Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

ARTICLE IN PRESS

Deep-Sea Research I 56 (2009) 2162–2174

Contents lists available at ScienceDirect

Deep-Sea Research I
journal homepage: www.elsevier.com/locate/dsri

Light absorption by phytoplankton, non-algal particles and dissolved


organic matter at the Patagonia shelf-break in spring and summer
Amábile Ferreira , Virginia M.T. Garcia, Carlos A.E. Garcia
Institute of Oceanography, Federal University of Rio Grande, Av. Italia, Km 8, Rio Grande, RS 96201-900, Brazil

a r t i c l e i n f o abstract

Article history: Satellite image studies and recent in situ sampling have identified conspicuous
Received 3 September 2008 phytoplankton blooms during spring and summer along the Patagonia shelf-break
Received in revised form front. The magnitudes and spectral characteristics of light absorption by total
26 July 2009
particulate matter (phytoplankton and detritus) and colored dissolved organic matter
Accepted 5 August 2009
Available online 14 August 2009
(CDOM) have been determined by spectrophotometry in that region for spring 2006 and
late summer 2007 seasons. In spring, phytoplankton absorption was the dominant
Keywords: optical component of light absorption (60–85%), and CDOM showed variable and
Absorption coefficient important contributions in summer (10–90%). However, there was a lack of correlation
Phytoplankton bloom
between phytoplankton biomass (chlorophyll-a concentration or [chl a]) and the non-
CDOM
algal compartment in both periods. A statistically significant difference was found
Detritus
Patagonian shelf-break between the two periods with respect to the CDOM spectral shape parameter (Scdom),
with means of 0.015 (spring) and 0.012 nm1 (summer). Nonetheless, the mean Scdm
values, which describe the slope of detritus plus CDOM spectra, did not differ between
the periods (average of 0.013 nm1). Phytoplankton absorption values in this work
showed deviations from mean parameterizations in previous studies, with respect to
[chl a], as well as between the two study periods. In spring, despite the microplankton
dominance, high specific absorption values and large dispersion were found
(a*ph(440) ¼ 0.0470.03 m2 mg [chl a]1), which could be attributed to an important
influence of photo-protector accessory pigments. In summer, deviations from general
trends, with values of a*ph(440) even higher (0.0970.02 m2 mg [chl a]1), were due to
the dominance of small cell sizes and also to accessory pigments. These results highlight
the difficulty in deriving robust relationships between chlorophyll concentration and
phytoplankton absorption coefficients regardless of the season period. The validity of a
size parameter (Sf) derived from the absorption spectra has been demonstrated and was
shown to describe the size structure of phytoplankton populations, independently of
pigment concentration, with mean values of 0.41 in spring and 0.72 in summer. Our
results emphasize the need for specific parameterization for the study region and
seasonal sampling approach in order to model the inherent optical properties from
water reflectance signatures.
& 2009 Elsevier Ltd. All rights reserved.

1. Introduction

It is well known that the upwelling of visible radiation


 Corresponding author. Tel.: +55 53 32336617; fax: +55 53 32336601. or reflectance (i.e. apparent optical properties (AOPs))
E-mail addresses: amabilefr@hotmail.com (A. Ferreira), from a natural water body is determined by the inherent
docvmtg@furg.br (V.M.T. Garcia), dfsgar@furg.br (C.A.E. Garcia). optical properties (IOPs, absorption and scattering) of the

0967-0637/$ - see front matter & 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.dsr.2009.08.002
ARTICLE IN PRESS
A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174 2163

optically active constituents, i.e., particulate and dissolved that the dominant cell size of the algal community,
material (Gordon et al., 1988). Understanding the pro- and consequent changes in pigment packaging,
cesses that cause changes in the optical behavior of these could largely explain the natural variability observed in
components and quantifying the variability in their IOPs the spectral shape of phytoplankton absorption. Models
should improve the accuracy of individual parameter have been proposed to retrieve information on cell size
retrieval by remote-sensing techniques (IOCCG, 2006). In from spectral absorption measurements (Ciotti et al.,
addition to quantifying these so-called optically active 2002) and from values of the specific absorption
components, IOPs have been used to estimate biogeo- coefficients (Devred et al., 2005). However, those studies
chemical variables such as organic carbon, chlorophyll-a, have not separated the relative contributions of
dissolved organic material and total suspended matter pigment packaging and accessory pigment effects but
(e.g. Twardowski et al., 2005), and primary productivity have relied on the co-variation between cell size and
(Marra et al., 2007; Prieto et al., 2008). Remote sensing accessory pigmentation (Ciotti et al., 1999, 2002; Trees
studies are increasingly focused on the understanding of et al., 2000).
both the spatial and temporal variability of primary Deviations in the relationship between optical proper-
production over large and regional scales (Behrenfeld et ties and [chl a] can be particularly important in high-
al., 2001). In addition, it is also desirable to quantify the chlorophyll Case I waters (O’Reilly et al., 1998), such as
role of both phytoplankton and dissolved organic matter offshore phytoplankton blooms. Morel et al. (2006)
(DOM) in the ocean carbon budget (Siegel et al., 2005; Hu analyzed bio-optical relationships in an upwelling area
et al., 2006). with no influence from continental runoff, comparing
The spectral light absorption is a major factor con- their natural variability with that previously observed in
tributing to the variation in optical properties of less productive waters. They concluded that the main
seawater. It can be quantified by the absorption coeffi- source of variability is the looseness of the co-variations
cient, which, in turn, consists of absorption by pure between the detrital (dissolved and particulate) compo-
seawater, phytoplankton, non-algal particles and colored nent and the algal stock.
dissolved organic matter (CDOM). Variability in the In the present work, a frontal region (Patagonia shelf-
absorption coefficient in oceanic (so-called ‘‘Case I’’) break front) relatively far from the coast, where extensive
waters has been thoroughly documented over the past phytoplankton blooms develop in spring-summer, has
decades (e.g. Morel, 1988; Bricaud et al., 1998; Reynolds et been selected for the study of optical properties of both
al., 2001; Morel and Maritorena, 2001). In such dissolved and particulate material, including phytoplank-
waters, phytoplankton and their derivative products are ton. The data provided in the study contribute to
optically dominant, and all other components, except documentation of absorption properties in relatively clear
seawater, are often assumed to co-vary with chlorophyll-a and high-chlorophyll waters.
concentration ([chl a]). However, parameterization of the The shelf-break front formed between Argentinean
absorption coefficient can show significant variability in shelf waters and the Malvinas Current (39–511S) in the
space and time, mainly where dissolved material or non- South Atlantic has been shown to be associated with a
algal particles are not proportional to phytoplankton conspicuous band of high chlorophyll throughout spring
absorption, and it must consider each of them separately and summer, detected by ocean color sensors (Acha et al.,
(‘Case 2 waters’). Specifically for phytoplankton, pigment 2004; Rivas et al., 2006; Romero et al., 2006). In situ
packaging (Duyens, 1956) or pigment composition gen- studies have found high surface [chl a] (between 1.8 and
erally can be an important cause of optical variability, 19.9 mg m3) and primary production rates (between 2.0
even for a given [chl a] (Bricaud et al., 1995, 2004). Thus, and 7.8 g C m2 d1) in spring that are comparable to
particular phytoplankton communities can show a given maximal seasonal values at eastern boundary currents
absorption signature, which may be characteristic of a (Garcia et al., 2008). The region presents a potentially
given region or time. At least in regions where phyto- significant biological control of gases such as O2 and CO2
plankton absorption is significant, these variations may in surface layers, as suggested by data in Bianchi et al.
affect the reflectance spectra (Carder et al., 1999; (2005, 2009) and Garcia et al. (2008). Despite the
Maritorena et al., 2002; Carder et al., 2004). Most semi- ecological and biogeochemical relevance of the Patagonia
analytical algorithm approaches relate reflectance to shelf-break front, there are no records of IOP measure-
absorption by phytoplankton and dissolved organic ments in the study area.
matter (e.g. Maritorena et al., 2002). Therefore, there is a In summary, the main objective of the present work is
need for regional in situ measurements of light absorption, to study the seasonal variation in magnitude and
in order to assess or to develop satellite ocean color spectral shape of the absorption coefficients of
algorithms. phytoplankton, detritus and CDOM in the Patagonian
The performance of semi-analytical models in describ- shelf-break front. The focus is on the absorption
ing relationships of ocean color to chlorophyll (Carder budget during two cruises, carried out in spring 2006
et al., 2004; Sathyendranath et al., 2001) has been and summer 2007, as part of the PATagonian EXperiment
shown to be sensitive to variations in the chlorophyll- (PATEX) project, conducted within the Brazilian
specific absorption coefficient (i.e. absorption coefficient Antarctic Program. Thus, the study seeks to contribute to
normalized to chlorophyll-a unit) due to changes in a better understanding of the phytoplankton absorption
phytoplankton cell size and pigment composition. properties in oceanic waters with high-chlorophyll
Ciotti et al. (2002) and Bricaud et al. (2004) have verified content.
ARTICLE IN PRESS
2164 A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174

