Potamogeton compared with
a control. Kemp et al. attributed about half the decrease in productivity
of treated systems to the accumulation of epiphytic solids on the plant
leaves. Further consideration of the microcosm data involved calculating
regressions between chlorophyll £ or filterable solids and light
attenuation coefficients. From these, it was concluded that in the
northern Bay, the effect of light attenuation by phytoplankton would be
small, however, the effect of suspended sediments on photosynthesis would
be significant.
In situ Studies of Community Response to Light
The effect of light on plant community metabolism was investigated in
upper and lower Chesapeake Bay grassbeds. In both areas, community
metabolism was estimated as oxygen production in large, transparent
incubation chambers. During these experiments, detailed measurements of
light energy (PAR) reaching the plants were made. In some experiments,
neutral density screens similar in design to the 14(j studies on
individual species were used to decrease available light.
A summary of the upper Bay Potamogeton community response to light is
presented in Figure 17, which includes estimates from both early (May) and
608
-------
AUGUST 1980
50-
40-
30-
20-
Ruppia BED
Ruppia mar Hi ma
MIXED BED
Ruppia maritime
Zostera BED
Zostera marina
MIXED BED
Zostera marina
LEAF AREA INDEX
Figure 14. Vertical2distribution of one-sided leaf area index (m'
plant m substrate) for Ruppia and Zostera at three
vegetated sites on the Eastern Shore, Virginia (from
Wetzel et al. 1982).
609
I
1
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
1
1
I
I
I
I
I
Ruppia
0
mg Chi. g fresh wt
Figure 15. Vertical distribution of total chlorophyll for Ruppia and
Zostera from a mixed bed area on the Eastern Shore, Virginia.
Values ± standard error, n = 3 (from Wetzel et al. 1982).
610
-------
•?'
0~
o>
co"
CO
LU
X
APPARENT PHOTOSY
<\J
'E
LU
a.
i-"
x
o
SS
wji 0 -
OOOmo
CONTROL
(C)
-(b)
ui
O
mg f
0
O
m
O
(Q )
PHOTOSYNTHETIC
RESPONSE TO
SEDIMENT LOADING
LIGHT AVAILABILITY
SUSPENDED SOLIDS
24 6 8 10
TIME OF EXPERIMENT, DAYS
Figure 16. Effect of (c) total suspended solids (TSS) on (b) light
availability and (a) rate of photosynthesis of Potamogeton
perfoliatus (from Kemp et al. 1981).
611
-------
I
I
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i
i •
-8
00
UJ
X
8 3
(D U-
O
tO
O
C\J
| i i 1 I |
O O
O
I
.-|M Ajp 6 .2Q Bui) AllAllOnaOyd
c
O
4J
c
>,
O
B
o
3
i-t
<*-(
si
oo
c
ca
r-l
a
3
o
ca
00
l-i
3
CO
rt
PQ
cfl
-------
I
late (August) periods in the growing season (Boynton, unpublished data). I
The Ic of the plant community occurs at about 200 uE m~2 sec~l, and
data suggest that the community is not light-saturated in the ranges of ^
measured in situ light flux. If the community were light-saturated, the •
rate of change would approach zero (Pmax) with the line in Figure 17
leveling off. An analysis of the seasonal trends suggests no differences
in the regression of light and community metabolism between seasons. •
Based on these and other studies, Kemp et al. (1981b) conclude that •
grass communities in the upper Bay are often light limited. For example,
actual subsurface light data and three theoretical light extinction M|
coefficients were used to calculate light penetration to a depth of 0.5 m •
above the substrate; a depth below which Potamogeton grows (Figures 18a,
18b). Photosynthetic parameters, Ic, I\, and Pmax were calculated
from a P-I curve (Figure I8c). These parameters are identified for each •
light penetration curve and suggest that for much of the daylight period, V
the plant community is light-limited or undersaturated, as it is not
operating at Pmax. At early morning and dusk periods of the day, the •
community is apparently heterotrophic (i.e., no net production). H
In the lower Bay, community metabolism studies were carried out in
three areas: Ruppia-dominated, Zostera-dominated, and a mixed _
Ruppia-Zostera area (Wetzel et al. 1982). These studies were conducted •
under a wide range of in situ light regimes and under artificial shading ™
conditions. The shallow Ruppia areas exhibited higher light and
temperature regimes than the deeper Zostera areas; the mixed bed was •
intermediate between the two. 9
Short-term shading experiments resulted in a general decrease in
community metabolism for both Ruppia and Zostera communities. For the •
Ruppia site, apparent productivity increased with increasing light to a •
midday peak and decreased during the early afternoon (Figure 19). Based on
P-I curves, Ruppia was light-saturated during much of the day and was not
photoinhibited. The unexplained afternoon depression that occurred while •
light was increasing may be due to increased community respiration rates •
under these high summer temperatures. A similar pattern was observed for
the Zostera site where shading also resulted in decreased apparent •
productivity (Figure 20). In contrast, the afternoon depression in £
productivity rates of the Zostera bed was not so dramatic as in the Ruppia
bed. This trend in Zostera seemed to follow the decreasing light —
availability unlike the response in Ruppia. These results are similar to •
those found throughout the study and support previous conclusions that the
two communities are physiologically (i.e., temperature and light response)
quite different. •
Plots of apparent productivity versus light flux at the top of the 9
canopy were used to compare all three habitats (Figure 21). Differences
among the three sites were characteristically observed for these summer M,
experiments. Both the Ruppia and the mixed bed areas showed decreases in B
apparent productivity at the highest light fluxes. The Zostera site, which
did not receive the high light that other sites received, showed no
decrease in rates. P-I curves for the seagrass species showed no •
photoinhibition, even at high summer temperatures, and suggested that the W
Pmax of Ruppia should be greater than Zostera at this time of the year.
As evidenced by its high apparent productivity rates, Zostera appears ft
adapted to lower light levels. The erratic pattern of data points and the ••
greater number of negative rates for Ruppia strongly suggest different
community behavior. At the community M
613
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
1500-
e
LU
co
CO
CO
o
o
I
O.
I-
LL)
OC
cf
1000-
500-
250-
500-
AIR-WATER INTERFACE
I
0900 1200 1500
TIME (hr)
200 600
1000
LIGHT FLUX (//Em"2 s"')
Figure 18. Diagramatic representation of (a) surface and underwater light
flux at Todds Cove, upper Chesapeake Bay calculated for three
light extinction (K) coefficients, (b) I , I and P
calculated from P-I curve of Potamogeton perfoliatus^from
Kemp et al. 1981c).
