You are on page 1of 11

Vol.

81: 257-267, 1992 '


l'
MARINE ECOLOGY PROGRESS SERIES
Mar. Ecol. Prog. Ser.
Published April 30

Phytoplankton photosynthesis, productivity, and


species composition in a eutrophic estuary:
comparison of bloom and non-bloom assemblages
Charles L. Gallegos
Smithsonian Environmental Research Center. PO Box 28. Edgewater. Maryland 21037, USA

ABSTRACT, During 2 periods in the late spring a n d early summer 1989, blooms in the Rhode River,
Maryland, USA, were triggered by nutrient inputs from local a n d remote watersheds. Parameters of the
photosynthesis-irradiance relationship increased in May when the bloom was dominated by Thalassios-
ira pseudonana, and decreased in J u n e when the bloom was dominated by large dinoflagellates.
Relative variability of quantum efficiency was a s great a s that of chlorophyll-specific absorption
coefficient. Light-saturated photosynthetic rates were predictable based on species counts and previ-
ously published size-dependent, nutrient-saturated growth rates for diatoms and dinoflagellates.
Dinoflagellate blooms occur in years of high hydrologic nutrient input, but, d u e to the reduced
photosynthetic efficiency of the dinoflagellate blooms, carbon fixation in the system varies pro-
portionately less than biomass.

INTRODUCTION Platt 1980, CBte & Platt, 1983, Sephton & Harris 1984).
In contrast to factors such as light and temperature, the
Since the possibility of estimating phytoplankton role of species composition has received little quantita-
productivity from Light and pigment data was first sug- tive treatment, because determination of photosynthe-
gested by Ryther & Yentsch (1957), considerable tic rates of individual species in mixed field assem-
research has been devoted to attempting to understand blages is a difficult task. In some cases, when densities
factors governing the variability in pigment-specific of large, robust cells are high, photosynthetlc rates of
parameters of the photosynthesis-irradiance relation- individual cells can be determined by micropipetting
ship (i.e. P-I parameters) (Harrison & Platt 1980, 1986, (Rvkin & Seliger 1981, Boulding & Platt 1986),but this
Falkowski 1981). Interest in estimation of productivity generally cannot be done with small or dehcate
from light and chlorophyll, which was originally con- species.
ceived a s a substitute for the time-consuming incuba- Recently, Joint & Pomroy (1988) used an allometric
tions necessary to estimate phytoplankton productivity approach to estimate the volumetric rate of productivity
directly, has been renewed recently by the widespread (i.e. not normalized to pigment concentration) at
availability of near-surface pigment data from satellite optimum depth. They used microscopic analyses of
remote sensing (Platt & Sathyendranath 1988). Pre- samples and the phylum- and size-dependent growth
sently, the most sophisticated remote sensing equations of Banse (1982) to estimate the maximal rate
algorithms (Sathyendranath & Platt 1989) combine re- of carbon increase. Here I extend their analysis to
gionally (and potentially seasonally) cataloged spectral predict the chlorophyll-specific light-saturated photo-
photosynthesis-irradiance (P-I) parameters with the synthetic rate, pmB,from estimates of species abun-
underwater light spectrum to calculate the depth- dances. To do so I combine the allometric growth
integral of phytoplankton production. equations of Banse (1982) with recent models relating
Discussions of factors governing P-I parameters typi- photosynthetic parameters to growth rate (Geider et al.
cally include species composition, usually by associat- 1986, Sakshaug et al. 1989, Cullen 1990).
ing a posteriori particular episodes of high or low para- Phytoplankton blooms in estuaries are frequently
meters with dominance by certain species (Harrison & dominated by one or a few species, are ephemeral in

O Inter-Research/Printed in Germany
258 Mar Ecol Prog Ser. 81 257-267, 1992

nature, a n d are often patchy in space (Paerl 1988). The ments on the western shore of Chesapeake Bay, Mary-
marked changes in species composition over short land, USA. It is 550 ha in area and averages 2 m deep,
times or distances associated with phytoplankton with a maximum of 4 m. The mean tidal range is 30 cm,
blooms afford an opportunity to examine the relation- but weather conditions often cause more extreme
ship between specles composition and photosynthetic changes in water level. Salinity varies seasonally from
parameters in greater isolation from other factors than 0 pp1 in spring at the head of Muddy Creek, the
can generally b e achieved in a seasonal or synoptic principal freshwater inflow to the Rhode River, to
study In 1989, 2 rainy penods during the s p n n g and almost 20 ppt at the mouth in fall during years with low
early summer resulted in elevated nutrient inputs and runoff.
unusually high chlorophyll concentrations in the Rhode Nutnent concentrations In the Rhode River are gen-
R v e r , a small tributary embayment of Chesapeake erally high (Jordan et al. 1991). Phosphate concen-
Bay, USA. Gallegos et al. (1992) demonstrated that trations typically peak in the summer at about 5 PM;
nutrient inputs from the local watershed produced ammonium is always present and concentrations are
localized blooms of short d u r a t ~ o nin the upper sub- relatively constant, averaging about 4 PM. Nitrate typi-
estuary, which were followed about 2 wk later by cally becomes undetectable in mid May (Jordan et al.
blooms throughout the sub-estuary resulting from nu- 1991); 1989 was unusual in that several late spring and
tnent inputs from the delayed arrival of the freshet early summer storms increased NO3- loading both from
down the mainstem of Chesapeake Bay Here I report the local watershed and from the Susquehanna River,
s p e c ~ e scomposition and photosynthetic parameters of the principal freshwater source to the upper
the blooms, and their effect on estimates of depth- Chesapeake Bay. The increased NOs- loading caused
integrated productivity. For each event, I first present phytoplankton blooms that produced chlorophyll con-
the response of the upper estuary to inputs from the centratlons well above the seasonal mean of 50 pg 1-'
local watershed, and then the response of the lower (Gallegos et al. 1992). Procedures for calculating NO3-
sub-estuary to the salinity drop at the mouth of the loading from the Muddy Creek watershed are given by
Rhode h v e r I then examine the predictions of the Gallegos et al. (1992).
allometnc relabonships for selected penods when Field sampling and experimental methods. Samples
photosynthetic parameters varied considerably over for phytoplankton chlorophyll, species composition,
short times or distances. and photosynthesis-irradiance curves were collected
along a transect of 6 stations from 1.4 km downstream
to 5.5 km upstream of the mouth of the Rhode River
MATERIALS AND METHODS (Fig. 1). The most upstream station was dominated by
advective flow due to the high runoff during the period
Study site. The Rhode River sub-estuary ( 3 8 3 2 'N, of interest, a n d will not be reported. Samples were
76'32 'W; Fig. I ) is one of several tributary embay- taken from the Secchi depth using a 2 1 Labline Teflon
sampler, emptied into 2 l polyethylene containers while
being shlelded from direct sunlight, and returned to the
laboratory in a cooler. Cruises were generally con-
ducted weekly; samples were taken daily during a ? d
penod following a large storm in July 1989. Station
numbering follows that of Gallegos et al. (1992);the 'B'
following station numbers designates 'bottle cast' to
d~stinguishthem from transect samples taken by pump
as part of a water quality monitoring program (not
reported here)
Vertical profiles of temperature and salinity were
made using a Beckman RS5-3 induction salinometer.
Vertical profiles of in vivo fluorescence were measured
uslng a Turner Designs model 10-005R fluorometer.
Underwater quantum scalar irradiance (400 to 700 nm)
was measured using a Biospherical Instruments
QSP-170 4n collector and QSR-250 integrator. Diffuse
attenuation coefficients for quantum scalar irradiance
Rg 1 Rhode fiver showlng station locations Inset shows were calculated from regression of log-transformed
locatlon of Rhode k v e r on Chesapeake Bay in Maryland readings against depth. Secchl depth was measured
(USA) with a 20 cm solid white disk.
Gallegos. Photosynthesi IS a n d specles composition 259