2. Material and methods determined in the three size classes, microplankton


(420 mm), nanoplankton (3–20 mm) and picoplankton
2.1. Sampling site (o3 mm), in order to estimate the contribution by each
class to the phytoplankton biomass. Although some
The vicinity of the Argentinean shelf-break (39–461S) traditional classification schemes (e.g. Sieburth et al.,
was sampled in October 26–29, 2006 (spring, PATEX II 1978) consider picoplankton as o2 mm, in this study,
cruise), when 26 stations were occupied, and in March because of logistical constraints (membrane filter avail-
25–29, 2007 (late summer, PATEX III cruise). There was no ability) the o3 mm fraction was considered as picoplank-
important difference in the location of the oceanographic ton. After filtration, the filters were wrapped in aluminum
stations between the two cruises. An additional transect foil and kept in the freezer until the end of the day,
was sampled in the northern section during the summer thereafter transferred to liquid nitrogen for later analysis.
cruise. Water sampling was always conducted during At the laboratory, the pigment was extracted in 90%
daytime, in groups of 6 or 7 stations, with the ship sailing acetone and fluorescence determined in a Turner Designs
southwards (PATEX II) and northwards (PATEX III). The TD-700 fluorometer (previously calibrated with Sigmas
sampling sections are presented in Fig. 1, where three chlorophyll-a standard), following the non-acidification
cross-shelf (two for PATEX II) and two along-shelf method of Welschmeyer (1994).
transects are shown on monthly SeaWiFS chlorophyll-a
composites from October 24 to 31, 2006, and from March 2.3. Absorption coefficients
22 to 29, 2007, corresponding to sampling periods of
PATEX II and PATEX III. In samples from the surface and from the fluorescence
peak, filtrations were carried out (Whatmans 25 mm
2.2. Chlorophyll-a concentration GF/F) for measurements of light absorption by parti-
culate material and CDOM during the two campaigns.
Discrete samples for total [chl a] were collected from During the PATEX III cruise, some samples were also
the surface and from selected depths (based on the collected from 5 m depth, and they are included as surface
fluorescence profile) and filtered onto 25-mm glass fiber samples.
filters (Whatmans GF/F). Chlorophyll size fractionation
was determined both at surface and at sub-surface 2.3.1. Particle absorption
fluorescence peak with a 20 mm mesh and then filtration Untreated seawater volumes of 300–700 ml, depending
throughout a 3-mm polycarbonate filter. Chlorophyll a was on particle load, were concentrated on GF/F filters that

Fig. 1. SeaWiFS chlorophyll-a image composites from October 24 to 31, 2006 and from March 22 to 29, 2007, showing the sampling stations (St) during (a)
the PATEX II cruise—October 26–29 (St. 201–226)—and (b) PATEX III cruise—March 25–29 (St. 301–330) along the Patagonian shelf-break.
ARTICLE IN PRESS
A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174 2165

were immediately stored in liquid nitrogen prior to spectral model:


analysis. In the laboratory, optical density values of the
aph ðlÞ ¼ a/phS ðlÞ½Sf a/picoS ðlÞ þ ½ð1  Sf Þa/microS ðlÞ;
particulate material were determined using a dual-beam
scanning spectrophotometer (Cary Model 1E). Values where a/picoS(l) and a/microS(l) are the ‘‘basis vectors’’ (or
corresponding to detritus were determined after extrac- absorption spectra normalized by their own average over
tion of pigment with pure methanol (Mitchell et al., 2000; the visible spectrum) corresponding to samples domi-
Kishino et al., 1985). Absorption coefficients (m1) of total nated by Prochlorococcus (picoplankton) and samples
particulate matter (first measurement: [apart(l)]) and of dominated by microplankton-sized organisms, and
detritus (measurement after extraction: [adet(l)]) were a/phS(l) is the scaling factor to be applied to the nor-
calculated by converting optical density at each wave- malized absorption spectrum (Ciotti et al., 2002). Here a
length to natural log, normalizing by filtered volume and picoplankton vector presented in Ciotti and Bricaud
applying a Beta factor, which corrects for the effect of (2006) was used. The size parameter Sf is a parameter
concentrating particles on filters (path-length amplifica- constrained to vary between 0 and 1 and to specify the
tion). The correction algorithm was determined by relative contributions of picoplankton and microplankton
comparing optical density (OD) in suspension and on to absorption, independently of [chl a] (Ciotti et al., 2002).
filters using four monospecific cultures (Skeletonema
costatum, 10 mm; Thalassiosira oceanica, 14 mm; Phaeo-
2.3.2. CDOM absorption
cystis sp., 4 mm; a coccoid cyanobacterium, 3 mm),
Filtrates from the particulate absorption sample filters
following procedures described in Mitchell (1990). The
(Whatman GF/F) were collected in clean borosilicate
coefficients derived from the beta equation were:
bottles that were previously treated according to Mitchell
ODsus ¼ 0.4743(ODfilt)2+0.4506ODfilt (r2 ¼ 0.98), where
et al. (2000). Briefly, sequential soaks and rinses in mild
ODfilt and ODsus correspond to OD due to material retained
detergent, freshly produced Mili-Q water, 10% HCl with
on filter and due to suspended material.
final copious rinse in Mili-Q water were made, following
Although there is evidence that the relationship
by combustion in aluminum foil covers at 450 1C for 4–6 h
between ODfilt and ODsus may exhibit significant differ-
(Mitchell et al., 2000). An initial sample volume of 100 ml
ences when various species are considered separately
was used for rinsing the GF/F filter and the filtration
(Moore et al., 1995; Bricaud et al., 1995), this was not the
apparatus. Although 0.2-mm membrane filters are recom-
case in the present work. Despite the large range in cell
mended for CDOM analyses, the use of GF/F filters ensures
sizes of the phytoplankton cultures, no significant differ-
consistency with the procedure for particulate material
ences were found in the algorithms for each separate
absorption measurement and the estimates of total
group (not shown). This suggests that variations in the
absorption (Ciotti and Bricaud, 2006).
Beta factor are more associated with different instruments
The spectral absorbance of the filtered water was
and methods than with phytoplankton groups or sizes. For
measured against air in a 10-cm quartz cuvette between
instance, different trends in the relation between ODfilt
300 and 750 nm. Freshly produced Mili-Q water was used
and ODsus, initially found between the groups, disap-
to zero the instrument. After conversion of optical density
peared after proper correction of ODsus for each spectrum
values into absorption coefficients (acdom(l), m1), the
with the ‘turbidity blank’ (value at 750 nm). Thus, the
same fitting procedure for detritus spectra was applied to
single relationship generated here appears to be effective
the 350–600 nm spectral absorption range for deriving the
to correct for path-length amplification in our samples,
CDOM spectral shape parameter Scdom (Babin et al., 2003).
even considering the different phytoplankton commu-
nities (in both cell sizes and taxonomic structure) during
the two sampled seasons. 3. Results
The final estimates of absorption of apart(l) and adet(l)
were obtained by subtracting the measured values of 3.1. Total and fractionated chlorophyll-a concentrations
apart(750) and adet(750) from all the measured spectral
values of apart(l) and adet(l) (Babin and Stramski, 2002). In spring (PATEX II cruise), surface [chl a] varied
The spectral absorption coefficients for phytoplankton between 0.92 and 11.85 mg m3. Lower values were
[aph(l)] were computed as the difference between apart(l) observed in the first longitudinal profile (north of the
and adet(l) estimates. region), where the maximum concentration was 3.56 mg
After conversion into absorption coefficients, the m3 (St. 205). Generally, higher values (45.0 mg m3)
adet(l) data from 350 to 600 nm were used to derive the were associated with the shelf-break front in the middle
spectral slope of detritus absorption (Sdet), following Babin of the sampling region (Fig. 1a), decreasing towards south
et al. (2003). In order to examine differences between (o4.5 mg m3, St. 223–226). Peaks in chlorophyll fluores-
phytoplankton absorption spectra, aph(l), values were cence were generally observed in the sub-surface, be-
normalized by their own average over the visible tween 11 and 26 m, reaching a maximum of 12.05 mg m3
spectrum range, according to Garver et al. (1994). at St. 212.
During PATEX III cruise (summer), surface [chl a] were
2.3.1.1. Estimation of a cell-size parameter. A cellular size between 0.24 and 2.16 mg m3 (Fig. 1b). Fluorescence peak
parameter (Sf) was estimated from each phytoplankton- depths were generally shallower than in the spring cruise,
normalized spectrum, following Ciotti et al. (2002). Every varying between 5 and 27 m. The lowest values were
normalized spectrum was decomposed using the mixed observed at the innermost station of the lower latitude
ARTICLE IN PRESS
2166 A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174

