A meta-analysis of oceanic DMS and DMSP cycling processes...

20
A meta-analysis of oceanic DMS and DMSP cycling processes: Disentangling the summer paradox Martí Galí 1,2 and Rafel Simó 1 1 Department of Marine Biology and Oceanography, Institut de Ciències del Mar (CSIC), Barcelona, Catalonia, Spain, 2 Takuvik Joint International Laboratory, Laval University (Canada) - CNRS (France), Département de biologie et Québec-Océan, Université Laval, Quebec, Quebec, Canada Abstract The biogenic volatile compound dimethylsulde (DMS) is produced in the ocean mainly from the ubiquitous phytoplankton osmolyte dimethylsulfoniopropionate (DMSP). In the upper mixed layer, DMS concentration and the daily averaged solar irradiance are roughly proportional across latitudes and seasons. This translates into a seasonal mismatch between DMS and phytoplankton biomass at low latitudes, termed the DMS summer paradox,which remains difcult to reproduce with biogeochemical models. Here we report on a global meta-analysis of DMSP and DMS cycling processes and their relationship to environmental factors. We show that DMS seasonality reects progressive changes in a short-term dynamic equilibrium, set by the quotient between gross DMS production rates and the sum of biotic and abiotic DMS consumption rate constants. Gross DMS production is the principal driver of DMS seasonality, due to the synergistic increases toward summer in two of its underlying factors: phytoplankton DMSP content (linked to species succession) and short-term community DMSP-to-DMS conversion yields (linked to physiological stress). We also show that particulate DMSP transformations (linked to grazing-induced phytoplankton mortality) generally contribute a larger share of gross DMS production than dissolved-phase DMSP metabolism. The summer paradox is amplied by a decrease in microbial DMS consumption rate constants toward summer. However, this effect is partially compensated by a concomitant increase in abiotic DMS loss rate constants. Besides seasonality, we identify consistent covariation between key sulfur cycling variables and trophic status. These ndings should improve the modeling projections of the main natural source of climatically active atmospheric sulfur. 1. Introduction The unimodal seasonal beat of upper-ocean DMS concentrations, with an annual maximum in summer, is a widespread feature from tropical to polar latitudes [Bates et al., 1987; Simó and Pedrós-Alió, 1999a; Lana et al., 2011]. In seasonally light-limited polar and subpolar latitudes, DMS peaks approximately in synchrony with phytoplankton biomass and with the concentration of its phytoplanktonic precursor, DMSP. As we move toward subtropical latitudes, the summer DMS peak tends to lag that of DMSP by some weeks and that of phytoplankton biomass by some months, thus setting the DMS summer paradox [Simó and Pedrós-Alió, 1999a]. These contrasting seasonal patterns across latitudes are also known as the bloom and the stress regimes for DMS dynamics [Toole and Siegel, 2004]. Global ecosystem models frequently struggle to reproduce DMS seasonality in the lower latitude, stress regime (or summer paradox) areas [Le Clainche et al., 2010; Vogt et al., 2010], which approximately contribute half of the oceanic DMS emission [Lana et al., 2011] despite their low productivity. There is wide consensus that seasonal phytoplankton succession is key in producing the summer paradox by favoring strong DMSP-producing taxa in highly irradiated, stratied, and nutrient-depleted waters [Stefels et al., 2007; Archer et al ., 2009]. Yet this alone cannot explain the temporal decoupling between DMS and its precursor DMSP, neither the larger amplitude of seasonal DMS variation compared to DMSP observed at oligotrophic locations [Dacey et al., 1998; Toole and Siegel, 2004; Vila-Costa et al., 2008]. This requires either an increase of community DMSP-to-DMS conversion yields toward summer, or a decrease of total DMS loss (due to bacterial oxidation, photolysis, and ventilation), or both. These mechanisms are not mutually exclusive and can even act synergistically, and considerable uncertainty remains in the literature as to their relative importance in producing the summer paradox. A further source of uncertainty concerns the contribution of particulateDMS production (generally attributed to phytoplankton) and dissolvedDMS production (generally GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 496 PUBLICATION S Global Biogeochemical Cycles RESEARCH ARTICLE 10.1002/2014GB004940 Key Points: We conducted a comprehensive meta-analysis of DMS(P) cycling processes Gross DMS production rather than DMS loss rate constants drives DMS seasonality Irradiance-driven community DMS yields can explain seasonal DMS-DMSP decoupling Supporting Information: Text S1 Table S1 Figure S1 Figure S2 Figure S3 Figure S4 Figure S5 Correspondence to: M. Galí, [email protected] Citation: Galí, M., and R. Simó (2015), A meta- analysis of oceanic DMS and DMSP cycling processes: Disentangling the summer paradox, Global Biogeochem. Cycles, 29, 496515, doi:10.1002/ 2014GB004940. Received 18 JUL 2014 Accepted 9 MAR 2015 Accepted article online 12 MAR 2015 Published online 28 APR 2015

Transcript of A meta-analysis of oceanic DMS and DMSP cycling processes...

Page 1: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

A meta-analysis of oceanic DMS and DMSP cyclingprocesses: Disentangling the summer paradoxMartí Galí1,2 and Rafel Simó1

1Department of Marine Biology and Oceanography, Institut de Ciències del Mar (CSIC), Barcelona, Catalonia, Spain, 2TakuvikJoint International Laboratory, Laval University (Canada) - CNRS (France), Département de biologie et Québec-Océan,Université Laval, Quebec, Quebec, Canada

Abstract The biogenic volatile compound dimethylsulfide (DMS) is produced in the ocean mainly fromthe ubiquitous phytoplankton osmolyte dimethylsulfoniopropionate (DMSP). In the upper mixed layer,DMS concentration and the daily averaged solar irradiance are roughly proportional across latitudes andseasons. This translates into a seasonal mismatch between DMS and phytoplankton biomass at low latitudes,termed the “DMS summer paradox,” which remains difficult to reproduce with biogeochemical models.Here we report on a global meta-analysis of DMSP and DMS cycling processes and their relationship toenvironmental factors. We show that DMS seasonality reflects progressive changes in a short-term dynamicequilibrium, set by the quotient between gross DMS production rates and the sum of biotic and abiotic DMSconsumption rate constants. Gross DMS production is the principal driver of DMS seasonality, due to thesynergistic increases toward summer in two of its underlying factors: phytoplankton DMSP content (linked tospecies succession) and short-term community DMSP-to-DMS conversion yields (linked to physiological stress).We also show that particulate DMSP transformations (linked to grazing-induced phytoplankton mortality)generally contribute a larger share of gross DMS production than dissolved-phase DMSP metabolism. Thesummer paradox is amplified by a decrease in microbial DMS consumption rate constants toward summer.However, this effect is partially compensated by a concomitant increase in abiotic DMS loss rate constants.Besides seasonality, we identify consistent covariation between key sulfur cycling variables and trophicstatus. These findings should improve the modeling projections of the main natural source of climaticallyactive atmospheric sulfur.

1. Introduction

The unimodal seasonal beat of upper-ocean DMS concentrations, with an annual maximum in summer, is awidespread feature from tropical to polar latitudes [Bates et al., 1987; Simó and Pedrós-Alió, 1999a; Lana et al.,2011]. In seasonally light-limited polar and subpolar latitudes, DMS peaks approximately in synchrony withphytoplankton biomass and with the concentration of its phytoplanktonic precursor, DMSP. As we movetoward subtropical latitudes, the summer DMS peak tends to lag that of DMSP by some weeks and that ofphytoplankton biomass by somemonths, thus setting the DMS summer paradox [Simó and Pedrós-Alió, 1999a].These contrasting seasonal patterns across latitudes are also known as the bloom and the stress regimes forDMS dynamics [Toole and Siegel, 2004]. Global ecosystem models frequently struggle to reproduce DMSseasonality in the lower latitude, stress regime (or summer paradox) areas [Le Clainche et al., 2010; Vogtet al., 2010], which approximately contribute half of the oceanic DMS emission [Lana et al., 2011] despitetheir low productivity.

There is wide consensus that seasonal phytoplankton succession is key in producing the summer paradox byfavoring strong DMSP-producing taxa in highly irradiated, stratified, and nutrient-depleted waters [Stefels et al.,2007; Archer et al., 2009]. Yet this alone cannot explain the temporal decoupling between DMS and its precursorDMSP, neither the larger amplitude of seasonal DMS variation compared to DMSP observed at oligotrophiclocations [Dacey et al., 1998; Toole and Siegel, 2004; Vila-Costa et al., 2008]. This requires either an increase ofcommunity DMSP-to-DMS conversion yields toward summer, or a decrease of total DMS loss (due to bacterialoxidation, photolysis, and ventilation), or both. These mechanisms are not mutually exclusive and can even actsynergistically, and considerable uncertainty remains in the literature as to their relative importance inproducing the summer paradox. A further source of uncertainty concerns the contribution of “particulate”DMS production (generally attributed to phytoplankton) and “dissolved” DMS production (generally

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 496

PUBLICATIONSGlobal Biogeochemical Cycles

RESEARCH ARTICLE10.1002/2014GB004940

Key Points:• We conducted a comprehensivemeta-analysis of DMS(P) cyclingprocesses

• Gross DMS production rather thanDMS loss rate constants drivesDMS seasonality

• Irradiance-driven community DMSyields can explain seasonalDMS-DMSP decoupling

Supporting Information:• Text S1• Table S1• Figure S1• Figure S2• Figure S3• Figure S4• Figure S5

Correspondence to:M. Galí,[email protected]

Citation:Galí, M., and R. Simó (2015), A meta-analysis of oceanic DMS and DMSPcycling processes: Disentangling thesummer paradox, Global Biogeochem.Cycles, 29, 496–515, doi:10.1002/2014GB004940.

Received 18 JUL 2014Accepted 9 MAR 2015Accepted article online 12 MAR 2015Published online 28 APR 2015

Page 2: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

attributed to bacteria) to gross DMS production. While some works estimated that conversion of dissolvedDMSP is of enough magnitude to sustain DMS turnover [e.g., Kiene and Linn, 2000a], other studies showedthat much higher DMS production rates and yields occur associated to particles [e.g., Scarratt et al., 2000a]. Allin all, the magnitude and the seasonal variability of DMS sources and sinks across ocean biomes, and theirunderlying factors, are poorly constrained [Vézina, 2004; Six and Maier-Reimer, 2006; Le Clainche et al., 2010]. Thisuncertainty translates into ecological model parameterizations and ultimately affects the conclusionsreached through numerical experiments and sensitivity analyses [Archer et al., 2002; Vézina, 2004; LeClainche et al., 2010].

The finding of a strong relationship between the daily averaged, vertically integrated irradiance in the uppermixed layer [the so-called solar radiation dose (SRD) index] and DMS concentrations over the global ocean[Vallina and Simó, 2007] emphasized the crucial role of sunlight in driving DMS seasonality. The SRD indexdepends directly on the daily solar irradiance and the water transparency, and inversely on the mixed layerdepth (MLD; see section 3). Therefore, it ultimately reflects the seasonal changes in physical forcing thatcontrol pelagic biological processes through the effects of light, turbulence, and nutrients. The SRD-DMSrelationship has been criticized either on statistical grounds [Derevianko et al., 2009] or based on the fact thatit does not explain mesoscale DMS distribution [Belviso and Caniaux, 2009]. However, SRD remains as thesingle variable explaining most of the seasonal DMS variability over large scales [Lana et al., 2012; Miles et al.,2009] and at specific sites [Vallina and Simó, 2007].