Subsamples for species identification were preserved aoPh(A),where h = wavelength (nm). Spectral mean
immediately in the field using 1 O/O acid Lugol's solution absorption, k, was then calculated as
and stored in 125 m1 polyethylene bottles kept in the
IdoO(A)
1 700
dark until counting (6 to 8 mo). For counting, 1 to 10 m1
were settled (minimum 4 h) and viewed at a magnifica-
*' = (700-400)
a ;h dk,

tion of 512x under an inverted microscope. Abundant Phytoplankton photosynthesls was measured as I4C
taxa were estimated from the number of fields required uptake using the 'photosynthetron' method of Lewis &
to count 200 individuals; density estimates of less Smith (1983). A 50 m1 subsample was inocculated with
abundant taxa were based on the number of individu- 370 to 1850 kBq NaH14C03, depending on biomass a s
als in 20 fields. Taxa comprising < 102 cells ml-' are anticipated based on fluorescence profiles. Subsamples
recorded as 'present'. All cells were identified to the of 1 m1 were dispensed into 26 (24 lighted, 2 darkened)
lowest taxonomic grouping possible, which was nearly 7 m1 glass scintillation vials placed in a n aluminum
always genus or lower. block bored with holes for lighting from below. Incuba-
Cellular carbon contents were estimated from cell tions were commenced within ca 2 h of sample collec-
volume (or plasma volume for diatoms) using the rela- tion a n d lasted 1 h. A range of light intensities was
tions of Strathman (1967). Volumes were estimated by supplied by a Westinghouse 400 W metal halide lamp.
assigning the taxa to regular geometric solids and Incubations were terminated by acidifying the samples
applying mensuration formulae using appropriate with 250 p1 of 6N HCI and shaking 1 h to drive off
linear dimensions determined from diagrams and unincorporated 14C (Lewis & Smith 1983). Samples
descriptions in keys or, in some cases, from measure- were then neutralized with NaOH and 5 m1 scintillation
ments on specimens from this site. cocktall (Ecolume, Packard) was added. Rates of 14C
Subsamples for chlorophyll analyses were filtered uptake were calculated by the equations of Strickland
onto Whatman GF/F glass fiber filters immediately & Parsons (1972),using a n isotope discrimination factor
upon return to the laboratory. Filters were extracted in of 1.06.
10 m1 of 90 % acetone overnight at 4 'C either immedi- Light intensities (400 to 700 nm) were measured a t
ately or after freezing for < 2 wk. Extracted chlorophyll the termination of the incubations using a Licor LI-185B
was estimated fluorometrically using a Gilson Spectra- quantum radiometer and 2 JC sensor inside a specially
Glo fluorometer calibrated against chlorophyll a deter- made adaptor that reproduces the geometry of the hole
mined periodically using the spectrophotometric equa- in the aluminum block and holds the sensor at the level
tions of Jeffrey & Humphrey (1975). of the sample meniscus. Comparisons of aB measured
Absorption spectra by particulate matter were deter- in the photosynthetron with measurements made in
mined with an EG & G Gamma spectrometer interfaced polystyrene flasks, in which hght intensity is more
to a LiCor 1800-12 integrating sphere (Gallegos et al. reliably measured, indicate that this procedure under-
1990). Water was filtered onto a 25 mm Whatman GF/F estimates light actually available to the phytoplankton
filter and scanned at 2 nm intervals from 400 to 750 nm. by a factor of 2.3, d u e most likely to a combination of
Filters were extracted overnight in methanol, then internal reflections, focusing by the bottom of the vial,
extracted a second time for 30 min before being rehy- and placement of the detector too high (Gallegos
drated and scanned again to determine absorption by unpubl.). Parameters dependent on light readings
nonpigmented particulate material (Kishino et al. (especially mB but sometimes also a photoinhibition
1985). parameter, see below) were adjusted accordingly.
Measured absorbances, after correcting for volume Data analysis. In the absence of photoinhibition, the
filtered and area of the filter, were divided by a path- P-I relationship was represented by the equation
length amplification factor (Gefer & SooHoo 1982) of (Jassby & Platt 1976)
1.8. This is lower than the pathlength amplification
factor of 3.31 used by Gallegos et al. (1990), who kept pB(1) = PmBtanh (5)
+ RB
pmB
filter loading low enough for absorption to be read at
full scale of 0.5 absorption units. In this work I used where the intercept, is allowed to b e either positive
heavier filter loading to minimize the wavelength or negative to avoid bias in calculation of aB.When
dependence of the pathlength amplification factor that photoinhibition was observed I used the equation (Platt
occurs at low absorbance (Mitchell & Kiefer 1984, et al. 1980)
Bricaud & Stramski 1990). Corrected absorbance of the
depigmented filters was subtracted from that of the
unextracted filters, and the difference was divided by
the chlorophyll concentration to give the chlorophyll- where p controls the degree of photoinhibition in the
specific absorption spectrum of the phytoplankton, fitted curve at high light intensities. Values of cuB esti-
260 Mar. Ecol. Prog. Ser. 81. 257-267, 1992