section (St. 301) and in the east of the front (St. 311), with of the cruises, discriminating surface and fluorescence
values of 0.24 and 0.56 mg m3, respectively. Along the peak depths.
continental slope, there was no evidence of considerable Fig. 2 shows a relationship between total and size-
latitudinal variation in [chl a]. The contrast between [chl fractionated [chl a] for the cruises. In spring (PATEX II,
a] in spring and summer periods was even more Fig. 2a), the range of total pigment concentration was
pronounced when the depth-integrated [chl a], which explained by changes in microplankton biomass, except at
ranged from 60 to 650 mg m2 in spring and from 12 to stations where low [chl a] values were observed (mainly
77 mg m2 in summer (not shown) were compared. The in the north). Excluding these stations (St. 201–206, see
SeaWiFS [chl a] images also illustrate the differences in Fig. 1), the mean dominance of microplankton was 69%. In
the phytoplankton biomass between the two sampling summer (PATEX III), a shift in algae size dominance from
periods (Fig. 1). Range of variation, mean and standard microplankton (spring, PATEX II) to nanoplankton biomass
deviation of [chl a] values are presented in Table 1 for both was observed (Fig. 2b). The nanoplankton fraction con-
tributed, on average, 66% to total [chl a], followed by
picoplankton (23%). The contribution of larger cells
Table 1
(420 mm) was only around 8%.
Ranges, means and standard deviations of chlorophyll-a concentrations
[chl a], and of absorption coefficients at 440 nm (m1), for colored
dissolved organic matter, acdom, total particulate matter, apart, detritus,
adet, phytoplankton pigments, aph and specific phytoplankton absorp- 3.2. Absorption coefficient magnitudes
tions coefficients for surface (above) and fluorescence peak depths
(below) spring (PATEX II cruise) (left) and summer (PATEX III cruise)
Table 1 shows separately the surface and fluorescence
(right) samples.
peak values of absorption coefficients (m1) at 440 nm
PATEX II PATEX III (wavelength at which all components are optically
important), of total particulate matter [apart(440)], phyto-
Range Mean Std. dev. Range Mean Std. dev.
plankton [aph(440)], detritus [adet(440)] and colored
Surface dissolved organic matter [acdom(440)] during in both
[Chl a] 0.920–11.85 5.578 3.167 0.240–2.160 1.234 0.455 cruises (PATEX II and PATEX III). In spring (PATEX II), all
acdom 0.014–0.126 0.044 0.029 0.020–0.345 0.121 0.088 apart(440) values varied from 0.074 to 0.375 m1 (Table 1,
apart 0.096–0.366 0.230 0.075 0.030–0.257 0.146 0.052 left), with lower values corresponding to the northern
adet 0.017–0.068 0.035 0.012 0.011–0.120 0.043 0.032
section, where low [chl a] were found (Table 1, left). In this
aph 0.067–0.351 0.191 0.071 0.002–0.182 0.102 0.043
a*ph 0.020–0.159 0.042 0.030 0.044–0.110 0.083 0.018 season, acdom(440) coefficients were relatively low and
varied within a narrow range (0.014–0.160 m1). There
Fluorescence peak depths
[Chl a] 0.920–12.05 5.708 3.690 0.570–2.530 1.314 0.461 was no significant difference between values at surface
acdom 0.017–0.160 0.046 0.032 0.015–0.144 0.057 0.031 and at the fluorescence peaks for any of the coefficients
apart 0.074–0.375 0.221 0.084 0.036–0.190 0.110 0.042 (Test-t, a ¼ 0.05).
adet 0.011–0.036 0.025 0.007 0.001–0.017 0.008 0.004 In summer (PATEX III), apart(440) were consistently
aph 0.074–0.339 0.190 0.075 0.035–0.172 0.099 0.038
a*ph 0.018–0.165 0.044 0.033 0.058–0.131 0.079 0.017
lower, with a mean value of 0.149 m1 at surface. During
this cruise, only the adet(440) values differed between

PATEX II
12 PATEX III

micro
10 nano 1.2
[chl a] per size class (mg m−3)

[chl a] per size class (mg m−3)

pico

8 1

0.8
6
0.6
4
0.4

2
0.2

0 0
0 2 4 6 8 10 12 14 0 0.5 1 1.5 2 2.5
total [chl a] (mg m−3) total [chl a] (mg m−3)

Fig. 2. Chlorophyll-a size-fractions (micro-, nano-e picoplankton) versus total [chl a] during PATEX II, spring (a) and PATEX III, summer (b) cruises.
Samples are from both surface and fluorescence peak depths. Note the different scales.
ARTICLE IN PRESS
A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174 2167

depths, with larger and variable values at the surface greater variation, notably due to differences in contribu-
(Table 1, right). acdom(440) magnitudes were also higher tion of CDOM versus that of phytoplankton. In summer, a
and more variable at the surface (average of group of points show a phytoplankton contribution
0.1270.09 m1) in summer, although not statistically around 60% and 90%, similarly to the spring cruise, but
different from the fluorescence peak (average of with small fractions of detritus (most points ranging
0.04570.032 m1), because of the high standard deviation between 0% and 10%). The remaining points from PATEX III
at the surface. showed a large dispersion throughout the range of
As seen in Table 1, phytoplankton was the dominant dissolved material contribution. In general, areas of higher
component of apart(l) in both PATEX II and PATEX III relative CDOM corresponded to low phytoplankton con-
cruises. In spring, phytoplankton contributed on average tribution and vice versa. The phytoplankton fraction
83% and 88% to total particulate absorption at 440 nm between 10% and 60% was associated with scattered
[(aph(440)/apart(440)] at surface and at the fluorescence contributions of CDOM and detritus. That indicates lack of
peak, respectively. In summer, the contributions were co-variation between non-algal components and phyto-
about 71% and 93%, respectively. On average, phytoplank- plankton, mainly in summer (PATEX III), and it suggests
ton accounted for approximately 83% of apart(440) in both that the contribution of non-algal components to light
periods. Despite an identical overall average contribution absorption must be highly variable, depending on the
of detritus in both periods (17%), the percent contribution bloom stage in the region.
of adet(440)/apart(440) at the surface was generally slightly
higher and more variable in summer (mean of 30715%) 3.3. Spectral slope parameters for detritus and CDOM
than in spring (mean of 1777%).
The absorption coefficients of particulate material In surface samples, the parameter that describes the
(phytoplankton+detritus) were statistically higher during spectral shape of detritus (Sdet) varied within a narrow
PATEX II (Test-t, a ¼ 0.05), clearly because of the higher range around 0.011 (70.0013) nm1 in PATEX II and 0.009
phytoplankton biomass, since highly variable values of (70.0007) nm1 in PATEX III (Table 2). In sub-surface
adet(440) showed no statistical difference between the peaks, a similar range was found for both PATEX II
periods of study. (0.01170.003) nm1 and PATEX III (0.00970.003) nm1,
The respective contributions of phytoplankton, detri- and therefore the parameters showed small variation
tus, and CDOM to total absorption at 440 nm (i.e. between depths and periods (Table 2).
apart(l)+acdom(l)) are displayed in a ternary plot (Fig. 3). Spectral slopes for CDOM (Scdom) at the surface showed
In spring (PATEX II, stars), the relative contribution of a statistically higher average in spring (0.0157
phytoplankton was between 60% and 80% in most 0.004 nm1) than in summer (0.01270.004 nm1) (Table
samples, while the contribution ranges of detritus and 2). However, at sub-surface peaks, there was no difference
CDOM were similar (between 10% and 30%). A distinct between periods (0.01470.004 nm1, PATEX II, 0.0137
pattern was clear in summer (PATEX III, triangles), with 0.004 nm1, PATEX III). Also, there was no significant
difference between surface and fluorescence peak values
between periods because of the high standard deviations
100 (Table 2). Overall, the average Scdom was 0.015 nm1
0
(70.004) during PATEX II and 0.012 nm1 (70.004)
90
10 during PATEX III. Clearly, values of Scdom showed greater
80 variation and higher magnitudes than those of detritus
20
(Sdet) (Table 2).
70
30 Fig. 4 illustrates the frequency distribution of the
%

exponential slopes, Scdm, of the CDM (CDOM+detritus)


