You are on page 1of 12

Deep-Sea Research Part II xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Deep-Sea Research Part II


journal homepage: www.elsevier.com/locate/dsr2

Bio-optical characterization of the northern Antarctic Peninsula waters:


Absorption budget and insights on particulate backscattering

Amabile Ferreiraa, , Áurea M. Ciottia, Carlos A.E. Garciab,c
a
Centro de Biologia Marinha, Universidade de São Paulo, Rod. Manuel Hypólito do Rego,Rego, Km 131.5, Praia do Cabelo Gordo, São Sebastião, SP 11612-109, Brazil
b
Instituto de Oceanografia, Universidade Federal do Rio Grande, Rio Grande, RS 96201-900, Brazil
c
Programa de Pós-Graduação em Oceanografia, Universidade Federal de Santa Catarina, SC 88900-900, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: We utilize a comprehensive set of bio-optical properties collected in the northern Antarctic Peninsula to con-
Phytoplankton tribute to the poor documentation of in situ bio-optical measurements in the Southern Ocean. The relative
cdom contributions of phytoplankton, colored dissolved organic matter (cdom) and detritus to light absorption high-
Detritus light the importance of phytoplankton, but cdom contribution is often important in the blue range of the spectra
Absorption
despite its roughly constant values at 443 nm along the three orders of magnitude of total chlorophyll-a con-
Scattering
centration, [TChl a], (0.019–2.91 mg m−3) observed in the upper 100 m. The particulate backscattering coef-
Backscattering
ficient, bbp(λ), was remarkably low if compared to other oceanic waters, but agrees with previous studies in the
Southern Ocean. Even with very low absorption, detritus was the component better correlated with particulate
backscattering in the Antarctic Peninsula, while phytoplankton cells (dominant in the particles pool) mostly
covaried with particulate scattering. Particulate backscattering ratios (bbp(λ) divided by the particulate scat-
tering coefficient, bp(λ)) were also below values observed in other oceanic waters. The spectral diffuse at-
tenuation coefficient, Kd(λ), was highly correlated to [TChl a] (R2=0.90 at 443 nm) and showed no dependence
on bbp(λ). Indeed, Kd(443) and non-water absorption coefficients at 443 nm were related by a 1 to 1 dependence.
The shape of the spectral remote sensing reflectance varied little responding mainly to variability in [TChl a],
while [TChl a] vs. maximum band ratios dependence deviated from global trends in a very similar fashion as in
other studies of the Southern Ocean, likely due to very low bbp(λ).

1. Introduction predict optical coefficients such as the particulate (e.g. Bricaud et al.,
1998) and phytoplankton (e.g., Bricaud et al., 2004; Bricaud et al.,
The optical properties of seawater can be reconstructed from the 1995) absorption, the particulate beam attenuation (e.g., Behrenfeld
additive contributions of water molecules, phytoplankton, nonalgal and Boss, 2006; Loisel and Morel, 1998), the particulate scattering (e.g.
particles (mainly detritus) and colored dissolved organic matter (cdom), Gordon and Morel, 1983; Huot et al., 2008) and the particulate back-
called bio-optical components. The magnitude and spectral shape of scattering (Antoine et al., 2011; Dall’Olmo et al., 2009; Huot et al.,
light absorption and scattering, inherent properties (Preisendorfer, 2008).
1961) are altered by changes in concentrations and compositions of While variations in the phytoplankton absorption coefficient are
these components (e.g., IOCCG, 2006). The characterizations of optical well documented, spectral absorption of cdom (e.g., Babin et al., 2003;
properties in different portions of the ocean are necessary steps for Bricaud et al., 1981) and detritus (e.g., Babin et al., 2003; Bricaud et al.,
using optical measurements, performed by in situ and remote platforms 1998; Bukata et al., 1995) have been less systematically investigated,
as tools for biogeochemical studies (e.g., Johnson et al., 2009). and because of that, frequently disregarded or simple assumed to
In case 1 waters, where phytoplankton and its co-varying particu- covary with chlorophyll-a concentration (Carder et al., 1991; but see
late and dissolved materials govern the optical signal (Gordon and Ciotti et al., 1999). In situ measurements of the beam attenuation
Morel, 1983; Morel and Prieur, 1977; Prieur and Sathyendranath, coefficients have long been carried out routinely as an index for par-
1981), parameterizations of bio-optical properties as functions of ticles (see Bartz et al., 1978), in contrast to the backscattering coeffi-
chlorophyll-a concentration have been useful for biogeochemical cient that only recently became systematically measured (e.g., Antoine
modeling (e.g., Morel, 1988). Chlorophyll-a concentrations are used to et al., 2011; Slade and Boss, 2015; Huot et al., 2008; Reynolds et al.,


Corresponding author.
E-mail address: amabilefr@gmail.com (A. Ferreira).

http://dx.doi.org/10.1016/j.dsr2.2017.09.007

0967-0645/ © 2017 Elsevier Ltd. All rights reserved.

Please cite this article as: Ferreira, A., Deep-Sea Research Part II (2017), http://dx.doi.org/10.1016/j.dsr2.2017.09.007
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

Fig. 1. Geographic locations of the sampling stations in the Gerlache Strait (GS, triangles), Bransfield Strait (BS, stars), northwestern Weddell Sea (NWW, circles). NWWHT denotes “high
transparency” for two stations sampled in 2014 (asterisks). In GS, BS and NWW, most of the stations of 2013 were sampled in 2014 and are indicated with the edge markers in black.

2016; Twardowski et al., 2007). Backscattering measurements may comparing to previous relationships found elsewhere. A complementary
provide information on particulate organic matter (Stramski et al., purpose of this study is to explore the prediction of chlorophyll-a
2008), phytoplankton carbon concentration (Graff et al., 2015, 2016), concentration and absorption coefficient from the diffuse attenuation
chlorophyll-a to autotrophic carbon ratios (Cetinić et al., 2015) and coefficient for downward irradiance, Kd(λ). This is of special interest
particle size distributions (Kostadinov et al., 2009; Loisel et al., 2006; because Kd(λ) is a relatively simple measurement to obtain in the field
Slade and Boss, 2015). and provides absorption (e.g., Loisel and Stramski, 2000), justified by
A large dispersion is generally observed in global empirical re- the dominance of absorption in the diffuse attenuation process (Morel
lationships developed for bio-optical properties or parameter as func- et al., 2007), particularly in waters where backscattering is low. Finally,
tions of chlorophyll-a concentration (e.g. Bricaud et al., 1998; O'Reilly we characterize the variability in the hyperspectral remote sensing re-
et al., 1998) that could be associated with methods, instruments, as well flectance, Rrs(λ), and evaluate the relationships between and Rrs(λ)
as actual regional differences (Morel and Maritorena, 2001). Indeed, ratios and chlorophyll-a concentration.
the empirical relationships between reflectance ratios and chlorophyll-a
concentration, tools for estimating phytoplankton biomass from remote 2. Material and methods
sensing, are significantly different in samples observed in Antarctica
compared to the other oceanic regions (e.g., Dierssen and Smith, 2000; 2.1. Sampling
Reynolds et al., 2001), also manifested as errors in the chlorophyll-a
retrievals when global algorithms of ocean color are applied (e.g., Data were collected during two oceanographic cruises in the
Garcia et al., 2005; Szeto et al., 2011). These differences have been northern tip of the Antarctic Peninsula in the austral summers of 2013
attributed to unique “packaged” phytoplankton absorption character- (14 February–3 March) and 2014 (8–24 February). The study area
istics (Dierssen and Smith, 2000; Mitchell and Holm-Hansen, 1991), but covered the Gerlache Strait (GS), which separates the Palmer
the contribution of distinct particulate backscattering coefficients Archipelago from the Peninsula, the Bransfield Strait (BS), between the
cannot be overruled (Brown et al., 2008). In fact, in situ particulate southern Shetland Islands and the Peninsula, and the northwestern
backscattering measurements in Antarctica have shown low magni- Weddell Sea (NWW). A total of 89 and 70 stations were sampled in
tudes (Dierssen and Smith, 2000; Reynolds et al., 2001), but they re- 2013 and 2014, respectively (Fig. 1). A detailed description of sampling
main too scarce to provide conclusive results (see Dierssen, 2010). strategy can be found in Ferreira et al. (in press).
Retrievals of chlorophyll-a concentration in the Antarctic Peninsula by Discrete samples for determination of chlorophyll-a concentration
global ocean color algorithms generally over and underestimate in situ and spectral absorption coefficients were collected at 5, 15, 25, 50, 75
low and high chlorophyll-a, respectively, and this aspect is generally and 100 m at night, and only at 5 m during the day. Profiles of inherent
related to a lower water leaving reflectance and a narrower variation in optical properties were acquired down to 100 m in all day and night
the blue-green band ratio (Zeng et al., 2016) as well as low back- stations. Above water hyperspectral reflectance's were recorded during
scattering coefficients in the green part of the spectrum at high chlor- the day. Radiometric profiles (see 2.5) to obtain the spectral Kd(λ) were
ophyll-a (Dierssen and Smith, 2000). However, the mechanisms which performed less often, because of the unfavorable sea conditions com-
control the ocean color variability in the Southern Ocean need to be monly found in Antarctic waters. Therefore, the number of data for
further understood. each dataset and analysis is variable and informed when pertinent. The
This work describes variations in the bio-optical properties in the data of discrete samples from 5 to 100 m were grouped in layers ac-
northern Antarctic Peninsula observed during late austral summers of cording to vertical variations of light. The first optical depth has im-
2013 and 2014. To the best of our knowledge, our work is the first to plications for remote sensing application, while the others keep con-
describe concurrent measurements of absorption, scattering, back- sistency with our previous work (Ferreira et al., in press). Briefly,
scattering, diffuse attenuation coefficients and remote sensing re- L > 37% corresponds to the layer between surface and the first optical
flectance in the Antarctic Peninsula. The main goal of this study is to depth, at which the downward irradiance 37% of its value just below
report the relative contributions of phytoplankton, cdom and detritus to the surface. The data within L > 37% contributes mostly to the ocean
light absorption and to access potential relationships between scat- color signal that corresponds only to the first portion of the photic zone
tering and backscattering coefficients and chlorophyll-a concentration, returning a weighted average value over this layer (Sathyendranath and