In search of a mechanistic explanation for the summer paradox and the SRD-DMS relationship, we assembled aglobal database of DMSP and DMS cycling rates within the upper mixed layer (UML). The data were statisticallyanalyzed within a common conceptual framework based on a minimal steady state UML DMS budget equation(Figure 1), thus disentangling the seasonal variability of the key DMS production and consumption terms andtheir modulation by environmental variables. Special emphasis was placed on stress regime or summer paradoxareas, defined as those displaying negative seasonal correlation between DMS and chlorophyll a (Chl).

2. Conceptual Framework

The variation of UML DMS concentrations ([DMS]) over time can be expressed by the following budget equation:

d DMS½ �=dt ¼ GPDMS � DMS½ � · kBC þ kphoto þ kvent� �

(1)

where GPDMS is the gross DMS production rate, and kBC, kphoto, and kvent represent the first-order rateconstants of bacterial DMS consumption, DMS photolysis, and DMS ventilation, respectively (see unitsin Table 1). Hereafter, we will call the (kBC + kphoto + kvent) term the total DMS loss rate constant kLOSS.

Figure 1. Scheme of the simplified DMS budget equation that articulates the meta-analysis. DMSO: dimethylsulfoxide; MeSH: methanethiol. DMSO-based DMSproduction is omitted for clarity. The budget terms are described in section 2 and the variables in Table 1. The arrow marked with an asterisk corresponds toparticulate-phase DMS production, which is not explicitly included in the meta-analysis.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 497

Page 3: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

Net turbulent DMS fluxes (horizontal and diapycnal) are, under most circumstances, orders of magnitudesmaller than kLOSS and can be neglected [Gabric et al., 2001; Herrmann et al., 2012; Yang et al., 2013].Significant vertical DMS fluxes have only been reported in situations where nighttime convection erodes asharp sub-pycnocline DMS maximum [Bailey et al., 2008].

GPDMS can be expressed as the product of total DMSP (DMSPt) concentration, the substrate-specific DMSPtconsumption rate constant (kDMSPt), and the community DMS yield, which is the fraction of DMSPt consumedthat is converted into DMS:

GPDMS ¼ DMSPt½ � · kDMSPt · Y t (2)

The DMS production rate resulting from dissolved DMSP (DMSPd) metabolism is expressed with theanalogous equation:

DPDMS ¼ DMSPd½ � · kDMSPd · Yd (3)

where [DMSPd] is operationally defined by the DMSP passing through a pore filter (0.2μm) or a GF/F (0.7μmnominal pore size) depending on the study, kDMSPd is the microbial DMSPd consumption rate constant, andYd is the corresponding dissolved-phase DMS yield. By definition, DPDMS cannot exceed GPDMS in a givenwater sample. Variable units and measurement techniques are summarized in Table 1, and a schematic of theoverall conceptual framework is represented in Figure 1.

[DMSPt] can be expressed as the product of phytoplankton carbon biomass (mol C L�1) and the cellularDMSP quota [mol DMSP-C (mol C)�1]. Substituting these variables, we obtain the following:

GPDMS ¼ Biomass½ � · DMSPt :Biomass½ � · kDMSPt · Y t (4)

Equation (4) splits gross DMS production into four driving factors, separately accounting for (1) phytoplanktondynamics (biomass); (2) mean DMSP content of the phytoplankton assemblage (DMSP quota); (3) thestrength of the trophic coupling between phytoplankton, microzooplankton, and osmotrophic DMSPconsumers (kDMSPt); and (4) the prevalence of DMS-producing pathways among community-level DMSPtransformations (Yt). These factors are usually ranked high in sensitivity analyses of sulfur ecosystemmodels [Vézina, 2004; Vallina et al., 2008; Vogt et al., 2010]. For reasons of data availability, here we willuse Chl to estimate phytoplankton carbon using the relationship of Sathyendranath et al. [2009] andwill report both measured DMSPt:Chl ratios and estimated DMSP-carbon quotas [see review by Stefelset al., 2007].

Table 1. Variables Definition, Units, and Corresponding Methods and Referencesa

Variable Abbreviation Units Method or Definition Reference

DMS productionGross DMS production GPDMS nmol L�1 d�1 Inhibitor (DMDS, MBE) Wolfe and Kiene [1993a]; Simó et al. [2000]DMSPt to Chl DMSPt:Chl nmol S (μg Chl)�1

DMSP-carbon quota CDMSPt:Cphyto mol C (mol C)�1

DMSPt consumption k kDMSPt d�1 Net loss curve Simó et al. [2000]DMSPp consumption k kDMSPp d�1 Dilution experiments Archer et al. [2001]; Saló et al. [2010]DMSPt to DMS conversion yield Yt % 100 · GPDMS/([DMSPt] · kDMSPt) Simó et al. [2000]GPDMS to DMSPt Y*t d�1 GPDMS/[DMSPt] Bailey et al. [2008]GPDMS to Chl YChl nmol S μg Chl�1 d�1 GPDMS/Chl Bailey et al. [2008]GPDMS to phyto carbon GPDMS-C:Cphyto mol C (mol C)�1 d�1 GPDMS/Cphyto This workDMSPd consumption k kDMSPd d�1 35S-DMSPd loss Kiene and Linn [2000b]DMSPd to DMS conversion yield Yd % 35S-DMSPd loss + 35S-DMS trapping Kiene and Linn [2000b]

DMS lossTotal DMS loss k kLOSS d�1 kBC + kphoto + kvent This workBacterial DMS consumption k kBC d�1 35S-DMS loss Kiene and Linn [2000b]

Inhibitor (DMDS, MBE) Simó et al. [2000]DMS photolysis k kphoto d�1 35S-DMS loss Toole et al. [2004]

Bulk Kieber et al. [1996]DMS ventilation k kvent d�1 Flux parameterization See sections 2, 3.1, and 3.4

Eddy covariance

ak refers to first-order rate constant.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 498

Page 4: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

DMSP consumption and DMS loss k values have all d�1 units but represent different underlying kinetics.Biological DMSP and DMS consumption have generally been found to follow Michaelis-Menten kinetics, asexpected in enzymatic reactions [Wolfe and Kiene, 1993a; Ledyard and Dacey, 1996]. The representation ofDMS or DMSP consumption rates with first-order kinetics (rate = [substrate] · k) is a valid simplification as longas, over the budget time scale, [substrate] remains (i) in the more linear portion of the Michaelis-Mentencurve (below the half-saturation concentration) or (ii) close to the initial (“steady state”) concentration at which kwas determined. DMS photolysis is a photosensitized process, whereby absorption of ultraviolet radiation (UVR)by colored dissolved organic matter (CDOM) or nitrate is required to generate the DMS oxidants. In this case,kphoto is a pseudo first-order rate constant that accounts for the concentration of oxidants resulting fromthe interaction between photosensitizers, the light field, and oxidant scavengers [Bouillon andMiller, 2005]. DMSventilation is a surface process, but for simplicity, we assume that the entire UML is exchanging with theatmosphere over the daily time scale. Dividing gas transfer coefficients (md�1) by MLD (m) renders kvent.

3. Methods3.1. DMS and DMSP Cycling Process Database

A database was assembled from published studies reporting: (i) gross DMS production rates (GPDMS); (ii) ratesor rate constants of biological DMS consumption, plus abiotic DMS loss due to photolysis and ventilationwhen available; (iii) dissolved and total DMSP consumption rates and/or rate constants, and/or thecorresponding DMSP-to-DMS conversion yields; and (iv) net DMS change rates in coherent water masses[supporting information Table S1]. Studies reporting only abiotic DMS losses were not included. Unpublisheddata obtained recently by our research group were also included. Since the upper mixed layer is the depthhorizon that regulates ocean-atmosphere exchange, the analysis focused only on UML DMS budgets. Allthe process rates or the corresponding first-order rate constants (k) were expressed as the UML averageover a 1 day period (after vertical propagation of photolysis when needed [e.g., Galí and Simó, 2010]).Only biological rates measured in dark incubations were included in the statistical analysis. Note that formost biological processes, a single daily measurement was assumed to be representative of the entire UML(90–100% of measurements). This assumption is an unavoidable source of uncertainty in our study dueto the possible existence of vertical and temporal variability within the UML. The limitation is partially overcomethrough a thorough assessment of uncertainty in the correlation and regression statistics (section 3.8).

DMS and DMSP cycling rates were matched to simultaneous measurements of environmental variables(temperature, salinity, MLD, nutrient concentrations, light attenuation coefficients, daily irradiance, andChl) when available. When the raw data could not be accessed, we used the G3Data Graph Analyzer opensoftware to digitize the information contained in figures (error <2%). Missing SRD data were filled withsatellite measurements (section 3.5). Missing nitrate data were filled only in open-ocean regions withthe WOA 2009 climatology (https://www.nodc.noaa.gov/cgi-bin/OC5/WOA09/woa09.pl), roughly doublingthe nitrate data available for statistical analyses. The data set covered a wide range of oceanographicregimes (Figure 2a), although temperate latitudes and the summer season were over-represented (supportinginformation Figure S1 and Table S1).

3.2. Net In Situ DMS Change in Lagrangian Studies

Daily net DMS production (d[DMS]/dt) was calculated in those studies where a coherent water mass hadbeen sampled over time, including Lagrangian studies and the sampling conducted on two consecutivedays on a monthly basis during 2003–2004 in the Blanes Bay Microbial Observatory [Vila-Costa et al., 2008].When only one measurement was available per day, d[DMS]/dt was calculated as Δ[DMS]/Δt. If moremeasurements were available, d[DMS]/dt was calculated as the slope of the linear regression between[DMS] and t (days). d[DMS]/dt was divided by [DMS] to obtain kNET (units of d

�1; Figure 2b). Note that thusdefined kNET is methodologically independent from simultaneous incubation-derived rates.

3.3. GPDMS Measurements Using the Inhibitor Technique

This approach takes advantage of the fact that in a dark gas-tight bottle, photochemical and physical DMSremovals are eliminated. If bacterial DMS removal is effectively stopped using a specific inhibitor, thevariation of DMS over time equals GPDMS [equation (1)]. If a dark unamended bottle is incubated in parallel,the variation of DMS over time corresponds to net biological DMS production rates (NPbioDMS) so that bacterial

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 499

Page 5: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

DMS consumption rates can be calculated as BCDMS =GPDMS�NPbioDMS. Bacterial DMS consumption rateconstants can be calculated as kBC = BCDMS / [DMS]. Frequently used inhibitor compounds are chloroformat a concentration of 500 nmol L�1, dimethyldisulfide at 200 nmol L�1, or methyl butyl-ether at 30 μmol L�1

[Wolfe and Kiene, 1993b; Simó et al., 2000]. Chloroform was shown to cause overestimation of GPDMS due toenhanced release of phytoplankton DMSP [Wolfe and Kiene, 1993a; Simó et al., 2000]. Therefore, GPDMS

measured with chloroform inhibition was not included in the meta-analysis. Dark GPDMS rates measuredwith the inhibitor technique generally carry a measurement uncertainty <25% [Vila-Costa et al., 2008;Galí and Simó, 2010; Galí et al., 2011]. Note that GPDMS rates measured using the inhibitor technique account forall DMS-producing processes occurring in dark incubations, not just bacterial metabolism, and also includeeventual DMSO reduction [Asher et al., 2011].