OM

60
ph

40
yto
CD

50
p

50 Table 2
lan
%

Ranges, means and standard deviations of spectral shape (nm1) of


kto

40
60 absorption coefficients of CDOM, Scdom, detritus, Sdet and non-algal
n

components (cdom+detritus), Scdm, for surface (above) and fluorescence


30
70 peak depths (below) spring (PATEX II cruise) (left) and summer (PATEX
20 III cruise) (right) samples.
80
10 PATEX II PATEX III
90
0 Range Mean Std. dev. Range Mean Std. dev.
1
0 10 20 30 40 50 60 70 80 90 100 Surface
% detritus Scdom 0.007–0.026 0.015 0.004 0.005–0.029 0.012 0.004
Sdet 0.008–0.013 0.011 0.001 0.008–0.011 0.009 0.001
Fig. 3. Ternary plot illustrating the relative contribution of CDOM, Scdm 0.010–0.015 0.013 0.002 0.006–0.019 0.011 0.002
phytoplankton and detritus to light absorption for all samples (surface Fluorescence peak depths
and fluorescence peak depths), at 440 nm, for PATEX II (spring, stars) and Scdom 0.005–0.022 0.014 0.004 0.005–0.019 0.012 0.003
PATEX III (summer, open triangles). The relative contribution, in Sdet 0.007–0.018 0.011 0.003 0.005–0.019 0.010 0.003
percentage, of a component to total absorption can be read on the Scdm 0.003–0.127 0.013 0.002 0.005–0.017 0.012 0.003
corresponding axis.
ARTICLE IN PRESS
2168 A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174

16 25

14
20
12

10
15
8
N

N
6 10
4

2 5

0
0.005 0.01 0.015 0.02 0.025 0
Scdm (nm−1) 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
aph* (440) m2 mg [chl a]−1
Fig. 4. Frequency distribution of the exponential slope, Scdm, of the CDM
(detritus+CDOM) for PATEX II (spring, grey) and PATEX III (summer, clear Fig. 5. Frequency distribution of the specific phytoplankton absorption
bars) cruises. coefficients at 440 nm (i.e. aph(440)/[chl a]) for PATEX II (spring, grey)
and PATEX III (summer, clear bars) cruises.
absorption spectra, including both surface and fluores-
cence peaks, during both cruises. The range of values was a*ph(676) values for all depths were constrained between
narrower during PATEX II, with a mean value of 0.013 0.009 to 0.055 m2 mg [chl a]1 (mean of 0.0197
(70.002) nm1. During PATEX III, it varied between 0.005 0.010 m2 mg [chl a]1) in PATEX II and 0.016 to
and 0.019 (mean ¼ 0.01270.003) m1. 0.051 m2 mg [chl a]1 (mean of 0.03170.006 m2 mg [chl
a]1) in PATEX III. The equation provided by least-squares
fitting was aph(676) ¼ 0.03 ([chl a])0.651 (r2 ¼ 0.73;
3.4. Phytoplankton absorption coefficients and chlorophyll-a n ¼ 47) for PATEX II (spring) and aph(676) ¼ 0.03 ([chl
concentration a])1.064 (r2 ¼ 0.84; n ¼ 55) for PATEX III (summer). PATEX
II values agree with the Bricaud et al. (1995) data set,
Fig. 5 shows the frequency distribution of the chlor- while PATEX III still departs slightly from that trend.
ophyll-a-specific absorption coefficient for phytoplankton,
a*ph(440) (m2 mg [chl a]1), for both cruises. Except for
two extreme spectra, most values ranged from 0.018 to 3.5. Phytoplankton absorption spectra
0.097 m2 mg [chl-a]1 (mean of 0.04370.03 m2 mg [chl
a]1) for PATEX II. The a*ph(440) values were statistically The average phytoplankton spectra and their standard
higher during PATEX III (Test-t, a ¼ 0.05), varying between deviations for both periods are illustrated in Fig. 6. In
0.04 and 0.13 m2 mg [chl a]1 (mean 0.0970.02 m2 mg spring (PATEX II), the vast majority of spectra (43 of 47)
[chl a]1). The means and standard deviations are also showed high absorption in the ultraviolet (UV) region,
shown in Table 1, discriminating the surface and fluores- with peaks centered at 330 nm (not shown). Fig. 6b
cence peak depths, which did not show statistical presents spectra for the PATEX III cruise, whose absolute
difference during either cruise. values contrasted with the spring data (note the different
The relations between [chl a] and aph(440) were scales and see Table 1 for values in the blue absorption
defined by a power function, and a least-square fitting peak). In that period, relatively high values in the UV
provided the following relationship for PATEX II: range were also observed in various spectra (not shown),
aph(440) ¼ 0.09 ([chl a])0.420 (r2 ¼ 0.52; n ¼ 47) and for but at a much lower magnitude than during PATEX II.
PATEX III: aph(440) ¼ 0.083 ([chl a])0.938 (r2 ¼ 0.77; Because of the high variability verified in the relation-
n ¼ 55). The slopes (i.e. exponents) are higher than those ship between phytoplankton absorption values and [chl a]
established by Bricaud et al. (1995) for a large data set (as seen throughout the a*ph(l) values), aph(l) values were
with [chl a] varying between 0.02 and 25 mg m3, with normalized by the spectral mean of each aph(l) spectrum,
greater difference in PATEX III (not shown). In that cruise, a/phS, to examine differences in shape between them.
the slope was still higher than the one found by Bricaud Figs. 7a and b show the mean-normalized absorption
et al. (2004), which established a new relationship based spectra for whole samples and their mean and standard
on a different data set from their previous work, in which deviation for PATEX II and PATEX III, respectively. This
a shift of aph(440) toward higher values was found, normalization removes the effect of concentration and
compared to the first Bricaud et al. study. allows computation of variability due solely to spectral
The scatter of the points in the relationship between shape (Roesler et al., 1989).
[chl a] and aph(676) is lower, and the difference between As expected for normalized spectra influenced
the cruises is consistently less pronounced (not shown), by pigment packaging, spectra for samples dominated by
although statistically significant (Test-t, a ¼ 0.05). The microplankton at most (see Section 3.1.) were relatively
ARTICLE IN PRESS
A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174 2169

Fig. 6. Mean spectrum and standard deviation (SD) obtained from phytoplankton absorption spectra for all samples from (a) PATEX II (spring) and (b)
PATEX III (summer). Note the different Y-axis scales.

2.5
2.5

2
2
aph(λ)/a<ph>

aph(λ)/a<ph>

1.5 1.5

1 1

0.5 0.5

0 0
400 450 500 550 600 650 700 400 450 500 550 600 650 700
Wavelength (nm) Wavelength (nm)

Fig. 7. Shape of phytoplankton absorption spectra for all samples (dashed-dotted lines) from (a) PATEX II (spring, n ¼ 43) and (b) PATEX III (summer,
n ¼ 55). Spectra were normalized to the average phytoplankton absorption between 400 and 700 nm (a/phS). The mean normalized spectra (bold line)
and standard deviation (values multiplied by 3 for visualization) (dashed line) are also shown.

flat (PATEX II, Fig. 7a), and spectra referring to smaller- variation in pigment composition can play a role in overall
size-class dominance showed higher normalized values shape of this data set.
and higher blue-to-red ratio (PATEX III, Fig. 7b). Therefore, The mean-normalized absorption spectra from PATEX
the discrepancies between the two cruises are more III (Fig. 7b) indicate there is less variability in the spectral
pronounced in the blue absorption range, where the shape than those from PATEX II (Fig. 7a). But distinctions
packaging effect is maximal (Figs. 7a and b). are found between 440 and 490 nm, higher between 505
In PATEX II, a similar standard deviation shape for and 560 nm, and still important variability between 570
a/phS(l) as the mean spectrum (Fig. 7a) indicates that the and 600 nm (see standard deviation in Fig. 7b). In summer,
majority of the spectra have a similar shape and that peaks centered at 412 nm were not observed, as seen in
variations of pigment packaging dominated the variance PATEX II. Standard deviation spectra for a/phS(l) from
(Magnuson et al., 2004). However, a peak in this spectrum PATEX III did not resemble the shape of the mean spectra
is found at 457 nm, and a pronounced peak appears (Fig. 7b). Since absorption around 460 nm is attributed
between 500 and 550 nm (Fig. 7a), suggesting that to accessory pigments, the peak at 463 nm in the SD
ARTICLE IN PRESS
2170 A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174

0.9 November and March) likely represent higher phytoplank-


ton biomass (PATEX II) and moderated values throughout
0.8 the summer (PATEX III) in the region.
A remarkable phytoplankton bloom was observed
0.7
along the Patagonian shelf-break in spring 2006, with
0.6 mean [chl a] of 5.6 mg m3 (PATEX II), similar to that
verified in 2004 by Garcia et al. (2008). The authors
pointed out that the main factors favoring the develop-
Sf

