generalized ocean color inversion model for retrieving

19
Generalized ocean color inversion model for retrieving marine inherent optical properties P. Jeremy Werdell, 1,2, * Bryan A. Franz, 1 Sean W. Bailey, 1,3 Gene C. Feldman, 1 Emmanuel Boss, 2 Vittorio E. Brando, 4 Mark Dowell, 5 Takafumi Hirata, 6 Samantha J. Lavender, 7 ZhongPing Lee, 8 Hubert Loisel, 9 Stéphane Maritorena, 10 Fréderic Mélin, 5 Timothy S. Moore, 11 Timothy J. Smyth, 12 David Antoine, 13 Emmanuel Devred, 14 Odile Hembise Fanton dAndon, 15 and Antoine Mangin 15 1 NASA Goddard Space Flight Center, Code 616, Greenbelt, Maryland 20771, USA 2 School of Marine Sciences, University of Maine, Orono, Maine 04401, USA 3 Futuretech Corporation, Greenbelt, Maryland 20770, USA 4 CSIRO Land & Water, Environmental Earth Observation Program, G.P.O. Box 1666, Canberra, Australia 5 Joint Research Centre, European Commission, Ispra 21027, Italy 6 Faculty of Environmental Earth Science, Hokkaido University, Sapporo 060-0810, Japan 7 Pixalytics Ltd., Plymouth PL6 8BX, UK 8 Environmental Earth and Ocean Sciences, University of Massachusetts at Boston, Boston, Massachusetts 02125, USA 9 Laboratoire dOcéanologie et de Géosciences, Université du LittoralCôte dOpale, Wimereux F-62930, France 10 Earth Research Institute, University of California Santa Barbara, Santa Barbara, California 93106, USA 11 Ocean Process Analysis Laboratory, University of New Hampshire, Durham, New Hampshire 03824, USA 12 Plymouth Marine Laboratory, Plymouth PL1 3DH, UK 13 CNRS, Laboratoire dOcéanographie de Villefranche, BP 08 06238 Villefranche sur mer, France 14 Unité mixte internationale Takuvik/Joint international Laboratory, Université Laval, Québec GV1 0AG, Canada 15 ACRI-ST, 06904 Sophia Antipolis Cedex, France *Corresponding author: [email protected] Received 20 December 2012; revised 7 February 2013; accepted 11 February 2013; posted 14 February 2013 (Doc. ID 182094); published 22 March 2013 Ocean color measured from satellites provides daily, global estimates of marine inherent optical proper- ties (IOPs). Semi-analytical algorithms (SAAs) provide one mechanism for inverting the color of the water observed by the satellite into IOPs. While numerous SAAs exist, most are similarly constructed and few are appropriately parameterized for all water masses for all seasons. To initiate community-wide discussion of these limitations, NASA organized two workshops that deconstructed SAAs to identify simi- larities and uniqueness and to progress toward consensus on a unified SAA. This effort resulted in the development of the generalized IOP (GIOP) model software that allows for the construction of different SAAs at runtime by selection from an assortment of model parameterizations. As such, GIOP permits isolation and evaluation of specific modeling assumptions, construction of SAAs, development of region- ally tuned SAAs, and execution of ensemble inversion modeling. Working groups associated with the 1559-128X/13/102019-19$15.00/0 © 2013 Optical Society of America 1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2019

Upload: others

Post on 01-May-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Generalized ocean color inversion model for retrieving

Generalized ocean color inversion model for retrievingmarine inherent optical properties

P. Jeremy Werdell,1,2,* Bryan A. Franz,1 Sean W. Bailey,1,3 Gene C. Feldman,1

Emmanuel Boss,2 Vittorio E. Brando,4 Mark Dowell,5 Takafumi Hirata,6

Samantha J. Lavender,7 ZhongPing Lee,8 Hubert Loisel,9

Stéphane Maritorena,10 Fréderic Mélin,5 Timothy S. Moore,11

Timothy J. Smyth,12 David Antoine,13 Emmanuel Devred,14

Odile Hembise Fanton d’Andon,15 and Antoine Mangin15

1NASA Goddard Space Flight Center, Code 616, Greenbelt, Maryland 20771, USA2School of Marine Sciences, University of Maine, Orono, Maine 04401, USA

3Futuretech Corporation, Greenbelt, Maryland 20770, USA4CSIRO Land & Water, Environmental Earth Observation Program, G.P.O. Box 1666, Canberra, Australia

5Joint Research Centre, European Commission, Ispra 21027, Italy6Faculty of Environmental Earth Science, Hokkaido University, Sapporo 060-0810, Japan

7Pixalytics Ltd., Plymouth PL6 8BX, UK8Environmental Earth and Ocean Sciences, University of Massachusetts at Boston, Boston, Massachusetts 02125, USA

9Laboratoire d’Océanologie et de Géosciences, Université du Littoral—Côte d’Opale, Wimereux F-62930, France10Earth Research Institute, University of California Santa Barbara, Santa Barbara, California 93106, USA

11Ocean Process Analysis Laboratory, University of New Hampshire, Durham, New Hampshire 03824, USA12Plymouth Marine Laboratory, Plymouth PL1 3DH, UK

13CNRS, Laboratoire d’Océanographie de Villefranche, BP 08 06238 Villefranche sur mer, France14Unité mixte internationale Takuvik/Joint international Laboratory, Université Laval, Québec GV1 0AG, Canada

15ACRI-ST, 06904 Sophia Antipolis Cedex, France

*Corresponding author: [email protected]

Received 20 December 2012; revised 7 February 2013; accepted 11 February 2013;posted 14 February 2013 (Doc. ID 182094); published 22 March 2013

Ocean color measured from satellites provides daily, global estimates of marine inherent optical proper-ties (IOPs). Semi-analytical algorithms (SAAs) provide one mechanism for inverting the color of thewater observed by the satellite into IOPs. While numerous SAAs exist, most are similarly constructedand few are appropriately parameterized for all water masses for all seasons. To initiate community-widediscussion of these limitations, NASA organized two workshops that deconstructed SAAs to identify simi-larities and uniqueness and to progress toward consensus on a unified SAA. This effort resulted in thedevelopment of the generalized IOP (GIOP) model software that allows for the construction of differentSAAs at runtime by selection from an assortment of model parameterizations. As such, GIOP permitsisolation and evaluation of specific modeling assumptions, construction of SAAs, development of region-ally tuned SAAs, and execution of ensemble inversion modeling. Working groups associated with the

1559-128X/13/102019-19$15.00/0© 2013 Optical Society of America

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2019

Page 2: Generalized ocean color inversion model for retrieving

workshops proposed a preliminary default configuration for GIOP (GIOP-DC), with alternative modelparameterizations and features defined for subsequent evaluation. In this paper, we: (1) describe thetheoretical basis of GIOP; (2) present GIOP-DC and verify its comparable performance to other popularSAAs using both in situ and synthetic data sets; and, (3) quantify the sensitivities of their output to theirparameterization. We use the latter to develop a hierarchical sensitivity of SAAs to various model pa-rameterizations, to identify components of SAAs that merit focus in future research, and to providematerial for discussion on algorithm uncertainties and future emsemble applications. © 2013 OpticalSociety of AmericaOCIS codes: 010.4450, 280.4991.

1. Introduction

Satellite ocean color instruments, such as the NASAModerate Resolution Imaging Spectroradiometer on-board Aqua (MODISA) and ESA Medium ResolutionImaging Spectrometer (MERIS), provide daily globalestimates of marine inherent optical properties(IOPs). Remotely sensed IOPs, namely the spectralabsorption and backscattering coefficients of thesurface water column and its particulate and dis-solved constituents, describe the contents of theupper ocean, information critical to furthering scien-tific understanding of biogeochemical processes, suchas carbon exchanges, phytoplankton biodiversityshifts, and responses to climatic disturbances. Assuch, the international community has investedsignificant effort in ensuring and improving thequality of remotely sensed IOPs. Recognizing this,the International Ocean Colour Coordinating Group(IOCCG) established a working group to compile andcompare popular methods for estimating IOPs fromocean color [1]. NASA recently extended this effortthrough a series of internationally attended work-shops aimed at: (a) deconstructing each method toidentify similarities and uniqueness; (b) identifyingstrategies to provide uncertainties for each method;and (c) achieving community-wide consensus on aunified method for generating global satellite IOPproducts [2].

Semi-analytical algorithms (SAAs) provide onemechanism for inverting ocean color into IOPsthrough a combination of empiricism and radiativetransfer theory. Many published SAAs attempt toretrieve spectral IOPs from spectral remote-sensingreflectances [Rrs�λ�; sr−1] and to further separate thetotal absorption and backscattering coefficients intosubcomponents [e.g., the absorption of algal, nonal-gal, and colored dissolved organic matter (CDOM)][3–11]. These SAAs generally assume spectral shapefunctions (or eigenvectors) of the constituent absorp-tion and scattering components and retrieve themagnitudes (or eigenvalues) of each constituentrequired to match the spectral distribution of Rrs�λ�.Thus, SAAs often differ only in the assumptionsemployed to define the eigenvectors and in themathematical methods applied to calculate the ei-genvalues. In lieu of this, and in support of the afore-mentioned workshops, the NASA OceanBiology Processing Group (OBPG) [12] developed

the generalized IOP (GIOP) framework to facilitatecontrolled evaluation of the varied SAA approacheswithin a satellite data processing environment [13].

Briefly, GIOP allows construction of different IOPmodels at runtime by selection from a wide assort-ment of published absorption and backscatteringeigenvectors [4,6–8,14–17]. As such, GIOP permitsisolation and evaluation of specific modeling assump-tions, construction of new SAAs, development ofregionally tuned SAAs (e.g., [18]), and ensemble in-version modeling. The OBPG implemented GIOP isstandard NASA ocean color processing code [19] andcurrently distributes GIOP to the research commu-nity via the Sea-viewing Wide Field-of-View Sensor(SeaWiFS) Data Analysis System (SeaDAS) [20]. Theworking group associated with the NASA GIOPworkshops (October 2008 and September 2010) [2]proposed the preliminary configuration of GIOP,with alternative model parameterizations and fea-tures defined for subsequent evaluation. In thispaper, we: (a) describe the theoretical basis of GIOP;(b) present its preliminary configuration; (c) evaluateand validate this configuration; and (d) quantify thesensitivities of output eigenvalues to their assumedeigenvectors. We provide a brief summary of themodel components within GIOP as compared toseveral common SAAs in Appendix A.