mated by Eq. (2) were divided by 1.20 { = tanh(l)/ (Banse 1982); for all others I used a = 2.20, b = -0.14
[l-exp(-l)]) to adjust for the tendency for the fitting (my interpolations).
routine to overestimate ruB to compensate for the lack of The relationship of PmB to growth rate is given by a
a well defined linear range at low light intensities in modification of the equation of Geider et al. (1986)
Eq. (2). After correcting for underestimation of light
intensities in the photosynthetron, photoinhibition was
rarely significant at ambient light intensities. where r (d-l) = respiration rate; D = h of daylight; and
Corrected values of uBand measured k, were used to 0 = cellular mg C (mg ch1)-'. Letting r = 0 . 1 5 ~(i.e.
estimate quantum efficiency, 4, by the relation energetic cost of synthesis slightly higher than respira-
tory O2 consumption; Geider et al. 1986), we can solve
for PmB of the jthspecies
PmBU)= 1.15a (j)C (j)b ( j l@D-' (5)
where the constant converts cuB to units of ~ m o C
l (mg
ch1)-'S-' (km01 quanta m-2 S-')-'. For the mixed field assemblage, Eq. (5) is summed
For calculation of the depth-integral of production over the species present, weighting the contribution of
over the day, J P ~Eq.
, (1) or (2) and estimated parame- each species by its cellular carbon. For 8, I used a
ters were used with measured chlorophyll concen- constant value of 40, in spite of known taxonomic
trations, diffuse attenuation coefficients and hourly dependence (see 'Discussion'). Values estimated from
averages of incident sunlight reduced by 10 % for field samples are highly variable, ranging from 4 to
surface reflection, to calculate photosynthesis as a 120, and averaging about 30. A higher value was used
function of depth and time; calculated photosynthetic because carbon estimates based on cell counts can be
rates were then integrated numerically. For standardi- expected to be biased low due to omission of present
zation, integrations in the vertical direction were but uncounted species.
always carried out to the photic depth (1 O/O penetra-
tion), although it should be understood that in the
upstream shallow regions, the photic depth sometimes RESULTS
exceeded the actual depth. Because the estimates are
based on short term (1 h) measurements and include 10 April to 10 June 1989
any excreted fixed carbon, they represent values closer
to gross daily photic zone production than to net pro- Following a week of high nitrate loading from the
duction. Furthermore, chlorophyll concentrations were local watershed of Muddy Creek in late April and early
assumed to be uniform with depth, which is usually May 1989, chlorophyll concentration at Stn 5B in the
true in this shallow system; but during June-July 1989, upper sub-estuary bloomed from about 30 to 120 yg 1-'
pronounced subsurface chlorophyll maxima were (Fig. 2A). The bloom consisted mainly of small phyto-
sometimes present a t or near the sampling depth (see plankton; changes in chlorophyll concentration a t Stn
e.g. Table 5 in Gallegos e t al. 1990). The calculated 5B generally paralleled changes in numerical abun-
daily production estimates should therefore be dance of the nanoplankton fraction (Fig. 2B).
regarded as potential rates for comparing bloom with Prior to the bloom, photosynthetic parameters, pmB
pre-bloom conditions. and ruB, averaged 3.7 mg C (mg ch1)-' h-' and 0.034 mg
Photosynthetic parameters and species composi- C (mg ch1)-' h-' (pm01 quanta m-' S - ' ) - ' ; Fig. 2C)
tion. I assumed all taxa present were growing at maxi- respectively. Following the N O , loading of 2 to 9 May
mal rates. For diatoms and dinoflagellates I used the 1989, PmB and aB initially dropped to < 2 and < 0.02,
size- und phylum-dependent relationships of Banse respectively, then rebounded to levels higher than ini-
(1982); he suggested that other microalgal taxa were tial values for the duration of the bloom. As the bloom
intermediate between diatoms and dinoflagellates, so declined between 23 May and 8 June, PmB and ruB
for those I used coefficients that were the median of the dropped to less than half their pre-bloom and mid-
corresponding coefficients for diatoms and dinoflagel- bloom values (Fig. 2C). Measured values of ruB were
lates. The allometric growth equations may be ex- more variable than measured k, (Fig. 2D), and varia-
pressed (Banse 1982) tions in the two were uncorrelated. Hence the variabil-
ity in nB was dominated by changes in calculated
values of 4.
where 11 (base e ; d - l ) = light- and nutrient-saturated At Stn 4B in the lower sub-estuary, chlorophyll con-
exponential growth rate; C (pg cell-') = cellular car- centration was about 80 (rg 1-' prior to the arrival of the
bon; and a and b a r e coefficients. For diatoms, a = 3.02, freshwater pulse from the Susquehanna, which was
b = -0.13; for dinoflagellates a = 1.38, b = -0.15 indicated by the drop in salinity at Stn 4B between 11
Gallegos: Photosynthesis and species composition 261

- - - - . _160
_
120,
l
7 A Chl.Sl.4B-.m---m--
.-C 5 \.#-a.
-
.-
d 3 -m 0
10 Apr 10 May 10 Jun
1o6
-- 105-
E 10"-
g 103.
S 10' - ' Large Dinoflagellates '.
10 Apr 10 May 10 Jun 10' -
10 Apr 10 May 10 Jun