0.5
ment and maintenance of these blooms are both nutrient
0.4 supply from the Malvinas Current and water-column
stability along the shelf-break front. In the present study,
0.3
phytoplankton biomass was much lower in summer than
0.2 in spring. These lower [chl a] (up to 2 mg m3) in summer
were associated with low (Pollery, R., personal commu-
0.1 nication), probably regenerated nutrients (decomposition
0 2 4 6 8 10 12 14 by bacteria and grazing), since upwelling does not seem to
[chl a] (mg m−3) occur in this period but is more frequent in spring
(Matano and Palma, 2008).
Fig. 8. Scatter plot of size spectral parameter (Sf) (non-dimensional) as a
Light absorption by water constituents in the Argen-
function of [chl a] for PATEX II (spring, stars) and PATEX III (summer,
open triangles). tinean shelf-break region was generally dominated by
phytoplankton during both periods (Fig. 3). There is a
considerable distance of the study area from waters of
continental origin, so phytoplankton is expected to be the
spectra suggests that pigment packaging (which is
most dominant particle in seawater. Both CDOM and
strongest in the chlorophyll-a absorption peak) was less
detritus absorption showed no relationship with [chl a].
important and that most of the variability in the blue band
Besides, CDOM absorption was relatively low and constant
of this dataset could be due to variations in pigment
in spring, while it was higher and more variable in
composition.
summer (Table 1 and Fig. 3). That suggests there is a time
lag between the phytoplankton bloom and CDOM in-
3.6. Size spectral parameter crease, consistent with other findings (Vodacek et al.,
1997; Siegel and Michaels, 1996; Bricaud et al., 1981).
Fig. 8 shows a plot of the size parameter for both Nelson et al. (1998) have shown that CDOM in the
cruises against [chl a], where an inverse relationship can Sargasso Sea follows a strong seasonal pattern that is
be seen. The spectral size parameter (Sf) varied between not correlated with [chl a], and have suggested that such a
0.21 and 0.79 in spring (PATEX II, mean of 0.4170.17), breakdown product of phytoplankton, resulting from
with higher values associated with small cell-size abun- microbial activity, is destroyed by photo-oxidation. The
dance in the northern section. Excluding this transect, relatively constant CDOM concentration in spring against
the mean value was about 0.35, corresponding to higher a great range of [chl a] (not shown) suggests that CDOM is
[chl a] values and to the dominance of the microplankton not produced by phytoplankton directly during the bloom.
fraction during PATEX II. In summer (PATEX III), the Moreover, high maximum irradiance levels verified in this
parameter values were between 0.48 and 0.87 (mean of season (1200 mE m–2 s–1, compared to E900 mE m–2 s–1 in
0.7270.08), in association with lower [chl a] values and a late summer, measured in our two cruises) must have an
small cell-size predominance during the period. important role in CDOM depletion in the upper layers.
Patagonia is an area influenced by the Antarctic polar
4. Discussion vortex and, thus, affected by air masses with low ozone
concentrations during October–November, in events of
The seasonal cycle of phytoplankton biomass has been variable duration (Orce and Helbling, 1997; Pérez et al.,
observed by recent analyses of ocean color data in the 1998). High levels of ultraviolet radiation and photo-
vicinity of the Patagonian shelf-break front (e.g. Garcia synthetically active radiation (PAR) have already been
et al., 2004; Rivas et al., 2006; Romero et al., 2006; observed in water bodies in the region, particularly in
Signorini et al., 2006). As presented by Romero et al. spring (Pérez et al., 1998), which must increase the photo-
(2006), strong [chl a] seasonal variability (44 mg m3) is oxidation. In fact, Bricaud et al. (1981) have demonstrated
observed in comparison with the open ocean (o1.5 mg that intense radiation in this spectral band reduced CDOM
m3) over the Patagonia shelf and shelf-break. Northward, absorption at 300 nm, about 75%, by indicating that a
high [chl a] begin in early austral spring, while southward fraction of the material is potentially sensible to photo-
blooms begin in late spring to early summer. The spring degradation.
maximum (43.5 mg m3) extends from the mid-shelf to Other studies have shown that a detritus absorption
the shelf-break, persisting at the last in summer (Romero peak occurs when a phytoplankton bloom starts to decline
et al., 2006). According to Rivas et al. (2006), peaks in (e.g. Sasaki et al., 2005). This probably happened in the
[chl a] over the shelf-break appear in October–November. period between our spring and summer cruises, although
Therefore, the timings of our cruises (beginning of an important variability in surface absorption values by
ARTICLE IN PRESS
A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174 2171

detritus could still be observed in summer (Table 1). In alterations in the material. Therefore, high acdom(l) values
addition, detritus is generated via planktonic food-web associated to lower Scdom could be evidence of photo-
interactions in grazing by zooplankton (as egested fecal oxidation (Vodacek et al., 1997) acting as an important
material), or as the result of decomposition of senescent regulator of CDOM concentrations in the region, as
cells (Nelson and Robertson, 1993). The importance of pointed out previously.
zooplankton biomass in the region is relatively well Semi-analytic models usually do not distinguish the
known (Vasilyev, 2007; Sabatini et al., 2004), and it is particulate and dissolved non-algal compartments, and
suggested that the herbivorous zooplankton must play an therefore utilize a mean parameter that describes the
important role in controlling biomass accumulations in shape of CDOM plus detritus spectra (e.g. Siegel et al.,
this ecosystem (Garcia et al., 2008). 2005; Maritorena et al. 2002). In the present work, there
was no statistical difference in Scdm values between
periods. That is an important result, as it indicates that
4.1. Spectral characteristics of absorption coefficients by this parameter does not present a considerable seasonal
non-algal components variation, at least in our study periods.

Information on seasonal variations in the absorption


coefficients of particulate and dissolved matter are 4.2. Phytoplankton absorption variability
required as inputs to optical models that seek to convert
satellite observations of water color into water-quality Specific absorption coefficients by phytoplankton are
information (Carder et al., 1999; Ciotti et al., 1999; Devred known to vary regionally and seasonally in response to
et al., 2005). Moreover, the spectral characteristics can changes in species composition, light and nutrient condi-
offer important information about the CDOM and detritus tions (e.g. Stuart et al., 2000; Sathyendranath et al., 1999;
composition and source (Twardowski et al., 2004; Babin Lutz et al., 1996). The non-linear relationship between
et al., 2003). aph(l) and [chl a] is referred to as the ‘pigment package
As verified in several studies, our Sdet values were effect’ (Duyens, 1956; Morel and Bricaud, 1981; Cleveland,
rather small and less variable than Scdom. In spring, 1995), which is caused by variations in cell size and
slightly higher Sdet values could be partly attributed to internal pigment concentration and/or overlapped pig-
silica frustules of the abundant diatoms (see taxonomic ments. At 440 nm, carotenoids, chlorophylls b and c and
information below). Nearly identical values of Sdet phicobilins can contribute to the absorption, resulting in
(0.011 nm1) were found for mineral particles by Bowers an increase in the specific absorption coefficient, in
et al. (1996) and, according to Bukata et al. (1995), the relation to a given [chl a], with important effects on the
parameter tends to be higher in elevated proportions of parameterization based only on the concentration of the
mineral to organic matter. Furthermore, in both periods, a main pigment (Lutz et al., 1996). At 676 nm, the absorp-
large variation in the detritus spectral parameter was tion is dominated by chlorophyll-a. Differences between
noted in fluorescence peak depths, probably reflecting a our values of aph(440) and those predicted by the Bricaud
greater heterogeneity of detritus at depth. These fluores- et al. (1995, 2004) regression can be partly explained by
cence peaks were generally close to the bottom of the possible differences in pigment compositions and cell size
mixing layer (not shown), suggesting that an important in our study region, relative to what is typically found in
component of the particle assembly at those depths could other Case 1 waters with similar [chl a]. In the Patagonian
be material that sedimented from surface layers, being shelf-break region, the spring bloom consisted of large
important areas of remineralization. diatoms dominated mainly by chains of Thalassiosira sp.
Based on comparison with ranges reported previously, (6–30 mm), large Corethron pennatum (long axis 187 mm)
the variability noted in our Scdom values was small and and Thallasiothrix antarctica (500 mm apical axis), large
close to what was found in other regions, even when both dinoflagellates (Gymnodiniales) (23–95 mm) and an im-
cruises were considered. Twardowski et al. (2004) have portant contribution of colonial Phaeocystis cf. antarctica
revised Scdom values reported in the literature and (Garcia et al., 2008; R. Comin, personal communication). It
suggested that the relatively broad range in Scdom values is well known that large phytoplankton can exhibit low
may be partly due to the fact that various methods were relative aph*(l) values. Therefore, a peculiar pigment
used to measure acdom(l) and to derive Scdom (such as composition may be responsible for the higher relative
optical setup, spectral range, baseline correction, fitting values found in the blue portion of the spectra in spring,
procedure). Although the same procedures for whole because the measured aph(676) when plotted as a function
samples were mantained, statistical differences were of [chl a] are consistent with the prediction from
found between our sampling periods, with higher Scdom Bricaud et al. (1995), and thus the size structure should
values in summer. That is in agreement with Carder et al. be quite similar (see Bricaud et al., 2004). Note that the
(1989) and others, who observed inverse relationships ‘package effect’ is determined not only by cell size, but
between acdom(l) and Scdom. As proposed by Twardowski also by photoacclimation processes which modify the
and Donaghay (2002), higher slopes found in oceanic intracellular pigment concentrations. High levels of both
waters, in contrast to coastal waters, must reflect a UV radiation and PAR observed in water bodies of the
cumulative effect of photo-bleaching. In addition, the Patagonia region, particularly in spring (Pérez et al., 1998),
spectral parameter tends to be higher at the surface in associated with high absorption in the UV shown here,
relation to deeper waters, as a result of chemical suggest the presence of mycosporine-like aminoacids
ARTICLE IN PRESS
2172 A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174