2. Methods

A. Model Development

Ocean color satellite instruments provide estimatesof Rrs�λ� after atmospheric correction. The method ofLee et al. [7] provides a convenient method to convertRrs�λ� to their subsurface values:

rrs�λ; 0−� �Rrs�λ�

0.52� 1.7Rrs�λ�: (1)

Subsurface remote-sensing reflectances relate tomarine IOPs via

rrs�λ; 0−� � G1�λ�u�λ� �G2�λ�u�λ�2; (2)

where the two G�λ� (sr−1) vary with illuminationconditions and geometry, sea surface properties,

2020 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 3: Generalized ocean color inversion model for retrieving

bidirectional effects, and the shape of the marinevolume scattering function. Most often, u�λ� isdescribed as

u�λ� � bb�λ�a�λ� � bb�λ�

; (3)

where bb is the total backscattering coefficient (m−1)and a is the total absorption coefficient (m−1) [21], butformulations such as u�λ� � bb�λ�∕a�λ� have alsobeen adopted [22]. Common methods for estimatingG�λ� include Gordon et al. [21], where G1 and G2 arespectrally fixed to 0.0949 and 0.0794 (see [7,23] foralternative coefficients), and the tabulated resultsof Morel et al. [22], where G1 is estimated using solarand sensor geometries and an estimate of algal bio-mass and G2 is set to 0. GIOP supports all of theseoptions. In practice, most SAAs first solve for u�λ�,then decompose (or, invert) u�λ� into its componentIOPs.

The absorption coefficient can be expanded asthe sum of all absorbing components. Further,each component can be expressed as the productof its concentration-specific absorption spectrum(eigenvector; a�) and its concentration or amplitude(eigenvalue; A):

a�λ� � aw�λ� �XNϕ

i�1

Aϕia�ϕi�λ�

�XNd

i�1

Adia�di�λ� �

XNg

i�1

Agia�gi�λ�; (4)

where the subscripts w, ϕ, d, and g indicate contribu-tions by water, phytoplankton, nonalgal particles(NAP), and CDOM. Both a�

d�λ� and a�g�λ� are com-

monly expressed as

a�d;g�λ� � exp�−Sd;gλ�; (5)

where Sd and Sg typically vary between 0.01 and0.02 nm−1 in natural waters [24]. As the spectralshapes of NAP and CDOM absorption differ onlyin their exponential slopes, the two componentsare typically combined for satellite applicationsand Eq. (4) becomes

a�λ� � aw�λ� �XNϕ

i�1

Aϕia�ϕi�λ� �

XNdg

i�1

Adgia�dgi

�λ�; (6)

which is the default option for GIOP.Total backscattering can be expanded to

bb�λ� � bbw�λ� �XNbp

i�1

Bbpib�bpi

�λ�; (7)

where the subscripts bw and bp indicate contribu-tions by seawater and particles. Bbp provides the

eigenvalue and a power function often representsthe eigenvector:

b�bp�λ� � λSbp ; (8)

where Sbp typically varies between −2 and 0 fromsmall to large particles. While commonly employedin the remote-sensing paradigm, we acknowledgethe validity of the power function for b�bp�λ� remainsdebatable [25–27]. Both aw�λ� and bbw�λ� areknown [28,29].

Using Rrs�λ� and eigenvectors as input, eigenval-ues for absorption (A) and backscattering (B) canbe estimated via linear or nonlinear least squaresinversion of Eqs. (1)–(3). For example, Roesler andPerry [3] used the nonlinear optimization schemeof Levenberg–Marquardt (LM), while Hoge and Lyon[4] showed the problem could be linearized and thusdirectly solved via linear matrix inversion. GIOPsupports both linear and nonlinear optimizationand matrix inversion schemes [13]. The optimizedeigenvalues represent the relative contributions ofeach defined absorbing and scattering constituent.In the special case where a�

ϕ�λ� is provided aschlorophyll-specific absorption (m2 mg−1), the eigen-value Aϕ provides an estimate of the chlorophyllconcentration, Ca (mgm−3). Note that this modeldescribes each component of absorption and back-scattering as a linear sum of subcomponents,presumably with unique spectral dependencies [sym-bolized by the summation over N in Eqs. (4), (6), and(7)]. In this way, the absorption characteristics fordifferent phytoplankton populations and the scatter-ing characteristics of multiple size distributions ofsuspended particles can be represented, or Eq. (6)can be re-expanded to Eq. (4).

B. Model Configuration

GIOP provides a satellite data processing frameworkwithin which an SAA can be constructed at runtimeby selection from an assortment of published eigen-vectors (Table 1). Franz and Werdell [13] provide adetailed description of its use within the SeaDASenvironment [20]. The general form of the GIOPinversion model (Section 2.A) is common to a numberof published approaches (e.g., [3–5,8,9,30]) whosedifferences reside in the choice of eigenvectorsemployed, the number of eigenvalues resolved, theoptimization method selected, and the number ofsensor wavelengths considered in the optimization.A unique instance of GIOP is therefore defined byspecifying eigenvectors for each optically significantconstituent assumed to exist in the water column.The working groups associated with the NASA GIOPworkshops [2] proposed the following preliminary,default configuration of GIOP, hereafter referred toas GIOP default configuration (GIOP-DC), to supportNASA’s production and distribution of global IOPdata products:

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2021

Page 4: Generalized ocean color inversion model for retrieving

• Relate rrs�λ; 0−� to IOPs using Eq. (2) and

u�λ��bbw�λ��Bbpb�bp�λ�

bbw�λ��Bbpb�bp�λ��aw�λ��Adga�dg�λ��Aϕa�

ϕ�λ�:

(9)

• Spectrally invariant G1 and G2 from Gordonet al. [21]• b�bp�λ� from Eq. (8) and Sbp from Lee et al. [7]• a�

dg�λ� from Eq. (5) and Sdg � 0.018 nm−1 [24]• a�

ϕ�λ� from Bricaud et al. [14], estimated usingOC-derived Ca [31] and normalized such thata�ϕ�443� � 0.055 m2 mg−1.

The latter normalization allows the spectral shape ofa�ϕ�λ� to change with an estimate ofCa, but constrains

its magnitude to an average value for oceanicwater [3,8,14].

Our choice of default parameterizations served twopurposes: to provide some consistency with previ-ously developed SAAs and to acknowledge theemerging quality of variable, dynamically selectedeigenvectors. In particular, we adopted G1 and G2from Gordon et al. [21] to be consistent with previ-ously published work [3–5,7–9] and variable Sbpand a�

ϕ�λ� to better represent heterogeneous naturalenvironments than would be possible with fixed val-ues. Our choice of a fixed Sdg represents a compro-mise between multiple published values [8,24],although we expect GIOP-DC to ultimately adopt avariable Sdg as methods for its dynamic calculationimprove. Note, we did not make these choices basedon comparisons with in situ or synthetic data sets,nor do we suggest that these parameterizations re-present all natural oceanic conditions at all times.Our choices simply represent a consistent, modernSAA configuration fromwhich to evolve as our under-standing of marine bio-optics improves.

In GIOP-DC, the three unknown eigenvalues(Bbp, Adg, and Aϕ) are optimized using LM, and wave-lengths between 400 and 700 nm are considered. Ourconvergence criterion tests both absolute and rela-tive changes in the optimized eigenvalues betweeniterations of the LM algorithm. Convergence isreached when the change in any eigenvalue, X ,between iteration i and iteration i� 1 satisfiesjXi�1 − Xij < 0.0001� 0.0001jXij. If the number ofiterations exceeds 50, the iteration is stopped andthe result is treated as algorithm failure. Note thatthese criteria do not require the results to be valid oreven that the model provides a good fit to the data.Despite this, we consider the IOP retrievals to bevalid when

• −0.05bbw�λ� ≤ bbp�λ� ≤ 0.05 m−1

• −0.05aw�λ� ≤ adg�λ� ≤ 5 m−1

• −0.05aw�λ� ≤ aϕ�λ� ≤ 5 m−1

• ΔRrs ≤ 33%,

where the mean relative Rrs�λ� difference, a measureof goodness of fit, is derived as

ΔRrs �100%Nλ

XNλ

i�1

jRrs�λi� − Rrs�λi�jRrs�λi�

; (10)

for 400 ≤ λ ≤ 600 nm with Rrs indicating remote-sensing reflectance reconstructed from the modeledIOPs. We consider slightly negative IOP values to beviable (rather, statistically equivalent to zero), asthe retrieved eigenvalues include some uncertainty.Weprovideabrief comparisonofGIOP-DC(andGIOP,in general) and several other SAAs commonly used insatellite ocean color applications in Appendix A.

C. Uncertainties

The LM optimization method minimizes a χ2 meritfunction defined as

Table 1. Summary of Eigenvectors Available for Use in GIOP (as of March 2011)a

Eigenvector Description Reference

User-provided a�ϕ�λ�

a�ϕ�λ� Maritorena et al. (2002) tabulated a�

ϕ�λ� [8]Bricaud et al. (1998)-derived a�

ϕ�λ� using OC-derived Ca [14]Ciotti and Bricaud (2006)-derived a�

ϕ�λ� using user-provided size fraction [17]

Eq. (5) with user-provided Sdg

a�dg�λ� Eq. (5) with Lee et al. (2002)-derived Sdg [7]

Eq. (5) with Franz and Werdell (2010)-derived Sdg [13]User-provided a�

dg�λ�Eq. (8) with user-provided Sbp

Eq. (8) with Hoge and Lyon (1996)-derived Sbp [4]Eq. (8) with Lee et al. (2002)-derived Sbp [7]Eq. (8) with Ciotti et al. (1999)-derived Sbp [15]

b�bp�λ� Eq. (8) with Morel and Maritorena (2001)-derived Sbp [22]Eq. (8) with Loisel and Stramski (2000)-derived Sbp [6]User-provided b�bp�λ�Loisel and Stramski (2000)-derived b�bp�λ� [6]Lee et al. (2002)-derived b�bp�λ� [7]

aBoldface indicates the eigenvector used in GIOP-DC.

2022 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 5: Generalized ocean color inversion model for retrieving

χ2 �XNλ

i�1

�Rrs�λi� − Rrs�λi��2σ2�λi�

; (11)

where σ�λi� are the input uncertainties onRrs�λi�. LMmakes use of the Jacobian matrix (J), which isderived by evaluating the partial derivatives of themerit function with respect to the optimized eigen-values at each λi. From the Jacobian, the covariancematrix (M) is then constructed as M � �JTJ�−1, re-sulting in a square matrix with dimensionality equalto the number of model eigenvalues. The square-rootof the diagonal elements of M represent the uncer-tainty on the optimized eigenvalues, taking intoaccount both the weighting and scale of the inputuncertainties. If reliable values of σ�λi� are notavailable, they are set to 1.0 and the optimizationis unweighted. In that case, the uncertainty is com-puted using the square-root of the diagonal terms ofthe variance–covariance matrix, σ2M, where

σ2 � 1Nλ

XNλ

i�1

�Rrs�λi� − Rrs�λi��2: (12)

The GIOP-DC configuration of GIOP currently uti-lizes an unweighted optimization with the variance–covariance matrix utilized to estimate uncertaintieson the model eigenvalues. The uncertainty on thederived IOPs is then determined by multiplyingthe uncertainty of the optimized eigenvalue by theassociated eigenvector.