2
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

Platt, 1989). L < 10% refers to the layer below the depth where the contributions regard the spectral non-water absorption, anw(λ).
downward irradiance falls to 10% of its value just below the surface.
The layer between the basis of L > 37% and top of L < 10% is referred to as 2.4. Particulate scattering and backscattering coefficient
L37–10%. The layers L > 37%, L37–10% and L < 10% comprised mostly
samples at 5 m, 15 and 25 m, and 50–100 m, respectively. This in- The inherent optical properties, IOPs, were measured from just
dicates that the light distribution along the vertical varied little among below the water surface down to 100 m of depth by deploying sensors
stations. This separation in optical layers indicated bio-optical char- in a frame that included: (i) one conductivity, temperature and pressure
acteristics very similar in L > 37% and L37–10%, while in L < 10% results sensor (SBE 37 SI, Sea-Bird Electronics Inc.); (ii) one multi-spectral
tended to depart from the main trends (see Ferreira et al., in press). photometer (AC-9, WET Labs Inc.) equipped with 25 cm pathlength
tubes measuring non-water absorption and attenuation, at the following
2.2. Chlorophyll-a concentration wavelengths: 412, 440, 488, 510, 532, 555, 650, 676 and 715 nm and
(iii) one backscattering sensor (HS-6, HOBI Labs Inc.) measuring in the
Total chlorophyll-a concentration [TChl a], representing the sum of wavelengths of 420, 442, 470, 510, 590 and 700 nm. Spikes of the
divinyl chlorophyll-a, chlorophyll-a epimers and allomers, chlor- optical profiles were firstly removed considering values that exceeded
ophyllide-a was measured by high-performance liquid chromatography 50 times (HS-6) and 30 times (AC-9) +1 the mean difference computed
(HPLC), using the procedure described by Mendes et al., (2012, 2007). between adjacent data along the vertical (function diff.m in MATLAB),
Volumes of 1–2 L of seawater were filtered onto 25-mm glass fiber fil- and then applying a third-order one-dimensional median filter (func-
ters (Whatman GF/F) for post-cruise analysis. tion medfilt.m, with the endpoint filtering ‘truncate’ in MATLAB). The
temperature, salinity and optical profiles were smoothed along depth
2.3. Absorption coefficients with a data window of 7 and finally binned to 1 m intervals.
Absorption and attenuation measurements recorded by the AC-9
Volumes of 1–2 L were concentrated on 25 mm GF/F filters and (previously calibrated in the laboratory using pure water) were first
immediately preserved in liquid nitrogen. Absorption of particulate corrected for temperature and salinity effects following the AC-9 m
material was determined following the TR method of Tassan and Ferrari Protocol Document, WET Labs Inc (http://wetlabs.com/ac-s). The re-
(1995) using a spectrophotometer (Lambda 35, Perkin Elmer) with an sulting absorption values were corrected for residual scattering effects
integrating sphere. Optical density (OD, no unit) of samples and blank using the proportional method of Zaneveld et al. (1994) and 715 nm as
filters were first measured against air from 390 to 750 nm at the en- the reference wavelength. The particulate scattering coefficients, bp(λ)
trance of the sphere and then placed directly against the exit of the in m−1, were obtained by subtracting the absorption coefficients from
sphere, backed with a light trap, for a second scan. Blanks and sample the attenuation coefficients.
filters were treated with a few drops of 0.5% NaClO for 10–15 min and The volume scattering function, β(λ), of water was computed using
then carefully washed with 0.2-μm filtered seawater to obtain the det- the concomitant temperature and salinity measurements following
ritus absorption (Tassan and Ferrari, 1995). The OD of extracted filters Zhang et al. (2009). Measured β(λ) values subtracted the β(λ) of water
were measured as described above. Using the correction of pathlength were firstly interpolated linearly to match the AC-9 wavelengths and
amplification described in Tassan et al. (2000), the OD values were extrapolated to particulate backscattering, bbp(λ), following Maffione
converted into absorption coefficients after subtraction of values aver- and Dana (1997) using the factor 1.13 for the angle of measurement of
aged between 745 and 750 nm (null correction) and correction for the 140° of the HB-6. The initial bbp(λ) values were corrected for absorption
filtered volume and the clearance area of the filter. The spectral ab- and scattering effects using the concurrent AC-9 absorption coefficients
sorption coefficients of phytoplankton, aph(λ) in m−1, were computed following the iterative method of Doxaran et al. (2016).
as the difference between absorption coefficients before (total particu-
late) and after (detritus, adet(λ)) the NaClO extraction. Detritus ab- 2.5. Diffuse Attenuation Coefficient
sorption associated to OD lower than 0.0005 (instrumental noise
of ± 0.0005) in the blue wavelengths was removed from the dataset (n Underwater radiometric measurements were obtained at stations
= 89). In these cases, adet(λ) is considered as zero (and so the adet(λ) during favorable sea conditions (n=17) with a hyperspectral profiling
contribution to non-water absorption), so apart(λ) equals to aph(λ). An radiometer (HyperOCR, Satlantic, Inc.), which measures the upwelling
exponential function was fitted by nonlinear regression to each adet(λ) radiance, Lu(λ,z), and the downward irradiance, Ed(λ,z), over the
spectrum between 350–600 nm, following the equation adet(λ) = 350–800 nm spectral range with a resolution of 3.3 nm (137 spectral
adet(λr) exp(-Sdet(λ – λr)), where λr (nm) is a reference wavelength, bands). The initial data processing was performed using the ProSoft ver.
443 nm, adet(λr) (m−1) is the absorption estimate at λr and Sdet(nm−1) 7.7.19 software. Dark offsets and the manufacturer's radiometric cali-
is the spectral slope of the adet(λ) spectrum (Babin et al., 2003). brations were applied to the raw data, and the spectral radiometric data
The filtrates obtained from the filtration described above were re- were binned every 0.1 m. Data were further analyzed in MATLAB. The
filtered through a Polycap AS capsule filter (0.2 µm) and collected in diffuse attenuation coefficient of downwelling irradiance, Kd(λ), at 5 m
amber Qorpak bottles previously cleaned with acid and sterilized. Light was calculated as the slope of a least-squares linear regression of the
absorption of cdom was measured throughout the UV and visible log-transformed Ed(λ,z) within the depth interval of 2–7 m.
spectral domains using a WPI UltraPath liquid waveguide spectro-
photometer along a 50 cm pathlength. The spectral absorbance of the 2.6. Remote Sensing Reflectance
filtered water was measured between 320 and 750 nm immediately or
within 3 h after sampling. Freshly produced, ultraviolet (UV) irradiated At 51 stations, hyperspectral radiometry was acquired above sea-
Milli-Q water was used as a reference. The spectral OD was converted water using a HyperSAS (Satlantic, Inc.) that measures the downward
into absorption coefficients, acdom(λ) in m−1. The slope for each irradiance, Es(λ), the water leaving radiance, Lt(λ), and the sky ra-
acdom(λ) spectrum, Scdom in nm−1, was obtained as for adet(λ), following diance, Li(λ), over the 350–800 nm spectral range with a resolution of
the equation acdom(λ) = acdom(λr) exp(-Scdom(λ – λr)) (Bricaud et al., 3.3 nm (137 spectral bands). The radiance sensors viewed the water
1981). and sky at 40° of zenith and 135° azimuth angles. The initial data
The relative contribution of each optical component to spectral processing was performed using the ProSoft ver. 7.7.19 software to
absorption was determined summing up the spectral absorption coef- obtain L3 level data, that contained averaging of Es(λ), Lt(λ) and Li(λ)
ficients aph(λ), adet(λ) and acdom(λ) and dividing each component by of the L2 level data (ProSoft manual). Data were further analyzed in
that sum, at the same wavelength. Therefore, the relative absorption MATLAB. Radiometric measurements were converted to above-water

3
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

remote-sensing reflectance, Rrs(λ), according to Rrs(λ) = (Lt(λ) - ρsky x algal material in the red spectral portion.
Li(λ)) / Es(λ), where ρsky represents the proportion of sky radiance that The fraction of absorption due to cdom can be equal to that of
is reflected off the surface of the water and is dependent upon wind phytoplankton for blue wavelengths, reaching contributions up to 50%
speed, solar zenith, sensor viewing angle, wavelength and cloud cover below 443 nm in L > 37% (mean of 32%, Fig. 2b). Similar cdom relative
in the sky radiance measurements. A look-up-table of ρsky was created contributions are observed in the green part of the spectrum (minimum
using Hydrolight. phytoplankton absorption), from 10% to 70% at around 560 nm. The
The Rrs(λ) spectra were quality-controlled as following: removal of cdom contribution progressively increases from 443 nm towards lower
data outside 10 degrees from the compass mean as well as spectra with wavelengths, making up 30–80% of non-water absorption near surface
negative values; removal of spectra within 50% of the highest observed and about 80% of absorption at the blue and green spectral ranges in
Rrs(780) to reduce glitter effects on the air–water interface. The final deeper waters. Because cdom absorption was relatively constant in all
Rrs(λ) for each station represents the average of the Rrs(λ) spectra depths (see later) and phytoplankton absorption varied over about
within 1 standard deviation. All Rrs(λ) spectra were interpolated line- three orders of magnitude (0.006–0.135 m−1 at 440 nm; Ferreira et al.,
arly to 1 nm intervals. in press), cdom contributions surpass that of phytoplankton in relatively
lower [TChl a] mainly in L < 10%, as phytoplankton biomass decreases
in depth.
3. Results and discussion Detritus makes a small impact to the non-water spectral absorption
in all layers, with contributions below 20% throughout most of the
3.1. Light absorption of phytoplankton, cdom and detritus spectrum and values nearly at the detection limit in many samples
(mean of 4% in L > 37%, Fig. 2c). Values of adet(443) were very low
Phytoplankton frequently dominated non-water absorption in the compared to observations in coastal waters (Babin et al., 2003), re-
blue part of the spectrum in L > 37%, with contributions from 40% to maining relatively constant and below 0.004 m−1, excepted for few
80% at 443 nm (mean of 65%), where the three components absorb cases exceeding this value (Fig. 3a). Two observed adet(443) values
importantly and is near the primary peak of chlorophyll-a absorption above 0.014 m−1 were associated with low cdom absorption but when
(Fig. 2a). It is noteworthy the appearance of absorption peaks of ac- [TChl a] were above 2 mg m−3, thus, are presumably related to the
cessory pigments, centered at 460 and 490 nm which are bands corre- increase in the phytoplankton biomass. Values of adet(443) showed no
sponded to the absorption by photoprotective pigments and fucox- trend with [TChl a] (Fig. 3a). Note that, however, values of adet(443)
anthin, respectively (see Ferreira et al., in press). The presence of increase with the increase of [TChl a] but no clear trend can be es-
fucoxanthin is also revealed at around 590 and 640 nm. The maximum tablished between these two variables even where [TChl
absorption contribution at 676 nm exhibits the secondary peak of a] > 1 mg m−3. Furthermore, the very low adet(λ) and adet(λ)/anw(λ)
chlorophyll-a absorption in combination to a very low influence of non-

1 1
a b

0.8 0.8
( )
( )

nw

0.6 0.6
nw

( )/a
( )/a

0.4 0.4
cdom
ph
a

0.2 0.2

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

1
c L>37%
L37-10%
0.8 L
<10%
adet ( ) / anw ( )

0.6

0.4

0.2

0
400 450 500 550 600 650 700
Wavelength (nm)
Fig. 2. Relative contributions of spectral absorption coefficients of a) phytoplankton, aph(λ) b) cdom, acdom(λ), and c) detritus, adet(λ), to non-water absorption, anw(λ), which refers to
aph(λ) + acdom(λ) + adet(λ), in L > 37%, L37–10% and L < 10%.