3.4. kLOSS Measurements

The three rate constants that make up kLOSS [equation (1), Table 1] were obtained as follows: kBC was determinedin unfiltered water dark incubations using either the 35S-DMS radiotracer or the inhibitor technique described

Figure 2. Summary of gross DMS production rates (GPDMS) and total DMS loss rate constants (kLOSS = kBC + kphoto + kvent)across contrasting biogeochemical regimes. (a) Geographical distribution of the studies included in the global database(listed in supporting information Table S1). The colors and symbols refer to the variables obtained from each study andthe corresponding methods (Table 1). (b) Histograms of daily net in situ DMS production (kNET, d

�1) in coherent watermasses, and DMS turnover rate constants deduced from gross DMS production (GPDMS/DMS, units of d�1) and total loss(kLOSS). (c) Gross DMS production vs. DMS concentration. (d) Total DMS loss rate constant vs. DMS concentration. Equalturnover rate constant lines (d�1) are shown on the background of Figure 2c (diagonal dotted lines) and d (horizontaldotted lines). Different subsets of kLOSS are distinguished in Figures 2b and 2d according to the method used to measurekBC: inhibitor (DMDS or MBE) or 35S-DMS radiotracer.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 500

Page 6: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

above; kphoto was determined by incubating filtered seawater under full spectrum sunlight and monitoringbulk DMS or 35S-DMS loss; and kvent was calculated as the quotient between sea-air DMS flux (calculatedwith wind-speed flux parameterizations) and MLD. Radiotracer techniques used to measure biologicaland photochemical DMS loss are generally assumed to be more precise than bulk methods, although inthe case of photolysis, they have been shown to be equivalent [Toole et al., 2004]. Potential differencesbetween the inhibitor and the radioisotope methods for measuring kBC [Wolfe and Kiene, 1993a] are the mainsource of uncertainty in kLOSS (see section 4.1). The use of different wind-speed DMS flux parameterizations is aminor source of uncertainty, despite the factor-of-two discrepancies among different models [Lana et al., 2011],because kvent generally contributes a minor portion of kLOSS (Figure 3). If we assign a (generous) fractionaluncertainty of 50% to each of the three k values, we obtain a propagated uncertainty of 36% for kLOSS.

3.5. Solar Radiation Dose (SRD) Estimation

The SRD index was calculated as the vertical integral of the irradiance profile within the UML [Vallina andSimó, 2007], from which the following equation is obtained:

SRD ¼ Ed · Kd · MLDð Þ�1 · 1� e�Kd ·MLD� �

(5)

where Ed is the mean daily shortwave irradiance above the water surface (Wm�2), Kd is the attenuationcoefficient (m�1) of downwelling photosynthetically available radiation (PAR), and MLD is the mixed layerdepth (m). Average irradiance of the prior 48 h was used when possible, except during ship translation wherethe 24 h average was preferred. We used the MLD values provided in each study, which possibly introducedsome noise in the SRD calculation due to the different criteria used to determine MLD. In about one third of

Figure 3. Histograms of the main variables considered in the meta-analysis (Table 1). Black (background): global database. Gray (foreground): stress regime areassubset (see section 3). IQF stands for interquartile factor, a dispersionmetric defined as the quotient between the third and first quartiles (IQF overlaid on the histogramfor stress regime areas). The medians and interquartile ranges are reported in the text. Note the logarithmic x axes.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 501

Page 7: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

the data points used for the statistical analysis, the MLD was available, but either Kd or Ed was missing. Inthese cases, the SRD index had to be estimated, combining measured MLDs with Ed and/or Kd deducedfrom satellite measurements and a bio-optical model, respectively. Satellite measurements (8-day 9 kmresolution L3bin SeaWiFS images; http://oceancolor.gsfc.nasa.gov/) of daily PAR were matched to thedatabase and used to complete the missing Ed measurements. Monthly climatological PAR was used when8-day data were not available. A conversion factor was calculated by comparing 8-day satellite PAR and insitu total shortwave irradiance measurements of the meta-analysis database (which showed reasonableagreement, Pearson’s r = 0.61, p< 10�12) and applied to satellite PAR measurements to ensure internalconsistency in the database. The missing Kd values were estimated using the Chl-based bio-optical modelof Morel et al. [2007].

3.6. Identification of Stress Regime Areas

Stress regime areas were initially identified as those regions comprised between the 45°N and 45°Slatitudes where a negative seasonal correlation existed between monthly DMS and Chl concentrations(supporting information Figure S2). Seasonal correlations (Spearman’s rank correlation) were calculated ona 1° × 1° pixel grid using the 1998–2009 SeaWiFS Chl climatology and the updated DMS climatology,following Lana et al. [2012]. The process studies identified as representative of the stress regime had beenconducted mostly in the subtropical NW Atlantic (Sargasso Sea), the NW Mediterranean Sea, and thesubtropical S Indian Ocean (Figure 2a, supporting information Table S1).

3.7. Monthly Climatology of Subtropical Oligotrophic Gyres

A monthly climatology of Chl, DMSPt, DMS, MLD, and SRD of the more oligotrophic stress regime areas,mostly corresponding to the cores of the subtropical gyres, was created by selecting the 1° × 1° pixelsdisplaying (i) negative DMS vs. Chl seasonal correlation significant at p< 0.10 level and (ii) annual medianChl< 0.25mgm�3 (supporting information Figure S2). Mixed layer depths (MLD) were obtained fromthe MIMOC climatology [Schmidtko et al., 2013] (http://www.pmel.noaa.gov/mimoc/). SRD was calculatedusing the satellite data and bio-optical models specified above. Median monthly DMS, DMSPt Chl, MLD,and SRD were calculated for each selected pixel and month, and the global median and interquartile rangewere finally calculated for each month using all the available monthly pixel medians. Note that DMS andDMSPt data used to compute monthly medians were directly obtained from the public sea-surface DMSdatabase (http://saga.pmel.noaa.gov/dms/) so that the number of monthly pixel medians available ranged34–184 for DMS and 3–41 for DMSPt. When needed, DMSPt was calculated as the sum of particulateand dissolved DMSP in each sample. Only measurements at depths shallower than 10m were used, sincethis is the shallowest MLD in the MIMOC climatology [Schmidtko et al., 2013].

3.8. Statistical Descriptors and Analyses

The variables of the meta-analysis had different probability distributions, generally ranging from roughlynormal to roughly lognormal (Figure 3). We report medians and interquartile ranges (IQR) throughoutand use preferentially non-parametric statistics to avoid the need for assessing statistical distributions.The relationship between environmental and sulfur cycling variables was assessed using thenon-parametric Spearman’s rank correlation coefficient (although the Pearson correlation coefficient wasoccasionally used). Confidence intervals (CI) for the correlation coefficients were calculated bybootstrapping (n= 2000) using the function “bootci” (Matlab® R2013b). Bootstrapped CIs have beenshown to provide more robust assessments of statistical significance than classical hypothesis testingbased on p-values [Erceg-Hurn and Mirosevich, 2008]. As an additional test, the correlations and CIs werere-calculated after contaminating the data with a normally distributed random noise of progressivelyincreasing magnitude (supporting information Table S2). This exercise indicated that most correlationsremained significant at the 99% CI level if a random error of at least 30% was added.

A resampling approach was also taken to estimate the fraction of gross DMS production arising fromdissolved DMSP metabolism (Fd). Fd was calculated as

Fd ¼ f · kDMSPd · Yd= kDMSPt · Y tð Þ (6)

where f= [DMSPd]/[DMSPt]. The “randsample” function (Matlab® R2013b) was used to resample the measuredkDMSPd, Yd, kDMSPt, and Yt. Ten thousand random combinations were used to calculate the probability

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 502

Page 8: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

distribution of Fd under two assumptions: f = 0.05 and f= 0.20. These values were chosen according to Kieneand Slezak [2006, Figure 6] and represent upper limits of dissolved DMSP fractions (f) in Phaeocystis-dominatedRoss Seawaters and picoplankton-dominated summer Sargasso Seawaters, respectively. Thus, the correspondingFd likely represent upper limits too.

Quantile regression [Koenker and Bassett, 1978] was used to determine the conditional distribution of sulfurcycling variables with respect to environmental variables, focusing on SRD. This technique depicts the shiftingdistribution of the y variable with respect to x for a given y–x functional relationship, thereby assessing thetrends in centrality and dispersion statistics simultaneously. Linear conditional quantile functions werecalculated in 5% increments between the 5% and 95% quantiles. The standard error at 95% confidence level ofthe quantile slopes and intercepts was obtained by bootstrapping (n=1000).

3.9. Steady State UML DMS Budget Model Optimization and Skill Metrics

Constrained nonlinear optimization [Glover et al., 2011] was used to explore the predictive power of theminimal steady state DMS budget model. To this end, community DMS yields and kLOSS were expressed asa function of SRD, while DMSPt was prescribed from real data. Three DMS-DMSPt-Chl-SRD time seriesrepresentative of stress regime areas were used for model optimization: the Dacey et al. [1998] study in theSargasso Sea (1992–1994), for which the original biweekly and depth-resolved measurements weremonthly averaged within the UML; the surface DMS(P) measurements conducted with (at least) monthlyfrequency in the Blanes Bay Microbial Observatory (BBMO) between 2003–2004 [Vila-Costa et al., 2008] and2008–2010 [see methods in Galí et al., 2013a]; and a monthly DMS climatology of the oligotrophic gyrecores (section 3.7). We underline that these numerical experiments were designed to complement ourobservations-based meta-analysis. A complete evaluation of all the sulfur cycling processes and theirinteractions in the multidimensional environmental variable space would require the use of fully coupledocean circulation and ecosystem models, which is beyond the scope of this study.

The equation parameters defining the SRD dependence of community DMS yields and kLOSS (supportinginformation Table S3) were iteratively optimized to obtain the best fit between modeled and observed

DMS. The “fmincon” function (Matlab® R2013b) was used to find a constrained (local) minimum of the

cost function, defined as the sum of squared residuals (modeled minus measured DMS). The results of themedian (50% quantile) regression were used as the initial estimate of the model parameters (slopes andintercepts in linear equations), and the parameter bounds were set to the first and third quartiles. Note thatwhen yields and kLOSS are simultaneously optimized (see section 4.5), there are potentially infinite solutionsthat can satisfy the error tolerance threshold. Therefore, inequality constraints were defined for eachparameter as the maximum range spanned by the first and third quartile regression lines within the SRDrange observed in each time series. The default trust-region-reflective algorithm was used and generallyconverged within a few tens of cost function evaluations. The goodness of model-data fits was evaluatedusing different statistics (Table 2): the coefficient of determination (R2), the relative bias (rbias), and therelative (mean-normalized) root mean squared error (rRMSE) [Glover et al., 2011].