0
10 Apr
1 10 May
0.00
10 Jun 01 m . ! 0.00
- 0.03
10 Apr 10 May 10 Jun

E
,
- 0.01 .
E -k, 'v' C
r" 0.00 0.00 6m
10 Apr 10 May 10 Jun

Fig. 2 . Response of biomass, populations, and photosynthetic


parameters to watershed ~ n p u t sfrom 12 Aprll to 8 June 1989, Fig. 3 ( A ) Response of chlorophyll at lower sub-estuary station
at the upper sub-estuary station (Stn 5B). (A) Chlorophyll (Stn 4B) to salinity drop due to arrival of freshwater pulse from
biomass in relation to NO3- loading from the local watershed; Susquehanna River from storm 2 w k earlier. (B), (C). (D) as
(B) cell concenstrations of nanoplankton and large dinoflagel- Fig. 2
lates; (C) photosynthetic maximum, pmB [mg C (mg ch1)-' h-']
and initial slope, a,' [mg C (mg ch1)-' h-' [ktmol m-' S - ' ] - ' ; ( D )
chlorophyll-specific absorption, k,, and quantum efficiency, Q,
/(m01 C fixed)(mol quanta absorbed)-'] by increases in chlorophyll however, the increment in
depth-integrated productivity should generally be less
than the increases in volumetric chlorophyll concen-
and 18 May 1989, i.e. about 9 d after the local rainfall tration, unless, a s was the case here, there are simul-
(Fig. 3A). After salinity stabilized at about 4.5 ppt, taneous changes in photosynthetic parameters (Fig. 4 ) .
chlorophyll Increased from 80 to 160 pg 1-l. As at the Thus, the initial increase in chlorophyll at the upper
upper sub-estuary station, the bloom consisted primar- sub-estuary station on 11 May had no impact on $Pd
ily of small phytoplankton, principally cryptomonads, (Fig. 4A) due to the initial declines in photosynthetic
Microcystis sp., and Thalassiosira pseudonana parameters. The following week JP, was maximal and
(Table l ) , and chlorophyll changes more closely the increment was roughly proportional to the incre-
tracked numbers of nanoplankton (Fig. 3B). mental increase in chlorophyll, due to the coincident
Photosynthetic parameters at the lower sub-estuary increases in cuB and PmB.The decrease in productivity
(Stn 4B; Fig. 3C) followed a very similar pattern to that on the next 2 sampling dates, in spite of continued high
at Stn 5B; i.e. typical values were observed prior to the photosynthetic parameters and increasing chlorophyll,
arrival of the freshwater pulse, followed by high values was due to increases in diffuse attenuation coefficient
near the end of May, then by lower values in early June from < l m-' on 11 May to 3.75 m-' on 24 May. Very
as numbers of nanoplankton decreased sharply similar patterns were observed a t the lower sub-
(Fig. 3C). During the period of dominance by nano- estuary station (Fig. 4B). At both stations, the effect of
plankton (10 April to 24 May) both k, [ca 0.007 to 0.014 day-to-day variations in cloud cover was assessed by
m2 (mg ch1)-'1 and [ca 0.015 to 0.03 m01 C (m01 calculating JPd for all days using incident irradiance
quanta)-'] varied by about a factor of 2 (Fig. 3D). measured on a cloudless day (Fig. 4 , dashed line).
Theory predicts that the maximum aerial standing While cloud cover reduced calculated productivity by
crop and potential depth-integrated productivity will 30 to 40 % on several days, the day-to-day changes
increase as depth decreases in a system (Takahashi & were much greater than this, and hence the temporal
Parsons 1972). Due to compression of the photic zone pattern was more strongly governed by other factors.
262 Mar Ecol. Prog. Ser. 81. 257-267, 1992

Table 1. Variation of phytoplankton abundance with time at Stn 4B, 27 April to 24 May 1989, and with location on 14 July 1989.
Units of abundance are cells ml-l P = present at < 102 ml-'

Stn 4B, 27 April-24 May 1989 Location on 14 July 1989


27 Apr l 1 May 18 May 22 May 24 May Stn l B Stn 2B Stn 3B Stn 4B Stn 5B

Nanoplankton
Cryptomonas sp. 360 3600 4300 1500 1200 1800 1700 800 1300 P
Thalassiosira pseudonana 3000 1200 34000 40000 60000 850 3300 4000 520 4700
Microcystis sp. P 1600 36000 8000 4500 11000 520 P
Microcystis aeruginosa 1500 1400 4100 250
Cyclotella glomerata 770 P P 900 P 920 1800 360 P P
Pyramimonas sp. P 650 P P 500 340 P 490 P
Gyrodinium estuariale 1100 4200 470 P P P P
Katodinium rotundaturn 1900 1600 P P
- P
Unidentified haptophyte P 770 650 P
Aphanizornenon sp. 1 P 1500 130
ChrysochromuLina sp. P P P
Trachelomonas sp P P 130 570 520
Eutreptia sp. P P P 410 520
Chaetoceros affinis 220 P P P
Large dinoflagellates
Gymnodinium sp. # 2 P 360 P P P 360 P P 5000
Gymnodjnium sp. # 3 1700 P P 700 130 2700 7700
Gyrodinium uncatenum 310 P P P 1100 460 1700 4600
Scrippsiella trochoidea P P 1300 670 1200 2200

5 June to 20 July 1989 1989; changes in chlorophyll most closely paralleled


changes in the large dinoflagellate cell numbers (Fig.
Two events resulted in high nitrate loading to the 5B). The second influx of NO3- from the Muddy Creek
upper sub-estuary during the second observation watershed was the result of a 2 d storm in early July.
period. The first was due to a sustained rainy period
during the first 3 wk in June (Fig. 5A, cf. Fig. 2A).
Chlorophyll at Stn 5B in the upper sub-estuary
bloomed from ca 60 pg I-' to > 500 pg 1-' on 19 June

10 Apr 10 May l0 Jun


10' l
5 Jun
15 Jun Jun 5 Jul 15 Jul
25
U

5 Jun 15 Jun 25 Jun 5 Jul


10 Apr 10 May 10 Jun

Fig. 4. Rates of gross daily production from 12 April to 8 June


1989, a s calculated from biomass and photosynthesis-
irradiance curves for (A) upper sub-estuary station (Stn 5B),
and (B) lower sub-estuary station (Stn 4B). (=p=) Calcu-
lated using incident irradiance measured on sampling day; (L -
- - G ) calculated using incident irradiance measured on a
clear-sky day Fig. 5. As Fig. 2 but for 8 June to 19 July 1989
Gallegos. Photosynthesi IS and s p e c ~ e scomposition 263