4
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

0.02
L a b
>37%
L
0.04
37-10%
0.015 L <10% 0.035
(443) (m )
-1

Sdet (nm-1)
0.03
0.01 0.025
0.02
det
a

0.005 0.015
0.01

0 0.005
0.05 0.1 1 4 0.001 0.01 0.03
-3
[TChl a] ( mg m ) a (443) (m -1)
det
Fig. 3. (a) Detritus absorption coefficient at 443 nm, adet(443), as a function of total chlorophyll-a concentration, [TChl a], in L > 37%, L37–10% and L < 10%. (b) Slope of the spectral
absorption coefficient of detritus, Sdet, as a function of adet(443).

values observed here suggests minor terrigenous influence in the therein; but see e.g., Bricaud et al., 2010). This uncoupling reflects
northern Antarctic Peninsula. different timescales associated with the creation and destruction of
Detritus corresponds to all non-algal particles, which may include cdom and the dynamics of phytoplankton blooms. Also, no dependence
zooplankton, bacteria, and actual detrital particles and phytoplankton was found between acdom(443) and salinity (not shown), so glacial
detritus (e.g., Iturriaga and Siegel, 1989; Kishino et al., 1985; Morel and meltwater in summer (Dierssen et al., 2002) is not a notable source for
Ahn, 1991; Stramski and Kiefer, 1998), and this pool may be also linked cdom in the Antarctic Peninsula. cdom is produced primarily as a by-
to zooplankton grazing on phytoplankton in the Antarctic Peninsula product of microbial metabolism and phytoplankton degradation
(Mendes et al., 2012). (Nelson et al., 2004; Yentsch and Reichert, 1962). One possible scenario
Values of Sdet varied largely (0.0071–0.0484 nm−1) but most fall in for nearly invariable cdom absorption is that the observed cdom consists
a range of 0.01–0.02 nm−1 (Fig. 3b), as reported in the literature (Babin of a labile fraction, which is rapidly degraded by microbes and any
et al., 2003; Bowers et al., 1996; Ferreira et al., 2014), decreasing input that exceed the constant pool of cdom is not necessarily detectable
continuously with adet(443), but this inverse trend is not normally ob- in the water column (Nelson et al., 2004). One must bear in mind,
served (Babin et al., 2003; Matsuoka et al., 2011) and its cause should however, that cdom represents only portion of dissolved organic
be investigated. The comprehension of causes in the variability of Sdet is matter (Nelson and Siegel, 2002).
limited, but it probably reveals the nature of particles (Babin et al., The spectral shape of cdom absorption hints changes in its compo-
2003; Babin and Stramski, 2004; Estapa et al., 2012), with a trend of sition. Scdom varied one order of magnitude in our dataset, from 0.001 to
higher values for mineral particles (e.g., Bukata et al., 1995). 0.022 but stayed often within 0.011 and 0.015 nm−1 (Fig. 4b), with a
The cdom absorption varied little in our observations, and general inverse trend with acdom(443) as observed elsewhere (e.g.,
acdom(443) values were typically around 0.011 and 0.070 m−1 (Fig. 4a) Matsuoka et al., 2011; Vodacek et al., 1997), mainly in L > 37%. The
in all sampled depths (excepted for three much higher values above Scdom values observed here are generally on the lower end of values of
0.1 m−1). Therefore, there is no relationship between acdom(443) and published data, that are 0.015–0.030 nm−1 for open ocean spectra
[TChl a] in < 100 m, contrasting with some observations for near- (e.g., Nelson and Siegel, 2013; Twardowski et al., 2004) and
coastal environments (e.g., Babin et al., 2003; Meler et al., 2016) and in 0.014–0.020 nm−1 for coastal waters (Babin et al., 2003). Scdom is
the Ross sea and Antarctic Polar Front Zone (Reynolds et al., 2001), but generally higher in the open ocean because of modifications in com-
agreeing with what is mostly reported for oceanic waters (Bricaud et al., position caused by photo-bleaching (Del Vecchio and Blough, 2004;
1981; Nelson and Siegel, 2013; Nelson et al., 1998; and references Goldstone et al., 2004; Swan et al., 2012). One possibility for relatively

0.25 0.025
a b

0.2 0.02
acdom (443) (m )
-1

(nm )
-1

0.15 0.015
cdom

0.1 0.01
S

0.05 0.005

0
0 0.01 0.03 0.1 0.2
0.05 0.1 1 4
-1
[TChl a] ( mg m-3 ) a (443) (m )
cdom

Fig. 4. (a) The cdom absorption coefficient at 443 nm, acdom(443), as a function of total chlorophyll-a concentration, [TChl a], in L > 37%, L37–10% and L < 10%. (b) Slope of the spectral
absorption coefficient of cdom, Scdom, as a function of acdom(443).

5
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

0.01 0.01
This study L >37% b
MM01
R01-Ross Sea L 37-10%
R01-APFZ L
<10%

b (442) (m -1)
b (442) (m -1)
B05
H08

bp
bp

0.001 0.001

a 0.0003
0.0003
0.05 0.1 1 4 0.001 0.01 0.04
-1
-3
[TChl a] ( mg m ) adet(443) (m )

Fig. 5. (a) The particulate backscattering coefficient at 442 nm, bbp(442), as a function of total chlorophyll-a concentration, [TChl a], in L > 37%, L37–10% and L < 10%. (b) bbp(442) as a
function of the absorption coefficient by detritus at 443 nm, adet(443). The solid lines represent the best fits of the linear regression obtained for L > 37% and L37–10%. In (a), the power laws
proposed by Morel and Maritorena (2001), MM01, and Huot et al. (2008), H08, and an invariant bbp(λ) mean value (near 0.0012 m−1) for chlorophyll concentration below 0.14 mg m−3
observed by Behrenfeld et al. (2005), B05, using satellite data are shown for comparison.

lower Scdom reported here is the combination of the location close to the 442 nm, bbp(442), against the values of 0.008 m−1 of pure sea water.
coast (Fig. 1) and low levels of light in Antarctica relatively to other These magnitudes are within the values found by Dierssen and Smith
areas in the globe (Bricaud et al., 2012), yet a general inverse trend of (2000) in the western Antarctic Peninsula and in the clearest waters
Scdom against acdom(443) may be indicative of photo-bleaching near the sampled in Arctic waters, bbp(λ) were below 0.005 m−1 in the visible
surface. Any thoughtful comprehension on the cdom sources and sinks spectral region (Reynolds et al., 2016). Yet, the bbp(λ) values observed
in Antarctica should be further addressed using chemical and experi- here are about the same order of magnitude of very low bbp(λ) mea-
mental analyses. surements made by Boss et al. (2007) in Crater Lake and by Stramska
We concluded that phytoplankton dominate light absorption near and Stramski (2005) in the Greenland Sea, and only higher (about one
surface in the Antarctic Peninsula waters. In L > 37%, 80% of our sam- order of magnitude) than the lowest particulate bbp(λ) (on the order of
ples (60 of 73) showed phytoplankton contribution of above 60% at 10−5 m−1) reported in the literature for the central South Pacific gyre
443 nm (Fig. 2). If we define case 1 waters as those where aph(443) (Twardowski et al., 2007).
represent more than 60% of the total non-water absorption (IOCCG, Different optical configurations of the instruments, calibration
2000), the northern Antarctic Peninsula will be classified as case 1 procedures and the processing of data can impact on final reported
waters. As expected, however, the relative contributions of absorption magnitudes of bbp(λ) among studies. For instance, correction of the
are highly dependent on the considered wavelength as, for instance, measured bbp(λ) signal for absorption and scattering losses was applied
cdom absorption at the green spectral region can be important. The to our data following the method developed by Doxaran et al. (2016)
partitioning of non-water absorption coefficients observed here is in that accounts losses due to absorption and scattering. The ability to
agreement to the proportion shown in IOCCG (2015) for the Southern resolve total backscattering structure in extremely low backscattering
Ocean (their Fig. 4.8). Such absorption budgets and magnitudes of waters with commercially available instruments has been demon-
absorption reported here could support the validation of current ap- strated, but these scattering levels push the limits of uncertainties in
proaches that provide satellite estimates of non-algal material and pure seawater, so improvements on this matter remain highly necessary
phytoplankton absorption for the considered area. This is noteworthy (Twardowski et al., 2007). The effect of other components such as
since very few data of absorption coefficients are available in literature bubbles can account for bbp(λ) (Stramski et al., 2004; Zhang et al.,
for the Southern Ocean (Szeto et al., 2011). 2002), which was predicted to contribute as much as 25% of the total
The large and small importance of cdom and detritus, respectively, particulate backscattering in the Southern Ocean (Dierssen et al., 2017)
to spectral absorption observed here is in agreement with observations and deservers further attention.
mostly for oceanic waters (Nelson and Siegel, 2013). One may consider When bbp(442) is presented as a function of [TChl a], a relationship
a background of cdom absorption in the northern Antarctic Peninsula statistically significant is found when considering L > 37% and L37–10%
for the purpose of regional bio-optical modeling, with a constant value (Fig. 5a), but there is a large amount of scatter around the fit that
of acdom(443) of about 0.030 m−1 and a general spectral slope, Scdom, of covered a restricted [TChl a] range (0.35–3.03 mg m−3). A power law
0.0136 nm−1 (the averages observed here). Despite the very low con- fit provided the dependence bbp(442) = 0.0013 [TChl a]0.864 (n = 109,
tribution of detritus absorption, one can consider an average of R2 = 0.32 for the log-transformed data; p < 0.00001), but a linear
0.0025 m−1 for adet(443) with a mean Sdet of 0.0151 nm−1. The model returned a similar degree of explanation (R2 = 0.30). In L < 10%,
variability in the spectral absorption coefficients of phytoplankton is bbp(442) ranged mostly from 0.0004 to 0.0012 m−1, but with a large
described elsewhere, including their possible impacts on bio-optical dispersion against [TChl a], but still with a significant dependence:
modeling (Ferreira et al., in press). The partition of the particulate bbp(442) = 0.0013 [TChl a]0.169 (n = 104, R2 = 0.12 for the log-
absorption into phytoplankton and detritus is useful for the inter- transformed data; p = 0.0003). Our bbp(442) values in the low [TChl a]
pretation of backscattering sources (see later), since the particulate range are below the prediction of an invariant bbp(λ) mean value (near
backscattering cannot be partitioned into the contribution of both 0.0012 m−1) for chlorophyll concentrations below 0.14 mg m−3
components. (Behrenfeld et al., 2005, their Fig. 5a). In accordance to the low
backscattering levels observed here, our bbp(442) values are con-
siderably below the relationship developed by Morel and Maritorena
3.2. Particulate backscattering coefficient (2001) for case 1 waters around the globe and in the eastern South
Pacific Gyre (Huot et al., 2008), but agrees with Reynolds et al. (2001)
The bbp(λ) values were low, ranging from 0.0004 to 0.0067 at