614
-------
AiiAiionaoyd iN3yvddv
xmj
CO ,-v
> <4-t CM
O CO
ft CTl
O 5-5 r-t
4-1 O
CO •
>> i-l
ft -O cfl
O C
C Cfl 4-1
CO 0)
a «
O rH
CD m a)
4= N
4-) f> 4-1
r-H CU
4J t^. JS
CO
x o o
3 O i-(
iH iH U-l
•u co a
x: o
W) W 4-1
•H 4-1
iH C >,
0) D-
T3 g O
C i-l C
Cfl !-< CO
II) O
>> ft
4-1 X t>
•H ni
4-J -H
s a
•B,^1
>-l "SO
ft O -H
i-l rH
4-1 UH
(3 4J
0) !>, CS
M CO 0)
CO "O -H
ft ^2
ft LH g
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
z--
o
o
m
O
O
O
O
ro
1
O
O
CM
O
O
O
O
O
O
CM
i
O
o
CO
1^
8
CM
8 8
00 (O
1H9I1
.8
o
o
CD
O <
.o a
S u.
o
UJ
O
•8
o
.o
o
o
'8
• o oo
en ro ,
oo (j a,
•r-l 0) O
CO
U
cfl «
4J -U
•H CO
> O
•H N
U >-i 4S
3 O bO
T3 u-i -H
O ^H
W >-.
a, to 4J
T3 S
W OJ
S U-i -H
M 6
« 0) nj
a, e
O, -H U-l
< 4-1 O
o
CM
0)
J-l
oo
•H
616
-------
o
00
-fO
-CM
• x:::.
O
o
oo
0>
I
JQ
-o
-I- ^
CM
• ro i
6
UJ
-CM X
-O
O
CO
on
O
«
N-
-ro
-O
O
o
in
c: x
(U -H «
^ e v
03 !-(
O
a
<; v
(U
J-i
3
60
co
iN3HVddV
617
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
level, the differences may be due to differences in community respiration
rates, plant species photorespiration rates, or the photosynthetic pattern
of other primary producers such as macro- and microalgae. The mixed bed
site shows an intermediate pattern, suggesting an interactive effect of the
presence of both species of seagrass. Under the influence of changes in
water quality, these data show that mixed beds would probably survive
better than a bed containing a single species.
A summary of linear regression analyses of apparent productivity versus
light flux at the top of the canopy for the three areas is presented in
Table 8. At the community level, the correlation coefficient, r, is
strongly influenced by season, with the lower values generally observed for
the winter months. These are the times of year of clearest water, and the
specific rate of Q£ productivity asymptotically approaches Pmax.
Therefore the linear relationship does not adequately describe the
Table 8. APPARENT 02 PRODUCTIVITY AND LIGHT: LINEAR REGRESSION
ANALYSIS FOR LOWER BAY STUDIES (FROM WETZEL ET AL. 1982)
[mg 02 m~2 h~l
DATE
14 Feb
21 Feb
19 Mar
29 Apr
2 May
2 Jun
5 Jun
9 Jul
16 Jul
19 Aug
23 Sep
7 May
11 Jul
21 Aug
25 Sep
26 Sep
AREA
80 Zostera
80
80 "
80
80 "
80
80 "
80 "
80
80
80 "
80 Ruppia
80 "
80
80 "
80
N
33
36
31
20
11
20
30
57
76
16
27
10
83
26
10
16
vs. uE m~2 If1 (AT CANOPY TOP)]
m
68
78
65
280
582
307
286
96
124
89
108
363
52
385
242
323
.1
.0
.4
.5
.2
.1
.5
.5
.2
b
86
157
105
-183
-267
-472
-309
-147
- 67
- 84
-159
-357
- 47
-434
- 79
-194
r uE m~2
.5
.1
.5
.8
.2
.1
.5
0.372
0.360
0.210
0.778
0.823
0.681
0.765
0.425
0.542
0.793
0.435
0.980
0.215
0.770
0.806
0.532
0
0
1
1
1
0
0
1
0
0
1
0
0
!IC
h"1 uE
-
-
-
.650
.459
.54
.08
.52
.541
.947
.48
.983
.899
.13
.326
.602
_9 _
m *-sec
-
-
-
181
127
427
300
423
150
203
411
273
250
313
90
167
1
.6
.2
618
-------
Table 8. (CONTINUED)
[mg 02 m~2 h'1 vs. uE m~2 h"1 (AT CANOPY TOP)]
DATE AREA N m b r uE m~2 h~l uE tn~2sec~l
5
14
May
Jul
80 Mixed
80 "
28
50
89
77
.7
.9
-189
- 48.9
0
0
.607
.553
2
0
.11
.627
585
174
1
N = number of observations
m = slope
b = y-intercept
r = correlation coefficient
Ic = estimated light compensation point (x-intercept)
photosynthetic response. This is true for all measures taken at or near
In the Zostera community, maximum rates occur in the spring and early
summer. Over this period, the estimated community light compensation point
progressively increases, because of increased respiration, to the point that
daily community production is negative. This corresponds to the
characteristic midsummer die off of Zostera in these areas (Wetzel et al.
1981). Except for the studies carried out in winter and early spring
(February and March), the community as a whole is light-limited.
The Ruppia community dominates the higher light and temperature areas of
the bed. Maximum rates of apparent photosynthesis occur during the summer,
and they corroborate the earlier conclusions that Ruppia has both higher
Pmax and Ic characteristics. Some data suggest that community respiration
increases in early afternoon during high light and temperature conditions.
These conditions are prevalent at midday low tides during July and August.
Overall, Ruppia-dominated communities in the lower Bay appear adapted to
increased light and temperature regimes and do not appear light-limited in the
Vaucluse Shores study area.
For Chesapeake Bay system as a whole, these data and similar studies
completed in upper-Bay communities suggest the extreme sensitivity of Bay
grasses to available light. These data also agree very well with information
on other geographical areas and species. The general conclusion is that light
and factors governing light energy availability to submerged aquatic vascular
plants are principal controlling forces for growth and survival.
619
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
SECTION 4
SUMMARY
The apparent optical properties of estuarine water create, in general,
a light-limited environment for the process of photosynthesis. Water in
itself, suspended particles, and dissolved compounds all interact to both
attenuate total photosynthetically active radiation as well as to
spectrally shift (selectively absorb) wavelengths most important for
autotrophic production. Plant pigment systems, in general, are adapted for
efficient light-energy capture in relatively narrow bands. In many cases,
it is precisely these wavelengths that are most rapidly attenuated in the
estuarine water column.