After an initial decline, chlorophyll rose over the next 3


d , peaked at nearly 1000 pg 1 - l , then decllned again.
Blooms throughout the sub-estuary durlng June-July
consisted primarily of the large dinoflagellates Gym-
nodinium sp. 2 and 3, Gyrodinium uncatenurn, and
Scripsiella trochoidea. The principal differences in
species composition at the peak of the bloom occurred
spatially, with large dinoflagellates being < 102 ml-' at
Stn l B , and increasing in numbers upstream (Table 1).
Large Dlnoflagellates
Although densities of nanoplankton and large dino- 10'
flagellates covaried, peak chlorophyll coincided with I
peak numbers of large dinoflagellates. 5 Jun 15 Jun 25 Jun 5 Jul 15 Jul
Photosynthetic parameters, which had declined at 0.04
I
the close of the first bloom period in May (cf. Fig. 2C)
declined even further, with pmBand aBreaching local
minima (PmB= 0.8, cuB = 0.008) on 19 June, the day of
the first peak in chlorophyll concentration (Fig. 5C).
Photosynthetic parameters varied roughly inversely
with chlorophyll, until the bloom that peaked on 14
July, when parameters were so low that changes were
- 0.03 0.06
Z1
F
undiscernable. In contrast with photosynthetic parame- 0.02
ters, k, was a s high and as variable during June-July a s E
during the 10 April to 10 June period. Photosynthetic
,
-
E
0.01 0.02 ;
C
m .
parameters at Stn 5B, especially pmBbut also cuB, were 1"0.00
@--m----orn * *
0.00 m 6
positively correlated with k , (Fig. 5 D ) . Therefore infer- 5 Jun 15 Jun 25 Jun 5 Jul 15 Jul
red values of @ were low and relatively uniform; @ Fig. 6.As Fig. 3, but for 8 June to 19 July 1989. Salinity at
exceeded 0.04 on one occasion, and was otherwise mouth of Rhode River is shown in (A) to indicate arrival of
< 0.03 with a minimum < 0.005. attenuated pulse from Susquehanna River on 27 June
At the lower sub-estuary station (4B; Fig. 6), salinity
8
was much lower than the seasonal average (Jordan et
7
al. 1991) and chlorophyll was at bloom levels through- 6
out most of the June-July period (Fig. 6A). Visually it is 5
4
difficult to relate chlorophyll changes to either of the 3
gross taxonomic groupings, but nanoplankton were 2
generally an order of magnitude less abundant and 1
0
large dinoflagellates a n order of magnitude more abun- 5 Jun 15 Jun 25 Jun 5 Jul 15 Jul
dant compared with the April-May period at the same
station (cf. Figs. 6B & 3B and Table 1).
Photosynthetic parameters at Stn 4B during the sec-
ond bloom period exhibited similar patterns as at Stn
5B; i.e. pmBand cuB were low and varied inversely with
chlorophyll (cf. Figs. 6C & 5C). Chlorophyll-specific
absorption was less variable at the lower sub-estuary
station, but calculated quantum efficiencies were simi- 01 . . . . . . . . . . . . . . . . . . . . . . .
5 Jun 15 Jun 25 Jun 5 Jul 15Jul
larly low and uniform, averaging 0.01 (Fig. 6D).
Variations in Jpd tracked changes in chlorophyll, but Fig. 7. As Fig. 4 but for 8 June to 19 July 1989. Note y-axis
were also influenced by changes in photosynthetic scale is half that in Fig. 4
parameters (Fig. 7). For example, in June chlorophyll
May bloom, even though chlorophyll levels were much
peaked on June 19, but JPd peaked the following
higher.
week, due to increases in both pmBand cuB. Similarly,
changes in JPd during the July bloom are clearly
related to chlorophyll changes. Note, however, that the Photosynthetic parameters and species composition
y-axis scale in Fig. 7 is half that in Fig. 4. Thus, due to
the much reduced photosynthetic parameters during Values of cellular carbon and predicted growth rates
the June-July period, JPd was lower than during the and assimilation number for the species observed to b e
2 64 Mar Ecol. Prog. Ser 81: 257-267, 1992

dominant over 5 sampling periods at Stn 4B during the its close coupling with PmB(Eq. 6) which, in turn, is
first bloom, and over 5 stations at the peak of the coupled with growth rate (Eq. 5).
second bloom, are given in Table 2. When the species-
specific pmBand carbon contents are combined with the
abundance data in Table 1, the predicted PmBfor the DISCUSSION
mixed assemblages ranged from < 1 to 7.1 and slightly
overestimated but correlated well with observed data The results demonstrate the potential impact of
(Fig. 8A; r2=0.??, slope of predicted versus ob- species composition on JPd. Photosynthetic parameters
served = 1.27, SE = 0.25). increased during the first bloom as Thalassiosira
While there is no basis in the equations of Banse pseudonana proliferated (Figs. 2 & 3, Table 1) with the
(1982) for predicting the value of aB from species com- result that calculated gross daily productivity increased
position, Cullen (1990), analysing the model of Sak- by a factor greater than the increment in chlorophyll
shaug et al. (1989), showed that (Fig. 4). Conversely, during the dinoflagellate-domi-
nated bloom of June-July the overall levels of JPdwere
lower than the previous period, despite higher
chlorophyll concentrations, due to the lower values of
where o = absorption cross section of the photosynthe- PmBand crBfor that assemblage. Nevertheless, predic-
tic unit; and t = turnover time of the photosynthetic tions of PmBduring both periods assumed the species
unit. No size- and taxa-dependent compilation of these present were growing at maximal rates (see below).
parameters has been made, but Eq. (6) provides a Surprisingly little of the difference in (uB between the
physiological basis for the correlation between aB and 2 periods was attributable to kc (cf. Figs. 2D, 3D, 5D &
PmBfrequently observed (e.g. Harding et al. 1985). 6D). This observation has implications for bio-optical
Such a correlation was observed here (Fig. 8B), and the approaches to estimating primary production whereby
same relationship between PmBand a Bwas maintained variability in aBis attributed to changes in k,, with @
across large changes in species composition and chlo- held constant. The degree to which a* is controlled by
rophyll concentrations (cf. Table 1, Figs. 2, 3, 5 & 6). kc rather than deserves further study.
These results suggest taxon-dependence of aB due to The dependence of PmBon species composition of the