6
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

for the Ross Sea. This shows that particulate assemblages in the Ross intriguing if we consider that detritus is too low in Antarctica and not
Sea and in our study region tend to backscatter less light than in the covaried with [TChl a] (Fig. 3a), which is the main variable considered
Antarctic Polar Front Zone (Reynolds et al., 2001). The scatter around for parametrizations of bbp(λ) for bio-optical modeling. Modeling re-
bbp(λ) vs. [TChl a] observed in our study is somewhat larger if we sults suggest that particles < 0.1 μm are responsible for the majority of
consider other statistical dependences of backscattering and chlor- backscattering the oceans (see revision of Stramski et al., 2004). In
ophyll-a concentration found elsewhere in the global ocean (e.g., particular, these particles include small-sized non-living particles with
Antoine et al., 2011; Huot et al., 2008; Reynolds et al., 2016, 2001; but perhaps sizable contributions of heterotrophic microbes (Morel and
see Loisel et al., 2001). In general, phytoplankton communities during Ahn, 1991; Stramski and Kiefer, 1991). The bacteria activity in Ant-
the surveys, estimated by pigment composition, were dominated by arctica is generally high (e.g., Delille, 2004; Thomson et al., 2010) and
diatoms and cryptophytes in the pico- and nanophytoplankton size uncoupled from phytoplankton populations (Karl et al., 1996).
fraction (Ferreira et al., in press). This implies that phytoplankton of All considerations above and the reasonable agreement of adet(443)
small sizes were not major contributors to bbp(λ) considering this da- and bbp(λ) (Fig. 5b) draw the conclusion that detritus is the main
taset. One hypothesis for low bbp(λ) and poor covariation with [TChl a] component for bbp(λ) in the Antarctic Peninsula, despite of the low
would be that photoadapted phytoplankton cells are poor backscatters, adet(443) compared to phytoplankton absorption (Fig. 2) and [TChl a]
as phytoplankton adapted to low light are expected to have low carbon (Fig. 3a). This is an interesting result that contributes to the under-
to chlorophyll ratios (Antoine et al., 2011). standing of backscattering sources not only in Antarctica but also in the
The covariation of bbp(442) with adet(443) (Fig. 5b) is considerably global ocean. While detritus is negligible for non-water absorption
stronger than that observed with [TChl a]. A linear model better de- (Fig. 2c), it is the main contributor to bbp(λ) in Antarctic waters. Even
scribed the dependence with the following equation (bbp(442) = 0.320 for intermediate to high [TChl a] found here (0.019–2.91 mg m−3), the
x adet(443) + 0.0007, n = 104, R2 = 0.78; p < 0.00001; L > 37% and very low levels of detritus surpass the effect of phytoplankton biomass
L37–10%), compared to a power law (R2 = 0.46 for log-transformed in bbp(λ). The fact that bbp(λ) in not well correlated to [TChl a] partially
data). Lower adet(443) (by half) occurred in L < 10%, and this layer was contrasts to numerical simulations (e.g., Bernard et al., 2009), labora-
not considered in the statistical adjustment for consistency with the fit tory measurements (e.g., Vaillancourt et al., 2004; Whitmire et al.,
for bbp(442) vs. [TChl a]. About 78% of the variability in bbp(442) is 2010) and field observational studies (e.g., Dall’Olmo et al., 2009,
explained by adet(443), against 32% by [TChl a] (Fig. 5a) and 45% for Antoine et al., 2011) that have suggested phytoplankton may be a
apart(443) (adet(443) + aph(443), not shown). This is somewhat significant contributor to bbp(λ). General better agreements have been

Fig. 6. (a) The particulate scattering coefficient at 650 nm, bp(650), as a function of total chlorophyll-a concentration, [TChl a], in L > 37%, L37–10% and L < 10%. The best fit obtained for all
layers in the form of a power law is also shown, along with the power laws provided by Loisel and Morel (1998), LM98, and Huot et al. (2008), H08, and the linear relationships derived
by Figueroa (2002), F02GR and F02Br. (b) bp(650) as a function of detritus absorption coefficient at 443 nm, adet(443). (c) The particulate backscattering ratio at 650 nm, bbp(650)/
bp(650), as a function of [TChl a], along with the best power law fit considering all the layers and the empirical relationships derived by Morel and Maritorena (2001), MM01,
Twardowski et al. (2001), T01, Whitmire et al. (2007), W07, and Huot et al. (2008), H08. (d) bbp(650)/bp(650) as a function of adet(443) and the best linear fit for L37% and L37–10% all
together.

7
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

Fig. 7. (a) The diffuse attenuation coefficient at 443 nm, Kd(443), as a function of total chlorophyll-a concentration, [TChl a], in L > 37%. The best linear fit and the power law developed
by Morel and Maritorena (2001), MM01, are also shown. (b) The non-water diffuse attenuation coefficient, Kdnw(443) (Kd(443) subtracted the water value of 0.00885 m−1) as a function
of the non-water absorption, anw(443), which refers to aph(443) + acdom(443) + adet(443), following a 1:1 line (dashed line). The inset graph shows the Kdnw(λ) spectra in the
400–560 nm range.

found for bbp(λ) against particulate organic carbon, POC, than for the particles pool contributing to bbp(λ) in the Antarctic Peninsula,
chlorophyll-a (e.g., Stramski et al., 2008; Reynolds et al., 2016). It while phytoplankton cells (dominant in the pool of particles) mostly
seems, therefore, that reasonable parametrizations for bio-optical contribute to bp(λ). Our results contrasts to the good agreement be-
modeling of bbp(λ) can be conducted using detritus rather than [TChl a] tween bp(λ) and bbp(λ) observed in diverse open ocean environments
in the northern Antarctic Peninsula waters. This aspect needs to be reported by Westberry et al. (2010).
investigated in other areas of the Southern Ocean and global ocean.
3.4. Particulate backscattering ratio
3.3. Particulate scattering coefficient
The lack of a clear pattern between the backscattering ratio at
Fig. 6a shows bp(λ) as a function of [TChl a] at 650 nm. This wa- 650 nm, bbp(650)/bp(650), and [TChl a] is not unexpected (Fig. 6c), as
velength is analyzed in most of studies that rely on beam attenuation phytoplankton is not the major component of the bbp(λ) but contribute
measurements, cp(λ), to estimate particulate scattering, because the efficiently to bp(λ). A power law was fitted considering the data from all
non-water absorption is considered negligible at this spectral portion. A three layers combined and provided: bbp(650)/bp(650) = 0.0037 x
good agreement between bp(650) and [TChl a] is observed and a power [TChl a]−0.405 (n = 219, R2 = 0.48, p < 0.0001). No clear pattern is
law provided the best statistical fit in the form of 0.0313 x [TChl observed between bbp(650)/bp(650) and adet(443) considering the three
a]0.7136 (R2=0.79 for the log-transformed data; n=455, p < 0.0001). layers (Fig. 6d), but a significant linear dependence was observed for
Differently to that for bbp(λ), the L < 10% data followed the same trend L37% and L37–10%: bbp(650)/bp(650) = 0.423 x adet(443) + 0.002 (n =
as in L > 37% and L37–10% and were included to the fit (see Loisel and 103, R2 = 0.56, p < 0.0001). Intuitively, if the bbp(λ) values observed
Morel, 1998). Our data is above the relationship provided by Loisel and here are lower than normally in the other oceanic waters, which is not
Morel (1998), but are rather well represented by the equation proposed the case for bp(λ) (Fig. 6a), than bbp(650)/bp(650) values will stay
by Huot et al. (2008). Intriguingly, the relationships found by Figueroa below the observations found elsewhere. It is noteworthy, however,
(2002) in areas of the northern Antarctic Peninsula (Gerlache - F02GR - that bbp(650)/bp(650) in L < 10% approximates to values found in other
and Bransfield - F02Br – Straits) are much above our data and less studies (Fig. 6c). One explanation may be that the detritus in depth is
dependent on chlorophyll-a variations. mostly composed of sunken mineral material and/or very small parti-
Interestingly, no pattern is observed between bp(650) and adet(443) cles, resulting thus in higher backscattering ratios (e.g. Boss et al.,
(Fig. 6b). This is consistent with our results that detritus is the main 2004).
responsible for bbp(λ), while mostly of bp(λ) can be attributed to algal The bbp(λ)/bp(λ) values observed in our data fall within the range
cells. Theoretical simulations for homogeneous spherical particles (i.e., indicative of the “soft,” water-filled organic particles one would expect
Mie theory) predict that bp(λ), cp(λ) and bbp(λ) are influenced by the in the open ocean (Twardowski et al., 2001). In fact, Twardowski et al.
size of the particles (Morel and Ahn, 1991; Stramski and Kiefer, 1991). (2006) concluded that the bbp(λ)/bp(λ) in particle assemblages domi-
As a specific example, Mie theory predicts that 50% of bp(λ) is due to nated by phytoplankton converge on a value around 0.5%, which clo-
particles < 3.2 μm, while 50% of bbp(λ) is from particles < 0.2 μm, but sely agrees with the measurements here. Indeed, our data is consistent
these proportions depend on the refraction index of particles (Stramski to observations of bbp(λ)/bp(λ) to be spectrally flat (not shown) in as-
and Kiefer, 1991). Thus, the relationship between bbp(λ) and [TChl a] sociation to “soft” particles (e.g., Morel and Maritorena, 2001;
may not be as strong as that for bp(λ) and [TChl a] and would depend Whitmire et al., 2007).
on the conservative nature of the particle size distribution (PSD) (i.e.,
that small particles covary with phytoplankton-sized particles). Any 3.5. Diffuse attenuation coefficient
covariation of bbp(λ) and chlorophyll-a concentration as generally ob-
served (e.g., Antoine et al., 2011; Dall’Olmo et al., 2009; Huot et al., Values of Kd(λ) at 443 nm, Kd(443) were linearly correlated with
2008) would be due to covariation between chlorophyll-a and particles [TChl a] (Kd(443) = 0.0449 x [TChl a] + 0.0384; n = 17, R2=0.90;
that contribute directly to bbp(λ). The uncoupling between adet(443) p < 0.00001, Fig. 7a). Interestingly, the offset obtained from our fit
and [TChl a] (Fig. 3a), the poor and good agreements between bbp(λ) (0.0384 m−1) minus the absorption coefficient of pure water described
and [TChl a] (Fig. 5a) and between bbp(λ) and adet(443) (Fig. 5b), re- in Pope and Fry (1997) of 0.00885 m−1 is 0.030 m−1, identical to the
spectively, and the good relationship between bp(λ) and [TChl a] mean acdom(443) found in this study. This implies that for extreme low
(Fig. 6a) reinforces the conclusion that detritus is the main portion of [TChl a], Kd(λ) is driven by light absorption of water and cdom (i.e.,