4. Results4.1. Macroscale Patterns in DMS Budgets

Gross DMS production rates measured with the inhibitor technique spanned two orders of magnitude,between 0.25 and 42nmol L�1 d�1 (Figure 2c). Total DMS loss rates, calculated as [DMS] · kLOSS, spanned0.19–18nmol L�1d�1 (Figure 2d). DMS turnover rate constants (kLOSS) were generally >0.25d�1 (Figures 2b–2d)or, what is equivalent, turnover times (τ) <4days. The magnitude of τ has important implications: it tells us thatthe rather smooth [DMS] seasonality observed at different locations [Dacey et al., 1998; Vila-Costa et al.,2008; this study] does not result from progressive DMS accumulation or loss over months, but froma dynamic equilibrium between production rates and consumption rate constants at the time scale ofdays. In other words, [DMS] will respond quickly to a perturbation in its controlling variables [equation (1)]due to ecological or physical dynamics. Unfortunately, very few studies exist where sources and sinkswere measured simultaneously (supporting information Table S1) [e.g., Merzouk et al., 2006; Galí and Simó,2010]. Therefore, in the following paragraphs, we will analyze separately the control exerted by GPDMS andkLOSS on DMS concentrations by means of statistical inference.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 503

Page 9: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

This analysis can be simplified by assuming that [DMS] is at steady state, d[DMS]/dt=0. Substituting inequation (1) and rearranging, we obtain the following:

DMS½ � ¼ GPDMS=kLOSS (7)

Equation (7) tells us that the GPDMS/kLOSS quotient controls steady state [DMS]. Steady state is a reasonableassumption as long as τ is much smaller than the time scale of variability we want to resolve (half a yearif we assume unimodal DMS seasonality). In our case, τ ≤ 4 days<< 180 days. The validity of the steady stateassumption can be tested by meta-analysis of temporal DMS evolution in coherent (Lagrangian) watermasses. As illustrated in Figure 2b, daily net DMS production is typically around zero, which impliesGPDMS ≈ [DMS] · kLOSS, except for particular conditions of thick phytoplankton bloom burst or decay. Moreprecisely, kNET has a median of 0.03 d�1, with an interquartile range (IQR) of �0.10–0.12 d�1 (n= 111).Net DMS change is smaller than ±25% of the UML stock per day in 72% of the observations, which indicatesthat assuming steady state [DMS] over the daily scale carries an unbiased error of magnitude comparableto typical GPDMS or kLOSS measurement uncertainties. As a corollary to these results, Figure 4 shows thatkNET is uncorrelated to SRD and thus introduces no seasonal bias regarding the steady state assumption(see also supporting information Figure S3).

Figure 2c shows that measured GPDMS is strongly correlated to in situ DMS concentrations [Spearman’sr= 0.74, p< 10�23, 99% confidence interval (CI) = 0.53–0.84, n= 159]. However, kLOSS is not significantlycorrelated to in situ [DMS] (Figure 2d; Spearman’s r =�0.16, p = 0.18, n = 70). DMS turnover time calculatedas τ = [DMS] / GPDMS has a median of 1.15 days and is constrained around an interquartile range (IQR)of 0.7–2.1 days (Figure 2b). This is a remarkably narrow range considering that DMS and Chl concentrationsspan more than two orders of magnitude. Turnover times calculated from total DMS loss as τ = 1 / kLOSSspread over a similar IQR of 1.0–2.3 days (Figures 2b and 3i), although their median is 40% larger (1.6 days).We will now examine the data subset comprising the oligotrophic subtropical gyres and the MediterraneanSea, which will be referred to as stress regime areas hereafter (section 3 and supporting informationFigure S2). In this subset, a strong relationship is again observed between [DMS] and GPDMS (Spearman’sr= 0.74, p< 10�13, 99% CI = 0.56–0.84, n= 76) but not between [DMS] and kLOSS (Spearman’s r = 0.03,p= 0.88, n= 34). These observations, together with the validation of the steady state assumption, supportthe notion that GPDMS drives large-scale [DMS] variability across ocean biomes, and particularly in stressregime areas.

Similar kLOSS vs. [DMS] relationships arise when kLOSS is split in two subsets according to the methods used(Figure 2d): one subset where kBC had been measured with 35S-DMS radiotracer additions (Spearman’sr=�0.17, p=0.38, n= 29) and the other where kBC had been measured with the use of competitive inhibitors(Spearman’s r=�0.18, p=0.27, n=41). This may indicate that these two approaches are consistent enoughfor the purpose of this analysis, although kLOSS is somewhat higher in samples where kBC was measuredwith inhibitor methods (median 0.66 d�1) compared to those where kBC was measured with radioisotope(median 0.56 d�1).

4.2. Factors Underlying Gross DMS Production

GPDMS is positively correlated to [DMSPt] both in the entire database (Spearman’s r=0.71, p< 10�19; Pearson’sr=0.64, p< 10�15, n= 125) and in stress regime areas (Spearman’s r= 0.51, p< 10�5; Pearson’s r = 0.50,p< 10�5, n = 76). Yet, the fact that 59% and 75% of the variance in GPDMS remain unexplained in the globaldatabase and stress regime subset, respectively, points at the crucial role of food web DMSP transformations.Indeed, GPDMS results from the addition of countless processes, including phytoplankton DMS leakageupon intracellular DMSP cleavage, and extracellular cleavage by algal and bacterial enzymes of the DMSPreleased through microzooplankton grazing, viral lysis, cell death, or active phytoplankton release. Nomethods have been established to measure these processes independently on a routine basis, and they aretherefore poorly constrained [Vézina, 2004; Six and Maier-Reimer, 2006; Vogt et al., 2010]. In our analysis,these transformations are embodied in DMSPt consumption k values and DMS yields, for which a small butsufficient number of observations exists.

In the global database, we find a median for total DMSPt consumption rate constant (kDMSPt) of 0.67 d�1,

with an IQR of 0.51–1.10 d�1 (n= 68). Very similar values are found in the stress regime subset (n= 48;

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 504

Page 10: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

Figure 3e). The few available data on particulate DMSP consumption due to grazing (kDMSPp), as deducedfrom dilution experiments (Table 1), display a median of 0.52 d�1 (IQR 0.42–0.69 d�1; n = 28). The figuresencountered for kDMSPt and kDMSPp are highly consistent with phytoplankton growth rates (~0.7 d�1) andmortality rates due to microzooplankton grazing (~0.45 d�1), respectively, measured in temperate andtropical oceans [Calbet and Landry, 2004]. At steady state, kDMSPt must be compensated by phytoplanktongrowth rates of similar magnitude, provided that the cell-associated DMSP pool is recycled at a similarpace than total phytoplankton carbon [Stefels et al., 2007]. This indicates that DMSP turnover is stronglylinked to herbivore microzooplankton grazing, generally recognized as the dominant fate of phytoplanktoncells [Calbet and Landry, 2004].

Community DMS yields (Yt) have a median of 14% in the global database (IQR 7–28%; n=68), with extremelysimilar figures in the stress regime subset (Figure 3f), confirming that non-DMS-producing pathways are thedominant fate of consumed DMSPt. Due to the limited number of Yt measurements available, we alsoexplored the quotient GPDMS:DMSPt (d�1), here noted as Y*t, which was recently proposed as an alternativemetric of community DMS yields [Bailey et al., 2008; Herrmann et al., 2012]. Y*t is dimensionally equivalent tokDMSPt · Yt (units of d

�1) so that equation (2) can be rewritten as follows:

GPDMS ¼ DMSPt½ � · Y�t (8)

A median Y*t of 0.10 d�1 (0.066–0.16 d�1; n = 124) is found in the ensemble of the database, with almost

identical values in the stress regime subset (n = 54).

It has to be underlined that community DMS yields (Yt) account for DMSP transformations occurring in theparticulate and the dissolved phase, thus pooling together processes occurring through different pathways,microorganisms, and characteristic time scales [Simó, 2004]. The distinction between community anddissolved (Yd) yields is sometimes unclear in the literature. Focusing now on dissolved-phase processes, we

Figure 4. Spearman’s rank correlation coefficients between DMS(P) cycling and environmental variables. (a) solar radiation dose (SRD) index; (b) nitrate + nitriteconcentration. Correlation coefficients are represented by stars if significant at p< 0.01 level, or by squares otherwise. Bars: confidence intervals (CI) of thecorrelation coefficient calculated by bootstrapping (n = 2000). Black: global database; gray: stress regime areas subset. The inner portion represents the 95% CI,and the outer portion (lighter color) the 99% CI. Numbers below each bar show sample n. Bootstrapped CIs provide a more stringent significance criterion thanp-values (significant CIs being those that do not include zero). Correlations with other variables are shown in supporting information Figure S3. See also supportinginformation Table S2. Note that Spearman correlations of Chl- or carbon-normalized DMSPt and GPDMS are the same because phytoplankton carbon was calculated as amonotonic function of Chl.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 505

Page 11: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

observe that DMSPd turns over aboutthree times faster than DMSPt, with amedian of 2.1d�1 (IQR 1.39–3.55 d�1;n=136) (Figure 3g). Dissolved DMSyields Yd are typically low, with amedian of 8.5% (IQR 6.4–14.4%; n=99),indicating that the sulfur fraction ofDMSP is mostly assimilated intomicrobial biomass or oxidized tonon-volatile compounds [Kiene andLinn, 2000b; Moran et al., 2012].

If we take the medians of total anddissolved k values and yields andassume that DMSPd/DMSPt is 0.10[f in equation (6)], it turns out that onlya median 20% of GPDMS arises fromdissolved DMSP transformations.To illustrate more thoroughly thevariability in the DPDMS/GPDMS

fraction [Fd in equation (6)], we tooka probabilistic approach. Toovercome the fact that total anddissolved DMSP transformationswere almost never simultaneously

measured, we assumed that any kDMSPt, kDMSPd,Yt, and Yd present in the database could randomly co-occurand calculated the corresponding Fd (DPDMS/GPDMS) frequency distributions (section 3.8). Figure 5 showsthe distributions deduced under this assumption and in two distinct scenarios, where DMSPd accountsfor either 5% or 20% of DMSPt [Kiene and Slezak, 2006]. In the first scenario, around 10–15% of GPDMS willarise from DMSPd, increasing to 20–30% in the second scenario. Although particulate DMSP (DMSPp) k’sand yields were not explicitly included in the budget equation (Figure 1) due to limited data availability, thisanalysis suggests that particulate-phase DMS production (GPDMS�DPDMS=[DMSPp] kDMSPpYp) contributes themajority of gross DMS production.

4.3. Drivers of Gross DMS Production4.3.1. Drivers of Phytoplankton DMSP ContentDMSPt:Chl ratios have a median of 65 nmolμg�1 (IQR 35–128nmolμg�1, n=311) in the global database. Instress regime areas, their distribution is left skewed (Figure 3b), with a median (IQR) of 122 (61–233nmolμg�1,n=129). Corresponding DMSP-C quotas have a median (IQR) of 0.048 (0.028–0.068) in the global databaseand 0.055 (0.031–0.072) in the stress regime subset, with an overall range of 0.005–0.21. These estimatesrepresent lower limits since the Chl-to-C conversion method we used provides upper limits for phytoplanktoncarbon [Sathyendranath et al., 2009]. On the other hand, some DMSP may be found in the dissolved phaseand in zooplankton [Besiktepe et al., 2004] and bacterial biomass [Kiene and Linn, 2000b], which compensatesthe underestimation of the quota due to the Chl-to-C conversion. Thus, calculated DMSP-C quotas are overallconsistent with the range measured in phytoplankton cultures [Stefels et al., 2007], although a large uncertaintymay affect individual estimates.