Table 2. Cellular carbon contents and predicted growth and light-saturated photosynthetic rates for dominant species during late
spring and early summer 1989 in the Rhode h v e r . Maryland. PmBcalculations based on C:chl. B = 40 rng C (mg chl a)-', and
daylength. D = 14 h

Species Carbon Growth PmB


(pg cell-') (d-') (mg C [mg chl a]-' h - ' )

Nanoplankton
Thalassiosira pseudonana
Cyclotella glomerata
Chaetoceros affin~s
Cryptomonas sp.
Microcystis aeruginosa
Unidentified haptophyte
Pyramimonas sp.
Microcystis sp
Aphanizomenon sp. 1
Chrysochromul~nasp.
Synura sp.
Trachelomonas sp.
Eutreptia sp.
Gyrodinium estuariale
Prorocen trum minim um
Katodinium rotundaturn
Large dinoflagellates
Gymnodinium sp. # 2
Gymnodinium sp. # 3
Gyrodinium uncatenurn
Scrippsiella trochoidea
Gallegos: Photosynthesis and specles composition 265

above by Eppley's (1972) equation but many points fall


well below the upper bound, as has been observed in
other coastal regions (Pennock & Sharp 1986, Keller
1989). CBte & Platt (1983),who observed positive corre-
lation of P,,,Bwith temperature, attributed the result to
changes in species assemblages, with species having
higher P-I parameters predominating during warmer
periods, rather than temperature effect on photosyn-
thesis per se. Here, the warmer temperatures favored
development of assemblages with lower photosynthe-
Observed pmB
tic capacity.
Light-shade adaptation is known to influence PmB
(Falkowslu 1981); vertical variations in photosynthetic
parameters develop under stratified conditions (e.g.
Gallegos et al. 1983) and temporal adaptations as cells
undergo displacement through the vertical light gra-
dient can influence produchon in well-mixed layers
(Lewis et al. 1984). The equation relating photosyn-
thesis to growth rate purportedly applies to adapted
rates of photosynthesis (Sakshaug et al. 1989), and is
not expected to describe plots of instantaneous P-I
curves as were measured here (Cullen 1990).
Fig. 8. (A) Light-saturated photosynthetic rates, predicted While laboratory studies have probed deeply into the
from species composition data and size and taxon-dependent physiological bases of light-shade adaptation (Perry et
growth rates versus measured (m) 27 April to 24 May 1989
al. 1981, Prezelin 1981), at the resolution of this study
at Stn 4 B ; (0) Stn 1B to 5 B on 14 July 1989. ( - - ) Least squares
regression; (-) 1 . 1 correspondence. (B) Relationship of crB the photoadaptive property of major concern is the
to pmB for the 2 bloom periods, and all stations. (c) 12 April to 8 carbon:chlorophyll ratio, 8 (Falkowski 1981, Geider et
J u n e 1989; ( :) 14 J u n e to 24 July 1989. Units of PmBare [mg C al. 1986). The use of a constant 19is the most tenuous of
(mg ch1)-' h-']: units of nB are [mg C (mg ch1)-l h - ' (~tmolm-2 the assumptions used to predict pmB from species com-
s-I)-']
position, d u e to the known tendency for 8 to b e higher
in dinoflagellates than in diatoms (e.g. Chan 1980,
phytoplankton assemblage has been noted in other Geider et al. 1986). It is possible, however, that adapta-
studies (Cbte & Platt 1983, Harris et al. 1983, Sephton & tion of 6' to the reduced light availability may have
Harris 1984). Joint & Pomroy (1988) predicted produc- offset the tendency for 0 to increase due to taxonomic
tivity maxima from species composition and taxon- and differences. Diffuse attenuation coefficients during
size-specific growth rates, but they did not attempt to April-May generally ranged from 1 to 2 m-', compared
predict pigment-normalized, light-saturated photosyn- with 3 to 5 m-' during the period of dinoflagellate
thesis. The success of that prediction in this analysis dominance. Thus the cells experienced much lower
may seem surprising in view of the large number of light intensities during June-July, which would create
factors known to influence photosynthetic parameters conditions for increased cellular chlorophyll concen-
(Falkowski 1981). Consideration of some commonly trations and consequent reduced carbon:chlorophyll
cited factors will help to circumscribe the results and to ratios.
evaluate the factors controlling primary productivity in Another assumption of the analysis is that the species
nutrient-rich estuaries. present are growing at the nutrient-saturated rates for
The light-saturated photosynthetic rate, being a n their size and coarse taxonomic groupings. The
enzymatic process, is temperature-dependent with Qlo assumption is based on the low N : P ratios that prevail
values clustering around 2 (Harris 1978). Between the 4 during the late spring a n d summer a n d the continual
May and 11 May sampling dates, temperature a t Stn 4 B presence of dissolved NH4+ a t 4 to 5 PM (Jordan et al.
decreased from 18 to 13OC. Otherwise temperature 1991). Based strictly on the continual presence of dis-
during the first period for which predictions were solved macronutrients, the assumption of nutrient suffi-
attempted varied only between 18 and 21 "C, which is ciency seems reasonable. Nevertheless, closer exami-
insufficient to account for the variation observed. nation of the data may reveal some evidence of nutrient
Temperature varied < l "C among the 5 stations on 14 limitation. The lowest predicted for any of the large
July. A plot of pmB versus temperature over several dinoflagellates (Table 2) was 1.3 mg C (mg ch1)-' h-',
seasons at this site (Gallegos unpubl.) is bounded close to values observed on 10 to 12 July (Figs. ?C &
266 Mar. Ecol. Prog. Ser 81: 257-267, 1992