8
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

detritus contribution to absorption is too small). The best fit obtained 0.01


with linear dependence, rather than a power law as that suggested by
Morel and Maritorena (2001), is probably due the small number of
0.008
Kd(λ) measurements (n = 17), unable to detect the nonlinearity of
Kd(λ) vs. [TChl a].
The Kd(λ) values for a given [TChl a] are appreciably below those

Rrs (sr )
0.006

-1
derived by the relationship of Morel and Maritorena (2001). In addi-
tion, no dependence of Kd(λ) on bbp(λ) was observed (not shown) and
Kd(443) minus the assumed diffuse attenuation coefficient due to water, 0.004
Kdbio(443), and non-water absorption coefficient, anw(443), are highly
correlated almost at a 1:1 correspondence (n = 15, Fig. 7b). This is 0.002
intuitive considering the theoretical relationship Kd(λ) = [a(λ) +
bb(λ)] / μd(λ) (where a(λ), bb(λ) and μd(λ) corresponds to total ab-
sorption, scattering and the average cosine for downward irradiance, 0
350 400 450 500 550 600 650 700
respectively, Kirk, 1994) and the very low magnitude of bbp(λ) relative
Wavelength (nm)
to that for anw(λ) observed here. Similarly, Organelli et al. (2017) found
Kd(λ) values generally lower for a given chlorophyll-a concentration in Fig. 8. Spectra of the remote sensing reflectance, Rrs(λ), within the 350 and 700 nm
the Southern Ocean (not including the Antarctic Peninsula). The au- range. The spectra associated to the lowest [TChl a] (0.120 and 0.137 mg m−3) are shown
thors attributed these observations to the presence of a higher pigment as dashed lines.

packaging and low specific absorption coefficients of phytoplankton


communities in the Southern Ocean than observed in other areas, in correlated. In fact, anw(λ) and bbp(λ) in individual spectral bands were
association to large phytoplankton cells. On contrary, the phyto- not well correlated in our study (not shown), consistent with our ob-
plankton communities during our surveys were mainly composed by servation that acdom(λ) and adet(λ) do not covary with [TChl a] as
small cells and specific absorption coefficients of phytoplankton similar bbp(λ) does (although not in a well-defined function).
to values found worldwide (Ferreira et al., in press), supporting low Despite the rigorous processing of the data and the resulting con-
bbp(λ) as the main justification of low Kd(λ) values in our dataset. sistent shapes of Rrs(λ), some small features are observed, mainly below
Two important points can be made from our results: 1) it corrobo- 400 and above 600 nm. These features were probably a result of loca-
rates the low levels of bbp(λ) observed in this study with independent lized noisy data combined to the smoothing procedures used in each
measurements, and 2) it reveals that measurements of Kd(λ) in the spectrum, and will not be discussed in here. The applied corrections for
sampled waters correspond essentially to total absorption. In phyto- radiometric data acquired above water (Section 2.6) may also promote
plankton-dominated waters, it essentially reveals the phytoplankton changes in magnitudes. While algorithms relying on spectral ratios of
absorption spectrum (inset in Fig. 7b). This is interesting because Kd(λ) reflectance or water-leaving radiance are probably less sensitive to
is obtained promptly in the field, compared to the very laborious these errors, closure exercises may be challenging in those cases, but
measurements of absorption at laboratory, often requiring different may be essential to verify the consistency of measured IOPs with those
instruments for particles and cdom. It also advantages absorption estimated by the inversion of the measured light field properties.
measurements obtained with sensors such as AC-9 and AC-s, as Kd(λ) For consistency with previous studies, the best agreement between
uncertainties are less sensitive to calibration issues. [Chl a] (instead of [TChl a], and Rrs(λ) ratios was observed for
Rrs(490)/Rrs(555) (Fig. 9a), as similarly observed in previous studies
3.6. Remote sensing reflectance aimed to develop regional bio-optical models for the Southern Ocean
(Dierssen and Smith, 2000; Garcia et al., 2005). Our data fall very close
Radiometric measurements in each station were mostly performed to those predicted by algorithms proposed for the Southern Ocean,
under complete or partially overcast skies, typical conditions in considering both the Rrs(490)/Rrs(555) (Fig. 9a) and Rrs(bluemax)/
Antarctica. Note that the consequent spectral variations are minimized Rrs(green) (Fig. 9b) ratios (see Fig. 9's caption). Therefore, [Chl a] va-
by normalization of water leaving radiance by above-water irradiance lues for a given Rrs(λ) ratio are above the global algorithms. This de-
to obtain Rrs(λ). A small part of the variability seen in Rrs(λ) spectral viation, commonly observed for the Southern Ocean, was presumed to
data can, therefore, be explained by the various illumination conditions be due to higher chlorophyll content per cell of phytoplankton in re-
encountered. Overall, the Rrs(λ) data set is characterized by important sponse to low light (Dierssen and Smith, 2000; Mitchell and Holm-
changes in magnitude (Fig. 8), for surface [TChl a] varying from 0.12 to Hansen, 1991; Szeto et al., 2011) and also due to the low backscattering
3.03 mg m−3. Two Rrs(λ) spectra considerably depart from the main (Dierssen and Smith, 2000). Our previous work, however, showed that
shape with the highest values in the blue, in association to the lowest in this dataset the absorption coefficients of phytoplankton and con-
[TChl a] (0.120 and 0.137 mg m−3). A common spectral feature is centration of accessory pigments do not differ from the global trends
observed: Rrs(λ)a primary peak near 400 nm and secondary peak (Ferreira et al., in press). Our results of bbp(λ) distinctly low and the
around 500 nm, consistent with the maximum phytoplankton absorp- deviations of reflectance ratios and [Chl a] from global trends (Fig. 9)
tion in the blue range. Indeed, the spectra with the lowest magnitudes may partially agree with the sensitivity analysis that investigated the
(higher [TChl a]) exhibit pronounced shoal at around 440 nm, mir- impact of variations in the inherent optical properties on the maximum
roring the phytoplankton absorption spectrum. Although noisy, the red band ratio of blue-to-green reflectance for the Southern Ocean (IOCCG,
end of the visible spectra reveals the chlorophyll-a fluorescence, cen- 2015). Their results showed that cdom absorption and bbp(λ) could
tered at about 683 nm in most spectra. Not surprisingly, no features as largely impact (> 70%) changes in the reflectance ratios, while phy-
maximums near 570 nm are observed, which would suggest an im- toplankton has a significant but mostly constant influence (50%). Since
portant role of bbp(λ) (e.g. Lubac and Loisel, 2007). The effect of cdom the phytoplankton absorption for our samples did not differ from global
absorption is not evidenced in the blue. As expected from the lack of trends (Ferreira et al., in press) and cdom absorption was relatively
relationships between bbp(λ) and Kd(λ), no dependence was found be- much less variable (Fig. 4), low bbp(λ) is likely the primary cause ex-
tween bbp(λ) and Rrs(λ) in any wavelength. This contrasts to the well plaining lower reflectance ratios for a given [Chl a] in our dataset
covariation of both variables observed by Stramski et al. (1999) for the compared to the global algorithms (Fig. 9).
Southern Ocean. The authors stated that the covariation of bbp(λ) and Another issue that deserves attention is the in situ [Chl a] data. Some
Rrs(λ) implies that backscattering and absorption coefficients are well

9
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

100 100
a b

10 10
[Chl a] ( mg m )

[Chl a] ( mg m )
-3

-3
1 1

0.1 0.1
This study Szeto2011
FURG OC2L SPGANT-GLIv3
DS 2000 Johson2013
0.01 0.01
0.1 1 10 0.3 1 10
R (490)/R (555) R (blue )/R (green)
rs rs rs max rs