However, data on spectral characteristics and specific waveband
attenuation in estuarine and coastal environments are lacking. Our summary
of available data, Section 2, indicates that few studies have been
completed that characterize these optical properties of estuarine waters
and even fewer that can evaluate the data in terms of potential control on
rates of photosynthesis. It is difficult, therefore, if not impossible at
the present time, to speculate as to the importance or generality of
specific waveband attenuation relative to photosynthesis and to autotrophic
production in Chesapeake Bay as well as in other estuaries. It has only
been within the past few years that submarine spectral irradiance studies
have become technologically feasible, and this is reflected in the general
paucity of information.
Studies in Chesapeake Bay indicate reductions in both light quality and
quantity at selected study sites and during various periods of the growing
season for submerged aquatic plants. Recent measures of diffuse
downwelling attenuation coefficients (Section 2) in lower Bay communities
indicate a severe attenuation of light energy in the photosynthetically
important violet blue (400 to 500 nm) region and in the near infrared (700
to 775 nm) region of the spectrum. Also for the March through July period
of study, there appears to be a progressive increase in attenuation in
these spectral regions.
Comparison of vegetated and non-vegetated areas in Chesapeake Bay with
regard to light quality and quantity suggests some improvement (lower
attenuation) in the vegetated areas, although the data are quite variable.
In the upper Bay, Kaumeyer et al. (1981) report significant differences for
one site and not for another. In the lower Bay, comparison of four sites
(two vegetated and two non-vegetated) indicates some differences in light
quality. There are at these lower Bay sites, some improvements in
attenuation in the 400 to 500 nm region in spring months (see recent report
by Wetzel et al. 1982 for an updated analysis of this and additional
data). The only definitive light quality differences between the sites was
reduced attenuation in the 500 to 700 nm region in vegetated areas during
spring, an important period in the growth of Zostera dominated
communities. Diffuse downwelling attenuation in some photosynthetically
sensitive spectral regions is severe. This, coupled with the general
increase in attenuation during the growing season and at higher
temperatures, indicates the plant communities are undoubtedly light
stressed.
620
-------
I
There is a much larger data base on plant response to total available B
light energy (PAR) for Chesapeake Bay as well as for other bodies of 9
water. The dominant plant species in the Bay show the classical,
hyperbolic photosynthetic response to increasing PAR. Specific plant •
response studies suggest physiological differences among species. The |
dominant upper Bay species, Myriophyllum spicatum and Potamogeton
perfoliatus, light-saturate between 600 and 800 uE m~2 Sec~L, but m
differ in Pmax and 1^. M. spicatum appears adapted to higher light •
conditions than ?_. perfoTTatus. In a similar manner, the dominant lower ™
Bay species, Ruppia maritima and Zostera marina, appear physiologically
different with regard to light response. R. maritima is adapted to high •
light and temperature; Z. marina is adapted to lower light regimes and is B
stressed at higher, summer temperatures.
In situ studies of entire plant communities in both Maryland and «
Virginia indicate that the communities are, in general, operating under I
sub-optimal light conditions. There was no apparent light saturation *
reached for upper-Bay communities; that is, net apparent community
productivity did not asymptotically approach a maximum value. Studies in B
lower-Bay communities suggest that Z^ marina is light-limited the majority B
of its growing seasons and only in more shallow R.. maritima areas did the
community photosynthetic response become light-saturated. These results •
indicate that, at least in terms of total PAR energy and probably because I
of the extreme attenuation in the 400 to 500 nm region noted earlier,
submerged plant communities in Chesapeake Bay as a whole are light-stressed.
Historical data relative to light (turbidity and indirectly, nutrients) B
and to past distribution and abundance on submerged aquatics indicate ™
progressive Bay-wide changes in systems structure and function. Heinle et
al. (1980) and Orth et al. (1971) discuss these in detail. In terms of Bay
grasses and the light environment, two overall conclusions of these reports
are particularily important. Heinle et al. (1980) note and document the
generalized increase in nutrients (and loadings) and chlorophyll »
concentrations in major tributaries of Chesapeake Bay over the past several fl
decades. Orth et al. (1981) conclude, for roughly the same time scale,
that the general pattern of disappearance of submerged plant communities
follows a "down-river" pattern. It also appears that upper-Bay and western B
shore lower-Bay communities have been the most severely impacted. These B
conclusions, together with our studies on the light environment and
photosynthesis-light relations in SAV ecosystems, suggest that total PAR •
and factors increasing diffuse downwelling attenuation in the 400-500 nm £
region are principal driving functions controlling plant growth and
survival. The specific factors at present that appear to have the greatest
impact are suspended particles, both organic and inorganic, which are B
controlled, in large part, by climatic conditions (runoff and nutrient ™
loading) and indirectly by associated changes in physical-chemical regimes
(salinity and temperature). B
In summary, it appears that Bay grasses are living in a marginal light B
environment, and that progressive changes in water quality as discussed by
Heinle et al. (1980) will further stress plant communities. To conclude m
I
621
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
that light has been singularily responsible for recent declines in the
vegetation goes beyond the data available. The data do indicate, however,
the extreme sensitivity of vegetation to both qualitative and quantitative
reductions of available light, and that over the past several decades water
quality throughout the Bay, particularily in the tributaries, has
progressively declined. Further changes in these parameters can only
affect Bay grasses in an adverse way. Results show that SAV can adapt to
changes in the availability of light. Long-term shading experiments (in
progress) will address this question further.
622
-------
LITERATURE CITED
Anderson, F. E. 1980. The Variation in Suspended Sediment and Water
Properties in the Flood-Water Front Traversing a Tidal Flat. Estuaries
3(l):28-37.
I
I
Backman, R. W., and D. C. Barilotti. 1976. Irradiance Reduction: Effects •
on Standing Crops of the Eelgrass Zostera marina in a Coastal Lagoon. I
Mar. Biol. 34:33-40. •
Beer, S., and Y. Waisel. 1979. Some Photosynthetic Carbon Fixation •
Properties of Seagrasses. Aquat. Bot. 7:129-138. (P
Biebl, R., and C. P. McRoy. 1971. Plasmatic Resistance and Rate of «
Respiration and Photosynthesis of Zostera marina at Different •
Salinities and Temperatures. Mar. Biol. 8:48-56.
Booth, C. R., and P. Dunstan. 1979. Diver-Operable Multiwavelength •
Radiometer. In: Measurements of Optical Radiations. Soc. •
Photo-optical Instrumentation Engineers. 196:33-39.
Bowes, G., T. K. Van, L. A. Barrard, and W. T. Haller. 1977. Adaption to Q
Low Light Levels by Hydrilla. J. Aquat. Plant Mgt. 15:32-35.