D. Data Acquisition

We acquired coincident observations of in situ andsynthesized Rrs�λ�, bbp�λ�, adg�λ�, aϕ�λ�, and Ca fromthe NASA bio-Optical Marine Algorithm Dataset (NOMAD) [32] and the IOCCG Ocean ColourAlgorithmsWorking Group synthetic data set [1]. Wealso acquired coincident Level-2 satellite-to-in situmatch-ups for the NASA SeaWiFS and MODISA in-struments from the OBPG (http://seabass.gsfc.nasa.gov/seabasscgi/validation_search.cgi). Satellite dataprocessing and quality assurance for the match-ups followed Bailey and Werdell [33]. Specifically:(1) temporal coincidence was defined as�3 h; (2) sat-ellite values were the filtered median (via the semi-interquartile range) of all unmasked pixels in a 5 × 5box centered on the in situ target; and (3) satellitevalues were excluded when the coefficient of varia-tion for the given product within this box exceeded0.15. SeaWiFS andMODISA data were processed fol-lowing their R2010.0 (September 2010) and R2012.0(May 2012) reprocessing configurations, respectively[34]. The in situ data used in the match-up analysesinclude a small subset of NOMAD, plus additionalRrs�λ�, Ca, bbp�λ�, adg�λ�, and aϕ�λ� archived in theNASA SeaWiFS Bio-optical Archive and StorageSystem (SeaBASS) [35]. Sample sizes for bbp�λ� differfrom a�λ�, adg�λ�, and aϕ�λ� in NOMAD and thematch-up data sets because of instrumental variabil-ity in field campaigns included in these compilations.

E. Validation Analyses

We adopted a validation approach based on model-measurement regression statistics and a spectralgoodness of fit metric. We compared modeled andground-truth IOPs, with the former calculated usingin situ, synthesized, and satellite Rrs�λ� from theNOMAD, IOCCG, and satellite-to-in situ match-updata sets. We generated Type II linear regressionstatistics and estimates of ΔIOP for all of the above,with ΔIOP defined as

ΔIOP � 200%Nλ

XNλ

i�1

j ^IOP�λi� − IOP�λi�j^IOP�λi� � IOP�λi�

; (13)

for 400 ≤ λ ≤ 600 nm. IOP and ^IOP indicate a mea-sured and modeled absorption or backscatteringcoefficient, respectively. Accordingly, we report ΔIOPelsewhere as Δbbp, Δa, Δadg, and Δaϕ. We generatedmodeled IOPs using in situ Rrs�λ� and calculatedΔRrs and ΔIOP distribution medians and semi-interquartile ranges for various populations (here,the semi-interquartile range is the range covered byvalues such that 50% occur with equal probability oneither side of the median). Furthermore, for theanalysis of these delta statistics, we stratified theground-truth stations into three trophic levels: oligo-trophic water has Ca ≤ 0.1 mgm−3; mesotrophicwater has 0.1 < Ca ≤ 1 mgm−3; and eutrophic waterhas Ca > 1 mgm−3. Table 2 provides calculations forother comparative statistics.

F. Sensitivity Analyses

We executed 12 additional unique instances of GIOPto evaluate the sensitivity of the GIOP-DC configu-ration to alternate selections of eigenvectors. Eachof the 12 instances included a single alternateparameterization: (1,2) Sdg � 33% (� 0.012 and0.024 nm−1), (3) Sdg dynamically calculated usingLee et al. [7], (4,5) Sbp from Lee et al. [7] �33%; (6,7)OC-derived Ca � 33% prior to input into Bricaudet al. [14]; (8) a�

ϕ�λ� from Bricaud et al. [14] withCa fixed at 0.18 mgm−3; (9) a�

ϕ�λ� from Ciotti andBricaud [17] with a size fraction of 0.5; (10) G�λ� fromMorel et al. [22], (11) optimization using linear ma-trix inversion, and (12) optimization considering only400 ≤ λ ≤ 600 nm. We quantified spectral changes inmodeled IOPs for each alternate parameterization(relative to the baseline GIOP-DC configuration) us-ing Type II regression statistics, estimates of ΔIOP,and Taylor [36] and Target [37] summary diagrams.The latter provide convenient means for simultane-ously considering magnitudes of deviations, correla-tions, and biases between the alternate runs andGIOP-DC. To minimize the impact of uncertaintiesinherent to the in situ IOP measurements, we onlyconsidered the synthetic IOCCG data set in theseanalyses (note, this does not remove uncertaintiesin the truth data set, but rather, limits them to thoseassociated with the inherent assumptions used togenerate this synthetic data set).

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2023

Page 6: Generalized ocean color inversion model for retrieving

3. Results

Direct comparisons of modeled and ground-truthIOPs provided estimates of the accuracy and preci-sion of GIOP-DC. Overall, spectral IOPs fromGIOP-DC (indicated by a carot hat below) repro-duced the in situ and synthetic values moderatelywell over the full dynamic range of the NOMADand IOCCG data sets, with r2 ranging from 0.49 to0.90 and 0.78 to 0.99, respectively (Fig. 1, Table 2).With the exception of aϕ�λ�, median percentdifferences (MPD; the caption for Table 2 providesits calculation) for NOMAD showed greater overallvariability than for the IOCCG data set, with valuesranging largely from 20%–40% compared to10%–30%. Both data sets showed less variabilityfor bbp�λ� and a�λ� than for the component absorptionproducts, adg�λ� and aϕ�λ� [38,39]. As demonstratedby their median ratios, bbp�λ� and a�λ� for NOMADshowed high and low biases (over and underesti-mates), respectively, while both approached unityfor the IOCCG data set. Relative to the ratios fora�λ�, ratios for adg�λ� and aϕ�λ� from both data setsshowed low and high biases, respectively, althoughthe IOCCG results better demonstrated this behav-ior. Interestingly, the absolute magnitudes of thebiases in component absorption products increasedwith increasing wavelength for both data sets, pos-sibly due to the increasing contribution of aw�λ� toa�λ� with increasing wavelength. The magnitudesof the regression slopes for bbp�λ� and adg�λ� from

NOMAD also showed strong, albeit opposite, spectraldependency.

Not surprisingly, comparisons of modeled andin situ IOPs for SeaWiFS and MODISA resulted insimilar statistics to those for NOMAD, as the in situIOPs used in the NOMAD and satellite comparisonsstemmed from similar sources (Fig. 2, Table 3). Thisalso implies, however, that the satellite and in situRrs�λ� carry similar spectral uncertainties or thatthe inversion process remains somewhat insensi-tive to uncertainties in Rrs�λ�. The SeaWiFS andMODISA samples sizes and dynamic ranges werefar lower than those for the NOMAD and IOCCGcomparisons, yet their MPD were higher, at leastfor the absorption coefficients. Similar to theNOMAD and IOCCG results, MPD for adg�λ� andaϕ�λ� exceeded those for bbp�λ� and a�λ�. The medianratios for adg�λ� and aϕ�λ� showed definitive spectraldependency, but only partial low and high biases rel-ative to the ratios for a�λ�. Similar to the NOMADand IOCCG results, the magnitudes of the slopesfor bbp�λ� and adg�λ� for SeaWiFS showed spectraldependencies. Such behavior was not obvious forMODISA, presumably given its small sample sizefor adg�λ�. As for the NOMAD and IOCCG results,GIOP-DC appeared to best retrieve a�λ�, echoingthe findings of an IOCCG working group [1].

GIOP-DC ran successfully on 90% of stations inNOMAD and the IOCCG data set, independent of tro-phic level (Table 4). The 10% failure rate resulted froma combination of ΔRrs > 33% and nonconvergence of

Table 2. Regression Statistics for GIOP-DC Using the NOMAD and IOCCG Data Setsa

NOMAD IOCCG

N r2 Slope (SE) Ratio MPD N r2 Slope (SE) Ratio MPD

412 217 0.49 0.99 (0.05) 1.26 28.2 437 0.92 1.08 (0.01) 0.98 10.1443 217 0.52 1.00 (0.05) 1.27 28.8 437 0.93 1.07 (0.01) 1.00 9.3

bbp 490 217 0.57 1.04 (0.05) 1.29 30.0 437 0.94 1.07 (0.01) 0.99 8.2555 217 0.60 1.06 (0.05) 1.30 32.8 437 0.95 1.09 (0.01) 0.91 11.8670 217 0.64 1.09 (0.05) 1.27 31.3 437 0.96 1.09 (0.01) 0.90 13.0

412 649 0.90 1.14 (0.01) 0.89 20.5 421 0.99 1.03 (0.00) 1.03 7.2443 657 0.90 1.13 (0.01) 0.88 21.8 427 0.99 1.03 (0.01) 0.99 6.1

a 490 657 0.88 1.13 (0.02) 0.84 24.6 432 0.98 1.03 (0.01) 0.98 8.7555 654 0.81 1.21 (0.02) 0.77 37.4 436 0.95 1.11 (0.01) 0.73 29.9670 609 0.80 1.18 (0.02) 1.11 37.4 424 0.86 1.01 (0.02) 1.50 58.6

412 654 0.83 1.12 (0.02) 0.84 28.8 426 0.98 1.02 (0.01) 1.01 13.1443 662 0.81 1.10 (0.02) 0.78 35.1 429 0.97 1.04 (0.01) 0.90 15.9

adg 490 662 0.76 1.06 (0.02) 0.65 43.2 435 0.95 1.07 (0.01) 0.80 25.0555 656 0.71 1.03 (0.02) 0.53 52.7 437 0.91 1.08 (0.02) 0.58 44.1670 621 0.61 0.96 (0.03) 0.32 69.8 426 0.82 1.05 (0.02) 0.36 64.2

412 676 0.76 1.23 (0.02) 0.94 27.5 419 0.81 1.18 (0.03) 1.16 32.8443 682 0.76 1.21 (0.02) 0.98 25.9 419 0.79 1.13 (0.03) 1.20 30.3

aϕ 490 682 0.76 1.23 (0.02) 1.08 27.2 422 0.78 1.13 (0.03) 1.34 41.4555 673 0.78 1.22 (0.02) 1.25 42.5 422 0.84 1.10 (0.02) 1.36 52.5670 680 0.82 1.12 (0.02) 1.33 43.3 422 0.81 1.00 (0.02) 2.01 100.8

aN, r2, and Slope (SE) are the sample size, coefficient of determination, and regression slope (standard error of the regression slope),respectively. Ratio is the median ratio calculated as median�Xi∕Xi� and MPD is the median percent difference calculated asmedian�100% � jXi∕Xi − 1j�, with Xi and Xi indicating each modeled and measured IOP, respectively. With the exception of Ratioand MPD, data were log-transformed.

2024 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 7: Generalized ocean color inversion model for retrieving

the inversion. When successful, ΔRrs fell below 2% onaverage for all trophic levels for both data sets(Fig. 3). While this indicates the eigenvectors can ac-curately reproduce Rrs�λ�, it does not necessarily sug-gest that GIOP-DC employs ideal eigenvectors. Incontrast to ΔRrs, ΔIOPs fell between 25%–50% forNOMAD (excluding the oligotrophic subset) and5%–35% for the IOCCG data set (excluding the eutro-phic, Δaϕ, Table 4). As will be discussed later, somevariability in the NOMAD results stems from thein situ data, most of which fail to accurately achieveradiometric closure betweenRrs�λ� and IOPs becauseof G�λ� or measurement error [see Eq. (2)]. TheIOCCG data set, however, maintains radiometric clo-sure by design and reported Δbbp and Δa distributionmeans and modes below 15% (Fig. 4). As for thematch-up results, delta values for the component ab-sorption products, Δadg and Δaϕ were degraded rel-ative to Δa. The latter is true for the full IOCCG dataset and for the meso- and eutrophic subsets ofNOMAD, which dominate its sampling distribution.For the IOCCG data set, Δadg remained consistentfor the three trophic levels (28%� 3%), while Δaϕ in-creased with increasing trophic level. This perhaps

resulted from the decreasing relative concentrationof aϕ�λ� to total absorption. The oligo-, meso-, and eu-trophic subsets of the IOCCG data set maintainmedian ratios for aϕ�443� to aϕ�443� � adg�443� of0.40, 0.31, and 0.23, respectively.