DMSPt:Chl displays a positive correlation with SRD and a negative correlation with NO3+NO2 concentrations(Figure 4, supporting information Figure S3). In stress regime areas, the quantile regression captures thepositive trend of DMSPt:Chl and also an increase in the dispersion as SRD increases (Figure 6a). A similar andtighter trend arises when the DMSP-carbon quota is plotted against SRD (Figure 6b). If we considered that SRDincreases from 40Wm�2 to 200Wm�2 fromwinter to summer, the approximately eightfold change of DMSPt:Chldeduced from the quantile regression would correspond to a three- to fourfold change in DMSP-carbon cellquotas, which reflects the decrease in Chl-to-C ratios toward summer (Figure 6b). These trends are consistent withthe underlying processes generally assumed to regulate DMSP concentrations: phytoplankton succession toward

Figure 5. Proportion of gross DMS production channeled through dissolvedDMSP (DPDMS/GPDMS). Probability distributions deduced through randomresampling (n = 10,000; see section 3) under the assumption that dissolved-to-total DMSP ratio is either (a) 0.05 or (b) 0.20 (see sections 2 and 3.8).Dashed black line: global database. Gray line: stress regime areas subset.Medians and means of each distribution are indicated with vertical lines.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 506

Page 12: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

strong DMSP producers (haptophytes, dinoflagellates, chrysophytes, or pelagophytes) in high light andstratified environments and, perhaps, physiological up-regulation of intracellular DMSP quotas in the faceof environmental stress [Sunda et al., 2002; Stefels et al., 2007]. Work is underway to better understand theenvironmental controls on DMSPt using much larger data sets (http://saga.pmel.noaa.gov/dms/).4.3.2. Drivers of DMSP-to-DMS ConversionTotal DMSP consumption rate constants (kDMSPt) showweak and non-significant correlations with environmentalvariables (Figure 4, supporting information Figure S3). This suggests that variations in DMSPt turnoverspeed generally play a secondary role in setting seasonal and spatial DMS patterns. DMSPd consumptionrate constants (kDMSPd) display significant correlations with salinity (negative) and Chl (positive). Thispoints to faster DMSPd turnover in productive environments characterized by salinity-driven stratificationsuch as some subpolar and polar regions [Levasseur, 2013], where DMSPd metabolism can potentiallycontribute a more relevant share of GPDMS than in prevailing global oceanic conditions (Figure 5). No clearrelationship is found between kDMSPd and SRD.

Community DMS yields (Yt) show a consistent covariation with environmental variables (Figure 4, supportinginformation Figure S3), particularly in stress regime areas, where Yt displays significant positive correlations withSRD and SST and negative correlations with NO3+NO2, Chl, and the “Days to Summer Solstice” (“D2SS”; aseasonal marker that decreases as the date approaches the summer solstice). Of the above relationships, only

Figure 6. Relationship between DMS(P) cycling variables and the solar radiation dose (SRD) index in stress regime areas. The results of the quantile regression areshown for the first and third quartiles (outer thin black lines) and for the median (inner thick black line). Seasonal medians and their interquartile ranges areshown in different colors for illustrative purposes only (not used in regressions), with symbol size proportional to n. Note that not all variables were measured inall samples, which causes the median seasonal SRD to move from one plot to another (especially in variables with small n). See also supporting informationFigures S4 and S5.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 507

Page 13: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

those between Yt and SRD and D2SS remain significant in the global database, reinforcing the view that similarirradiance-related processes influence community DMS yields across ocean regimes.

In stress regime areas, Yt typically doubles from a median of 6.5 (IQR 1.7–12%) at an SRD of 40Wm�2 to15 (9.9–31%) at an SRD of 200Wm�2 (Figure 6f ). Y*t, calculated as GPDMS:DMSPt, displays similar althoughweaker relationships with environmental variables, and for the same winter-to-summer SRD range, itincreases from 0.070 (IQR 0.046–0.11 d�1) to 0.12 (0.074–0.18 d�1) (Figure 6g). These observations argueagainst the constancy of Y*t, which was proposed to be constrained around 0.06 ± 0.01 d�1 [Herrmann

et al., 2012]. The slightly weaker response of Y*t (Figure 6g) to SRD compared to Yt seems to result from the slightlynegative slope of kDMSPt vs. SRD, which counteracts the positive Yt trend (remember that Y*t≈ kDMSPt · Yt).Another variable that is worth exploring here is the GPDMS:Chl ratio, which we will note as YChl. This ratioallows an even more simplified expression of gross DMS production:

GPDMS ¼ Chl½ � · YChl (9)

Across the same SRD range as above, median (IQR) YChl increases from 2.3 (0.6–6.1) to 19 (11–40) nmol DMSd�1 (μg Chl)�1 in stress regime areas (Figure 6c). It is also interesting to calculate the quotient betweenGPDMS (in carbon units) to phytoplankton carbon (Table 1). This calculation indicates that the portion ofphytoplankton carbon cycled daily through DMS production increases from about 0.06% in winter to0.3% in summer in stress regime areas, eventually exceeding 1% at SRD> 180Wm�2 (Figure 6d). Thelinear fits to the first and third quartiles of community DMS yields (expressed as Yt or Y*t) display positiverelationships with SRD: as SRD increases, the uncertainty envelopes also show positive trends. The quantileregression results are extended in supporting information Figure S4 to all 5% quantile intervals. The trendsof the variables shown in Figure 6 are also displayed after binning each y variable with respect to SRD insupporting information Figure S5, lending further support to the patterns encountered.

4.4. Drivers of Total DMS Loss

Total DMS loss rate constants (kLOSS) are calculated as kBC + kphoto + kvent (Table 1). kBC has a median of 0.37(IQR 0.15–0.77 d�1; n= 196), with similar statistics of 0.34 (0.12–0.66 d�1; n= 95) in stress regime areas(Figure 3j). Stress regime areas hold significant correlations between kBC and NO3 +NO2 and Chl (positive),and between kBC and SRD and SST (negative). Thus, kBC decreases toward oligotrophic and highly irradiatedwaters. Only the correlations with SRD and Chl remain significant (as well as of the same sign) when analyzedon the global data set, again pointing at the key role of sunlight exposure [Toole et al., 2006].

Abiotic DMS losses generally represent minor sinks compared to bacterial consumption, with a median (IQR)of 0.13 (0.073–0.23 d�1; n= 66) for kphoto and 0.050 (0.024–0.12 d�1; n=120) for kvent in the global data set,and similar figures in stress regime areas (Figure 3). In general terms, the variations in kphoto and kvent resultingfrom environmental forcing are better mechanistically described than those of biological DMS removal. SinceDMS photolysis is a photosensitized process, changes in photosensitizer concentration (CDOM or nitrate)and photoreactivity (in the case of CDOM) can alter DMS photolysis quantum yields [Bouillon and Miller, 2004;Toole et al., 2003; Taalba et al., 2013]. Our observations of significant relationships between kphoto and SRD(Figures 4 and 6, supporting information Figures S3–S5) indicate that eventual variations in photolysisquantum yields are overridden by the large seasonal variations in shortwave UVR exposure within the UML[Toole et al., 2003]. Since volumetric ventilation (kvent) was calculated by dividing interfacial DMS transfercoefficients (md�1) by MLD (m), a strong negative covariation with MLD was expected, thus a positivecovariation with SRD. Mean daily wind speeds in the database are estimated at 6.6 ± 3.5m s�1, well withinglobal-ocean estimates for wind speed statistical distribution [Monahan, 2006]. Strong wind conditions(>10m s�1) are underrepresented, which may strengthen the kvent-MLD covariation. Note, however, thatstrong winds are often associated with deepening MLD, which will decrease kvent [e.g., Simó and Pedrós-Alió,1999b; Yang et al., 2013]. The significant relationship between kvent and SST reflects, besides upper-oceanstratification, the positive effect of temperature on DMS diffusivity.

DMS budget studies done in contrasting environments show that abiotic DMS sinks take over biological DMSconsumption as the UML shoals [Toole et al., 2006; Galí and Simó, 2010]. This may explain the relatively small

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 508

Page 14: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

variability in DMS turnover times and the small change in kLOSS with increasing SRD (Figure 6). In stressregime areas, the median regression predicts that at SRD= 40Wm�2, kBC, kphoto, and kvent will account for88%, 8%, and 4% of total DMS loss, respectively (with kLOSS = 0.66 d�1); at SRD=200Wm�2, kBC, kphoto, andkvent will account for 50%, 34%, and 16% of total DMS loss, respectively (kLOSS = 0.55 d�1). Some regions coulddepart from this big (and rough) picture, for example, the Southern Ocean, where stronger-than-averagewinds prevail [Monahan, 2006]. Recently, Yang et al. [2013] calculated abiotic DMS rate constants in aSouthern Ocean (“bloom regime”) Lagrangian study. They estimated that despite moderate-to-high windspeeds, abiotic DMS loss (photolysis, ventilation, and mixing) represented at most 20% of kLOSS. Since weestimate that SRD was 30–40Wm�2 during Yang et al.’s study (MLD= 50), their figures agree with ourpredicted ranges.

Figure 7. Illustration of the minimal steady state budget model optimized to predict DMS seasonality in stress regime areas. (a) Idealized seasonality of DMS and itsdrivers. The temporal mismatch between DMS and DMSPt and the magnitude of DMS seasonality are modulated by SRD-dependent Y*t and kLOSS. A strongerpositive (negative) SRD dependence of Y*t (kLOSS) enhances the summer DMS peak (vertical dotted lines) and delays it with respect to the Chl and DMSPt peaks. DMShas been normalized to its minimum, and the remainder variables to their respective maximum. (b, c, and d) SRD dependence of Y*t , YChl, and kLOSS. Black line:median regression line (as in Figure 4) with the first and third quartiles depicted by the gray areas. Red and purple lines: SRD dependence of each variable afteroptimization for different sites. DMS, DMSPt, Chl, and SRD time series are shown for (e) the ensemble of the oligotrophic gyres (dashed line), (f ) the Sargasso Sea(dash-dotted line), and (g) BBMO in the NW Mediterranean (solid line). The texture of the DMS line in Figures 7e–7g corresponds to the optimized SRD dependenceof the (b and c) yields and (d) kLOSS. Colored areas represent the interquartile range for each month in Figure 7e and the standard deviation within the UML for eachmonth in Figure 7f. Measurements at BBMO generally had monthly frequency, so no means and standard deviations are calculated.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 509

Page 15: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

4.5. A Minimal Steady State UML Model

To further understand the control of DMS sources and sinks on DMS seasonality, a number of numericalexperiments were performed using the output of the quantile regression and three time series of[DMS] measurements representative of the stress regime: the multi-year studies in the Sargasso Sea(Hydrostation-S of the Bermuda Atlantic Time-series Study, BATS) and the NW Mediterranean Sea [BlanesBay Microbial Observatory (BBMO)], and a monthly climatology of the subtropical oligotrophic gyre cores(see section 3.7).