(MAAs) in the phytoplankton populations of the region inance of microplankton during the spring cruise, PATEX II
and consequently photoacclimation processes. That (see Section 3.1), some spectra were dominated by small
may result in high relative levels of photo-protector cells. Because of that, spectra were also grouped according
carotenoids, and it could partially explain the relatively to the dominant [chl a] size fractions in that sample,
high specific absorption coefficients in the spring (PATEX throughout the criteria that assumed the dominant size to
II). The fact that summer spectra (PATEX III) showed be the one containing more than 50% of the total [chl a]
relatively high absorption in the UV range (not shown) (see Fig. 2). This separation showed lower blue-to-red
also suggests the presence of the MAAs. However, peak ratios in samples dominated by microplankton and
accumulations of these compounds are generally attrib- higher ratios in the few spectra dominated by small cells
uted to larger-size groups, probably because of greater (not shown). That indicates size was a major factor
storage capacity (Hannach and Sigleo, 1998; Helbling controlling the shape of phytoplankton absorption spectra
et al., 1996), which can explain the high absorption in the in such a data set, with a great magnitude of the ‘package
UV range in PATEX II spectra. Nonetheless, this has effect’ in the majority of samples dominated by bigger
also been observed in cyanobacteria (Garcia-Pichel and cells (see Section 3.1), consistently with the increase in
Castenholz, 1993), indicating the need for further research pigment packaging (Fig. 7a). This is in agreement with the
to confirm this fact in our study. dominance by big cells in the period, as already men-
In summer (PATEX III), small phytoplankton cells tioned. Carotenoid effects are also suggested by the fact
should result in a weaker ‘package effect’, revealed by that the standard deviation peak does not correspond to
much higher aph(l), if compared to spring values. Also, the the chlorophyll-a absorption peak in the blue region
fact that summer values are still slightly above those (440 nm). In addition, a distinct peak around 412 nm,
found by Bricaud et al. (1995) at 676 nm suggests that the where phaeopigment absorption is maximal (Roesler
phytoplankton cell-size structure found in this study may et al., 1989), suggests that the bloom could be in an
be the main factor responsible for departures found in the intermediate to late stage of development, which is an
general spectra. In fact, dominance of small nanoflagel- important observation, in the sense that this blue portion
lates (o10 mm), Phaeocystis cf. antarctica and high con- is usually attributed to CDOM absorption.
centrations of Synechococcus sp. were verified by both The whole spectra in PATEX III corresponded to
microscopy and flow cytometry (R. Comin and M. Tenório, dominance by nanoplankton (considering the same
personal communication) in the summer samples. Since criteria based on percentage of size-fractionated [chl a]).
departures from the regression obtained by Bricaud et al. Thus, the variability within the period could be attributed
(1995) are higher at 440 nm, both differences in pigment to variations in pigment composition. The weak ‘package
composition and size structure seem to contribute to the effect’ comparing to PATEX II spectra suggests that it
differences between summer data (PATEX III) and the might be appropriate to replace the global parameteriza-
other studied regions. This is in agreement with increas- tions by locally and seasonally adapted relationships in
ing trends in non-photosynthetic pigments-to-chl a ratio order to apply bio-optical models for the region.
in small cells (Bricaud et al., 1995). The great difference between size parameter values
Another important point to note is that differences from both periods derived from phytoplankton absorption
between results from this study and those of Bricaud et al. spectra emphasize the strong ‘package effect’ of phyto-
could also be attributed to the choice of the Beta plankton cells in spring (PATEX II), in contrast to the weak
correction algorithm. By recalculating our particulate effect in the summer organisms (PATEX III), as seen
absorption values, using the Beta algorithm from their throughout the absorption spectra (Fig. 7). Moreover, the
studies (Bricaud and Stramski, 1990), it could be seen that agreement between these values and information about
the values provided by our factor are approximately the taxonomic composition obtained by traditional meth-
25–30% higher, at most, in the blue spectral band. This ods (microscopy, citometry) indicates the potential for
is comparable to the differences between distinctive assessing changes in cell size from optical measurements.
algorithms presented in the literature, and it would not Ciotti and Bricaud (2006) have shown a good agreement
explain the departures by the present data set itself. An between in situ Sf values and those obtained from
important source of discrepancies is the fact that Bricaud phytoplankton absorption retrieved from water-leaving
et al. used HPLC-derived pigment concentrations, which radiances at SeaWiFS channels. This inversion method can
can differ markedly from Turner-fluorescence-derived [chl be an important tool to characterize the remarkable
a], especially in diatom-dominated waters. Indeed, HPLC seasonality in the phytoplankton community size struc-
[chl a] measurements were conducted in a later PATEX ture in the Patagonian shelf-break region from remote
cruise, and they revealed lower values by a factor sensing. This information can contribute to understanding
of approximately 1.5 compared to fluorescence derived the biogeochemical processes that are associated with
[chl a] (Mendes, C.R., personal communication). In this the different community structures. As presented by
case, our phytoplankton absorption values would be even Garcia et al. (2008), the primary production rates and
higher for a given [chl a], resulting in higher a*ph(l) values diatom dominance verified in spring indicate a potentially
compared to HPLC-derived [chl a]. significant biological control of gases such as O2 and
The analyses of spectra normalized by the average CO2 in surface layers, as opposed times when small
phytoplankton absorption between 400 and 700 nm (Fig. 7) flagellates are the prevailing phytoplankton group
revealed the same trends described above throughout (Schloss et al., 2007), which was the case in our summer
specific wavelength absorption values. Despite the dom- cruise.
ARTICLE IN PRESS
A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174 2173