A hierarchy emerges from the sensitivity analysesperformed on the IOCCG data set using alternate pa-rameterizations of eigenvectors. Under- and overesti-mation of Sbp produced eigenvalues that differedfrom GIOP-DC by only 2%–8% (MPD), with slightlyelevated ΔIOPs (Table 5). This indicates the re-trieved eigenvalues to be fairly insensitive to Sbp,particularly given that GIOP-DC employs a dynamiccalculation of this eigenvector [7]. Modifying the Caused as input into Bricaud et al. [14] to retrieve a�

ϕ�λ�produced eigenvalues that differed fromGIOP-DC byonly 1%–7% (MPD), with very comparable ΔIOPs.We suspect this results from stability in a�

ϕ�λ� fromthe Bricaud model over narrow ranges of Ca (�33%).Not surprisingly, the use of a fixed a�

ϕ�λ�, from eitherBricaud et al. [14] or Ciotti and Bricaud [17], resultedin eigenvalues that differed more significantly fromGIOP-DC, particularly in aϕ�λ�. This less dynamicapproach to assigning eigenvectors cannot as effi-ciently represent all water types at all times and,as such, these two runs appear as outliers with re-gard to standard deviations in the Taylor and Targetdiagrams (Figs. 5–7; blue circle and green square).However, while these two fixed a�

ϕ�λ� runs returnedsomewhat elevated Δbbp, Δa, and Δadg relative toGIOP-DC, they returned improved Δaϕ. The choiceof Sdg appears to be the most critical within the con-text of this experiment, at least with regard to sepa-rating total absorption into its algal and nonalgalcomponents [38]. The three changes to Sdg producedthe three most significant departures from GIOP-DCin retrieved eigenvalues. Reducing Sdg to 0.012 nm−1

produced the highest Δbbp and Δa. This run appearsas an outlier with regard to standard deviation andcorrelation in the Taylor and Target diagrams(Figs. 5–7; blue cross). In contrast, elevating Sdg to0.024 nm−1 produced the highest Δadg and Δaϕ.Dynamically selecting Sdg via Lee et al. [7] producedan equivalentΔIOP to GIOP-DC and the lowest over-all Δadg of all the runs, emphasizing a potential ben-efit from dynamically assigning Sdg.

Choices made for the inversion itself also impactedthe retrieved eigenvalues, a discussion of which in-frequently appears in previous literature (Table 5)[9,39]. Using a linear matrix inversion method inlieu of the nonlinear LM optimization resulted indepartures from GIOP-DC of only 2%–7% (MPD),with comparable ΔIOPs. Excluding Rrs(670) from theinversion (that is, using only Rrs from 400 to 600 nm)resulted in negligible differences from GIOP-DC andsimilar ΔIOP. Intuitively, we expected larger depar-tures given the significance of red Rrs for SAAapplications in turbid waters [1,40]. This analysisemployed a synthesized data set, however, whichdoes not comprehensively represent highly turbid,highly scattering environments. As will be explored

bbp(443)

10−4 10−3 10−2 10−110−4

10−3

10−2

10−1

GIO

P−

DC

(m

−1 )

a(443)

10−3 10−2 10−1 100 10110−3

10−2

10−1

100

101

GIO

P−

DC

(m

−1 )

adg(443)

10−3 10−2 10−1 100 10110−3

10−2

10−1

100

101

GIO

P−

DC

(m

−1 )

aφ(443)

10−3 10−2 10−1 100 101

in situ (m−1)

10−3

10−2

10−1

100

101

GIO

P−

DC

(m

−1 )

bbp(443)

0 1 202468

1012

Fre

quen

cy (

%)

a(443)

0 1 20

5

10

15

Fre

quen

cy (

%)

adg(443)

0 1 202468

10

Fre

quen

cy (

%)

aφ(443)

0 1 2 GIOP−DC / in situ

02468

10

Fre

quen

cy (

%)

Fig. 1. Comparison of GIOP-DC and ground truth (in situ) IOPsat 443 nm fromNOMAD (black) and the IOCCG data set (red). Theleft column shows scatter plots for regression analyses. The rightcolumn shows ratios of GIOP-DC to ground truth. See Table 2 foraccompanying statistics.

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2025

Page 8: Generalized ocean color inversion model for retrieving

in detail later, an alternate choice of G�λ� [22] pro-duced eigenvalues that deviated more significantlyfrom GIOP-DC [6%–14% (MPD)] and resulted inelevated ΔIOP.

4. Discussion

We developed GIOP to provide a software environ-ment for developing and applying SAAs, not tointroduce a novel SAA. While GIOP-DC (one configu-ration instance of GIOP) provides a viable SAA foruse in global applications, it remains imperfect,and we recommend that this “default configuration”evolve over time as the community develops novelmethods and achieves advanced insights into bio-optical modeling and Rrs�λ�-IOP closure. We initiatedthis exercise to illustrate that differences in mostSAAs are limited to only the selection of eigenvectorsand the inversion or optimization approach, and thata consolidated software framework such as GIOPprovides a simple mechanism for moving ocean colorscience forward by allowing controlled SAA evalu-ation, regional tuning, dynamic eigenvector parame-terization based on optical water types (OWTs), andensemble inversion modeling. While GIOP-DC itselfperforms decently for the data sets under inves-tigation (Tables 2 and 3) and comparably to otherestablished algorithms [41] (see Appendix A for aquantitative discussion), we cannot expect reliabilityat all times in all water types, as is true for all glob-ally parameterized SAAs. We focus the following dis-cussion on lessons learned from its development andevaluation with the goal of recommending (outlining)directions for subsequent community research.

Table 3. Regression Statistics for GIOP-DC Using the SeaWiFS and MODISA Match-Up Data Sets

SeaWiFS MODISA

N r2 Slope (SE) Ratio MPD N r2 Slope (SE) Ratio MPD

412 123 0.30 0.77 (0.07) 0.93 25.2 56 0.69 0.92 (0.07) 1.00 13.2443 123 0.31 0.79 (0.07) 0.92 25.0 56 0.69 0.95 (0.08) 1.02 17.2

bbp 490 123 0.32 0.82 (0.07) 0.90 24.2 56 0.69 1.00 (0.08) 0.99 18.2555a 123 0.32 0.84 (0.07) 0.87 25.2 56 0.68 1.05 (0.09) 0.99 18.8670b 123 0.31 0.87 (0.07) 0.83 28.5 56 0.63 1.06 (0.09) 1.04 23.0

412 192 0.74 1.12 (0.04) 0.87 30.9 21 0.45 0.76 (0.14) 0.89 36.9443 192 0.81 1.07 (0.03) 0.81 25.5 21 0.73 0.77 (0.10) 0.88 16.9

a 490 192 0.80 1.01 (0.03) 0.76 29.3 21 0.84 0.79 (0.07) 0.79 21.1555a 192 0.67 1.03 (0.05) 0.68 42.3 21 0.86 0.74 (0.07) 0.75 28.9670b 180 0.69 1.19 (0.05) 0.87 45.0 17 0.47 0.91 (0.19) 1.88 87.8

412 192 0.51 1.12 (0.06) 0.86 45.7 20 0.07 0.96 (0.27) 0.88 40.2443 192 0.51 1.08 (0.06) 0.78 49.8 20 0.09 0.96 (0.27) 0.81 34.8

adg 490 192 0.48 1.01 (0.06) 0.64 54.2 20 0.11 0.96 (0.26) 0.68 41.8555a 191 0.42 0.93 (0.06) 0.50 62.5 20 0.12 0.96 (0.26) 0.46 55.3670b 183 0.44 0.82 (0.05) 0.34 72.0 20 0.13 0.98 (0.26) 0.32 67.7

412 195 0.72 1.17 (0.05) 0.73 35.5 25 0.85 1.14 (0.09) 0.91 22.5443 197 0.68 1.14 (0.05) 0.80 31.5 25 0.82 1.20 (0.11) 0.90 31.0

aϕ 490 197 0.68 1.12 (0.05) 0.86 30.1 25 0.81 1.13 (0.11) 0.93 33.2555a 186 0.71 1.17 (0.05) 0.95 44.8 25 0.82 1.14 (0.10) 0.97 36.6670b 195 0.73 1.05 (0.04) 1.11 48.6 24 0.81 1.24 (0.12) 1.35 49.3

aindicates that the wavelength is 547 nm for MODISA.bindicates that the wavelength is 667 nm for MODISA.

bbp(443)

10−4 10−3 10−2 10−110−4

10−3

10−2

10−1

GIO

P−

DC

(m

−1 )

a(443)

10−3 10−2 10−1 100 10110−3

10−2

10−1

100

101

GIO

P−

DC

(m

−1 )

adg(443)

10−3 10−2 10−1 100 10110−3

10−2

10−1

100

101

GIO

P−

DC

(m

−1 )

aφ(443)

10−3 10−2 10−1 100 101

in situ (m−1)

10−3

10−2

10−1

100

101

GIO

P−

DC

(m

−1 )

Fig. 2. Comparison of GIOP-DC and ground truth (in situ) IOPsat 443 nm from SeaWiFS (black) and MODISA (red). See Table 3for accompanying statistic.