In a first set of experiments, [DMS] was predicted in each time series from concurrently measured DMSPt andSRD using the following expression:

DMS½ � ¼ DMSPt½ � · Y�t=kLOSS (10)

where Y*t and kLOSS were expressed as a linear function of SRD. This allowed us to assess how SRD modulatesthe timing and the magnitude of the decoupling between [DMS] and [DMSPt] (Figure 7a). The initial modelparameters, directly derived from the quantile regression (model mDMSP0), were subsequently optimized toobtain the best fit to observed DMS in each time series (mDMSP1–mDMSP4; Figures 7b–7d, Table 2, andsupporting information Table S3). The initial model (mDMSP0) explained 47–55% of [DMS] variability,increasing to 50–63% after optimization (mDMSP1) with general improvements in the other model skill

Table 2. Skill Metrics of Minimal UML Steady State Models Optimized for DMS Prediction in Stress Regime Areasa

Model

Oligotrophic Gyres Sargasso Sea Mediterranean Sea

rbias rRMSE R2 rbias rRMSE R2 rbias rRMSE R2

DMSPt-based models [equation (10)]mDMSP0. Initial Y*t and kLOSS 0.06 0.41 0.48 �0.43 0.62 0.47 0.11 0.73 0.55mDMSP1. Optim. Y*t and kLOSS �0.11 0.36 0.50 �0.08 0.46 0.50 �0.05 0.66 0.60mDMSP2. Optim. Y*t, constant kLOSS �0.10 0.37 0.46 �0.07 0.47 0.44 �0.04 0.71 0.53mDMSP3. Optim. kLOSS, constant Y*t �0.09 0.37 0.41 �0.26 0.55 0.35 �0.04 0.74 0.49mDMSP4. Constant Y*t and kLOSS �0.08 0.41 0.21 �0.21 0.59 0.17 �0.08 0.88 0.29

Chl-based models [equation (11)]mCHL0. Initial YChl and kLOSS 0.33 0.51 0.50 �0.47 0.70 0.28 2.14 3.23 0.17mCHL1. Optim. linear YChl �0.05 0.28 0.59 �0.14 0.53 0.42 0.23 1.12 0.21mCHL2. Optim. power YChl �0.05 0.28 0.61 �0.15 0.50 0.51 �0.17 0.98 0.18mCHL3. Optim. quadratic YChl �0.06 0.28 0.66 �0.17 0.48 0.56 0.11 1.05 0.21

aBest model-data fits for each time series are marked in bold. All variables (Y*t, YChl, and kLOSS) are expressed as a linear function of SRD (y = a + b · SRD) unlessotherwise noted in column 1 (last two rows). Power functions: YChl = a + b · SRDc; quadratic functions: YChl = a + b · SRD + c · SRD2.

Table 3. Winter-to-Summer Changes in SRD and DMSPt, and Corresponding Changes in Y*t, GPDMS, and kLOSS in theUpper Mixed Layer Deduced From the SRD-Forced Modela

Oligotrophic Gyres Sargasso Sea Mediterranean Sea

SRD DJF 63 ± 5 18 ± 6 30 ± 18JJA 170 ± 13 184 ± 15 233 ± 47

DMSPt DJF 8.4 ± 4.3 6.5 ± 3.0 12.8 ± 11.6JJA 13.4 ± 5.0 10.7 ± 2.3 28.6 ± 12.9

Change factor 1.6 1.7 2.2Y*t DJF 0.072 ± 0.001 0.096 ± 0.002 0.064 ± 0.007

JJA 0.091 ± 0.002 0.146 ± 0.004 0.148 ± 0.019Change factor 1.3 1.5 2.3

GPDMS DJF 0.61 ± 0.32 0.62 ± 0.29 0.81 ± 0.71JJA 1.25 ± 0.48 1.56 ± 0.36 4.27 ± 2.10

Change factor 2.1 2.5 5.3kLOSS DJF 0.73 ± 0.01 0.66 ± 0.01 0.99 ± 0.03

JJA 0.56 ± 0.02 0.40 ± 0.02 0.67 ± 0.08Change factor 1.3 1.7 1.5

aWinter and summer values are December–January–February (DJF) and June–July–August (JJA) means ± standarddeviation, respectively. SRD has Wm�2 units. Other units as in Table 1.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 510

Page 16: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

metrics. The same model was optimized but now constraining either Y*t or kLOSS to be constant (i.e.,slope = 0). As a result, the predictions degraded, with better results with a constant kLOSS (mDMSP2) thanwith a constant Y*t (mDMSP3). Further degradation of the predictions was observed when both Y*t andkLOSS were held constant (mDMSP4). These results suggest that seasonally variable Y*t and kLOSS are bothrequired to obtain a realistic DMS-DMSPt decoupling and that expressing Y*t and kLOSS as a linear functionof SRD can explain most of the [DMS] variability not accounted for by [DMSPt].

In a second set of numerical experiments named mCHL0–mCHL3, [DMS] was predicted from measured [Chl]and SRD using the following expression:

DMS½ � ¼ Chl½ � · YChl=kLOSS; (11)

thus, bypassing DMSPt and omitting the taxonomic information it carries. The initial parameters (mCHL0)provided reasonable predictions in the Sargasso Sea and the ensemble of oligotrophic gyres, but poorermodel-data fits at the BBMO site. This time, the SRD dependence of kLOSS was fixed using the output of theprior optimization (mDMSP1; see section 3), and the SRD dependence of YChl was optimized using linear andnonlinear functions. A quadratic YChl vs. SRD function (mCHL3) generally produced the best fits (Table 2).YChl-based models overestimated [DMS] when elevated SRD and [Chl] co-occurred, which explains thepoorer fit at BBMO, a coastal location with higher [Chl] variability than open-ocean locations (Figure 7g).Nevertheless, themodel captured the interannual variability in peak DMS concentrations due to the combinationof [Chl] and SRD as predictor variables, which would not be possible using SRD as the sole predictor.

Table 3 shows that modeled GPDMS increased by two- to fivefold between winter and summer months.Besides variations in phytoplankton biomass and their DMSP quota, this was due to an increase of Y*tby a factor of 1.3–2.3, mirrored by a 1.3–1.7 fold decrease in kLOSS. The negative relationship between kLOSS andSRD in stress regime areas is not identified by the quantile regression analysis. This is possibly due to the smallamount of kLOSS measurements, which is limited by kphoto measurements. Instead, a slightly negative kLOSS vs.SRD trend is obtained when kLOSS is calculated by adding the kBC, kphoto, and kvent median regression lines(Figures 6i–6k), which show more robust relationships with SRD (Figure 4). The increase (decrease) of GPDMS

(kLOSS) toward summer implied by the optimization experiments, as well as their different magnitudesamong sites, is consistent with previous observations at BBMO [Vila-Costa et al., 2008] and diagnosticmodel results in the Sargasso Sea [Toole et al., 2008].

5. Discussion

The results of the meta-analysis portray a seasonal DMS trend driven mostly by DMS production processes(Figure 2). In stress regime areas, the concerted increase of phytoplankton DMSP quotas and community DMSyields toward summer (Table 3) overcomes the low summertime phytoplankton biomass to produce thesummer paradox (Figure 7). A concomitant decrease in microbial DMS loss (kBC) toward summer is athird important factor allowing the summer paradox to occur. However, this effect is counteracted by asimultaneous increase of abiotic DMS loss (kphoto + kvent), which buffers the seasonal variations of totalDMS loss rate constant kLOSS (Figures 4 and 6). Our analysis indicates that similar seasonal variations of thebiotic factors may occur in bloom regime areas (Figure 4), where they will add to a phytoplankton phenologythat is already favorable to the summer DMS peak. Extremely high DMS concentrations of some tens of nmol L�1

can only build up when large GPDMS co-occurs with small kLOSS, a transient situation that has been observedin the Ross Sea at the onset of large Phaeocystis blooms [Del Valle et al., 2009] or in purposeful iron fertilizationexperiments [Merzouk et al., 2006] (data points at the lower right corner in Figure 2d).

Both light stress and nutrient limitation have been invoked to explain the decoupling between DMS andDMSPt through DMS yields, but since these environmental factors generally vary in concert, their effects onplankton cannot be separated. There is experimental evidence that phytoplankton increase their DMSproduction rates upon increased exposure to UVR [Sunda et al., 2002; Archer et al., 2010] and their synergisticeffects with nitrogen limitation [Sunda et al., 2007]. In concordance, the implementation of light-mediated“phytoplankton” DMS release in sulfur cycling models has proved critical to better predict the summerparadox [Toole et al., 2008; Vallina et al., 2008; Vogt et al., 2010]. Heterotrophic bacteria have been suggestedto increase their DMS yields (Yd) from released DMSP when photoinhibition or nutrient scarcity limits theirsulfur demands [Kiene et al., 2000]. This hypothesis is supported by field studies conducted during strong

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 511

Page 17: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

bloom events [Merzouk et al., 2006; Steiner and Denman, 2008], but meta-analysis of observations does notshow any clear dependence of Yd on SRD or season (Figure 4) [Lizotte et al., 2012]. Rather it is DMSP-sulfurassimilation, and not conversion into DMS, that increases with irradiance both in the short term [del Valleet al., 2012] and through seasons [Vila-Costa et al., 2007]. Recent modeling studies suggest an importanteffect of nutrient limitation (either nitrogen or phosphorus) on bacterial DMS yields. Although theproposed mechanisms seem plausible, those studies overestimated the magnitude of dissolved-phase DMSproduction due to either too high dissolved DMSP concentrations [Polimene et al., 2012] or DMS yields (Yd)that were well beyond the observations range [Belviso et al., 2012] reported here (Figure 3h).

Our study identifies the seasonal changes in particulate DMS yields as an essential contributor to the summerparadox because, although particulate-phase DMS productionmeasurements were not available, dissolved DMSPmetabolism is shown to support in most instances a minor fraction of GPDMS (Figure 5). In agreement with thisview, the few studies that attempted tomeasure size-resolved DMSP cleavage observed thatmost of the potentialenzymatic activity occurred in particles>2μm [Levine et al., 2012] or even>10μm in some environments [Scarrattet al., 2000a, 2000b; Steinke et al., 2002]. Therefore, variations in DMSPd turnover and fate should no longer beregarded as themain control on GPDMS inmost oceanic settings, as is often implicitly or explicitly assumed [Cursonet al., 2011; Moran et al., 2012], and more efforts should be devoted to understand the drivers of particulateDMSP turnover [Archer et al., 2002; Saló et al., 2010; Archer et al., 2011]. It must be noted that particulate DMSproduction, usually attributed to phytoplankton, can also arise from particle-attached bacteria, which areexposed to DMSP concentrations that are orders-of-magnitude higher than in bulk seawater [Scarratt et al.,2000b; Seymour et al., 2010]. Further work is needed to resolve the partitioning of particulate DMS productionbetween phytoplankton cells and phycosphere bacteria, and their contribution to community DMS yields.