Buesa, R.J. 1975. Population Biomass and Metabolic Rates of Marine •
Angiosperms on the Northwestern Cuban Shelf. Aquat. Bot. 1:11-25. *
Burt, W. V. 1953. Extinction of Light by Filter Passing Matter in •
Chesapeake Bay Waters. Science 118:386-387. |
Burt, W. V. 1955a. Interpretation of Spectrophotometer Readings on m
Chesapeake Bay Waters. J. Mar. Res. 14:33-46. •
Burt, W. V. 1955b. Distribution of Suspended Materials in Chesapeake Bay. _
J. Mar. Res. 14:47-62. I
Burt, W. V. 1958. Selective Transmission of Light in Tropical Pacific
Waters. Deep-Sea Res. 5:51-61. •
Capone, D.G., P.A. Penhale, R.S. Oremland, and B.F. Taylor. 1979.
Relationship Between Productivity and ^2^2^-2^ Fixation in a • •
Thalassia testudinum Community. Limnol. Oceanogr. 24:117-125. M
Champ, M. A., G. A. Gould, III, W. E. Bozzo, S. G. Ackleson, K. C. Vierra.
1980. Characterization of Light Extinction and Attenuation in •
Chesapeake Bay, August, 1977. In: Estuarine Perspectives. V. S. 9
Kennedy, ed. Academic Press Inc., NY 263-267 pp.
Clarke, G. L., and G. C. Ewing. 1974. Remote Spectroscopy of the Sea for £
Biological Production Studies. In: Optical Aspects of Oceanography.
N. G. Jerlov and E. Steemann-Nielsen, eds. Academic Press, NY 389-413 M
PP- I
623
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Clarke, G. L., and H. R. James. 1939. Laboratory Analysis of the Selective
Absorption of Light by Seawater. J. Opt. Soc. Am. 29:43-55.
Congdon, R. A., and A. J. McComb. 1979. Productivity of Ruppia: Seasonal
Changes and Dependence on Light in an Australian Estuary. Aquat. Bot.
6:121-132.
den Hartog, C. 1970. The Seagrasses of the World. North Holland,
Amsterdam. 275 pp.
den Hartog, C., and P. J. G. Polderman. 1975. Changes in the Seagrasses
Population of the Dutch Waddenzee. Aquat. Bot. 1:141-147.
Dennison, W. 1979. Light Adaptation of Plants: A Model Based on the
Seagrass Zostera marina L. M. S. Thesis, Univ. Alaska, Fairbanks. 69
pp.
Drew, E. A. 1978. Factors Affecting Photosynthesis and Its Seasonal
Variation in the Seagrasses Cymodocea nodosa (Ucria) Aschers, and
Posidonia oceanica (L.) Delile in the Mediterranean. J. Exp. Mar.
Biol. Ecol. 31:173-194.
Drew, E. A. 1979. Physiological Aspects of Primary Production in
Seagrasses. Aquat. Bot. 7:139-150.
Drew, E. A., and B. P. Jupp. 1976. Some Aspects of the Growth of
Posidonia oceanica in Malta. In: Underwater Research. E. A. Drew, J.
N. Lythgoe, and J. D. Woods, eds. Academic Press, London, pp. 357-367.
Dubinsky, Z. and T. Herman. 1979. Seasonal Changes in the Spectral
Composition of
Downwelling Irradiance in Lake Kinneret (Israel). Limnol. Oceanogr.
24(4):652-663.
Evans, G.C. 1972. The Quantitative Analysis of Plant Growth. Univ.
California Press, Berkeley. 734 pp.
Fleischer, P., T. A. Gosink, W. S. Hanna, J. C. Ludwick, D. E. Bowker and
W. G. White. 1976. Correlation of Chlorophyll, Suspended Matter, and
Related Parameters of Waters in the Lower Chesapeake Bay Area to
Landsat-1 Imagery. Inst. of Oceanography, Old Dominion University
Technical Report No. 28, Norfolk, VA. 125 pp.
Flemer, D. A., D. H. Hamilton, C. W. Keefe, and J. A. Mihursky. 1970.
Final Report to Office of Water Resources Research on the Effects of
Thermal Loading and Water Quality on Estuarine Primary Production.
Contract No. 14-01-0001-1979. Office of Water Resources Research.
U.S. Dept. of Interior, Washington, D.C. NRI Ref. No. 71-6, Chesapeake
Biological Laboratory, Solomons, MD.
Gates, D. M. 1971. The Flow of Energy in the Biosphere. Sci. Amer.
224(3):88-103.
624
-------
I
Ginsburg, R. N., and H. A. Lowenstam. 1957. The Influence of Marine Bottom I
Communities on the Depositional Environment of Sediments. J. Geol. B
66:310-318.
Gross, M. G., M. Karweit, W. B. Cronin, and J. R. Schubel. 1978. Suspended I
Sediment Discharge of the Susquehanna River to Northern Chesapeake Bay,
1966 to 1976. Estuaries 1:106-110. _
Gullizzoni, P. 1977 Photosynthesis of the Submergent Macrophyte *
Ceratophyllum demeroum in Lake Wingia. Wise. Acad. Sci. Arts Lett.
65:152-162. •
Heinle, D. R., C. F. D'Elia, J. L. Taft, J. S. Wilson, M. Cole-Jones,
A. B. Caplins and L. E. Cronin. 1980. Historical Review of Water .
Quality and Climatic Data from Chesapeake Bay with Emphasis on Effects •
of Enrichment. Report to U.S. Environmental Protection Agency, *
Chesapeake Bay Program. Chesapeake Research Consortium, Inc. Pub. No.
84. Univ. MD. Ctr. Environmental and Estuarine Studies. No. 80-15 CBL. I
Hurlburt, E. A. 1945. Optics of Distilled and Natural Water. J. Opt.
Soc. Amer. 35:689-705. •
Idso, S. B., and R. G. Gilbert. 1974. On the Universality of the Poole
and Atkins Secchi Disk-Light Extinction Equation. J. Appl. Ecol. _
11:399-401. •
Inada, K. 1976. Action Spectra for Photosynthesis in Higher Plants. Plant
Cell Physiol. 17:355-365. •
Jacobs, R. P. W. M. 1979. Distribution and Aspects of the Production and
Biomass of Eelgrass, Zostera marina L., at Roscoff, France. Aquat. M
Bot. 7:151-172. J
James, H. R., and E. A. Birge. 1938. A Laboratory Study of the Absorption
of Light by Lake Waters. Trans. Wise. Acad. Sci. 31:154 pp. •
Jassby, A.D., and T. Platt. 1976. Mathematical Formulation of the
Relationship Between Photosynthesis and Light for Phytoplankton. •
Limnol. Oceanog. 21:540-547. 0
Jerlov, N. C. 1976. Marine Optics. Elseview Oceanography Series. 14. ^
Elsevier Scientific Publ. Co., NY 231 pp. •
Johnstone, I. M. 1979. Papua New Guinea Seagrasses and Aspects of the
Biology and Growth of Enhalus acoroides (L.f.) Royle. Aquat. Bot. •
7:197-208. •
Kalle, K. 1966. The Problem of Gelbstoff in the Sea. Oceanogr. Mar.