2026 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 9: Generalized ocean color inversion model for retrieving

GIOP provides a resource for better understandingthe sensitivity of an SAA to its assigned eigenvectors.Our sensitivity analyses permitted generation of apreliminary hierarchy of eigenvector significance,which can be used to focus future research on the pa-rameterizations of most importance. This approachcannot effectively identify an optimal set of eigenvec-tors and methods, as we sequentially varied just asingle parameter and combinations of certain eigen-vectors mutually compensate for their respectiveerrors. Spectral shapes for bbp�λ�, adg�λ�, and aϕ�λ� de-crease from 440 to 660 nm, for example, and there-fore transfer variances when their shapes changewithin realistic bounds. We also expect adg�λ� andaϕ�λ� to covary under certain environmental condi-tions, such as the open ocean where the detrital poolincludes derivative products of phytoplankton, suchas cells walls and cytoplasm. That said, our approachidentifies the eigenvectors and methods with the

largest impact on eigenvalue retrievals, and thus,reveals those requiring additional attention by thecommunity. If we assume that the goal of any SAAis the best simultaneous retrieval of bbp�λ�, a�λ�,adg�λ�, and aϕ�λ� (e.g., minimizing ΔIOP), the choiceof Sbp appears to be the least critical (Table 5). In con-trast, given the robust ability of SAAs to retrievebbp�λ� and a�λ�, and their reduced capacity to decon-volve total absorption into adg�λ� and aϕ�λ�, the choiceof Sdg and a�

ϕ�λ� appears to be highly significant.Using ΔIOP as the performance metric, the use ofdynamically evolving Sdg and a�

ϕ�λ� produced supe-rior results to the use of static eigenvectors (Table 5).Despite this, spectral dependencies in the retrievedeigenvalues revealed in the match-up results(Tables 2 and 3) indicates compensation for imperfecteigenvectors. Elevated Δadg and Δaϕ relative to Δahighlights the challenge of using SAA approachesin isolation to provide information on phytoplankton

Table 4. Delta Statistics for Various Trophic Levelsa

All

NOMAD IOCCG

N% Ntotal Med SIQR N% Ntotal Med SIQR

ΔRrs 90 964 1.68 1.05 90 500 1.04 0.49Δbbp 97 223 26.94 15.64 87 500 8.52 6.80Δa 88 735 27.75 17.28 87 500 8.56 5.90Δadg 90 735 51.02 28.38 87 500 27.25 20.19Δaϕ 87 780 29.32 19.17 84 500 35.83 23.25

Oligotrophic

NOMAD IOCCG

N% Ntotal Med SIQR N% Ntotal Med SIQR

ΔRrs 98 51 1.14 1.38 100 100 0.59 0.32Δbbp 100 1 93.27 NA 100 100 5.61 3.00Δa 94 37 53.56 28.94 100 100 6.26 2.60Δadg 97 37 99.83 36.75 100 100 28.38 16.06Δaϕ 93 43 20.67 11.75 100 100 22.81 11.42

Mesotrophic

NOMAD IOCCG

N% Ntotal Med SIQR N% Ntotal Med SIQR

ΔRrs 90 477 1.56 0.76 100 125 1.22 0.37Δbbp 97 127 33.65 15.79 100 125 6.51 3.84Δa 93 326 25.71 16.23 100 125 6.59 3.22Δadg 96 326 46.53 28.81 100 125 30.72 22.95Δaϕ 94 353 23.93 13.22 100 125 31.44 19.20

Eutrophic

NOMAD IOCCG

N% Ntotal Med SIQR N% Ntotal Med SIQR

ΔRrs 88 436 1.91 1.51 82 275 1.29 0.70Δbbp 96 95 19.90 12.06 77 275 13.31 11.04Δa 82 372 28.25 17.31 76 275 14.56 8.77Δadg 84 372 48.47 25.94 77 275 25.00 21.75Δaϕ 80 384 40.27 30.36 71 275 54.49 34.19aN% provides the percentage of valid retrievals returned by GIOP-DC calculated as 100% �Nvalid∕Ntotal, where Nvalid is the

number of valid retrievals (not shown). Ntotal, Med and SIQR provide the distribution sample size, median and semi-interquartile range, respectively.

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2027

Page 10: Generalized ocean color inversion model for retrieving

community structure [for example, when multiplea�ϕ�λ� are employed to retrieve eigenvalues for multi-

ple phytoplankton groups]. Each successive de-composition of an IOP into component IOPs addsvariability and ambiguity.

The selection of G�λ�, and Rrs�λ�-IOP closurein general, merits additional consideration by theresearch community. Ambiguity in the retrieval ofIOPs from Rrs�λ� remains a known issue (that is be-yond the scope of this work) and combinations ofsuccessfully retrieved eigenvalues may not makegeophysical sense, despite numerical closure in theinversion [42,43]. That aside, the choice of G�λ� fromGordon et al. [21] versus Morel et al. [22] resulted ineigenvalues that differed by 6%–14% (Table 5). Anunexpected relationship between G�λ�, aϕ�λ�, andCa emerged when evaluating these results (Fig. 8).The use of the Gordon quadratic expression [G1�λ� �0.0949 and G2�λ� � 0.0794] resulted in aϕ�443� withlittle bias across the full dynamic range of NOMADstations (see the red best-fit line with a slope nearunity in panel B), but Ca with a clear trophic bias(see the red best-fit line with a slope less than unityin panel A). In contrast, the look-up-table (LUT) ap-proach of Morel [G1�λ� from the LUT and G2�λ� � 0]

produced Ca with little bias (panel C), but aϕ�443�with a definitive slope greater than unity (panel D).Such variability in derived optical and biogeochem-ical products calls into question the ability of existingapproaches to universally estimate G�λ� and relateaϕ�λ� to Ca within the SAA paradigm [39,44]. Withregard to G�λ�, the authors acknowledge the limita-tions in their approaches. The quadratic coefficientsfrom Gordon are valid for open ocean conditions andsolar zenith angles ≥20°. The LUTs from Morel arevalid for conditions where only phytoplankton andtheir derivative products shape the marine lightfield. Furthermore, navigating these LUTs requiresan estimate of Ca, a remotely sensed data productknown to be valid in the open ocean, but less soin coastal or optically complex conditions. Severalalternative methods for estimating G�λ� in opticallycomplex environments now exist [7,23]. While notconsidered in this study, GIOP provides a frameworkfor their systematic comparision in subsequent stud-ies. With regard to estimating phytoplankton bio-mass, a paucity of comprehensive, global data setswith which to obtain robust relationships confoundsour ability to universally toggle between aϕ�λ� and Cain this paradigm. As Ca and aϕ�443� remain propor-tionally constant in GIOP-DC [aϕ�443� � 0.055 Ca],we expect much of the variability shown in Fig. 8to stem from variability in the in situmeasurements.However, repeating this analysis without normal-izing a�

ϕ�443� to 0.055 m2 mg−1 further amplifiedthe disconnect between G�λ�, aϕ�λ�, and Ca, par-ticularly forCa > 1 mg m3 (results not shown). With-out evolving to more sophisticated approaches(explored below), an end-user may have to decidebetween tuning an SAA to optics [aϕ�λ�; our choice]or biogeochemistry (Ca) (Fig. 8), both of which havemerit depending on the science question(s) underconsideration.

Clearly, the use of fixed eigenvectors within anSAA (or dynamic eigenvectors developed using lim-ited data sets) cannot capture their natural variationacross different optical and biogeochemical condi-tions. Optical properties of the oceans vary over or-ders of magnitude globally and remain neitherspatially uniform nor constant in time [45–49]. Novelmethods now exist to constrain regional variabilityin optical parameters, several of which port easilyinto the GIOP framework. OWT classification ap-proaches provide one avenue for capturing spatialand temporal variations in optical parameters frommeasured Rrs�λ� [45,49]. The approach of Moore et al.[45], for example, assigns individual satellite pixelsto specific OWTs and subsequently selects OWT-specific eigenvectors or algorithms. GIOP providestwo paths to estimate the final eigenvalues for agiven pixel: (1) the selected parameterizations areblended using OWT-weighted fuzzy membershipfunctions and used to construct a single executableSAA instance; or (2) the selected parameterizationsare used to construct multiple executable SAAinstances and the resulting eigenvalues are blended

0 2 4 6 8 100

5

10

15

20

Fre

quen

cy (

%)

0 2 4 6 8 100

5

10

15

20F

requ

ency

(%

)

0 2 4 6 8 100

5

10

15

20

Fre

quen

cy (

%)

0 2 4 6 8 10∆Rrs (%)

0

5

10

15

20

Fre

quen

cy (

%)

All

Oligotrophic

Mesotrophic

Eutrophic

Fig. 3. Frequency distributions of ΔRrs from NOMAD (black) andthe IOCCG data set (red) for the all available data (“All”) and sub-setted into three trophic levels. The main text provides definitionsfor the oligo-, meso-, and eutrophic subsets. See Table 4 for accom-panying statistics.

2028 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 11: Generalized ocean color inversion model for retrieving

using OWT-weighted fuzzy membership functions.Ensemble (or, bootstrapping) approaches provideanother vehicle to allow for spatial and temporalvariations in optical parameters [9,39]. The approach

of Wang et al. [9], for example, successively iterateson a single satellite pixel using ranges of eigenvec-tors (e.g., Sdg from 0.01 to 0.02nm−1, and Sbp from−2 to 0 at assigned intervals) and outputs themedianeigenvalues (and their standard deviations) for allvalid retrievals. GIOP provides a mechanism toexecute the successive iterations. Both series ofmethods remain unconstrained by geographic boun-daries and can be easily updated to accommodateadditional OWT-specific parameterizations as addi-tional in situ data become available.

We must consider the data sets employed inthis investigation to fully interpret our results.The elevated (>1) and reduced (<1) ratios of bbp�λ�and a�λ� suggest incomplete Rrs�λ�-IOP closure inNOMAD (Table 2). The Rrs�λ�, bbp�λ�, and a�λ� werecollected using different instruments and methods,and uncertainties associated with varied measure-ments and methodologies are neither negligiblenor spectrally independent [32]. The divergent ratiosof adg�λ� and aϕ�λ� with increasing wavelength likelyresult from low Rrs�λ� and absorption signals at longwavelengths and imperfect parameterization of Sdgand Sbp, both of which control the rate of declineof their eigenvectors. None of the data sets we consid-ered perfectly encapsulate all bio-optical conditionsat all times, nor do they represent the relative globaldistribution of such conditions. NOMAD remainsunevenly distributed among trophic levels, showingdominance in meso- and eutrophic waters, with spa-tial and temporal biases [32]. The SeaWiFS andMODISA match-ups results are equally biased.The IOCCG dataset achieves optical closure as itwas synthesized using a forward model with simpleoptical relationships (e.g., Raman and fluorescenceeffects were not included), but suffers similarly inhow it numerically represents the relative spatialdistribution of coincident bio-optical properties.Likewise, no single validation method or statisticadequately captures the full performance of anSAA. In practice, the evaluation of any algorithm

Table 5. Delta Statistics for the Sensitivity Analysesa

MPD Median

Run N bbp a adg aϕ ΔRrs Δbbp Δa Δadg Δaϕ

GIOP-DC 437 NA NA NA NA 1.04 8.52 8.56 27.25 35.83Sbp − 33% 440 5.19 5.17 7.58 2.98 0.99 11.23 11.70 32.14 34.69Sbp � 33% 436 5.65 5.70 8.82 2.90 1.14 11.40 10.70 23.51 39.12Sdg − 33% 448 18.96 33.44 101.73 46.59 1.61 16.27 19.08 32.94 31.95Sdg � 33% 399 3.77 8.41 40.10 32.92 1.23 9.44 8.95 79.90 59.32Sdg from [7] 439 3.20 5.33 20.40 14.58 1.10 8.65 9.80 22.25 34.42Ca − 33% in [14] 419 2.02 2.92 1.48 7.25 1.19 8.79 8.83 28.62 31.10Ca � 33% in [14] 437 1.56 2.28 1.14 5.90 1.10 8.12 9.17 26.79 40.09Fixed Ca in [14] 369 4.57 7.89 2.60 21.68 1.46 11.30 11.53 30.97 26.70a�ϕ from [17] 357 8.33 12.72 7.04 22.23 1.20 14.26 16.75 38.30 23.13

G from [22] 422 9.99 6.15 7.49 14.12 1.16 11.50 13.64 37.49 36.24Matrix inversion 475 4.60 3.68 2.24 7.41 1.73 9.15 9.43 24.79 36.82400 ≤ λ ≤ 600 nm 424 0.23 0.21 0.08 0.38 0.92 8.76 8.78 31.94 36.55aN is the sample size. MPD is the average spectral median percent difference between GIOP-DC and each alternate run, as calculated

in Tables 2 and 3. Medians of the ΔIOP frequency distributions are also presented, as presented in Table 4.

