<<

EGU Journal Logos (RGB) Open Access Open Access Open Access Advances in Annales Nonlinear Processes Geosciences Geophysicae in Geophysics Open Access Open Access Natural Hazards Natural Hazards and System and Earth System Sciences Sciences Discussions Open Access Open Access Atmospheric Atmospheric Chemistry Chemistry and Physics and Physics Discussions Open Access Open Access Atmospheric Atmospheric Measurement Measurement Techniques Techniques Discussions Open Access Open Access

Biogeosciences Biogeosciences Discussions Open Access Open Access Climate Climate of the Past of the Past Discussions Open Access Open Access Earth System Earth System Dynamics Dynamics Discussions Open Access Geoscientific Geoscientific Open Access Instrumentation Instrumentation Methods and Methods and Data Systems Data Systems Discussions Open Access Open Access Geoscientific Geoscientific Model Development Model Development Discussions Open Access Open Access Hydrology and Hydrology and Earth System Earth System Sciences Sciences Discussions Open Access Ocean Sci., 9, 193–216, 2013 Open Access www.ocean-sci.net/9/193/2013/ Ocean Science doi:10.5194/os-9-193-2013 Ocean Science Discussions © Author(s) 2013. CC Attribution 3.0 License. Open Access Open Access Solid Earth Solid Earth Discussions

CO2

Global surface-ocean p and sea–air CO2 flux variability fromOpen Access an Open Access observation-driven ocean mixed-layer scheme The Cryosphere The Cryosphere Discussions C. Rodenbeck¨ 1, R. F. Keeling2, D. C. E. Bakker3, N. Metzl4, A. Olsen5,6,7, C. Sabine8, and M. Heimann1 1Max Planck Institute for Biogeochemistry, Jena, Germany 2Scripps Institution of , University of California, San Diego, USA 3School of Environmental Sciences, University of East Anglia, Norwich Research Park, Norwich, UK 4LOCEAN-IPSL, CNRS, Paris, France 5Institute of Marine Research, Bergen, Norway 6Uni Bjerknes Centre, Bergen, Norway 7Bjerknes Centre for Climate Research, Bergen, Norway 8NOAA Pacific Marine Environmental Laboratory, Seattle, USA

Correspondence to: C. Rodenbeck¨ ([email protected])

Received: 27 April 2012 – Published in Ocean Sci. Discuss.: 27 June 2012 Revised: 23 January 2013 – Accepted: 29 January 2013 – Published: 1 March 2013

Abstract. A temporally and spatially resolved estimate of 1 Introduction the global surface-ocean CO2 field and the sea–air CO2 flux is presented, obtained by fitting a simple The oceans are considered the dominant player in the data-driven diagnostic model of ocean mixed-layer biogeo- global on long timescales, e.g. in the glacial– chemistry to surface-ocean CO2 partial pressure data from interglacial cycles (e.g. Sigman and Boyle, 2000). On a the SOCAT v1.5 database. Results include seasonal, interan- multi-millennial timescale, the oceans will be the sink for nual, and short-term (daily) variations. In most regions, es- 80–95 % of the anthropogenic CO2 emissions, and 70–80 % timated seasonality is well constrained from the data, and on a timescale of several hundred years (Archer et al., 1997). compares well to the widely used monthly by Currently, the oceans take up about 25 % of the emissions Takahashi et al. (2009). Comparison to independent data ten- (Sarmiento et al., 2010). Concerns exist, however, that the tatively supports the slightly higher seasonal variations in our sink efficiency may decrease in the coming decades as a estimates in some areas. We also fitted the diagnostic model consequence of anthropogenic , as suggested to atmospheric CO2 data. The results of this are less robust, by model projections (e.g. Sarmiento and Le Quer´ e´, 1996; but in those areas where atmospheric signals are not strongly Matear and Hirst, 1999; Joos et al., 1999) and tentatively con- influenced by land flux variability, their seasonality is nev- firmed by data analysis (e.g. Le Quer´ e´ et al., 2007, 2010). As ertheless consistent with the results based on surface-ocean a prerequisite to understanding the involved processes, one data. From a comparison with an independent seasonal cli- needs to quantify sea–air CO2 fluxes, their variability, and matology of surface-ocean nutrient concentration, the diag- their response to forcing. nostic model is shown to capture relevant surface-ocean bio- Currently, two data streams are used to estimate the vari- geochemical processes reasonably well. Estimated interan- ability of sea–air CO2 fluxes: nual variations will be presented and discussed in a compan- – Based on measurements of the CO2 partial pressure ion paper. CO (p 2 ) of surface water, the sea–air CO2 flux is cal- culated through a gas exchange parameterization. As pCO2 measurements only exist at discrete locations and ship tracks, spatial and temporal interpolation is needed. The following interpolation methods have been