Ann. Rev. 4:91-104.
I
Kaumeyer, K., W. R. Boynton, L. Lubbers, K. Staver, S. Bunker, W. M. Kemp
and J. C. Means. 1981. Metabolism and Biomass of Submerged Macrophyte •
P.nmmiin i f* i o c in Nr»i-t-ho i-n P.Jid tzanf* a Vo Rflv. Tn* CIilhmPTaoH A/inafir* ^*
Communities in Northern Chesapeake Bay. In: Submerged Aquatic
625
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Vegetation in Chesapeake Bay: Its Ecological Role in Bay Ecosystems
and Factors Leading to Its Decline. W.M. Kemp, W.R. Boynton, J.C.
Stevenson, and J.Means, eds. Report to U.S. Environmental Protection
Agency, Chesapeake Bay Program.
Kemp, W. M., M. R. Lewis, T. W. Jones, J. J. Cunningham, J. C. Stevenson,
and W. R. Boynton. 1981a. Measuring productivity of submerged aquatic
macrophytes: a comparison of methodologies. Final Report, U.S. EPA,
Annapolis, MD.
Kemp, W.M., J.J. Cunningham, M.R. Lewis, R.Hanson, R. Twilley, T.W. Jones,
J.C. Stevenson, J. Perron, S. Grarbrandt. 1981b. Experimental
Observations of Turbidity/Light Relations and Their Influence on
Submerged Macrophytes in Northern Chesapeake Bay. In: Submerged
Aquatic Vegetation in Chesapeake Bay: Its Ecological Role in Bay
Ecosystems and Factors Leading to Its Decline. W.M. Kemp, W.R.
Boynton, J.C. Stevenson, and J.Means, eds. Final Report to U.S.
Environmental Protection Agency, Chesapeake Bay Program.
Kemp, W.M., W.R. Boynton, J.C. Stevenson, and J. Means. Eds. 1981c.
Submerged Aquatic Vegetation in Chesapeake Bay: Its Ecological Role in
Bay Ecosystems and Factors Leading to its Decline. Final Grant Report
to U.S. Environmental Protection Agency, Chesapeake Bay Program,
Annapolis, MD.
Kiefer, D.A., and R.W. Austin. 1974. The Effect of Varying Phytoplankton
Concentration on Submarine Light Transmission in the Gulf of
California. Limnol. Oceanor. 19:55-64.
Kiley, K.P. 1980. The Relationship between Wind and Current in the York
River Estuary, Virginia, April 1973. M.A. Thesis, School of Marine
Science, The College of William and Mary, VA. 195 pp.
Kiorbe, T. 1980. Production of Ruppia cirrhosa (Petagna) Grande in Mixed
Beds in Ringkobing Fjord (Denmark). Aquat. Bot. 9:125-143.
Kranck, K. 1980. Variability of Particulate Matter in a Small Coastal
Inlet. Can. J. Fish. Aquat. Sci. 37:1209-1215.
McCluney, W. R. 1975. Radiometry of Water Turbidity Measurements. J.
Water Pollut. Control Fed. 47:252-266.
McMillan, C., and R. C. Phillips. 1979. Differentiation in Habitat
Response Among Populations of New World Seagrasses. Aquat. Bot.
7:185-196.
McRoy, C.P. 1974. Seagrass Productivity: Carbon Uptake Experiments in
Eelgrass, Zostera marina. Aquaculture. 4:131-137.
626
-------
I
McRoy, C.P., and C. McMillan. 1979. Production Ecology and Physiology of •
Seagrasses. In: Seagrass Ecosystems: A Scientific Perspective. (|
McRoy, C. P., and C. Helferrich, eds. Marcel Dekker, Inc., NY. pp.
53-87. .
Mihursky, J.A., and W. R. Boynton. 1978. Review of Patuxent River Data
Base. Univ. Md. Ctr. Environmental and Estuarine Studies Ref. No.
78-157-CBL.
I
Mukai, H., K. Aioi, and Y. Ishida. 1980. Distribution and Biomass of
Eelgrass Zostera marina L. and Other Seagrasses in Odawa, Central •
Japan. Aquat. Bot. 8:337-342. |
Nash, C. B. 1947. Environmental Characteristics of a River Estuary. J. .
Mar. Res. 6:147-174. •
Nienhuis, P.H., and B.H.H. DeBree. 1977. Production and Ecology of
Eelgrass (Zostera marina L.) in the Grevelingen Estuary, The I
Netherlands, Before and After the Closure. Hydrobiologia 52:55-66. |
Nixon, S.W. and C. Oviatt. 1973. Preliminary Measurements of Midsummer M
Metabolism in Beds of Eelgrass, Zostera marina. Ecology. 53:150-153. •
Orth, R. J. 1977. Effect of Nutrient Enrichment on Growth of the Eelgrass
Zostera marina in the Chesapeake Bay, Virginia, U.S.A. Mar. Biol. •
44:187-194. 9
Orth, R.J., K.A. Moore, M.H. Silberhorn and H.H. Gordon. 1979. The •
Biology, Propagation and Impact of Herbicides on Zostera Marina. US |
EPA Chesapeake Bay Program.
Parsons, T.R., M. Takahashi, and B. Hargrave. 1977. Biological Oceano- •
graphical Processes. Pergamon Press, NY. 332 pp.
Patten, B.C., R.A. Mulford, and J.E. Warriner. 1963. An Annual B
Phytoplankton Cycle in Chesapeake Bay. Chesapeake Sci. 4:1-20.
I
Penhale, P.A. 1977. Macrophyte-Epiphyte Biomass and Productivity in an
Eelgrass (Zostera marina L.) Community. J. Exp. Mar. Biol. Ecol.
26:211-224.
Penhale, P.A., and W.O. Smith, Jr. 1977. Excretion of Dissolved Organic •
Carbon by Eelgrass (Zostera marina) and Its Epiphytes. Limnol. *
Oceanogr. 22:400-407.
Phillips, G. L., D. Eminson, and B. Moss. 1978. A Mechanism to Account for •
Macrophyte Decline in Progressively Eutrophicated Freshwaters. Aquat.