9C). As the bloom peaked on 14 July, however, Understanding the Life history patterns determining
obsewed values were only 0.62 mg C (mg ch1)-' h-'. growth and loss rates of dominant phytoplankters in
Although the observed value is about half that pre- relation to environmental variability at local and re-
dicted, the difference is small compared with the range gional scales may, therefore, be of equal importance in
of observed values (0.62 to 6.0). Thus it appears that the predicting production dynamics in such environments
prediction of PmBfrom species composition worked as is a deeper understanding of the mechanistic bases
here because the predicted growth rates for large dino- of photosynthesis.
flagellates found here conform to the size dependent
relationships of Banse (1982) and were so much lower Acknowledgements. I thank S. Hedrick for assistance with
than those of small diatoms (Table 2); further reduc- field work and analysis of phytoplankton samples. Portions of
tions due to nutrient limitation, if it indeed occurred, the data on phytoplankton carbon content were made avail-
able by K. Sellner. Comments by R. Geider, T Malone, and T.
were of second-order importance here. Jordan improved the manuscript. This study was funded by
Changes in species abundance is the time-integral the Smithsonian Environmental Sciences Program.
(after correcting for horizontal mixing and exchange) of
the net growth rate, i.e. the difference between gross
growth rate and the sum of all loss rates. The foregoing LITERATURE CITED
analysis assumed that differences PmBwere attribut- Banse, K. (1982) Cell volumes, maximal growth rates of
able to species-specific differences in maximal growth unicellular algae and cihates, and the role of ciliates in the
rates. The observation that slower growing species marine pelagial. Limnol. Oceanogr. 27- 1059-1071
(large dinoflagellates) sometimes bloomed while faster Boulding, E. G., Platt, T. (1986). Variation in photosynthetic
rates among individual cells of a marine dinoflagellate.
growing species (nanoplankton) declined implies that
Mar. Ecol. Prog. Ser. 29: 199-203
species-specific differences in loss rates also occurred, Bricaud, A., Stramslu, D. (1990). Spectral absorption coeffi-
with slower-growing species necessarily having lower cients of living phytoplankton and nonalgal biogenous
loss rates during such periods. Although loss rates were matter. a comparison between the Peru upweUlng area
not observed in this study, Reynolds et al. (1982) and the Sargasso Sea. Limnol. Oceanogr. 35: 562-582
Chan, A. T (1980).Comparative physiological study of marine
demonstrated that selective loss processes were a &atoms and dinoflagellates in relation to irradiance and
major determinant of species abundance patterns in cell size. 11. Relationship between photosynthesis, growth,
large limnetic enclosures. Given the quantitative con- and carbon/chlorophyll a ratio. J. Phycol. 16: 428-432
nection demonstrated here between pmBand the C6te, B., Platt, T (1983). Day-to-day variations in the spring-
summer photosynthetic parameters of coastal marine phy-
species present, lt is worth re-emphasizing the import-
toplankton. L~mnol.Oceanogr. 28: 320-344
ance of loss processes in determining both the produc- Cullen, J. J. (1990). On models of growth and photosynthesis
tion rate of the phytoplankton community and its even- in phytoplankton. Deep Sea Res. 37: 667-683
tual fate (e.g. Harris 1986). Day, J. W., Hall. C. A. S., Kemp, W. M., Yanez-Arancibia, A.
When watershed inputs favored development of (1989). Estuarine phytoplankton. In: Estuarine ecology.
John Wiley and Sons, New York, p. 147-187
dinoflagellate blooms, photosynthetic parameters were Eppley, R. W. (1972). Temperature and phytoplankton growth
reduced, and the relative fluctuations in rates of pro- in the sea. Fish. Bull. U.S. 70: 1063-1085
duction were lower than those of pigment biomass. Falkowski, P. G. (1981). Light-shade adaptation and assimila-
Thus carbon input to the plankton foodweb would tion numbers J . Plankton Res. 3: 203-216
Gallegos, C. L., Correll, D. L., P~erce,J. W. (1990). Modeling
appear to be at least partially buffered against large
spectral diffuse attenuation, absorption, and scattering coef-
interannual variations in hydrologic forcing at local and ficientsina turbid estuary. Limnol. Oceanogr. 35: 1486-1502
regional scales. Due to large changes in phytoplankton Gallegos, C. L.. Jordan, T E., Correll, D. L. (1992). Event-scale
community structure, however, qualitative effects on response of phytoplankton to watershed inputs in a subes-
foodweb structure may be pronounced. tuary: timing, magnitude and location of blooms. Limnol
Oceanogr (in press)
In environments meeting the criteria of nutrient- Gallegos, C. L., Platt, T., Harrison, W. G., Irwin, B. (1983).
sufficiency and having species assemblages conform- Photosynthetic parameters of arctic marine phytoplankton:
ing to the size-growth relationships of Banse (1982), vertical variations and time scales of adaptation. Limnol
these results suggest that data on species composit~on Oceanogr. 28: 698-708
Geider, R. J., Platt, T., Raven, J. A. (1986).Size dependence of
may be of value in making quantitative predictions of
growth and photosynthesis in diatoms: a synthesis. Mar
P-I parameters; their utility in remote sensing models Ecol. Prog. Ser. 30: 93-104
follows directly. While the seasonal periodicity of some Harding, L. W., Jr, Meeson, B. W., Fisher, T R., Jr (1985).
species and the temporal persistence of others (e.g. Photosynthesis patterns in Chesapeake Bay phytoplank-
Riley 1967, Day et al. 1989) may facilitate such predic- ton: short- and long-term responses of P-I curve parame-
ters to light. Mar Ecol. Prog. Ser. 26: 99-11 1
tions, interannual variability in species composition, Hanis, G. P. (1978).Photosynthesis. productivity and growth:
particularly in relation to the occurrence of sporadic the physiological ecology of phytoplankton. Arch. Hy-
blooms, can be expected to be large (Mallin et al. 1991). drobiol. Beih. Ergebn. Limnol 10: 1-171
Gallegos: Photosynthesis and s p e c ~ e scomposition 267