The present work quantified the magnitudes and relationship with fronts and chlorophyll distribution. Journal of
spectral characteristics of light absorption by total Geophysical Research 114, C03018, doi:10.1029/2008JC004854.
Bowers, D.G., Harker, G.E.L., Stephan, B., 1996. Absorption spectra of
particulate (phytoplankton and detritus) and CDOM in inorganic particles in the Irish Sea and its relation to tidal stirring.
two distinct bloom stages along the Patagonia shelf-break International Journal of Remote Sensing 19, 2789–2805.
front. Regarding the absorption properties, the phyto- Bricaud, A., Morel, A., Prieur, L., 1981. Absorption by dissolved organic
matter of the sea (yellow substance) in the UV and visible domains.
plankton was the dominant component during both
Limnology and Oceanography 26, 43–53.
cruises. Although CDOM seems to be an important Bricaud, A., Stramski, D., 1990. Spectral absorption coefficients of living
component of the absorption budget, it is probably a phytoplankton and nonalgal biogenous matter: a comparison
between the Peru upwelling area and the Sargasso Sea. Limnology
phytoplankton by-product. It must be addressed that
and Oceanography 35 (3), 562–582.
CDOM could contribute to absorption in a higher degree Bricaud, A., Babin, M., Morel, A., Claustre, E.H., 1995. Variability in the
verified here, depending on other bloom stages. To a chlorophyll-specific absorption coefficients of natural phytoplank-
certain extent, the region could be considered as a ‘Case I’ ton: analysis and parameterization. Journal of Geophysical Research
100 (C7), 13,321–13,332.
water. However, it seems reasonable not to consider it as a Bricaud, A., Morel, A., Babin, M., Allali, K., Claustre, H., 1998. Variations of
final conclusion, given that only two periods were light absorption by suspended particles with chlorophyll a concen-
sampled. Important information was generated from two tration in oceanic (case 1) waters: analysis and implications for
bio-optical models. Journal of Geophysical Research 103 (C13),
distinct periods of the phytoplankton seasonal cycle,
31033–31044.
which in association with other determinant optical Bricaud, A., Claustre, H., Ras, J., Oubelkheir, K., 2004. Natural variability of
properties, such as scattering, attenuation and reflectance phytoplanktonic absorption in oceanic waters: influence of the size
(Garcia et al., in preparation) shall contribute to optical structure of algal populations. Journal of Geophysical Research 109
(C11010).
parameterizations for using remote-sensing techniques. Bukata, R.P., Jerome, J.H., Kondratyev, K.Y., Pozdnyakov, D.V., 1995. Optical
Properties and Remote Sensing of Inland and Coastal Waters. CRC
Press, Boca Raton, FL.
Carder, K.L., Steward, R.G., Harvey, G.R., Ortner, P.B., 1989. Marine humic
and fulvic acids: their effects on remote sensing of ocean chlorophyll.
Acknowledgements Limnology and Oceanography 36, 68–81.
Carder, K.L., Chen, F.R., Lee, Z.P., Hawes, S.K., Kamykowski, D., 1999.
Semianalytic moderate-resolution imaging spectrometer algorithms
The PATagonian EXperiment (PATEX) is a multidisci- for chlorophyll-a and absorption with bio-optical domains based on
plinary project as part of Group of high latitude oceano- nitrate depletion temperatures. Journal of Geophysical Research 104,
graphy (GOAL) activities in the Brazilian Antarctic 5403–5421.
Carder, K.L., Chen, F.R., Cannizzaro, J.P., Campbell, J.W., Mitchell, B.G.,
Program. We thank the crew of the Brazilian Navy 2004. Performance of the MODIS semi-analytical ocean color
research ship ‘‘Ary Rongel’’ for their assistance during the algorithm for chlorophyll-a. Advances in Space Research 33,
field sampling. The project was sponsored by CNPq 1152–1159.
Ciotti, A.M., Cullen, J.J., Lewis, M.R., 1999. A semi-analytical model of the
(Brazilian National Council on Research and Development)
influence of phytoplankton community structure on the relationship
through Grant 557305/2005-5. Ocean color images were between light attenuation and ocean color. Journal of Geophysical
provided by GSFC/NASA. A. Ferreira was granted a Research 104 (C1), 1559–1578.
graduate scholarship from the Brazilian agency Coorde- Ciotti, A.M., Lewis, M.R., Cullen, J.J., 2002. Assessment of the relationships
between dominant cell size in natural phytoplankton communities
nac- ão de Aperfeic- oamento de Pessoal de Ensino Superior and the spectral shape of the absorption coefficient. Limnology and
(CAPES). Thanks are due to Áurea Maria Ciotti for Oceanography 47, 404–417.
providing MATLABs scripts used to calculate the ‘size Ciotti, A.M., Bricaud, A., 2006. Retrievals of a size parameter for
phytoplankton and spectral light absorption by colored detrital
parameter’, Sf. We also thank three anonymous reviewers matter from water-leaving radiances at SeaWiFS channels in a
for their reviews and helpful comments. continental shelf region off Brazil. Limnology and Oceanography-
Methods 4, 237–253.
Cleveland, J.S., 1995. Regional models for phytoplankton absorption as a
References function of chlorophyll a concentration. Journal of Geophysical
Research 100 (7), 13333–13344.
Acha, M.E., Mianzan, H.W., Guerrero, R.A., Favero, M., Bava, J., 2004. Devred, E., Fuentes-Yaco, C., Sathyendranath, S., Caverhill, C., Platt, T.,
Marine fronts at the continental shelves of South America. Physical Mass, H., 2005. A semi-analytical, seasonal algorithm to retrieve the
and ecological processes. Journal of Marine Systems 44 (1/2), chlorophyll-a concentration in the Northwest Atlantic from SeaWiFS
83–105. data. Indian Journal of Marine Sciences 34 (4), 356–367.
Babin, M., Stramski, D., 2002. Light absorption by aquatic particles in the Duyens, L.N.M., 1956. The flattening of the absorption spectra
near-infrared spectral region. Limnology and Oceanography 47, of suspensions as compared to that of solutions. Biochimica et
911–915. Biophysica Acta 19, 1–12.
Babin, M., Stramski, D., Ferrari, G.M., Claustre, H., Bricaud, A., Obolensky, Garcia-Pichel, F., Castenholz, R.W., 1993. Occurrence of UV-absorbing,
G., Hopepffner, N., 2003. Variations in the light absorption coeffi- mycosporine-like compounds among cyanobacterial isolates and an
cients of phytoplankton, nonalgal particles, and dissolved organic estimate of their screening capacity. Applied and Environmental
matter in coastal waters around Europe. Journal of Geophysical Microbiology 59, 163–169.
Research 108 (C7), 3211. Garcia, C.A.E., Sarma, Y.V.B., Mata, M.M., Garcia, V.M.T., 2004. Chlorophyll
Behrenfeld, M.J., Randerson, J.T., McClain, C.R., Feldman, G.C., Los, S.O., variability and eddies in the Brazil-Malvinas Confluence region.
Tucker, C.J., Falkowski, P.G., Field, C.R., Frouin, C.B., Esaias, W.E., Deep-Sea Research Part II-Topical Studies in Oceanography 51,
Kolber, D.D., Pollack, N.H., 2001. Biospheric primary production 159–172.
during an ENSO transition. Science 291, 2594–2597. Garcia, V.M.T., Garcia, C.A.E., Mata, M.M., Pollery, R.C., Piola, A.R.,
Bianchi, A., Bianucci, L., Piola, AR., Pino D.R., Schloss, I., Poisson A., Signorini, S., Mcclain, C.R., 2008. Environmental factors controlling
Balestrini, C.F., 2005. Vertical stratification and air-sea CO2 fluxes in the phytoplankton blooms at the Patagonia shelf-break in spring.
the Patagonian shelf. Journal of Geophysical Research 110, C07003, Deep-Sea Research Part I-Oceanographic Research Papers 55,
doi:10.1029/2004JC002488. 1150–1166.
Bianchi, A.A., Pino, D.R., Perlender, H.G.I., Osiroff, A.P., Segura, V., Lutz, V., Garver, S., Siegel, D.A., Mitchell, B.G., 1994. A statistical analysis of
Clara, M.L., Balestrini, C.F., Piola, A.R., 2009. Annual balance and particulate absorption spectra: What can a satellite ocean color
seasonal variability of sea-air CO2 fluxes in the Patagonia Sea: Their imager see? Limnology and Oceanography 39, 1349–1367.
ARTICLE IN PRESS
2174 A. Ferreira et al. / Deep-Sea Research I 56 (2009) 2162–2174