Recent research demonstrated that sunlight exposure, and especially UVR, stimulates daily communityGPDMS up to 80% [Galí et al., 2011, 2013b] in an irradiance and spectrum-dependent manner [Galí et al.,2013a]. GPDMS was also shown to vary through diel cycles in summer [Galí et al., 2013b], driven by a daytimeincrease in Yt. Notably, the magnitude of the daily variation of Yt and Y*t was of similar magnitude to theseasonal variations reported here (Figure 6g). The same study observed that GPDMS determined in darkincubations reflected the stimulatory effects of recent light exposure, whichmay partially explain why Yt and Y*tdeduced from dark incubations were correlated to SRD in the present meta-analysis. Thus, variations in recentlight exposure within the UML may partly explain the scatter around the median Yt vs. SRD relationship, withthe remainder possibly due to distinct microbial communities and food web interactions.

In summary, we showed that the SRD index effectively captures the seasonal variations in key ecosystemsulfur cycling processes that drive the summer paradox (Figures 4–7). Besides light-driven processes, the SRDindex likely accounts for other effects linked to vertical mixing and the nutritional status of the microbialcommunity that may affect biological DMS(P) cycling. Significant variations in DMS(P) cycling across spatialgradients of productivity and nutrient limitation are also identified (Figure 4) and deserve further attention.Yet, the reasonable fit obtained through parameter optimization in the ensemble of oligotrophic gyre coressuggests that an SRD-forced process-based model would have widespread applicability in stress regimeareas. Indeed, the process-based model we present was purposely kept as simple as possible to illustrate themechanistic basis of the SRD-DMS relationship. A diagnostic model based on the same principles wouldbenefit from a proper parameterization of abiotic DMS loss processes and the use of remotely sensedphysical and biogeochemical variables relevant to DMS cycling.

These findings should enhance the performance of prognostic sulfur modules in Earth system models byproviding better-constrained parameters. In the near future, the routinely implementation of cutting-edgemethods will surely provide new insights. For example, the use of multiple isotopic tracers will help us tounderstand and parameterize microscale processes [Asher et al., 2011], and automated high-frequency DMSmeasurement will allow us mapping DMS concentration at high vertical-temporal resolution in Lagrangiansettings [Royer et al., 2014]. This meta-analysis emphasizes the need for more comprehensive process-oriented studies of the biogenic sulfur cycle in the ocean.

ReferencesArcher, S. D., C. E. Stelfox-Widdicombe, P. H. Burkill, and G. Malin (2001), A dilution approach to quantify the production of dissolved

dimethylsulfoniopropionate and dimethyl sulphide due to microzooplankton herbivory, Aquat. Microb. Ecol., 23, 131–145.

AcknowledgmentsFunding was provided by the (former)Spanish Ministry of Science andInnovation through projects ATOS(POL2006-00550/CTM), SUMMER(CTM2008-03309/MAR), PEGASO(CTM2012-37615) and Malaspina 2010(CSD2008-00077). M.G. acknowledgesfunding through a JAE PhD scholarship(CSIC, Spain), Takuvik and the CanadaExcellence Research Chair in RemoteSensing of Canada’s New ArcticFrontier, and a Beatriu de Pinóspostdoctoral fellowship (Generalitat deCatalunya). Emmanuel Devred and EricRehm provided support with the useof satellite data. The paper benefitedfrom suggestions made by MathieuArdyna, Martine Lizotte, Michel Lavoie,and Sarah-Jeanne Royer. We thankNASA’s Ocean Color webpage forproviding satellite data. The commentsof two anonymous reviewers helpedimprove the first version of the paper.We apologize for not being able to citeall relevant contributions because ofspace restrictions. This is a contributionof the Research Group on MarineBiogeochemistry and Global Change,supported by the Generalitat deCatalunya, and of Takuvik JointCNRS-Université Laval InternationalLaboratory. Supporting information isavailable. Correspondence and requestsfor materials should be addressed toM.G. or R.S.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 512

Page 18: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

Archer, S. D., F. J. Gilbert, P. D. Nightingale, M. V. Zubkov, A. H. Taylor, G. C. Smith, and P. H. Burkill (2002), Transformation ofdimethylsulphoniopropionate to dimethylsulphide during summer in the North Sea with an examination of key processes via amodelling approach, Deep Sea Res., Part II, 49, 3067–310.

Archer, S. D., D. G. Cummings, C. A. Llewelyn, and J. R. Fishwick (2009), Phytoplankton taxa, irradiance and nutrient availability determine theseasonal cycle of DMSP in temperate shelf seas, Mar. Ecol. Prog. Ser., 394, 111–124.

Archer, S. D., M. Ragni, R. Webster, R. L. Airs, and R. J. Geider (2010), Dimethylsulfoniopropionate and dimethyl sulfide production in responseto photoinhibition in Emiliania huxleyi, Limnol. Oceanogr., 55, 1579–1589.

Archer, S. D., K. Safi, A. Hall, D. G. Cummings, andM. Harvey (2011), Grazing suppression of dimethylsulfoniopropionate (DMSP) accumulationin iron-fertilised, sub-Antarctic waters, Deep Sea Res., Part II, 58, 839–850.

Asher, E. C., J. W. H. Dacey, M. M. Mills, K. R. Arrigo, and P. D. Tortell (2011), High concentrations and turnover rates of DMS, DMSP and DMSOin Antarctic sea ice, Geophys. Res. Lett., 38, L23609, doi:10.1029/2011GL049712.

Bailey, K. E., D. A. Toole, B. Blomquist, R. G. Najjar, B. Huebert, D. J. Kieber, R. P. Kiene, P. Matrai, G. R. Westby, and D. A. del Valle (2008),Dimethylsulfide production in Sargasso Sea eddies, Deep Sea Res., Part II, 55, 1491–1504.

Bates, T. S., J. D. Cline, R. H. Gammon, and S. R. Kelly-Hansen (1987), Regional and seasonal variations in the flux of oceanic dimethylsulfide tothe atmosphere, J. Geophys. Res., 92, 2930–2938, doi:10.1029/JC092iC03p02930.

Belviso, S., and G. Caniaux (2009), A new assessment in North Atlantic waters of the relationship between DMS concentration and the uppermixed layer solar radiation dose, Global Biogeochem. Cycles, 23, GB1014, doi:10.1029/2008GB003382.

Belviso, S., et al. (2012), DMS dynamics in the most oligotrophic subtropical zones of the global ocean, Biogeochemistry, 110,215–241.

Besiktepe, S., K. Tang, M. Vila, and R. Simó (2004), Dimethylated sulfur compounds in seawater, seston and mesozooplankton in the seasaround Turkey, Deep Sea Res., 51, 1179–1197.

Bouillon, R.-C., and W. L. Miller (2004), Determination of apparent quantum yield spectra of DMS photodegradation in an in situ iron-inducedNortheast Pacific Ocean Bloom, Geophys. Res. Lett., 31, L06310, doi:10.1029/2004GL019536.

Bouillon, R.-C., and W. L. Miller (2005), Photodegradation of dimethyl sulfide (DMS) in natural waters: Laboratory assessment of thenitrate-photolysis-induced DMS oxidation, Environ. Sci. Technol., 39, 9471–9477.

Calbet, A., and M. R. Landry (2004), Phytoplankton growth, microzooplankton grazing, and carbon cycling in marine systems, Limnol.Oceanogr., 49, 51–57.

Curson, R. J., J. D. Todd, M. D. Sullivan, and A. W. B. Johnston (2011), Catabolism of dimethylsulphoniopropionate: Microorganisms, enzymesand genes, Nat. Rev. Microbiol., 9, 849–859.

Dacey, J. W. H., F. A. Howse, A. F. Michaels, and S. G. Wakeham (1998), Temporal variability of dimethylsulfide and dimethylsulfoniopropionatein the Sargasso Sea, Deep Sea Res., Part I, 45, 2085–2104.

del Valle, D. A., D. J. Kieber, D. A. Toole, J. Brinkley, and R. P. Kiene (2009), Biological consumption of dimethylsulfide (DMS) and its importancein DMS dynamics in the Ross Sea, Antarctica, Limnol. Oceanogr., 54, 785–798.

del Valle, D. A., R. P. Kiene, and D. M. Karl (2012), Effect of visible light on dimethylsulfoniopropionate assimilation and conversion todimethylsulfide in the North Pacific Subtropical Gyre, Aquat. Microb. Ecol., 66, 47–62.

Derevianko, G. J., C. Deutsch, and A. Hall (2009), On the relationship between ocean DMS and solar radiation, Geophys. Res. Lett., 36, L17606,doi:10.1029/2009GL039412.

Erceg-Hurn, D. M., and V. Mirosevich (2008), Modern robust statistical methods: An easy way to maximize the accuracy and power of yourresearch, Am Psychol., 63(7), 591–601.

Gabric, A., W. Gregg, R. G. Najjar, D. Erickson, and P. Matrai (2001), Modeling the biogeochemical cycle of dimethylsulfide in the upper ocean:A review, Chemosphere Global Change Sci., 3, 377–393.

Galí, M., and R. Simó (2010), Occurrence and cycling of dimethylated sulfur compounds in the Arctic during summer receding of the ice edge,Mar. Chem., 122, 105–117.

Galí, M., V. Saló, R. Almeda, A. Calbet, and R. Simó (2011), Stimulation of gross dimethylsulfide (DMS) production by solar radiation, Geophys.Res. Lett., 38, L15612, doi:10.1029/2011GL048051.

Galí, M., C. Ruiz-González, T. Lefort, J. M. Gasol, C. Cardelús, C. Romera-Castillo, and R. Simó (2013a), Spectral irradiance dependence ofsunlight effects on plankton dimethylsulfide production, Limnol. Oceanogr., 58, 489–504.

Galí, M., R. Simó, M. Vila-Costa, C. Ruiz-González, J. M. Gasol, and P. Matrai (2013b), Diel patterns of oceanic dimethylsulfide (DMS) cycling:Microbial and physical drivers, Global Biogeochem. Cycles, 27, 620–636, doi:10.1002/gbc.20047.

Glover, D. M., W. J. Jenkins, and S. C. Doney (2011), Modeling Methods for Marine Science, Cambridge Univ. Press, Cambridge, U. K.Herrmann, M., et al. (2012), Diagnostic modelling of dimethylsulfide production in coastal water west of the Antarctic Peninsula, Cont. Shelf

Res., 32, 96–109.Kieber, D. J., J. Jiao, R. P. Kiene, and T. S. Bates (1996), Impact of dimethylsulfide photochemistry on methyl sulphur cycling in the equatorial

Pacific Ocean, J. Geophys. Res., 101, 3715–3722, doi:10.1029/95JC03624.Kiene, R. P., and D. Slezak (2006), Low dissolved DMSP concentrations in seawater revealed by small-volume gravity filtration and dialysis

sampling, Limnol. Oceanogr. Methods, 4, 80–95.Kiene, R. P., and L. J. Linn (2000a), Distribution and turnover of dissolved DMSP and its relationship with bacterial production and

dimethylsulfide in the Gulf of Mexico, Limnol. Oceanogr., 45, 849–861.Kiene, R. P., and L. J. Linn (2000b), The fate of dissolved dimethylsulfoniopropionate (DMSP) in seawater: Tracer studies using

35S-DMSP,

Geochim. Cosmochim. Acta, 64, 2797–2810.Kiene, R. P., L. J. Linn, and J. A. Bruton (2000), New and important roles for DMSP in marine microbial communities, J. Sea Res., 43, 209–224.Koenker, R., and G. Bassett (1978), Regression quantiles, Econometrica, 46, 33–50.Lana, A., et al. (2011), An updated climatology of surface dimethlysulfide concentrations and emission fluxes in the global ocean, Global

Biogeochem. Cycles, 25, GB1004, doi:10.1029/2010GB003850.Lana, A., R. Simó, S. Vallina, and J. Dachs (2012), Re-examination of global emerging patterns of DMS concentration, Biogeochemistry, 110, 173–182.Le Clainche, Y., et al. (2010), A first appraisal of prognostic ocean DMS models and prospects for their use in climate models, Global

Biogeochem. Cycles, 24, GB3021, doi:10.1029/2009GB003721.Ledyard, K. M., and J. W. H. Dacey (1996), Microbial cycling of DMSP and DMS in coastal and oligotrophic seawater, Limnol. Oceanogr., 41(1),

33–40.Levasseur, M. (2013), Impact of Arctic meltdown on the microbial cycling of sulphur, Nat. Geosci., 6, 691–700.Levine, N. M., V. A. Varaljay, D. A. Toole, J. W. H. Dacey, S. C. Doney, and M. A. Moran (2012), Environmental, biochemical and genetic drivers of

DMSP degradation and DMS production in the Sargasso Sea, Environ. Microbiol., 14, 1210–1233.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 513

Page 19: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

Lizotte, M., M. Levasseur, S. Michaud, M. G. Scarratt, A. Merzouk, M. Gosselin, J. Pommier, R. B. Rivkin, and R. P. Kiene (2012), Macroscalepatterns of the biological cycling of dimethylsulfoniopropionate (DMSP) and dimethylsulfide (DMS) in the Northwest Atlantic,Biogeochemistry, 110, 183–200.