Fig. 9. (a) The chlorophyll-a concentration, [Chl a] as a function of the Rrs(490)/Rrs(555) ratio, and the best linear fit in the form (log10([Chl a]) = −2.121 x log10(Rrs(490)/Rrs(555)) +
0.5709, R2=0.62, n=51, p < 0.00001). The grey points in (a) and (b) represent the NASA's Bio-optical Marine Algorithm Data Set (NOMAD; Werdell and Bailey, 2005). The linear fits of
Garcia et al. (2005) (FURG OC2L) and Dierssen and Smith (2000) (DS 2000) are also shown. (b) [Chl a] as a function of the ratio between the maximum Rrs(λ) among the blue
wavelengths, Rrs(bluemax), and the green wavelength, Rrs(green). The fourth-polynomials of Szeto et al. (2011) (Szeto, 2011) and Mitchell and Kahru (2009) (SPGANT-GLIv3) and the
third-polynomial for MODIS-Aqua (their Eq. 5) of Johnson et al. (2013) (Johson2013) are also shown. The maximum band ratio in our study ranged between 1 and 3 (excluding the
outlying data points of 3.19 and 3.44 collected in NWWHT), thus the Mitchell and Kahru (2009)’s algorithm blending was not applied (see Kahru and Mitchell, 2010).

studies have shown that fluorescence-derived chlorophyll-a can be phytoplankton cells (dominant in the particles pool) mostly covaried
overestimated up to 2-fold in the Southern Ocean (e.g., Marrari et al., with particulate scattering. This is an interesting result that contributes
2006) when compared to HPLC. Thus, it is important to note that all to the comprehension of the backscattering sources not only in
studies considered for comparison here (Fig. 9), excluding Johnson Antarctica, but also other regions of the global ocean.
et al. (2013), used fluorescence-derived chlorophyll-a. The slightly de- The very low values of backscattering as previously found in
part of our data and Garcia et al. (2005)’s equation could be associated Antarctica waters was confirmed and corroborated using independent
with differences in the phytoplankton communities. Indeed, the two measurements of the diffuse attenuation coefficient, which corre-
lowest [Chl a] values (0.11 and 0.13 mg m−3) (Fig. 1, NWWHT) de- sponded almost 1 to 1 to the absorption coefficient. The shape of the
parting from the main trend (Figs. 9a and 9b), contained a very distinct spectral remote sensing reflectance varied mainly to variability in [TChl
phytoplankton community (composed by Phaeocystis antarctica) and a], while [TChl a] vs. maximum band ratios dependence deviated from
spectral absorption (Ferreira et al., in press). These results reveal that, global trends in a very similar fashion as in other studies of the
despite the proximity of those two stations and the Antarctic Peninsula Southern Ocean, likely due to very low bbp(λ). The present study con-
(Fig. 1), the bio-optical properties of the northern Weddell Sea differ stitutes the most comprehensive description of concurrent apparent and
considerably and the usage of distinct algorithms among regions in the inherent optical properties measured in the Antarctic Peninsula and
proximities of Antarctic Peninsula is likely necessary. contributes for further bio-optical studies and ocean color applications.
Our Rrs(λ) dataset is not large enough to have sufficient re-
presentation of the northern Antarctic Peninsula (n=51), but it is no- Acknowledgments
teworthy the agreement of our data to previous studies in the Southern
Ocean and, specifically, in the Antarctic Peninsula. Also, from our This work is part of a multidisciplinary program under the GOAL
knowledge, this is the first report of Rrs(λ) in Antarctic waters using (Brazilian High Latitudes Oceanography Group) activities in the
Satlantic SAS sensors. In fact, Zeng et al. (2016) measured Rrs(λ) above Brazilian Antarctic Program (PROANTAR). Financial support was pro-
water around the Antarctica Peninsula using an ASD Fieldspec Hand- vided by CNPq (National Council for Research and Development) for
Held 2, showing less satisfactory agreements among Rrs(λ) ratios and the projects POLARCANION (CNPq Grant 556848/2009-8) and PRO-
[TChl a] than those observed here. OASIS (CNPq Grant 565040/2010-3). Amabile Ferreira was supported
Despite being out the scope of our study, a complementary ex- by FAPESP (Research Support Foundation of the State of São Paulo,
planation for the lack of a well-defined function between bbp(λ) and Grants 2013/16957-2 and 2016/01601-6). Áurea M. Ciotti (PQ
Rrs(λ) is the influence of inelastic radiative processes, such as Raman 310985/2013-7) and Carlos A. E. Garcia (PQ 311943/2015-2) thank
scattering by water molecules and fluorescence by colored dissolved CNPq for the research fellowship. The analysis of samples for algal
organic matter and chlorophyll-a, which is not generally considered in pigments was performed at FURG. The authors thank the crew of the
the equation Rrs(λ) ~ bbp(λ)/a(λ), but shown to have an influence on Polar Vessel “Almirante Maximiano” of the Brazilian Navy and several
ocean color inverse models (e.g., Westberry et al., 2013; McKinna et al., investigations on the cruises for assistance. We acknowledge the help
2016). Further efforts to elucidate their role in the optical properties are provided by Bianca Tocci and Maria Fernanda C. Giannini during the
recommended (e.g., Lee and Huot, 2014; Li et al., 2016). field work. The discussions with David Doxaran and Vincenzo Vellucci
on the processing of AC-9 and backscattering data is much appreciated.
4. Conclusions Special thanks to Inia Soto that provided the MATLAB code for the SAS
data processing, developed at FURG. This research is a contribution of
Phytoplankton contribute to more than 60% of the total non-water the NP-BioMar, University of São Paulo. We acknowledge the two
absorption in the blue wavelengths in the northern Antarctic Peninsula, anonymous reviewers for their valuable contributions that have greatly
which may be classified as case 1 waters. The cdom absorption can be improved the manuscript.
equal to that of phytoplankton in the blue wavelengths, detritus has
very low spectral absorption and none covaries with chlorophyll-a References
concentration. Even with very low absorption, detritus was the com-
ponent that better correlated with particulate backscattering, while Antoine, D., Siegel, D.A., Kostadinov, T., Maritorena, S., Nelson, N.B., Gentili, B., Vellucci,