0 20 40 60 80 100∆bbp (%)

02468

1012

Fre

quen

cy (

%)

0 20 40 60 80 100∆a (%)

0

5

10

15

Fre

quen

cy (

%)

0 20 40 60 80 100∆adg (%)

0

2

4

6

8

Fre

quen

cy (

%)

0 20 40 60 80 100∆aφ (%)

0

2

4

6

8

Fre

quen

cy (

%)

Fig. 4. Frequency distributions of ΔIOP from NOMAD (black)and the IOCCG data set (red) for all available data. See Table 4for accompanying statistics.

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2029

Page 12: Generalized ocean color inversion model for retrieving

must consider the science question(s) to be ad-dressed, which subsequently leads to the definitionof validation metrics. Inevitably, trade-offs emerge,such as: (1) the value in retrieving IOPs highly ac-curately at a single wavelength and poorly at otherwavelengths versus moderately accurately at allwavelengths; (2) the value of retrieving IOPs withhigh accuracy and poor spatial and temporal cover-age versus moderately accurately with high spatialand temporal coverage; and (3) the value of retriev-ing some products with high accuracy [say, bbp�λ�]

and others with less accuracy [say, adg�λ�] versusall products [say, both bbp�λ� and adg�λ�] with mod-erate accuracy. Choices made in performance met-rics ultimately affect the use and interpretationof cumulative statistics. Interpreting ΔRrs, forexample, provides little value for SAAs designedto accurately retrieve IOPs at single wavelengthsat the expense of quality at other wavelengths(Fig. 9).

Our choices in cost function for the inversion andgoodness-of-fit metrics for the validation effort affect

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

ncorrelation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

b bp(

412)

a(41

2)a d

g(41

2)a φ

(412

)

Fig. 5. Taylor and Target diagrams for IOPs at 412 nm from the IOCCG data set for the 12 alternate parameterizations of GIOP com-pared to GIOP-DC. uRMSD is the unbiased root mean square difference. Symbols indicate the following: blue cross � Sdg − 33%(� 0.012 nm−1); red cross � Sdg � 33% (� 0.024 nm−1); green circle � Sdg dynamically calculated using Lee et al. [7]; blue square �Sbp from Lee et al. [7] −33%; black square � Sbp from Lee et al. [7] �33%; red circle � OC-derived Ca − 33% prior to input into Bricaudet al. [14]; black circle � OC-derived Ca � 33% prior to input into Bricaud et al. [14]; green square � a�

ϕ�λ� from Bricaud et al. [14] with Ca

fixed at 0.18 mgm−3; blue circle � a�ϕ�λ� from Ciotti and Bricaud [17] with a size fraction of 0.5; black cross � G�λ� from Morel et al. [22];

orange cross � optimization using linear matrix inversion; and green cross � optimization considering only 400 ≤ λ ≤ 600 nm.

2030 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 13: Generalized ocean color inversion model for retrieving

full interpretation of our results. The χ2 cost function[Eq. (11)] uses absolute values and considers wave-lengths from 400 to 700 nm. The ΔRrs [Eq. (10)] andΔIOP [Eq. (13)] metrics report relative values (%)and consider only wavelengths from 400 to 600 nm.We knowingly adopted this inconsistency for thiswork. In red wavelengths (>600 nm), open oceanRrs�λ� and IOPs can be sufficiently low, such that rel-ative (fractional) differences are large, but absolutedifferences are small. Conversely, in blue wave-lengths, relative differences are small, while absolutedifferences are large. We recommend that the com-bined consideration of absolute and relative spectraluncertainties be adopted in future studies. Criteriafor the cost function and ΔRrs, for example, could bebased on maxima of assigned absolute values andrelative fractions of Rrs�λ� for all wavelengths. Doing

so would not only ensure consistency between the costfunction and the validation goodness-of-fit metrics,but also eliminate the need to constrain the latterto 400–600 nm.

The choice of input uncertainties for Rrs�λ� affectsboth SAA performance and the derived uncertaintieson IOP retrievals. Reliable uncertainties forsatellite-derived Rrs�λ� remain difficult to accuratelydetermine. The use of signal-to-noise ratio (SNR)estimates for the satellite instrument and of statis-tical measures based on agreement between satelliteretrievals and in situ measurements have both beenexplored [50–52]. However, these approaches arerestricted to the assignment of a single, global uncer-tainty estimate to each wavelength (even the latterapproach, due to the limited availability and geo-graphic distribution of such match-ups). In practice,

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

ncorrelation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

b bp(

443)

a(44

3)a d

g(44

3)a φ

(443

)

Fig. 6. As in Fig. 5, but for IOPs at 443 nm.

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2031

Page 14: Generalized ocean color inversion model for retrieving

errors in the atmospheric correction process andchanges in the instrument radiometric performanceover time complicate uncertainties in satellite Rrs�λ�.Remote measurements collected at large viewingangles (large atmospheric path lengths) or throughelevated aerosol loads, where the water-leaving sig-nal is a much smaller portion of the total observedradiance at the sensor, have higher Rrs�λ� uncertain-ties than those collected through shorter atmosphericpaths or clear atmospheres dominated by simpleRayleigh scattering. Turbid or highly reflectivewaters require additional corrections to separate theatmospheric signal from the water signal, which addsadditional assumptions and inherent uncertainties[53]. Similarly, Rrs�λ� retrievals obtained in the vicin-ity of land, clouds, or sun glint may be contaminatedby stray light or atmospheric adjacency effects.

Finally, the radiometric sensitivity of a satelliteinstrument often degrades over time as the satelliteages, leading to decreased SNRs and increasedinstrumental uncertainties [54]. In practice, uncer-tainties for Rrs�λ� vary spatially and temporally, andthis variability includes changes in both their magni-tude and spectral shape.

The primary impact of uncertainties for Rrs�λ� onthe IOP model optimization process, as implementedfor GIOP, is to change the spectral weighting of the χ2minimization [Eq. (11)]. For example, if the assignedvariation does not represent the true relative uncer-tainty at each wavelength, the minimization processwill be skewed to improperly favor specific wave-lengths and ignore others, and the resulting modelwill not necessarily reproduce the spectral depend-ence of Rrs�λ� to within their assigned uncertainties.

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

ncorrelation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.00.20.40.60.81.01.21.4

norm

aliz

ed s

tand

ard

devi

atio

n

correlation

0.200.40

0.60

0.80

0.900.95

0.99

−2 −1 0 1 2normalized uRMSD

−2

−1

0

1

2

norm

aliz

ed b

ias

b bp(

555)

a(55

5)a d

g(55

5)a φ

(555

)

Fig. 7. As in Fig. 5, but for IOPs at 555 nm.

2032 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 15: Generalized ocean color inversion model for retrieving

Spectrally constant error in the magnitudes of uncer-tainties for Rrs�λ� will have no impact on the fittingresults. The derived IOP uncertainties, however,are directly proportional to the magnitude of theassigned uncertainties for Rrs�λ�.

Currently, GIOP-DC does not assign uncertaintiesto Rrs�λ� and uses the variance–covariance matrix toderive the IOP uncertainties. This solution effec-tively assumes a spectrally flat distribution for Rrs�λ�uncertainties. In practice, such an unweighted ap-proach yields very similar IOP retrieval results aswould using global mean uncertainties for Rrs�λ�from match-up analyses, as the latter tends towardrelatively flat spectral dependence in the blue–greenregime [51,52]. The use of the variance–covariancematrix likely understates the IOP retrieval uncer-tainties, since it effectively uses spectral agreementbetween the model and the measurements to esti-mate inherent variability in the Rrs�λ� measure-ments and, thus, ignores spectrally independenterror in Rrs�λ�. The community ultimately needs areliable method for assigning Rrs�λ� uncertaintiesto each remote sensing observation based on a de-tailed error budget of the instrument calibrationand atmospheric correction process. Characterizingthe Rrs�λ� match-up statistics based on viewinggeometry, aerosol optical thickness, turbidity, orother relevant variables, within the limitations ofthe available satellite-to-in situ match-up database,provides an intermediate option [55,56]. Until we ac-quire improved knowledge of Rrs�λ� uncertainties,the assumption of global uncertainties or the useof an unweighted optimization provide viableoptions.

5. Conclusions and Future Directions

GIOP provides a consolidated software environmentfor developing, applying, and evaluating ocean colorSAAs to retrieve marine IOPs. While our primarycontribution remains the GIOP framework itself,its community-recommended default configuration,

0.01 0.10 1.00 10.00 100.00

NOMAD Ca (mg m−3)

0.01

0.10

1.00

10.00

100.00

GIO

P−

DC

Ca

(mg

m−

3 )

A0.01 0.10 1.00 10.00 100.00

NOMAD Ca (mg m−3)

0.01

0.10

1.00

10.00

100.00

C

0.001 0.010 0.100 1.000

NOMAD aφ(443) (m−1)

0.001

0.010

0.100

1.000

GIO

P−

DC

aφ(

443)

(m

−1 )

B

0.001 0.010 0.100 1.000

NOMAD aφ(443) (m−1)

0.001

0.010

0.100

1.000

D

Gordon et al. (1988) Morel et al. (2002)

Fig. 8. Comparison of GIOP-DC and in situmeasurements of Ca (panels A and C) and aϕ�443� (panels B and D) from NOMAD using G�λ�parameterized by Gordon et al. [21] (panels A and B) andMorel et al. [22] (panels C and D). The black line indicates a 1∶1 relationship. Thered line indicates the best fit. The slopes of the best fit lines are 0.84, 1.06, 0.91, and 1.16 for panels A–D, respectively.

Fig. 9. MODISAΔRrs for GIOP-DC (panel A) and GSM (run usingGIOP; panel B). GIOP was applied to the monthly MODISA level-3bin file for March 2010. Units are nondimensional (0.1 � 10%).

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2033

Page 16: Generalized ocean color inversion model for retrieving

GIOP-DC, produces global IOPs of comparable qual-ity to other common algorithms [41]. As GIOP caneasily accommodate new eigenvectors and advancedapproaches as the research community evolves, weanticipate and recommend that GIOP-DC be up-dated routinely. Several features recently included,or in the queue for inclusion, into GIOP include:(1) temperature and salinity dependent aw�λ� andbbw�λ� [29]; (2) alternate mathematical inversionapproaches; (3) alternate eigenvector parameteriza-tions; (4) alternate Rrs�λ�-IOP relationships [e.g.,G�λ� from Lee [23]]; (5) consideration of Raman in-elastic scattering; (6) ensemble solution methods,such as Wang et al. [9] and Brando et al. [39]; and(7) dynamic configuration based on detected OWTs,such as Vantrepotte et al. [49] and Moore et al. [45](provisional parameters for which were recently de-rived). We hope our sensitivity analyses and sub-sequent discussion will provide the communityguidance on future directions for in situ data collec-tion and algorithm refinement to support advancingSAAs (through GIOP) and their application. We alsoexpect the GIOP framework to facilitate analyses as-sociated with new mission planning. Its inherentability to operate on any array of wavelengths, for ex-ample, provides a resource for identifying new chan-nels to be added to forthcoming satellite instruments(e.g., ultraviolet bands).