Harris. G. P. (1986). Phytoplankton ecology: structure, func- Paerl, H. W. (1988). Nuisance phytoplankton blooms in coast-
tion and fluctuation. Chapman and Hall, New York al, estuarine, and inland waters. Limnol. Oceanogr. 33:
Harris, G. P., Piccinin, B. B., van Ryn, J . (1983). Physical 823-847
variability and phytoplankton communities: V. cell size, Pennock, J . R., Sharp, J . H. (1986). Phytoplankton production
niche diversification and the role of competition. Arch. in the Delaware Estuary: temporal and spatial variability.
Hydrobiol. 98: 215-239 Mar Ecol. Prog. Ser 34. 143-155
Harrison, W G., Platt, T (1980). Variations in assimilation Perry, M. J , Talbot, M. C., Alberte, R. S. (1981). Photoadapta-
number of coastal marine phytoplankton: effects of tion in marine phytoplankton: response of the photo-
environmental CO-variates.J Plankton Res. 2: 249-260 synthetic unit Mar Biol. 62: 91-101
Harrison, W. G., Platt, T (1986). Photosynthesis-irradiance Platt, T., Gallegos, C . L., Harrison, W. G . (1980). Photoinhibi-
relationships in polar and temperate phytoplankton popu- tion of photosynthesis in natural assemblages of m a n n e
l a t i o n ~ Polar
. Biol. 5: 153-164 phytoplankton J. Mar. Res. 38: 687-701
Jassby, A. D., Platt, T. (1976). Mathematical formulation of the Platt, T., Sathyendranath, S . (1988). Oceanic primary produc-
relationship between photosynthesis and light for phyto- tion: estimation by remote sensing at local and regional
plankton. Limnol. Oceanogr 21: 540-547 scales. Science 24 1: 1613-1620
Jeffrey, S. W., Humphrey, G. F. (1915).New spectrophotornet- Prezelin, B. B. (1981). Light reactions in photosynthesis. In:
ric equation for determining chlorophyll a, b, c', and c2in Platt, T (ed.) Physiological bases of phytoplankton ecol-
higher plants, algae, a n d natural phytoplankton. Biochem. ogy. Can. Bull. Fish. Aquat. Sci. 210: 1-43
Physiol. Pflanz. 167: 191-194 Reynolds. C. S., Thornpson, J . M,, Ferguson, A. J. D., Wise-
Joint, I. R., Pornroy, A. J. (1988). Allometric estimation of the man, S. W. (1982). Loss processes in the population
productivity of phytoplankton assemblages. Mar. Ecol. dynamics of phytoplankton maintained in closed systems.
Prog. Ser. 47: 161-168 J. Plankton Res. 4: 561-600
Jordan, T E., Correll, D. L., Miklas, J., Weller, D. E. (1991). k l e y , G. A. (1967). The plankton of estuaries. In: Lauff, G. H.
Nutrients and chlorophyll at the interface of a watershed ( e d . ) Estuaries. American Assoc. for the Advancement of
and a n estuary. Limnol. Oceanogr. 36. 251-267 Science, Washington, D.C., p. 316-326
Keller, A. A. (1989).Modeling the effects of temperature, light, h v l u n , R. B., Seliger, H. H. (1981). Liquid scintillation count-
and nutrients on primary productivity. a n empirical and ing for 14C uptake of single algal cells isolated from natural
mechanistic approach compared. Llnlnol. Oceanogr 34: samples. Limnol. Oceanogr 26: 780-785
82-96 Ryther, J. H., Yentsch, C . S . (1957). The estimation of phyto-
Kiefer, D. W . , SooHoo, J. B. (1982). Spectral absorption by plankton production in the ocean from chlorophyll and
marine particles of coastal waters of Bala Cahfornia. Lim- light data. Limnol. Oceanogr 2: 281-286
nol. Oceanogr. 27: 492-499 Sakshaug, E., Kiefer, D. A., Andresen, K. (1989). A steady
Kishino, M,, Takahashi, M., Okami, N., Ichimura, S. (1985). state description of growth and light absorption in the
Estimation of the spectral absorption coefficients of phyto- marine planktonic diatom Skeletonerna costaturn. Limnol.
plankton in the sea. Bull. mar. Sci. 37: 634-642 Oceanogr. 34: 188-197
Lewis. M. R., Horne, E. P. W., Cullen, J. J . , Oakey, N. S., Platt, Sathyendranath, S., Platt, T (1989). Computation of aquatic
T (1984). Turbulent motions may control phytoplankton primary production: extended formalism to include effect
photosynthesis in the upper ocean. Nature, Lond. 311. of angular and spectral distributions of light. Lirnnol.
49-50 Oceanogr 34: 188-197
Lewis, M. R., Smith, J. C. (1983). A small volume, short- Sephton, D. H . , Harris. G. P. (1984). Physical variability and
incubation-time method for measurement of photosyn- phytoplankton communities. VI. Day to day changes in
thesis as a function of incident irradiance. Mar. Ecol. Prog. primary productivity and species abundance. Arch
Ser. 13: 99-102 Hydrobiol. 102 155-175
Mallin, M. A., Paerl, H. W., Rudek, J. (1991) Seasonal phyto- Strathmann, R. R. (1967) Estimating the organic carbon con-
plankton composition, productivity, and biomass in the tent of phytoplankton from cell volume or plasma volume
Nuese River estuary, North Carolina. Estuai-. coast. Shelf Lirnnol. Oceanogr 12: 41 1-418
Sci. 32: 609-623 Strickland, J. D. H., Parsons, T R. (1972). A practical hand-
Mitchell, B. G., Kiefer, D. A. (1984). Determination of absorp- book of seawater analysis, 2nd edn. Bull. Fish. Res. Bd
tion and fluorescence excitation spectra for phytoplankton. Can., Ottawa
In. Holm-Hansen, O., Bolis, L., Gilles. R. (eds.) Marine Takahashi, M., Parsons, T. R. (1972). The maximization of the
phytoplankton and productivity. Springer-Verlag, New standing stock a n d primary productivity of phytoplankton
York, p. 157-169 under natural conditions. Indian J . mar Sci. 1: 61-62

This article was presented by 7.Platt, Dartrnouth, N.S., Manuscript first received: J u l y 9, 1991
Canada Revised verslon accepted: March 4, 1992

You might also like