Gordon, H.R., Brown, O.B., Evans, R.H., Brown, J.W., Smith, R.C., Baker, K.S., Prieto, L., Vaillancourt, R.D., Hales, B., Marra, J., 2008. On the relationship
Clark, D.K., 1988. A semianalytic radiance model of ocean color. between carbon fixation efficiency and bio-optical characteristics of
Journal of Geophysical. Research 93, 10,909–10,924. phytoplankton. Journal of Plankton Research 30 (1), 43–56.
Hannach, G., Sigleo, A.C., 1998. Photoinduction of UV-absorbing com- Reynolds, R.A., Sstramski, D., Mitchell, E.B.G., 2001. A chlorophyll-dependent
pounds in six species of marine phytoplankton. Marine Ecology semianalytical reflectance model derived from field measurement of
Progress Series 174, 207–222. absorption and backscattering coefficients within the Southern Ocean.
Helbling, E.W., Chalker, B.E., Dunlap, W.C., Holm-Hansen, O., Villafañe, Journal of Geophysical Research 106 (C4), 7125–7138.
V.E., 1996. Photoacclimation of antarctic marine diatoms to solar Rivas, A.L., Dogliotti, A.I., Gagliardini, D.A., 2006. Seasonal variability in
ultraviolet radiation. Journal of Experimental Marine Biology and satellite-measured surface chlorophyll in the Patagonian Shelf.
Ecology 204, 85–101. Continental Shelf Research 26, 703–720.
Hu, C., Lee, Z., Muller-Karger, F.E., Carder, K.L., Walsh, J.J., 2006. Ocean color Roesler, C., Perry, M., Carder, K., 1989. Modeling in situ phytoplankton
reveals phase shift between marine plants and yellow substance. IEEE absorption from total absorption spectra in productive inland marine
Geoscience and Remote Sensing Letters 3 (2), 262–266. waters. Limnology and Oceanography 34, 1510–1523.
IOCCG, 2006. Remote sensing of inherent optical properties: funda- Romero, S.I., Piola, A.R., Charo, M., Garcia, C.A.E., 2006. Chlorophyll-a
mentals, tests of algorithms, and applications. In: Lee, Z.P. (Ed.), variability off Patagonia based on SeaWiFS data. Journal of
Reports of the International Ocean-Colour Coordinating Group, No. 5. Geophysical Research 111, C05021.
IOCCG, Dartmouth. Sabatini, M., Reta, R., Matano, R., 2004. Circulation and zooplankton
Kishino, M., Takahashi, N., Okami, N., Ichimura, S., 1985. Estimation of the biomass distribution over the southern Patagonian shelf during late
spectral absorption coefficients of phytoplankton in the sea. Bulletin summer. Continental Shelf Research 24, 1359–1373.
of Marine Science 37, 634–642. Sasaki, H., Miyamura, T., Saitoh, S.-i., Ishizaka, J., 2005. Seasonal variation
Lutz, V.A., Sathyendranath, S., Head, J.H., 1996. Absorption coefficient of of absorption by particles and colored dissolved organic matter
phytoplankton: regional variations in the North Atlantic. Marine (CDOM) in Funka Bay, southwestern Hokkaido, Japan. Estuarine and
Ecology Progress Series 135, 197–213. Coastal Shelf Science 64, 447–458.
Magnuson, A., Harding, L.W., Mallonee, M.E., Adolf, J.E., 2004. Bio-optical Sathyendranath, S., Stuart, V., Irwin, B.D., Maass, H., Savidge, G., Gilpin, L.,
model for Chesapeake Bay and the middle Atlantic bight. Estuarine, Platt, T., 1999. Seasonal variations in bio-optical properties of
Coastal and Shelf Science 61, 403–424. phytoplankton in the Arabian Sea. Deep-Sea Research Part II-Topical
Maritorena, S., Siegel, D.A., Peterson, A.R., 2002. Optimization of a Studies in Oceanography 46, 633–653.
semianalytical ocean color model for global-scale applications. Sathyendranath, S., Cota, G., Stuart, V., Maass, H., Platt, T., 2001. Remote
Applied Optics 41, 2705–2714. sensing of phytoplankton pigments: a comparison of empirical and
Marra, J., Trees, C.C., O’Reilly, J.E., 2007. Phytoplankton pigment absorption: theoretical approaches. International Journal of Remote Sensing 22,
a strong predictor of primary productivity in the surface ocean. Deep- 249–273.
Sea Research Part I-Oceanographic Research Papers 54, 155–163. Schloss, I.R., Ferreyra, G.A., Ferrario, M.E., Almandoz, G.O., Codina, R.,
Matano, R.P., Palma, E.D., 2008. On the upwelling of downwelling currents. Bianchi, A.A., Balestrini, C.F., Ochoa, H.A., Ruiz Pino, D., Poisson, A.,
Journal of Physical Oceanography, doi:10.1175/2008JPO3783.1. 2007. Role of plankton communities in sea–air variations in pCO2 in
Mitchell, B.G., 1990. Algorithms for determining the absorption coeffi- the SW Atlantic Ocean. Marine Ecology Progress Series 332, 93–106.
cient of aquatic particulates using the quantitative filter technique Sieburth, M.McN., Smetacek, V., Lenz, J., 1978. Pelagic ecosystem
(QFT). Ocean Optics X 1302, 137–148. structure: heterotrophic compartments of the plankton and their
Mitchell, B.G., Bricaud, A., Carder, K., Cleveland, J., Ferrari, G.M., Gould, R., relationships to plankton size fractions. Limnology and Oceanogra-
Kahru, M., Kishino, M., Maske, H., Moisan, T., Moore, L., Nelson, N., phy 23, 1256–1263.
Phinney, D., Reynolds, R.A., Sosik, H., Stramski, D., Tassan, S., Trees, C., Siegel, D.A., Michaels, A.F., 1996. Quantification of non-algal attenuation
Weidemann, A., Wieland, J.D., Vodacek, A., 2000. Determination of in the Sargasso Sea: implications for biogeochemistry and remote
spectral absorption coefficients of particles, dissolved material and sensing. Deep-Sea Research Part II-Topical Studies in Oceanography
phytoplankton for discrete water samples. In: Fargion, G.S., Mueller, 43, 321–345.
J.L., McClain, C.R. (Eds.), Ocean Optics Protocols for Satellite Ocean Siegel, D.A., Maritorena, S., Nelson, N.B., Behrenfeld, M.J., Mcclain, C.R.,
Color Sensor Validation, Revision 2. NASA, Goddard space flight 2005. Colored dissolved organic matter and its influence on the
center, Greenbelt, Maryland, pp. 125–153. satellite-based characterization of the ocean biosphere. Geophysical
Moore, L.R., Goericke, R., Chisholm, S.W., 1995. Comparative physiology Research Letters 32, L20605.
of Synechococcus and Prochlorococcus: influence of light and Signorini, S.R., Garcia, V.M.T., Piola, A.R., Garcia, C.A.E., Mata, M.M.,
temperature on growth, pigments, fluorescence and absorptive Mcclain, C.R., 2006. Seasonal and interannual variability of cocco-
properties. Marine Ecology Progress Series 116, 259–275. lithophore blooms in the vicinity of the Patagonian shelf break
Morel, A., Bricaud, A., 1981. Theoretical results concerning light (381S–521S). Geophysical Research Letters 33, L16610.
absorption in a discrete medium, and application to specific Stuart, V., Sathyendranath, S., Head, E.J.H., Platt, T., Irwin, B., Maass, H., 2000.
absorption of phytoplankton. Deep-Sea Research 28, 1375–1393. Bio-optical characteristics of diatom and prymnesiophyte populations
Morel, A., 1988. Optical modeling of the upper ocean in relation to its in the Labrador Sea. Marine Ecology Progress Series 201, 91–106.
biogenous matter content (Case I waters). Journal of Geophysical Trees, C.C., Clark, D.K., Bidigare, R.R., Ondrusek, M., Mueller, J., 2000.
Research 93 (C9), 10,749–10,768. Accessory pigments versus chlorophyll a concentrations within
Morel, A., Maritorena, S., 2001. Bio-optical properties of oceanic waters: a euphotic zone: a ubiquitous relationship. Limnology and Oceano-
reappraisal. Journal of Geophysical Research 106, 7163–7180. graphy 45, 1130–1143.
Morel, A., Gentili, B., Chami, M., Ras, J., 2006. Bio-optical properties of Twardowski, M.S., Donaghay, P.L., 2002. Photobleaching of aquatic dissolved
high chlorophyll Case 1 waters, and of yellow-substance dominated materials: absorption removal, spectral alteration, and their interrela-
Case 2 waters. Deep-Sea Research Part I-Oceanographic Research tionship. Journal of Geophysical Research 107 (C8), 3091.
Papers 53, 1439–1559. Twardowski, M.S., Boss, E., Sullivan, J.M., Donaghay, P.S., 2004. Modeling
Nelson, N.B., Siegel, D.A., Michaels, A.F., 1998. Seasonal dynamics of the spectral shape of absorption by chromophoric dissolved organic
colored dissolved organic material in the Sargasso Sea. Deep-Sea matter. Marine Chemistry 89, 69–88.
Research Part I-Oceanographic Research Papers 45, 931–957. Twardowski, M.S., Lewis, M., Barnard, A., Zaneveld, J.R.V., 2005. In-water
Nelson, J.R., Robertson, C.Y., 1993. Detrital spectral absorption: Labora- instrumentation and platforms for ocean color remote sensing
tory studies of visible light effects on phytodetritus absorption, applications. In: Miller, R., Del Castillo, C., McKee, B. (Eds.), Remote
bacterial spectral signal, and comparison to field measurements. Sensing of Coastal Aquatic Waters. Springer Publishing, Dordrecht,
Journal of Marine. Research 51, 181–207. Netherlands, pp. 69–100.
O’Reilly, J., Maritorena, S., Mitchell, B.G., Siegel, D., Carder, K.L., Garver, S., Vasilyev, V.I., 2007. Zooplankton of the Patagonian slope and adjacent
Kahru, M., Mcclain, C., 1998. Ocean color chlorophyll algorithms for waters in the autumn. Bulletin of Moscow University, Biological
SeaWiFS. Journal of Geophysical Research 103, 24937–24953. Sciences/Vestnik Moskovskiin Universitet Biologiia 62 (1), 44–48.
Orce, V.L., Helbling, E.W., 1997. Latitudinal uvr-par measurements in Vodacek, N.V., Blough, M.D., Degrandpre Peltzer, E.T., Nelson, R.K., 1997.
Argentina: extent of the ‘‘ozone hole’’. Global and Planetary Change Seasonal variation of CDOM and DOC in the Middle Atlantic Bight:
15, 113–121. terrestrial inputs and photooxidation. Limnology and Oceanography
Pérez, A., Aguirre de Carcer, I., Tocho, J.O., Crino, E., Ranea, Sandoval, H.F., 42, 674–686.
Berni, M.E., 1998. The extent of the ozone hole over South America Welschmeyer, N.A., 1994. Fluorometric analysis of chlorophyll-a in the
during the spring of 1993, 1994 and 1995. Journal of Physics presence of chlorophyll-b and pheopigments. Limnology and
D-Applied Physics 31, 812–819. Oceanography 39 (8), 1985–1992.

You might also like