Bot. 4:103-126. m
627
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Pierce, J.W., D.L. Correll, M.A. Faust, W.H. Klein, and B. Goldberg.
1981. Spectral Quality of Underwater Light in a Turbid Estuary, Rhode
River, Maryland, U.S.A. Unpublished Manuscript.
Prieur, L., and S. Sathyendranath. 1981. An Optical Classification of
Coastal and Oceanic Waters Based on the Specific Spectral Absorption
Curves of Phytoplankton Pigments, Dissolved Organic Matter, and Other
Particulate Materials. Liranol. Oceanogr. 26:671-689.
Riaux, C., and J. L. Douville. 1980. Short-Term Variations in
Phytoplankton Biomass in a Tidal Estuary in Northern Brittany.
Estuarine Coastal Mar. Sci. 10:85-92.
Sand-Jensen, K. 1977. Effect of Epiphytes on Eelgrass Photosynthesis.
Aquat. Bot. 3:55-63.
Sauberer, F., and F. Ruttner. 1941. Die Strahlungsverhaltnisse der
Binnengewasser. Akademic Verlag, Berlin.
Scoffin, T. P. 1970. The Trapping and Binding of Subtidal Carbonate
Sediments by Marine Vegetation in Bimini Lagoon, Bahamas. J. Sed.
Petrol. 40:249-273.
Scott, B. D. 1978. Phytoplankton Distribution and Light Attenuation in
Port Hacking Estuary. Aust. J. Mar. Freshwater Res. 29:31-44.
Seliger, H.H., and M.E. Loftus. 1974. Growth and Dissipation of Phyto-
plankton in Chesapeake Bay. II. A Statistical Analysis of
Phytoplankton Standing Crops in the Rhode and West Rivers and an
Adjacent Section of the Chesapeake Bay. Chesapeake Sci. 15:185-204.
Spence, B.H.N., and J. Chrystal. 1970a. Photosynthesis and Zonation of
Fresh-Water Macrophytes. I. Depth Distribution and Shade Tolerance.
New Phytol. 69:205-215.
Spence, B.H.N., and J. Chrystal. 1970b. Photosynthesis and Zonation of
Fresh-Water Macrophytes. II. Adaptability of Species of Deep and
Shallow Water. New Phytol. 69:217-227.
Steeman-Nielsen, E. 1975. Marine Photosynthesis with Special Emphasis on
the Ecological Aspects. Elsevier, Amsterdam.
Stevenson, J.C., and N.M. Confer. 1978. Summary of Available Information
on Chesapeake Bay Submerged Vegetation. Fish and Wildlife Service
Biological Services Program. FWS/OBS/78/66, U.S. Dept. of Interior.
335 pp.
Stirban, M. 1968. Relationship Between the Assirailatory Pigments, the
Intensity of Chlorophyll Fluroescence and the Level of the
Photosynthesis Zone in Zostera marina L. Rev. Rovm. Biol. Serv. Bot.
13:291-295.
628
-------
I
Tailing, J. F. 1957. Photosynthetic Characteristics of Some Fresh-Water •
Plankton Diatoms in Relation to Underwater Radiation. New Phytol. •
56:29-50.
Thompson, M.J., L.E. Gilliland, and L.F. Rosenfeld. 1979. Light I
Scattering and Extinction in a Highly Turbid Coastal Inlet. Estuaries ™
2(3):164-171.
Titus, J. E., and M, S. Adams. 1979. Coexistence and the Comparative Light m
Relations of the Submersed Macrophytes Myriophyllum spicatum L. and
Vallisneria americana Michx. Oecologia (fieri) 40:373-386. •
Van, T.K., W.T. Haller, and G. Bowles. 1976. Comparison of the Photo-
synthetic Characteristics of Three Submersed Aquatic Plants. Plant
Physiol. 58:761-768. •
van Tine, R. 1981. Ecology of Seaweeds and Seagrasses in a Thermally
Impacted Estuary in the Eastern Gulf of Mexico. In: Proc. Vlllth •
Intl. Seaweed Symp., Aug. 1974. G. E. Fogg, and W. D. Jones, eds. The |
Marine Science Labs Menai, Bridge, Univ. College of North Wales, Wales,
U.K. pp. 499-506. _
Westlake, C.F. 1967. Some Effects of Low Velocity Currents on the •
Metabolism of Aquatic Macrophytes. J. Exp. Bot. 18:187-205.
Wetzel, R.G., and P.A. Penhale. 1979. Transport of Carbon and Excretion •
of Dissolved Organic Carbon by Leaves and Roots/Rhizomes in Seagrass
and Their Epiphytes. Aquat. Bot. 6:149-158. •
Wetzel, R.L., K.L. Webb, P.A. Penhale, R.J. Orth, D.F. Boesch, G.W.
Boehlert, and J.V. Merriner. 1979. The Functional Ecology of _
Submerged Aquatic Vegetation in the Lower Chesapeake Bay. Annual Grant •
Report, U.S. EPA R805974, Chesapeake Bay Program, Annopolis, MD. •
Wetzel, R.L., P.A. Penhale, R.F. van Tine, L. Murray, A. Evans, and K.L.
Webb. 1982. Primary Productivity, Community Metabolism, and Nutrient
Cycling. In: Functional Ecology of Submerged Aquatic Vegetation in
the Lower Chesapeake Bay. R.L. Wetzel, ed. Final Report U.S. M
Environmental Protection Agency, Chesapeake Bay Program, Annapolis, MD. •
Wiginton, J.R., and C. McMillan. 1979. Chlorophyll Composition under
Controlled Light Conditions as Related to the Distribution of •
Seagrasses in Texas and the U.S. Virgin Islands. Aquat. Bot. •
6:171-184.
Williams, J. 1980. If Increased Turbidity is a Contributing Factor in (|
SAV Degradation, What are the Major Causes of Increased Turbidity?