Merzouk, A., M. Levasseur, M. G. Scarratt, S. Michaud, R. B. Rivkin, M. S. Halec, R. P. Kiene, N. M. Price, and W. K. W. Li (2006), DMSP and DMSdynamics during a mesoscale iron fertilization experiment in the Northeast Pacific—Part I: Biological cycling, Deep Sea Res., Part II, 53,2353–2369.

Miles, C. J., T. G. Bell, and T. M. Lenton (2009), Testing the relationship between the solar radiation dose and surface DMS concentrationsusing in situ data, Biogeosciences, 6, 1927–1934.

Monahan, A. H. (2006), The probability distribution of sea surface wind speeds. Part I: Theory and seawinds observations, J. Clim., 19, 497–520.Moran, M. A., C. R. Reisch, R. P. Kiene, and W. B. Whitman (2012), Genomic insights into bacterial DMSP transformations, Annu. Rev. Mar. Sci., 4,

523–542.Morel, A., Y. Huot, B. Gentili, P. J. Werdell, S. B. Hooker, and B. A. Franz (2007), Examining the consistency of products derived from various

ocean color sensors in open ocean (Case 1) waters in the perspective of a multi-sensor approach, Remote Sens. Environ., 111, 69–88.Polimene, L., S. D. Archer, M. Butenschön, and J. I. Allen (2012), A mechanistic explanation of the Sargasso Sea DMS “summer paradox”,

Biogeochemistry, 110, 243–255.Royer, S.-J., M. Galí, E. S. Saltzman, C. A. McCormick, T. G. Bell, and R. Simó (2014), Development and validation of a shipboard system for

measuring high-resolution profiles of aqueous dimethylsulfide concentrations using chemical ionisation mass spectrometry, Environ.Chem., 11(3), 309–317.

Saló, V., R. Simó, and A. Calbet (2010), Revisiting the dilution technique to quantify the role of microzooplankton in DMS(P) cycling:Laboratory and field tests, J. Plankton Res., 32, 1255–1267.

Sathyendranath, S., V. Stuart, A. Nair, K. Oka, T. Nakane, H. Bouman, M.-H. Forget, H. Maass, and T. Platt (2009), Carbon-to-chlorophyll ratio andgrowth rate of phytoplankton in the sea, Mar. Ecol. Prog. Ser., 383, 73–84.

Scarratt, M. G., G. Cantin, M. Levasseur, and S. Michaud (2000a), Particle size-fractionated kinetics of DMS production: Where does DMSPcleavage occur at the microscale?, J. Sea Res., 43, 245–252.

Scarratt, M. G., M. Levasseur, S. Schultes, S. Michaud, G. Cantin, A. Vézina, M. Gosselin, and S. J. de Mora (2000b), Production and consumptionof dimethylsulfide (DMS) in North Atlantic waters, Mar. Ecol. Prog. Ser., 204, 13–26.

Schmidtko, S., G. C. Johnson, and J. M. Lyman (2013), MIMOC: A global monthly isopycnal upper-ocean climatology with mixed layers,J. Geophys. Res. Oceans, 118, 1658–1672, doi:10.1002/jgrc.20122.

Seymour, J. R., R. Simó, T. Ahmed, and R. Stocker (2010), Chemoattraction to dimethylsulfoniopropionate throughout the marine microbialfood web, Science, 329, 342–345.

Simó, R. (2004), From cells to globe: Approaching the dynamics of DMS(P) in the ocean at multiple scales, Can. J. Fish. Aquat. Sci., 61,673–684.

Simó, R., and C. Pedrós-Alió (1999a), Role of vertical mixing in controlling the oceanic production of dimethyl sulphide, Nature, 402, 396–399.Simó, R., and C. Pedrós-Alió (1999b), Short-term variability in the open ocean cycle of dimethylsulfide, Global Biogeocehm. Cycles, 13,

1173–1181.Simó, R., C. Pedrós-Alió, G. Malin, and J. O. Grimalt (2000), Biological turnover of DMS, DMSP and DMSO in contrasting open-sea waters,

Mar. Ecol. Prog. Ser., 203, 1–11.Six, K. D., and E. Maier-Reimer (2006), What controls the oceanic dimethylsulfide (DMS) cycle? A modeling approach, Global Biogeochem.

Cycles, 20, GB4011, doi:10.1029/2005GB002674.Stefels, J., M. Steinke, S. Turner, G. Malin, and S. Belviso (2007), Environmental constraints on the production and removal of the climatically

active gas dimethylsulphide (DMS) and implications for ecosystem modelling, Biogeochemistry, 83, 245–275.Steiner, N. S., and K. Denman (2008), Parameter sensitivities in a 1-D model for DMS and sulphur cycling in the upper ocean, Deep Sea Res.,

Part I, 55, 847–865.Steinke, M., G. Malin, S. D. Archer, P. H. Burkill, and P. S. Liss (2002), DMS production in a coccolithophorid bloom: Evidence for the importance

of dinoflagellate DMSP lyases, Aquat. Microb. Ecol., 26, 259–270.Sunda, W., D. J. Kieber, R. P. Kiene, and S. Huntsman (2002), An antioxidant function for DMSP and DMS in marine algae,Nature, 418, 317–319.Sunda, W. G., R. Hardison, R. P. Kiene, E. Bucciarelli, and H. Harada (2007), The effect of nitrogen limitation on cellular DMSP and DMS release

in marine phytoplankton: Climate feedback implications, Aquat. Sci., 69, 341–351.Taalba, A., H. Xie, M. G. Scarratt, S. Bélanger, and M. Levasseur (2013), Photooxidation of dimethylsulfide (DMS) in the Canadian Arctic,

Biogeosciences, 10, 6793–6806.Toole, D. A., and D. A. Siegel (2004), Light-driven cycling of dimethylsulfide (DMS) in the Sargasso Sea: Closing the loop, Geophys. Res. Lett., 31,

L09308, doi:10.1029/2004GL019581.Toole, D. A., D. J. Kieber, R. P. Kiene, D. A. Siegel, and N. B. Nelson (2003), Photolysis and the dimethylsulfide (DMS) summer paradox in the

Sargasso Sea, Limnol. Oceanogr., 48, 1088–1100.Toole, D. A., D. J. Kieber, R. P. Kiene, E. M. White, J. Bisgrove, D. A. del Valle, and D. Slezak (2004), High dimethylsulfide photolysis rates in

nitrate-rich Antarctic waters, Geophys. Res. Lett., 31, L11307, doi:10.1029/2004GL019863.Toole, D. A., D. Slezak, R. P. Kiene, D. J. Kieber, and D. A. Siegel (2006), Effects of solar radiation on dimethylsulfide cycling in the western

Atlantic Ocean, Deep Sea Res., Part I, 53, 136–153.Toole, D. A., D. A. Siegel, and S. C. Doney (2008), A light-driven, one-dimensional dimethylsulfide biogeochemical cycling model for the

Sargasso Sea, J. Geophys. Res., 113, G02009, doi:10.1029/2007JG000426.Vallina, S. M., and R. Simó (2007), Strong relationship between DMS and the solar radiation dose over the global surface ocean, Science, 315,

506–508.Vallina, S. M., R. Simó, T. R. Anderson, A. Gabric, R. Cropp, and J. M. Pacheco (2008), A dynamic model of oceanic sulfur (DMOS) applied to the

Sargasso Sea: Simulating the dimethylsulfide (DMS) summer paradox, J. Geophys. Res., 113, G01009, doi:10.1029/2007JG000415.Vézina, A. F. (2004), Ecosystem modelling of the cycling of marine dimethylsulfide: A review of current approaches and of the potential for

extrapolation to global scales, Can. J. Fish. Aquat. Sci., 61, 845–856.Vila-Costa, M., J. Pinhassi, C. Alonso, J. Pernthaler, and R. Simó (2007), An annual cycle of dimethylsulfoniopropionate-sulfur and leucine

assimilating bacterioplankton in the coastal NW Mediterranean, Environ. Microbiol., 9, 2451–2463.Vila-Costa, M., R. Simó, and R. P. Kiene (2008), Seasonal variability of the dynamics of dimethylated sulfur compounds in a coastal northwest

Mediterranean site, Limnol. Oceanogr., 53, 198–211.Vogt, M., S. M. Vallina, E. T. Buitenhuis, L. Bopp, and C. Le Quéré (2010), Simulating dimethylsulphide seasonality with the Dynamic Green

Ocean Model PlankTOM5, J. Geophys. Res., 115, C06021, doi:10.1029/2009JC005529.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 514

Page 20: A meta-analysis of oceanic DMS and DMSP cycling processes ...digital.csic.es/bitstream/10261/116529/1/Gali_et_al_2015.pdf · A meta-analysis of oceanic DMS and DMSP cycling processes:

Wolfe, G. H., and R. P. Kiene (1993a), Radioisotope and chemical inhibitor measurements of dimethyl sulfide consumption rates and kineticsin estuarine waters, Mar. Ecol. Prog. Ser., 99, 261–269.

Wolfe, G. H., and R. P. Kiene (1993b), Effects of methylated, organic, and inorganic substrates on microbial consumption of dimethylsulfide inestuarine waters, Appl. Environ. Microbiol., 59, 2723–2726.

Yang, M., S. D. Archer, B. W. Blomquist, D. T. Ho, V. P. Lance, and R. J. Torres (2013), Lagrangian evolution of DMS during the Southern Oceangas exchange experiment: The effects of vertical mixing and biological community shift, J. Geophys. Res. Oceans, 118, 6774–6790,doi:10.1002/2013JC009329.

Global Biogeochemical Cycles 10.1002/2014GB004940

GALÍ AND SIMÓ ©2015. American Geophysical Union. All Rights Reserved. 515