Appendix A

In practice, most common, published SAAs fall intothree broad classes, hereafter referred to as spectraloptimization, spectral deconvolution, and bulk inver-sion. In this Appendix, we briefly introduce eachclass with attention to the interoperability of theGIOP framework and several widely use SAAs fromeach class. Note, this classification includes so-calledinversion algorithms [those that derive IOPs fromRrs�λ� via inverse solutions to Eqs. (2) and (3)] anddoes not explicitly consider empirical (statistical)or neural-network approaches.

SAAs in the spectral optimization class operate inthe manner described for GIOP in Section 2B. Thatis, eigenvectors are predefined [e.g., for b�bp�λ�, a�

dg�λ�,and a�

ϕ�λ�] and simultaneous solutions for theeigenvalues (e.g., for Bbp, Adg, and Aϕ) are achievedvia linear (matrix) or nonlinear (least squares) opti-mization of Eq. (9). The system is overdetermined ifNλ exceeds the number of unknowns. Examples in-clude the SAAs described in Roesler and Perry [3],Hoge and Lyon [4], Garver and Siegel [5], Maritorenaet al. [8], Wang et al. [9], and Devred et al. [10]. TheseSAAs predominantly differ in their choice of eigen-vectors and inversionmethod and, in principle, GIOPcan be configured to mimic each of them. For exam-ple, the Garver–Siegel–Maritorena (GSM) algorithmcan be executed within the GIOP framework byassigning user-defined Sbp, Sdg, and a�

ϕ�λ� fromMaritorena et al. [8], G�λ� from Gordon et al. [21],and LM optimization. IOPs derived using GSMand GIOP with a GSM-like configuration compared

extremely well for the NOMAD, IOCCG, and match-up data sets (results not shown). Validation resultsfor GIOP-DC and GSM also compared favorably withMPD for GSM-derived bbp�443�, a�443�, adg�443�,and aϕ�443� at 22%, 27%, 29%, and 40% for NOMADand 22%, 26%, 20%, and 52% for the IOCCG data set(Figs. 10–12; see Table 2 for equivalent GIOP-DC sta-tistics). The nuances in quality assurance metrics,success and failure conditions, and fail-safe behaviorthat accompany each SAA listed above, however,may not be currently available within the GIOPframework.

SAAs in the spectral deconvolution class simi-larly assign eigenvectors, but operate in a step-wisefashion to determine the spectral backscattering andabsorption coefficients, rather than optimizing si-multaneous solutions of the eigenvalues. Examplesinclude the SAAs described in Lee et al. [7] [thequasi-analytic algorithm (QAA)], Smyth et al. [11],and Pinkerton et al. [57]. Broadly speaking, SAAsin this class operate via the following steps:

(1) Assign Sbp, Sdg, and ϵϕ [a partial eigenvectorfor a�

ϕ�λ� defined as a�ϕ�412�∕a�

ϕ�443�].(2) Estimate bbp�λ0�, where λ0 is typically a green

wavelength.(3) Calculate bbp�λ� as the product of bbp�λ0�

and Eq. (8) (requires Sbp).

Fig. 10. MODISA bbp�443� for GIOP-DC (panel A), GSM (run us-ing GIOP; panel B), and QAA (panel C). The algorithms were ap-plied to the monthly MODISA level-3 bin file for March 2010.Units are m−1.

2034 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 17: Generalized ocean color inversion model for retrieving

(4) Calculate a�λ� using bbp�λ� and Eq. (2)[requires G�λ�].

(5) Estimate adg�λ0�, where λ0 is typically a bluewavelength (requires Sdg and ϵϕ).

(6) Calculate adg�λ� as the product of adg�λ0� andEq. (5) (requires Sdg).

(7) Calculate aϕ�λ� as a�λ� − aw�λ� − adg�λ�.Note that bbp�λ0� and adg�λ0� are equivalent to Bbpand Adg, respectively, in Eq. (9). These SAAs differin their assignment of Sbp, Sdg, and ϵϕ, and in theirtreatment of steps 2 and 4. A complete review of thedifferences exceeds the scope of this paper, however,several merit mentioning. Both Smyth et al. [11] andPinkerton et al. [57] assign constant values for Sbp,Sdg, and ϵϕ, whereas Lee et al. [7] dynamically esti-mates all three using empirical relationships basedon Rrs�λ�. GIOP supports the Lee et al. [7] estimatesof Sbp and Sdg, with the former included as part ofGIOP-DC (Table 1). Both Smyth et al. [11] andPinkerton et al. [57] adopt iterative approaches toderiving bbp�λ0� (step 2) and G�λ� (step 4). Using aLUT for G�λ� that is keyed on environmental geom-etries and absorption and scattering coefficients,both SAAs iterate until either the selectedG�λ� or de-rived a�λ� stabilize. In contrast, Lee [7] adopts G�λ�from Gordon et al. [21] with modified coefficientsand dynamically estimates a�λ0� using an empiricalrelationship based on Rrs�λ�, which is in turn used toderive bbp�λ0� via rearrangement of Eqs. (2) and (3).Note that SAAs in this class can be halted at step 4

to enable testing or application of alternate ap-proaches to decompose a�λ� into its component parts(e.g., [17] and [58]). At this time, GIOP does not sup-port the step-wise deconvolution approach typicalof this SAA class. However, validation results forGIOP-DC and QAA compared favorably with MPDfor QAA-derived bbp�443�, a�443�, adg�443�, andaϕ�443� at 38%, 21%, 30%, and 28% for NOMADand 16%, 14%, 19%, and 61% for the IOCCG dataset (Figs. 10–12; see Table 2 for equivalent GIOP-DC statistics). Per their design, SAAs in this classalways report ΔRrs � 0, unlike SAAs in the spectraloptimization class.

SAAs in the bulk inversion class do not assign ei-genvectors, that is, they do not predefine spectralshapes for the absorption or scattering coefficients.The approach introduced in Loisel and Stramski(LAS) [6] provides a widely used example. Briefly,LAS exploits a relationship between the diffuse at-tenuation coefficient for downwelling irradiance,Kd�λ� (m−1), solar zenith angle, and the absorptionand scattering coefficients (e.g., [59]). As such, it re-quires the remote estimation of Kd�λ� from Rrs�λ�.LAS sequentially estimates bbp�λi� and a�λi� at eachwavelength λi using Kd�λi�, environmental geom-etries, and LUTs derived from radiative transfer(Monte Carlo) simulations. By estimating IOPs ateach wavelength independently, spectral shape func-tions, such as Sbp, can be calculated dynamicallyand considered output products [60]. GIOP supportsthe use of LAS-derived Sbp and (tabulated) b�bp�λ� as

Fig. 11. As in Figure 10, but for adg�443�. Units are m−1. Fig. 12. As in Figure 10, but for aϕ�443�. Units are m−1.

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2035

Page 18: Generalized ocean color inversion model for retrieving

assignable eigenvectors (Table 1). GIOP does notsupport, however, the spectrally independent inver-sion approach typical of this SAA class.

We thank Mike Behrenfeld, Paula Bontempi,Catherine Brown, Yannick Huot, Paul Lyon,Constant Mazeran, and Jill Schwarz for their helpfuladvice and participation in the NASA GIOP work-shops. We also thank an anonymous reviewer foruseful comments that improved this manuscript.Support for this work was provided through theNASA MODIS Science Team (P.J.W., B.A.F., S.W.B.)and the CSIRO Wealth from Oceans Flagship (V.B.).

References1. IOCCG, “Remote sensing of inherent optical properties:

fundamentals, tests of algorithms, and applications,” Reportsof the International Ocean-Colour Coordinating Group No. 5(IOCCG, 2006).

2. P. J. Werdell, “Global bio-optical algorithms for ocean colorsatellite applications,” EOS Trans. AGU 90, 4 (2009).

3. C. S. Roesler and M. J. Perry, “In situ phytoplankton absorp-tion, fluorescence emission, and particulate backscatteringspectra determined from reflectance,” J. Geophys. Res. 100,13279–13294 (1995).

4. F. E. Hoge and P. E. Lyon, “Satellite retrieval of inherentoptical properties by linear matrix inversion of oceanic radi-ance models: an analysis of model and radiance measurementerrors,” J. Geophys. Res. 101, 16631–16648 (1996).

5. S. A. Garver and D. A. Siegel, “Inherent optical propertyinversion of ocean color spectra and its biogeochemical inter-pretation 1. Time series from the Sargasso Sea,” J. Geophys.Res. 102, 18607–18625 (1997).

6. H. Loisel and D. Stramski, “Estimation of the inherent opticalproperties of natural waters from the irradiance attenuationcoefficient and reflectance in the presence of Raman scatter-ing,” Appl. Opt. 39, 3001–3011 (2000).

7. Z.-P. Lee, K. L. Carder, and R. Arnone, “Deriving inherentoptical properties from water color: a multi-band quasi-analytical algorithm for optically deep waters,” Appl. Opt.41, 5755–5772 (2002).

8. S. Maritorena, D. A. Siegel, and A. Peterson, “Optimization ofa semi-analytic ocean color model for global scale applica-tions,” Appl. Opt. 41, 2705–2714 (2002).

9. P. Wang, E. Boss, and C. S. Roesler, “Uncertainties of inherentoptical properties obtained from semi-analytical inversions ofocean color,” Appl. Opt. 44, 4074–4085 (2005).

10. E. Devred, S. Sathyendranath, and T. Platt, “Inversion basedon a semi-analytical reflectance model,” in Reports of theInternational Ocean-Colour Coordinating Group No. 5, Z.-P.Lee, ed. (IOCCG, 2006), pp. 87–94.

11. T. J. Smyth, G. F. Moore, T. Hirata, and J. Aiken, “Semiana-lytical model for the derivation of ocean color inherent opticalproperties: description, implementation, and performanceassessment,” Appl. Opt. 45, 8116–8131 (2006).

12. C. R. McClain, G. C. Feldman, and S. B. Hooker, “An overviewof the SeaWiFS Project and strategies for producing a climateresearch quality global ocean bio-optical time series,” DeepSea Res. II 51, 5–42 (2004).

13. B. A. Franz and P. J. Werdell, “A generalized framework formodeling of inherent optical properties in ocean remote sens-ing applications,” in Proceedings of Ocean Optics, Anchorage,Alaska, 27 September–1 October 2010, 2010.

14. A. Bricaud, A. Morel, M. Babin, K. Allali, and H. Claustre,“Variations in light absorption by suspended particles withchlorophyll a concentration in oceanic (case 1) waters: analy-sis and implications for bio-optical models,” J. Geophys. Res.103, 31033–31044 (1998).