Oral Presentation at Atlantic Estuarine Research Society Meetings, Nov. M
6-8, 1980. Virginia Beach, VA. •
Williams, S. F. 1977. Seagrass Productivity: The Effects of Light on
Carbon Uptake. M. S. Thesis. University of Alaska, Fairbanks. 95 pp. B
Yentsch, C. S. 1960. The Influence of Phytoplankton Pigments on the Colour
of Seawater. Deep-Sea Res. 7:1-9. M
I
629
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Distribution of Submerged
Vascular Plants in
the Chesapeake Bay
Distribution and Abundance
of SAV in the Lower
Chesapeake Bay, 1978, 1979
Distribution of Submerged
Aquatic Vegetation in
Chesapeake Bay, Maryland
Biostratigraphy of the
Chesapeake Bay and its
Tributaries
Zostera Marina; Biology
Propagation and Impact
Herbicides
Submerged Aquatic
Vegetation in the
Chesapeake Bay: Its Role
in the Bay Ecosystem and
Factors Leading to its
Decline
The Functional Ecology of
Submerged Aquatic
Vegetation in the Lower
Chesapeake Bay
Value of Vegetated Habitats
and Their Roles as Nursery
Areas and Shelter from
Predation
Assessment of Potential
Impact of Industrial
Effluents on Submerged
Aquatic Vegetation
Effects of Recreational
Boating on Turbidity and
Sedimentation Rates in
Relationship to Submerged
Aquatic Vegetation
Factors Affecting, and
Importance of Submerged
Aquatic Vegetation in
Chesapeake Bay
Appendix A
CBP Submerged Aquatic Vegetation
Richard R. Anderson
Robert J. Orth
Robert J. Macomer
Grace Brush
Robert J. Orth
J. Court Stevenson
W.M. Kemp
W.R. Boynton
Projects
American University
Virginia Institute of
Marine Science
Chesapeake Bay
Foundation
John Hopkins University
Virginia Institute of
Marine Science
U. of MD, Center for
Environmental and
Estuarine Studies
R.L. Wetzel
R.J. Orth
J.V. Merriner
K.L. Heck, Jr.
G.E. Walsh
Jerome Williams
Herman Gucinski
V. Valentine
Virginia Institute of
Marine Science
Academy of Natural
Sciences of
Philadelphia
U.S. E.P.A. Gulf
Breeze Environmental
Research Labs.
U.S. Naval Academy
U.S. Fish and Wildlife
Service
630
-------
I
Submerged Aquatic Vegetation: Robert J. Orth Virginia Institute of ™
Distribution and Abundance Marine Science
in the Lower Chesapeake Bay •
and Interactive Studies of |
Light, Epiphytes, and
Grazers •
Environmental Regulation R. L. Wetzel Virginia Institute of
of Zostera marina and Marine Science
Ruppia maritima: Growth •
and Metabolism V
Synthesis of Ecological W.M. Kemp U. of MD, Center for •
Research from U.S. EPA's W.R. Boynton Environmental and •
Chesapeake Bay Program: J.D. Stevenson Esturarine Studies
A Continuing Effort J.C. Means _
1981-1982. •
631
I
I
I
I
I
I
I
I
I
I
I
I
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
SUMMARY AND CONCLUSIONS
As previously indicated, the rationale for conducting an intensive
study of SAV was founded in the perceived fact that the distribution and
abundance of the Bay grasses had significantly declined during the early
1970s and the intuitive feeling that the Chesapeake Bay ecosystem was
healthier when the grasses were more abundant. The general feeling was
that an overall degradation of the quality of the Bay's estuarine and
riverine waters was, in some way, involved in the decline. Overall, the
SAV research was based on a series of questions explained in the
introduction to this part. This summary highlights the findings and
conclusions from the CBP-SAV research as synthesized in the previous
chapter, and attempts to answer these questions.
Although there is no scientific way to measure the exact distribution
of SAV throughout the Bay some 50 or 100 years ago for use as a baseline
against which to compare current populations, selected areas were studied
using archival aerial photography and biostratigraphic analysis of bottom
cores. Essentially, this work revealed that an unprecedented decline in
SAV populations occurred during the period of 1965 to 1980. The decline
was not species-specific, and therefore, was not felt to be the result of
disease or some similar natural perturbation.
Overall, the pattern of the decline appears to have been "down river",
(from up river, down to the lower estuarine portion, and from up-estuary to
down-estuary). The significance of this pattern is that these up-estuary
regions have, over time, been the areas subjected to the most rapid
urbanization and development.
Additionally, personal communications of Dr. Robert Orth, who conducted
the majority of the SAV distribution studies, suggests little evidence that
a simultaneous decline has occurred in other areas along the east coast of
the United States. Still, there does appear to be growing indications that
throughout the world, SAV communities are becoming increasingly stressed in
areas where there is extensive industrial and/or urban development.
Having documented that there has been a decline in SAV distribution and
abundance in the Chesapeake Bay, the next critical question is "are the
grasses a valuable component of the ecosystem?" The Chesapeake Bay Program
sponsored research that investigated the role and value of SAV in the
context of Bay grasses (l) contribution of organic matter to local food
webs, (2) habitat to infaunal and juvenile nekton species, (3) role as a
sink for sediments, and (4) role in nearshore nutrient dynamics.
The contribution of SAV to heterotrophic food webs is by either direct
grazing of living plants, or by consumption of detritus. It is known that
SAV serves as a food source for several waterfowl species. With the
decline in SAV populations, those waterfowl have switched to another food
source or occur in reduced numbers in the Bay.
Studies indicate that most SAV material enters the food webs through
detrital pathways. Data indicate that large predator fish feed in SAV
beds, and that their food items, e.g., amphipods, shrimp, are detrital
feeders whose food source probably includes some fraction of SAV. Many
epifaunal species in the estuary, which are important food items for many
consumers, are closely associated with SAV.
632
-------
I
weeks following initial contact. Moreover, herbicides degrade rapidly in •
the estuarine environment, with half-lives measured in days and weeks, and
residual concentrations do not appear to build up in sediments. The M
hypothesized mechanisms of increasing SAV exposure to herbicides through •
concentration of the compounds in epiphytic sediments or surface-layer
films, do not appear to represent significant factors. The one caveate
which remains unresolved is the fact that very little is known about •
estuarine concentrations and SAV toxicities of major herbicide m
metabolites. The de-ethylated daughter products of atrazine degradation do
tend to persist for months under estuarine conditions, and the weed-control •
literature attributes "carry-over" toxicity (after atrazine application) to •
this metabolite.
Ephemeral herbicide concentration in excess of five ppb do occur _
periodically in some estuarine water that once contained extensive SAV •
beds. In general, such concentrations cause losses in SAV productivity of »
10 to 25 percent, even when exposures are brief (about an hour); recovery
may take days to weeks even without ambient herbicides. The effects of •
repeated, brief exposures to such concentrations are not known. However, |
if the time interval between runoff events (which might yield such
concentrations) is greater than SAV recovery time, then partial loss of M
photosynthesis may persist. Such reductions in SAV productivity will •
definitely add to the generally-stressed conditions that these plants
currently experience in the estuary. Herbicide-induced loss of
productivity could act in concert with many of these stressors to create •
intolerable conditions for SAV existence. B
Being plants, SAV require light to grow and survive, and the apparent
optical properties of estuarine water create, in general, a light-limited •
environment for photosynthesis. Chesapeake Bay studies indicate reductions |
in both light quality and quantity during SAV growing season. Diffuse
downwelling attenuation coefficients in lower Bay communities indicate a _
severe attenuation of light energy in the photosynthetically-important blue •
(400 to 500 nm) region, and in the near infrared (700 to 775 nm) region of ™
the spectrum.