10
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

V., Guillocheau, N., 2011. Variability in optical particle backscattering in contrasting doi.org/10.1002/2017JC012964.
bio-optical oceanic regimes. Limnol. Oceanogr. 56 (3), 955–973. Ferreira, A., Ciotti, Á.M., Giannini, M.F.C., 2014. Variability in the light absorption
Babin, M., Stramski, D., 2004. Variations in the mass-specific absorption coefficient of coefficients pf phytoplankton, non-algal particles, and colored dissolved organic
mineral particles suspended in water. Limnol. Oceanogr. 49, 756–767. matter in a subtropical bay (Brazil). Estuar. Coast Shelf Sci. 139, 127–136.
Babin, M., Stramski, D., Ferrari, G.M., Claustre, H., Bricaud, A., Obolensky, G., Hoepffner, Figueroa, F.L., 2002. Bio-optical characteristics of Gerlache and Bransfield strait waters
N., 2003. Variations in the light absorption coefficients of phytoplankton, non-algal during an Antarctic cruise. Deep Sea Res. Part II 49, 675–691.
particles, and dissolved organic matter in coastal waters around Europe. J. Geophys. Garcia, C.A.E., Garcia, V.M.T., McClain, C., 2005. Evaluation of SeaWiFS chlorophyll
Res. 108 (3211), 3210 (1029/2001JC000882). algorithms in the Southwestern Atlantic and Southern Oceans. Remote Sens. Environ.
Bartz, R., Zaneveld, Z.R.V., Park, H., 1978. A transmissometer for profiling and moored 95, 125–137.
observations in water. Proc. SPIE 160, 102–108 in Ocean Optics V, M. B. White, ed. Goldstone, J.V., Del Vecchio, R., Blough, N.V., Voelker, B.M., 2004. A multicomponent
Behrenfeld, M.J., Boss, E., 2006. Beam attenuation and chlorophyll concentration as al- model of chromophoric dissolved organic matter photobleaching. J. Photochem.
ternative optical indices of phytoplankton biomass. J. Mar. Res. 64, 431–451. Photobiol. 80, 52–60.
Behrenfeld, M.J., Boss, E., Siegel, D.A., Shea, D.M., 2005. Carbon-based ocean pro- Gordon, H.R., Morel, A., 1983. In: Barber, R.T., Mooers, C.N.K., Bowman, M.J.,
ductivity and phytoplankton physiology from space. Glob. Biogeochem. Cycles 19, Zeitzschel, B. (Eds.), Remote Assessment of Ocean Color for Interpretation of Satellite
GB1006. http://dx.doi.org/10.1029/2004GB002299. Visible Imagery: A Review. Springer- Verlag, New York.
Bernard, S., Probyn, T.A., Quirantes, A., 2009. Simulating the optical properties of phy- Graff, J.R., Westberry, T.K., Milligan, A.J., Brown, M.B., Dall’Olmo, G., Reifel, K.M.,
toplankton cells using a two-layered spherical geometry. Biogeosci. Discuss. 6, Behrenfeld, M.J., 2016. Photoacclimation of natural phytoplankton communities.
1497–1563. Mar. Ecol. Prog. Ser. 542, 51–62.
Boss, E., Collier, R., Larson, G., Fennel, K., Pegau, W.S., 2007. Measurements of spectral Graff, J.R., Westberry, T.K., Milligan, A.J., Brown, M.B., et al., 2015. Analytical phyto-
optical properties and their relation to biogeochemical variables and processes in plankton carbon measurements spanning diverse ecosystems. Deep Sea Res. Part I
Crater Lake National Park, OR. Hydrobiologia 574, 149–159. 102, 16–25.
Boss, E., Pegau, W.S., Lee, M., Twardowski, M.S., Shybanov, E., Korotaev, G., Baratange, Huot, Y., Morel, A., Twardowski, M.S., Stramski, D., Reynolds, R.A., 2008. Particle optical
F., 2004. The particulate backscattering ratio at LEO-15 and its use to study particle backscattering along a chlorophyll gradient in the upper layer of the eastern South
composition and distribution. J. Geophys. Res. 109, C0101410 (1029/ Pacific Ocean. Biogeosciences 5, 495–507.
2002JC001514). IOCCG - Internacional Ocean-Color Coordinating Group, 2015. Ocean colour remote
Bowers, D.G., Harker, G.E.L., Stephan, B., 1996. Absorption spectra of inorganic particles sensing in polar seas. In: Babin, M., Arrigo, K., Belanger, S., Forget, M.-H. (Eds.),
in the Irish Sea and their relevance to remote sensing of chlorophyll. Int. J. Remote IOCCG Report Series, No. 16. IOCCG, Dartmouth, Canada.
Sens. 17 (12), 2449–2460. IOCCG - Internacional Ocean-Color Coordinating Group, 2006. Remote sensing of ocean
Bricaud, A., Babin, M., Claustre, H., Ras, J., Tièche, F., 2010. Light absorption properties colour in Coastal, and other optically-complex, waters. In: Lee, Z.-P. (Ed.), Reports of
and absorption budget of Southeast Pacific waters. J. Geophys. Res. 115, C08009. the International Ocean-Colour Coordinating Group, No. 5. IOCCG, Dartmouth,
http://dx.doi.org/10.1029/2009JC005517. Canada.
Bricaud, A., Claustre, H., Ras, J., Oubelkheir, K., 2004. Natural variability of phyto- IOCCG - Internacional Ocean-Color Coordinating Group, 2000. Remote sensing of in-
plankton absorption in oceanic waters: influence of the size structure of algal po- herent optical properties: fundamentals, tests of algorithms, and applications. In:
pulations. J. Geophys. Res. 109, C11010. http://dx.doi.org/10.1029/2004JC002419. Sathyendranath, S. (Ed.), Reports of the International Ocean-Colour Coordinating
Bricaud, A., Ciotti, A.M., Gentili, B., 2012. Spatial-temporal variations in phytoplankton Group, No. 3. IOCCG, Dartmouth, Canada.
size and colored detrital matter absorption at global and regional scales, as derived Iturriaga, R., Siegel, D.A., 1989. Microphotometric characterization of phytoplankton and
from twelve years of SeaWiFS data (1998–2009). Glob. Biogeochem. Cycles 26, detrital absorption properties in the Sargasso Sea. Limnol. Oceanogr. 34, 1706–1726.
GB1010. Johnson, K.S., Berelson, W.M., Boss, E., Chase, Z., Claustre, H., Emerson, S.R., Gruber, N.,
Bricaud, A., Morel, A., Babin, M., Allali, K., Claustre, H., 1998. Variations of light ab- Körtzinger, A., Perry, M.J., Riser, S.C., 2009. Observing biogeochemical cycles at
sorption by suspended particles with chlorophyll a concentration in oceanic (case 1) global scales with profiling floats and gliders: prospects for a global array.
waters: analysis and implications for bio-optical models. J. Geophys. Res. 103, Oceanography 22, 216–225.
31,033–31,044. Kahru, M., Mitchell, B.G., 2010. Blending of ocean colour algorithms applied to the
Bricaud, A., Babin, M., Morel, A., Claustre, H., 1995. Variability in the chlorophyll spe- Southern Ocean. Remote Sens. Lett. 1, 119–124.
cific absorption coefficients of natural phytoplankton: analysis and parameterization. Karl, D.M., Christian, J.R., Dore, J.E., Letelier, R.M., 1996. Microbiological oceanography
J. Geophys. Res. 100, 13,321–13,332. in the region west of the Antarctic Peninsula: microbial dynamics, nitrogen cycle and
Bricaud, A., Morel, A., Prieur, L., 1981. Absorption by dissolved organic matter of the sea carbon flux. In: Ross, R.M., Hofmann, E.E., Quetin, L.B. (Eds.), Foundations for
(yellow substance) in the UV and visible domains. Limnol. Oceanogr. 26, 43–53. Ecological Research West of the Antarctic Peninsula. American Geophysical Union,
Brown, C.A., Huot, Y., Werdell, P.J., Gentili, B., Claustre, H., 2008. The origin and global Washington, D. C.
distribution of second order variability in satellite ocean color and its potential ap- Kirk, J.T.O., 1994. Light and Photosynthesis in Aquatic Ecosystems (509 pp.). Cambridge
plications to algorithm development. Rem. Sens. Environ. 112, 4186–4203. University Press, New York.
Bukata, R.P., Jerome, J.H., Kondratyev, K.Y., Pozdnyakov, D.V., 1995. Optical Properties Kishino, M., Takahashi, M., Okami, N., Ichimura, S., 1985. Estimation of the spectral
and Remote Sensing of Inland and Coastal Waters. CRC Press, London. absorption coefficients of phytoplankton in the sea. Bull. Mar. Sci. 37 (2), 634–642.
Carder, K.L., Hawes, S.K., Baker, K.A., Smith, R.C., Steward, R.G., Mitchell, B.G., 1991. Kostadinov, T.S., Siegel, D.A., Maritorena, S., 2009. Retrieval of the particle size dis-
Reflectance model for quantifying chlorophyll a in the presence of productivity de- tribution from satellite ocean color observations. J. Geophys. Res. 114, CO9015.
gradation products. J. Geophys. Res. 96 (C11), 20,599–20,611. Lee, Z., Huot, Y., 2014. On the non-closure of particle backscattering coefficient in oli-
Cetinić, I., Perry, M.J., D'Asaro, E., Briggs, N., Poulton, N., Sieracki, M.E., Lee, C.M., 2015. gotrophic oceans. Opt. Express 22, 29223–29233.
A simple optical index shows spatial and temporal heterogeneity in phytoplankton Li, L., Stramski, D., Reynolds, R.A., 2016. Effects of inelastic radiative processes on the
community composition during the 2008 North Atlantic Bloom Experiment. determination of water-leaving spectral radiance from extrapolation of underwater
Biogeosciences 12, 2179–2194. near-surface measurements. Appl. Opt. 55, 7050–7067.
Ciotti, A.M., Cullen, J.J., Lewis, M.R., 1999. A semi-analytical model of the influence of Loisel, H., Nicolas, J.M., Sciandra, A., Stramski, D., Poteau, A., 2006. Spectral dependency
phytoplankton community structure on the relationship between light attenuation of optical backscattering by marine particles from satellite remote sensing of the
and ocean color. J. Geophys. Res. 104 (C1), 1559–1578. global ocean. J. Geophys. Res. 111, C09024.
Dall’Olmo, G., Westberry, T.K., Behrenfeld, M.J., Boss, E., Slade, W.H., 2009. Significant Loisel, H., Bosc, E., Stramski, D., Oubelkheir, K., Deschamps, P.Y., 2001. Seasonal
contribution of large particles to optical backscattering in the open ocean. variability of the backscattering coefficient in the Mediterranean Sea based on sa-
Biogeosciences 6, 947–967. tellite SeaWiFS imagery. Geophys. Res. Lett. 28, 4203–4206.
Del Vecchio, R., Blough, N.V., 2004. On the origin of the optical properties of humic Loisel, H., Stramski, D., 2000. Estimation of the inherent optical properties of natural
substances. Environ. Sci. Technol. 38, 3885–3891. waters from the irradiance attenuation coefficient and reflectance in the presence of
Delille, D., 2004. Abundance and function of bacteria in the Southern Ocean. Cell. Mol. Raman scattering. Appl. Opt. 39, 3001–3011.
Biol. 50, 543–551. Loisel, H., Morel, A., 1998. Light scattering and chlorophyll concentration in Case 1
Dierssen, H., Randoph, K., Garaba, S., 2017. Interpreting backscattering in the Southern waters: a reexamination. Limnol. Oceanogr. 43, 847–858.
Ocean, International Ocean Colour Science Meeting 2017, Lisbon, Portugal. Lubac, B., Loisel, H., 2007. Variability and classification of remote sensing reflectance
Dierssen, H.M., 2010. Perspectives on empirical approaches for ocean color remote sen- spectra in the eastern English Channel and southern North Sea. Remote Sens.
sing of chlorophyll in a changing climate. Proc. Nat. Acad. Sci. 107, 17073–17078. Environ. 110, 45–58.
Dierssen, H., Smith, R., Vernet, M., 2002. Glacial meltwater dynamics in coastal waters McKinna, L.I.W., Werdell, P.J., Proctor, C.W., 2016. Implementation of an analytical
west of the Antarctic Peninsula. Proc. Natl. Acad. Sci. USA 99, 1790–1795. Raman scattering correction for satellite ocean-color processing. Opt. Express 24
Dierssen, H.M., Smith, R.C., 2000. Bio-optical properties and remote sensing ocean color (14), A1123–A1137.
algorithms for Antarctic Peninsula waters. J. Geophys. Res. 105, 26301–26312. Maffione, R.A., Dana, D.R., 1997. Instruments and methods for measuring the backward-
Doxaran, D., Leymarie, E., Nechad, B., Dogliotti, A., Ruddick, K., Gernez, P., Knaeps, E., scattering coefficient of ocean waters. Appl. Opt. 36, 6057–6067.
2016. Improved correction methods for field measurements of particulate light Marrari, M., Hu, C., Daly, K., 2006. Validation of SeaWiFS chlorophyll a concentrations in
backscattering in turbid waters. Opt. Express 24 (4), 3615–3637. the Southern Ocean: a revisit. Remote Sens. Environ. 105 (4), 367–375.
Estapa, M.L., Boss, E., Mayer, L.M., Roesler, C.S., 2012. Role of iron and organic carbon in Matsuoka, A., Hill, V., Huot, Y., Babin, M., Bricaud, A., 2011. Seasonal variability in the
mass-specific light absorption by particulate matter from Louisiana coastal waters. light absorption properties of western Arctic waters: parameterization of the in-
Limnol. Oceanogr. 57, 97–112. dividual components of absorption for ocean color applications. J. Geophys. Res. 116,
Ferreira, A., Ciotti, Á.M., Mendes, C.R.B., Uitz, J., Bricaud, A., 2017. Phytoplankton light C02007.
absorption and the package effect in relation to photosynthetic and photoprotective Meler, J., Kowalczuk, P., Ostrowska, M., Ficek, D., Zabłocka, M., Zdun, A., 2016.
pigments in the northern tip of Antarctic Peninsula. J. Geophys. Res. 122. http://dx. Parameterization of the light absorption properties of chromophoric dissolved