15. A. M. Ciotti, J. J. Cullen, and M. R. Lewis, “A semi-analyticalmodel of the influence of phytoplankton community structureon the relationship between light attenuation and oceancolor,” J. Geophys. Res. 104, 1559–1578 (1999).

16. A. Morel and S. Maritorena, “Bio-optical properties of oceanicwaters: a reappraisal,” J. Geophys. Res. 106, 7163–7180(2001).

17. A. M. Ciotti and A. Bricaud, “Retrievals of a size parameter forphytoplankton and spectral light absorption by colored detri-tal matter fromwater-leaving radiances at SeaWiFS channelsin a continental shelf region off Brazil,” Limnol. Oceanogr. 4,237–253 (2006).

18. A. Magnuson, L. W. Harding, Jr., M. E. Mallonee, and J. E.Adolf, “Bio-optical model for Chesapeake Bay and the middleAtlantic bight,” Estuar. Coast. Shelf. Sci. 61, 403–424 (2004).

19. B. A. Franz, “l2gen: the Multi-Sensor Level-1 to Level-2Generator,” http://oceancolor.gsfc.nasa.gov/WIKI/OCSSW(2f)l2gen.html (2010).

20. K. Baith, R. Lindsay, G. Fu, and C. R. McClain, “Data analysissystem developed for ocean color satellite sensors,” EOS.Trans. AGU 82, 202 (2001).

21. H. R. Gordon, O. B. Brown, R. H. Evans, J. W. Brown, R. C.Smith, K. S. Baker, and D. K. Clark, “A semianalytic radiancemodel of ocean color,” J. Geophys. Res. 93, 10909–10924(1988).

22. A. Morel, D. Antoine, and B. Gentili, “Bidirectional reflectanceof oceanic waters: accounting for Raman emission and varyingparticle scattering phase function,” Appl. Opt. 41, 6289–6306(2002).

23. Z.-P. Lee, K. Du, K. J. Voss, G. Zibordi, B. Lubac, R. Arnone,and A. Weidemann, “An inherent-optical-property-centeredapproach to correct the angular effects in water-leaving radi-ance,” Appl. Opt. 50, 3155–3167 (2011).

24. C. S. Roesler, M. J. Perry, and K. L. Carder, “Modeling in situphytoplankton absorption from total absorption spectra inproductive inland marine waters,” Limnol. Oceanogr. 34,1510–1523 (1989).

25. A. Bricaud and A. Morel, “Light attenuation and scattering byphytoplanktonic cells: a theoretical modeling,” Appl. Opt. 25,571–580 (1986).

26. D. Doxaran, K. Ruddick, D. McKee, B. Gentili, D. Tailliez, M.Chami, and M. Babin, “Spectral variations of light scatteringby marine particles in coastal waters, from visible to nearinfrared,” Limnol. Oceanogr. 54, 1257–1271 (2009).

27. W. H. Slade, E. Boss, and C. Russo, “Effects of particleaggregation and disaggregation on their inherent opticalproperties,” Opt. Express 19, 7945–7959 (2011).

28. R. M. Pope and E. S. Fry, “Absorption spectrum (380–700 nm)of pure water. II. Integrating cavity measurements,” Appl.Opt. 36, 8710–8723 (1997).

29. X. Zhang, L. Hu, and M.-X. He, “Scattering by pure seawater:effect of salinity,” Opt. Express 17, 5698–5710 (2009).

30. S. Sugihara and M. Kishino, “An algorithm for estimating thewater quality parameters from irradiance just below the seasurface,” J. Geophys. Res. 93, 10857–10862 (1988).

31. J. E. O’Reilly, S. Maritorena, B. G. Mitchell, D. A. Siegel, K. L.Carder, S. A. Garver, M. Kahru, and C. R. McClain, “Oceancolor chlorophyll algorithms for SeaWiFS,” J. Geophys. Res.103, 24937–24953 (1998).

32. P. J. Werdell and S. W. Bailey, “An improved in-situ bio-opticaldata set for ocean color algorithm development and satellitedata product validation,” Remote Sens. Environ. 98, 122–140(2005).

33. S. W. Bailey and P. J. Werdell, “A multi-sensor approach forthe on-orbit validation of ocean color satellite data products,”Remote Sens. Environ 102, 12–23 (2006).

34. B. A. Franz, “NASA ocean color data reprocessing,” http://oceancolor.gsfc.nasa.gov/WIKI/OCReproc.html (2010).

35. P. J. Werdell, S. Bailey, G. Fargion, C. Pietras, K.Knobelspiesse, G. Feldman, and C. McClain, “Unique datarepository facilitates ocean color satellite validation,” EOSTrans. AGU 84, 377 (2003).

36. K. E. Taylor, “Summarizing multiple aspects of model perfor-mance in a single diagram,” J. Geophys. Res. 106, 7183–7192(2001).

37. J. K. Jolliff, J. C. Kindle, I. Shulman, B. Penta, M. A. M.Friedrichs, R. Helber, and R. A. Arnone, “Summary diagramsfor coupled hydrodynamic-ecosystemmodel skill assessment,”J. Mar. Syst. 76, 64–82 (2009).

2036 APPLIED OPTICS / Vol. 52, No. 10 / 1 April 2013

Page 19: Generalized ocean color inversion model for retrieving

38. Z. P. Lee, R. Arnone, C. Hu, P. J. Werdell, and B. Lubac,“Uncertainties of optical parameters and their propagationsin an analytical ocean color inversion algorithm,” Appl.Opt. 49, 369–381 (2010).

39. V. E. Brando, A. G. Dekker, Y. J. Park, and T. Schroeder, “Adap-tive semianalytic inversion of ocean color radiometry in opti-cally complex waters,” Appl. Opt. 51, 2808–2833 (2012).

40. P. J. Werdell, S. W. Bailey, B. A. Franz, L. W. Harding, Jr.,G. C. Feldman, and C. R. McClain, “Regional and seasonalvariability of chlorophyll-a in Chesapeake Bay as observed bySeaWiFS and MODIS-Aqua,” Remote Sens. Environ. 113,1319–1330 (2009).

41. R. J. W. Brewin, S. Sathyendranath, D. Müeller, C.Brockmann, P.-Y. Deschamps, E. Devred, R. Doerffer, N.Fomferra, B. Franz, M. Grant, S. Groom, A. Horseman,C. Hu, H. Krasemann, Z.-P. Lee, S. Maritorena, F. Mélin,M. Peters, T. Platt, P. Regner, T. Smyth, F. Steinmetz, J.Swinton, J. Werdell, and G. N. White III, “The ocean colourclimate change initiative: a round-robin comparison of in-water bio-optical algorithms,” Rem. Sens. Environ. (2012)(to be published).

42. M. Sydor, R. W. Gould, R. A. Arnone, V. I. Haltrin, and W.Goode, “Uniqueness in remote sensing of the inherent opticalproperties of ocean water,” Appl. Opt. 43, 2156–2162 (2004).

43. M. Defoin-Platel and M. Chami, “How ambiguous is the in-verse problem of ocean color in coastal waters?” J. Geophys.Res. 112, C03004 (2007).

44. R. J. W. Brewin, E. Devred, S. Sathyendranath, S. J. Lavender,and N. J. Hardman-Mountford, “Model of phytoplanktonabsorption based on three size classes,” Appl. Opt. 50,4535–4549 (2011).

45. T. S. Moore, J. W. Campbell, and M. D. Dowell, “A class-basedapproach to characterizing and mapping the uncertaintyof the MODIS ocean chlorophyll product,” Remote Sens.Environ. 113, 2424–2430 (2009).

46. H. Loisel, B. Lubac, D. Dessailly, L. Duforet-Gaurier, and V.Vantrepotte, “Effect of inherent optical properties variabilityon the chlorophyll retrieval from ocean color remote sensing:an in situ approach,” Opt. Express 18, 20949–20959 (2010).

47. M. Szeto, P. J. Werdell, T. S. Moore, and J. W. Campbell, “Arethe world’s oceans optically different?,” J. Geophys. Res. 116,C00H04 (2011).

48. M. J. Sauer, C. S. Roesler, P. J. Werdell, and A. Barnard,“Under the hood of satellite empirical chlorophyll-a algo-rithms: revealing the dependencies of maximum band ratio

algorithms on inherent optical properties,” Opt. Express 20,20920–20933 (2012).

49. V. Vantrepotte, H. Loisel, D. Dessailly, and X. Mériaux,“Optical classification of constrasted coastal waters,” RemoteSens. Environ. 123, 306–323 (2012).

50. M. L. Estapa, “Photochemical reactions of particulate organicmatter,” Ph.D. dissertation (University of Maine, 2011).

51. S. Maritorena, O. H. F. d’Andon, A. Mangin, and D. A. Siegel,“Merged satellite ocean color data products using a bio-opticalmodel: characteristics, benefits, and uses,” Rem. Sens. Envi-ron. 114, 1791–1804 (2010).

52. F. Mélin, G. Zibordi, and S. Djavidnia, “Merged series ofnormalized water leaving radiances obtained from multiplesatellite missions for the Mediterranean Sea,” Adv. SpaceRes. 43, 423–437 (2009).

53. S. W. Bailey, B. A. Franz, and P. J. Werdell, “Estimation ofnear-infrared water-leaving reflectance for satellite oceancolor data processing,” Opt. Express 18, 7521–7527 (2010).

54. G. Meister, B. A. Franz, E. J. Kwiatkowska, and C. R.McClain,“Corrections to the calibration of MODIS aqua ocean colorbands derived from SeaWiFS data,” IEEE Trans. Geosci.Remote Sens. 50, 310–319 (2012).

55. F. Mélin, G. Zibordi, and J. F. Berthon, “Assessment of satelliteocean color products at a coastal site,” Remote Sens. Environ.110, 192–215 (2007).

56. G. Zibordi, F. Mélin, and J. F. Berthon, “Intra-annual varia-tions of biases in remote sensing primary ocean color productsat a coastal site,” Remote Sens. Environ. 124, 624–636 (2012).

57. M. H. Pinkerton, G. F. Moore, S. J. Lavender, M. P. Gall, K.Oubelkheir, K. M. Richardson, P. W. Boyd, and J. Aiken, “Amethod for estimating inherent optical properties of NewZealand continental shelf waters from satellite ocean colourmea-surements,” N. Z. J. Mar. Freshwater Res. 40, 227–247 (2006).

58. E. Devred, S. Sathyendranath, V. Stuart, H. Maass, O. Ulloa,and T. Platt, “A two-component model of phytoplanktonabsorption in the open ocean: Theory and applications,”J. Geophys. Res. 111, C03011 (2006).

59. J. T. O. Kirk, “Dependence of relationship between inherentand apparent optical properties of water on solar altitude,”Limnol. Oceanogr. 29, 350–356 (1984).

60. H. Loisel, J.-M. Nicolas, A. Sciandra, D. Stramski, and A.Poteau, “Spectral dependency of optical backscattering bymarine particles from satellite remote sensing of the globalocean,” J. Geophys. Res. 111, C09024 (2006).

1 April 2013 / Vol. 52, No. 10 / APPLIED OPTICS 2037