Historical data relative to light (turbidity, and indirectly, •
nutrients) and past distribution and abundance on SAV, indicate progressive V
Bay-wide changes in systems structure and function. In terms of Bay
grasses and the light environment, two overall conclusions are •
particularily important. It has been noted and documented that a •
generalized increase in nutrients and chlorophyll a_ concentrations in major
tributaries of the Chesapeake Bay has occurred over the past several _
decades. It has also been concluded that for roughly the same time scale, •
the general pattern of disappearance of submerged plant communities follows ™
a "down-river" pattern. It also appears that upper-Bay and western-shore-
lower-Bay communities have been affeced most severely. •
These conclusions, together with our studies on the light environment |
and photosynthesis-light relations in SAV ecosystems, suggest that total
PAR and factors increasing diffuse downwelling attenuation in the 400-500
nm region are principal driving functions controlling plant growth and
survival. The specific factors that, at present, appear to have the
greatest impact are suspended particles, both organic and inorganic, that
are largely controlled by climatic conditions (runoff and nutrient
loading), and indirectly by associated changes in physical-chemical regimes
(i.e. salinity and temperature).
633
-------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
In summary, it appears that Bay grasses are living in a marginal light
environment, and that progressive changes in water quality will further
stress the plant communities. To conclude that light has been singularly
responsible for recent declines in the vegetation goes beyond the data
available. The data do indicate, however, the extreme sensitivity of the
vegetation to both qualitative and quantitative measures of available
light. The data further imply that over the past several decades water
quality throughout the Bay, and particularily in the tributaries, has
progressively declined. More changes in these parameters can only affect
Bay grasses in an adverse way.
Following three years of research, we conclude that SAV exists in a
stressed environment. The sources of those stresses include such natural
factors as salinity extremes, waterfowl grazing, uprooting by cownose rays
and major storm events, as well as man-induced stresses such as
water-column turbidity, accumulation of epiphytic materials resulting from
nutrient enrichment and exposure to agricultural chemicals. The natural
stresses do not appear to be responsible for the presently reduced
populations of Bay grasses, because SAV has always been subjected to these
pressures and the historic record, as we have been able to reconstruct it,
does not reveal previous declines of such magnitude.
The issue as far as light is concerned is not simply of suspended
material, both inorganic and organic, in the water. Recent observations
and studies indicate that the growing nutrient enrichment of the Bay's
waters is stimulating the growth of epiphytic material. Combined, the
increased epiphytes and suspended materials may be the most significant
cause of the reduced SAV populations. At this time the results of
investigations into this issue are being analyzed.
634
-------
I
I
I
i
I
I
1
I
I
I
I
i
I
I
DISCLAIMER
This report has been reviewed by the Office of Research and Development
and Office of Water Programs, U.S. Environmental Protection Agency, and
approved for publication. Mention of trade names or commercial products
does not constitute endorsement or recommendation for use.
J.S. GOVERNMENT PRINTING OFFICE. 1983 - 606-490
-------
I
and Sn, factors are largely less than two or close to baseline factors
throughout the Bay proper. Seaward of the Bay Bridge (Annapolis) factors I
generally diminish, but Cd, Pb, and Zn are greater than two. The •
longitudinal distribution of values does not display a maximum in the Bay
near Baltimore, an expected increase if metals were emanating from •
Baltimore. Instead, the values mainly decrease from the Susquehanna River •
mouth, suggesting a river source (Helz et al. 1981). If the Susquehanna
watershed is not naturally enriched compared to average crust, then the _
enrichment is affected by direct contamination from industrial and •
municipal sources or from acid mine drainage. ™
Bed sediments within the Patapsco River, Baltimore Harbor, are markedly
enriched in Co, Cr and Zn (Sinex et al. 1981). Longitudinal distributions •
of enrichment factors, show that Cr increases with distance landward, and |
Zn is enriched throughout the Harbor. The Elizabeth River, Hampton Roads,
is notably enriched in Zn with Zn/Al ratios of six to 25 (Sinex et al. H
1981). •
Enrichment factors for Cd, Cu, Pb, and Zn in surface suspended material
of the central Bay are much greater than in bed sediments of the northern
Bay. .Metal/Fe ratios range from 10-118 for Cd, 12-27 for Cu, 37-51 for Pb, •
and 16-74 for Zn. The high enrichment factors in the central Bay are •
associated with high percentages of organic matter, probably produced by
plankton metabolism. Additionally, the metal content of central Bay •
suspended material exceeds the content of oceanic phytoplankton more than J[
nine times for Cd and Zn, and more than 19 times for Cu, Ni, and Pb.
Historic Metal Input Recorded in Sediments •
Some sediments in the Bay reveal trends in metal enrichment. In
sediments deposited in anoxic waters, no benthic macrofauna are present. •
Therefore, the sediments remain relatively undisturbed and may record the •
history and rate of change of metal influx. When a core of such sediments
is analyzed for trace metals and dated by ^lOp^ chronology, the vertical •
changes reveal variations in metal input. This approach assumes no •
diagenetic migration of metals through the length of the core. In oxic
environments, however, burrowing activities of benthic organisms can _
disturb the record of sedimentary sequences, create an "artificial" 210pjj •
distribution, and influence vertical trace metal distributions. ™
The vertical distribution of ^lOp^ and metal concentrations (Helz et
al. 1981) and the degree of bioturbation have been carefully examined for •
selected sediments of the Bay. Cores 4, 18, and 60 (Figure 19) exhibit |
exponential ^lOpj, profiles, low ^lOpj, depth-integrated concentrations,
and low or moderate bioturbation. They also show no metal peaks and «
display a relatively uniform rock structure. In addition, core 4 has •
1^'Cs data that verify the ^lOp^ sedimentation rate. Metal/aluminum
ratios for the three cores, and 210pb chronology are presented in Figure
19. All three cores show Zn enrichment in the Zn/Al ratios near the core •
surface, with maximum enrichment occurring at about 1940 in core 40 and •
about 1960 in cores 18 and 60. The first appearance of excess
concentrations is also temporally displaced down the Bay from 1890 in core •
4, to 1920 in cores 18 and 60. If the source of this excess Zn is fluvial Jj
(or anthropogenic) and up-Bay, then it takes about 20 years for the metals
to be transported 80 kilometers between core 4 and core 18, a nominal rate .
327
I
I
-------