11
A. Ferreira et al. Deep-Sea Research Part II xxx (xxxx) xxx–xxx

organic matter in the Baltic Sea and Pomeranian lakes. Ocean Sci. 12, 1013–1032. Stramski, D., Boss, E., Bogucki, D., Voss, K.J., 2004. The role of seawater constituents in
Mendes, C.R.B., de Souza, M.S., Garcia, V.M.T., Leal, M.C., Brotas, V., Garcia, C.A.E., light backscattering in the ocean. Prog. Oceanogr. 61, 27–56.
2012. Dynamics of phytoplankton communities during late summer around the tip of Stramski, D., Reynolds, R.A., Kahru, M., Mitchell, B.G., 1999. Estimation of particulate
the Antarctic Peninsula. Deep-Sea Res., I 65, 1–14. organic carbon in the ocean from satellite remote sensing. Science 285, 239–242.
Mendes, C.R., Cartaxana, P., Brotas, V., 2007. HPLC determination of phytoplankton and Stramski, D., Kiefer, D.A., 1998. Can heterotrophic bacteria be important to marine light
microphytobenthos pigments: comparing resolution and sensitivity of a C18 and a C8 absorption? J. Plankton Res. 20, 1489–1500.
method. Limnol. Oceanogr. Methods 5, 363–370. Stramski, D., Kiefer, D.A., 1991. Light scattering by microorganisms in the open ocean.
Mitchell, B., Holm-Hansen, O., 1991. Bio-optical properties of Antarctic Peninsula waters: Prog. Oceanogr 28, 343–383.
differentiation from temperate ocean models. Deep Sea Res., Part A 38 (8–9), Swan, C.M., Nelson, N.B., Siegel, D.A., Kostadinov, T.S., 2012. The effect of surface ir-
1009–1028. radiance on the absorption spectrum of chromophoric dissolved organic matter in the
Mitchell, B.G., Kahru, M., 2009. Bio-optical algorithms for ADEOS-2 GLI. J. Rem. Sens. global ocean. Deep-Sea Res. I 63, 52–64.
Soc. Jpn. 29, 80–85. Szeto, M., Werdell, P.J., Moore, T.S., Campbell, J.W., 2011. Are the worlds oceans opti-
Morel, A., Claustre, H., Antoine, D., Gentili, B., 2007. Natural variability of bio‐optical cally different? J. Geophys. Res. 116, C00H04.
properties in case 1 waters: attenuation and reflectance within the visible and Tassan, S., Ferrari, G.M., Bricaud, A., Babin, M., 2000. Variability of the amplification
near‐UV spectral domains, as observed in South Pacific and Mediterranean waters. factor of light absorption by filter-retained aquatic particles in the coastal environ-
Biogeosciences 4, 913–925. ment. J. Plankton Res. 22, 659–668.
Morel, A., Maritorena, S., 2001. Bio-optical properties of oceanic waters: a reappraisal. J. Tassan, S., Ferrari, G.M., 1995. An alternative approach to absorption measurements of
Geophys. Res. 106, 7163–7180. aquatic particles retained on filters. Limnol. Oceanogr. 40, 1358–1368.
Morel, A., Ahn, Y.-H., 1991. Optics of heterotrophic nanoflagellates and ciliates: a ten- Thomson, P.G., Davidson, A.T., vandenEnden, R., Pearce, I., Seuront, L., Paterson, J.S.,
tative assessment of their scattering role in oceanic waters compared to those of Williams, G.D., 2010. Distribution and abundance of marine microbes in the Southern
bacterial and algal cells. J. Mar. Res. 49, 177–202. Ocean between 30 and 80°E. Deep Sea Res. II 57, 815–827.
Morel, A., 1988. Optical modeling of the upper ocean in relation to its biogenous matter Twardowski, M.S., Claustre, H., Freeman, S.A., Stramski, D., Huot, Y., 2007. Optical
content (case 1 water). J. Geophys. Res. 93, 10,749–10,768. backscattering properties of the ‘clearest’ natural waters. Biogeosciences 4,
Morel, A., Prieur, L., 1977. Analysis of variations in ocean color. Limnol. Oceanogr. 22, 1041–1058.
709–722. Twardowski, M.S., Sullivan, J.S., Dierssen, H., Zaneveld, J.R.V., 2006. Sources of back-
Nelson, N.B., Siegel, D.A., 2013. The global distribution and dynamics of chromophoric scattering in marine waters, Ocean Sciences Meeting, Honolulu, HI, February 20-24,
dissolved organic matter. Annu. Rev. Mar. Sci. 5, 447–476. Abstract OS32M-02.
Nelson, N.B., Carlson, C.A., Steinberg, D.K., 2004. Production of chromophoric dissolved Twardowski, M.S., Boss, E., Sullivan, J.M., Donaghay, P.L., 2004. Modeling the spectral
organic matter by Sargasso Sea microbes. Mar. Chem. 89, 273–287. shape of absorption by chromophoric dissolved organic matter. Mar. Chem. 89,
Nelson. N.B., Siegel, D.A. 2002. Chromophoric DOM in the open ocean. Hansell & Carlson 69–88.
2002, pp. 547–578. Twardowski, M., Boss, E., Macdonald, J.B., Pegau, W.S., Barnard, A.H., Zaneveld, J.R.V.,
Nelson, N.B., Siegel, D.A., Michaels, A.F., 1998. Seasonal dynamics of colored dissolved 2001. A model for estimating bulk refractive index from the optical backscattering
material in the Sargasso Sea. Deep-Sea Res. I 45, 931–957. ratio and the implications for understanding particle composition in case I and case II
O'Reilly, J.E., Maritorena, S., Mitchell, B.G., Siegel, D.A., Carder, K.L., Garver, S.A., waters. J. Geophys. Res. 106, 14,129–14,142.
Kahru, M., McClain, C.R., 1998. Ocean color chlorophyll algorithm for SeaWiFS. J. Vaillancourt, R.D., Brown, C.W., Guillard, R.R.L., Balch, W.M., 2004. Light backscattering
Geophys. Res. 103, 24937–24953. properties of marine phytoplankton: relationships to cell size, chemical composition
Organelli, E., Claustre, H., Bricaud, A., Barbieux, M., Uitz, J., D’Ortenzio, F., Dall’Olmo, and taxonomy. J. Plank. Res. 26 (2), 191–212.
G., 2017. Bio-optical anomalies in the world's oceans: an investigation on the diffuse Vodacek, A., Blough, N.V., DeGrandpre, M.D., Peltzer, E.T., Nelson, R.K., 1997. Seasonal
attenuation coefficients for downward irradiance derived from Biogeochemical Argo variation of CDOM and DOC in the Middle Atlantic Bight: terrestrial inputs and
float measurements. J. Geophys. Res. 122. photooxidation. Limnol. Oceanogr. 42, 674–686.
Preisendorfer, R.W., 1961. Application of Radiative Transfer Theory to Light Werdell, P., Bailey, S., 2005. An improved in‐situ bio‐optical data set for ocean color
Measurements in the Sea. 10. Institut géographique national, pp. 11–30. algorithm development and satellite data product validation. Remote Sens. Environ.
Prieur, L., Sathyendranath, S., 1981. An optical classification of coastal and oceanic 98 (1), 122–140.
waters based on the specific spectral absorption curves of phytoplankton pigments, Westberry, T.K., Boss, E., Lee, Z., 2013. The influence of Raman scattering on ocean color
dissolved organic matter, and other particulate materials. Limnol. Oceanogr. 26, inversion models. Appl. Opt. 52, 55525561.
671–689. Westberry, T., Dall’Olmo, G., Behrenfeld, M.J., Moutin, T., 2010. Coherence of particulate
Reynolds, R.A., Stramski, D., Neukermans, G., 2016. Optical backscattering by particles in beam attenuation and backscattering coefficients in diverse open ocean environ-
Arctic seawater and relationships to particle mass concentration, size distribution, ments. Opt. Express 18 (15), 15,419–15,425.
and bulk composition. Limnol. Oceanogr. 61, 1869–1890. Whitmire, A.L., Boss, E., Cowles, T.J., Pegau, W.S., 2007. Spectral variability of the
Reynolds, R.A., Stramski, D., Mitchell, B.G., 2001. A chlorophyll dependent semi- particulate backscattering ratio. Opt. Express 15, 7019–7031.
analytical reflectance model derived from field measurements of absorption and Whitmire, A.L., Pegau, W.S., Karp-Boss, L., Boss, E., Cowles, T.J., 2010. Spectral back-
backscattering coefficients within the Southern Ocean. J. Geophys. Res. 106, scattering properties of marine phytoplankton cultures. Opt. Express 18,
7125–7138. 15073–15093.
Sathyendranath, S., Platt, T., 1989. Remote sensing of ocean chlorophyll: consequence of Yentsch, C.S., Reichert, C.A., 1962. The interrelationship between water-soluble yellow
nonuniform pigment profile. Appl. Opt. 28 (3), 490–495. substances and chloroplastic pigments in marine algae. Bot. Mar. 3, 65–74.
Slade, W.H., Boss, E., 2015. Spectral attenuation and backscattering as indicators of Zaneveld, J.R.V., Kitchen, J.C., Moore, C.M., 1994. The scattering error correction of
average particle size. Appl. Opt. 54 (24), 7264–7277. reflecting-tube absorption meters. Proc. SPIE 2258, 44–55 Ocean optics XII.
Stramski, D., Reynolds, R., Babin, M., Kaczmarek, S., Lewis, M., Roettgers, R., Sciandra, Zeng, C., Xu, H., Fischer, A.M., 2016. Chlorophyll-a estimation around the antarctica
A., Stramska, M., Twardowski, M., Franz, B., et al., 2008. Relationships between the peninsula using satellite algorithms: hintshints from field water leaving reflectance.
surface concentration of particulate organic carbon and optical properties in the Sensors 16, 2075.
eastern South Pacific and eastern Atlantic Oceans. Biogeosciences 5 (1), 171–201. Zhang, X., Hu, L., He, M., 2009. Scattering by pure seawater: effect of salinity. Opt.
Stramska, M., Stramski, D., 2005. Variability of particulate organic carbon concentration Express 17 (7), 5698–5710.
in the north polar Atlantic based on ocean color observations with Sea-viewing Wide Zhang, X., Lewis, M., Lee, M.E.-G., Johnson, B., Korotaev, G.K., 2002. The volume scat-
Field-of-view Sensor (SeaWiFS). J. Geophys. Res. 110, 1–16. tering function of natural bubble populations. Limnol. Oceanogr. 47, 1273–1282.

12

You might also like