Published by Copernicus Publications on behalf of the European Geosciences Union. CO 194 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability

proposed: (1) statistical interpolation combined with an In view of the above-mentioned problem of the atmospheric advection–diffusion equation (Takahashi et al., 2009); CO2 inversions, most of these studies use the sea–air flux cli- (2) purely statistical interpolation with error quantifica- matology by Takahashi et al. (2009) (or earlier versions) as tion (Jones et al., 2012b); (3) multi-linear regressions a Bayesian prior (in some cases, the long-term fluxes from between pCO2 and ocean state variables (e.g. Park et al., ocean-interior inversions are used in the prior as well, e.g. 2010; Chen et al., 2011); (4) neural networks learn- Rodenbeck¨ et al. (2003), or formalized as joint inversion ing the relationship between pCO2 and oceanic state by Jacobson et al. (2007)). Out of this context, the data- variables available from remote sensing platforms and based sea–air CO2 fluxes presented here have been devel- ocean reanalysis projects (e.g. Lefevre` et al., 2005; Tel- oped aiming to (1) provide not only a monthly seasonal cy- szewski et al., 2009); (5) assimilation of the pCO2 data cle but variability also on short-term (daily) and interannual into a process model of ocean biogeochemistry (Valsala timescales, (2) use an assimilation scheme that can easily and and Maksyutov, 2010; While et al., 2012). These meth- self-consistently compare or even combine the observational ods are complementary in certain aspects: the more sta- information of atmospheric and surface-ocean data, and (3) tistical methods are dominated by the data themselves use a framework that can be extended to also incorporate con- but are strongly affected by spatial and temporal gaps, straints from other tracers, such as oxygen. while the more complex methods more easily spread The presented results are based on the newly available Sur- the information but are dependent on driver data sets face Ocean CO2 Atlas (SOCAT) database (version 1.5) of or even on model formulations. CO2 fugacity measurements (Pfeil et al., 2012, http://www. socat.info/). We describe the estimation method and test its – The CO exchange between the Earth surface and the at- 2 performance, also using independent data. As a first step of mosphere can be quantified based on atmospheric CO 2 analysis, this study mainly focuses on the mean seasonal cy- mixing ratio measurements (e.g. Conway et al., 1994) cle, because on this timescale (1) we can compare to the by the atmospheric transport inversion (e.g. Newsam widely used monthly pCO2 climatology by Takahashi et al. and Enting, 1988; Rayner et al., 1999; Bousquet et al., (2009), (2) we can investigate mutual consistency between 2000; Rodenbeck¨ et al., 2003; Baker et al., 2006) (see the surface-ocean pCO2 constraint and that by atmospheric Sect. 2.1). This estimation does not involve any param- CO data as the signal/noise ratio is large, and (3) we can eterization of gas exchange. However, ocean fluxes are 2 test the plausibility of our process representations invoking difficult to detect by this method in most parts of the a seasonal surface-ocean phosphate (PO ) climatology. Esti- globe, because their imprint on the atmospheric mixing 4 mated interannual variations of the sea–air CO flux are pre- ratio records is small compared to that of the much more 2 sented and discussed in a companion paper (Rodenbeck¨ et variable land fluxes: Even if ocean-internal processes al., 2013). (biology, transport) cause mixed-layer carbon sources and sinks of comparable variability as the land bio- sphere, the resulting sea–air CO exchange is much less 2 2 Method variable and smoothed out in time because the carbon- ate chemistry of seawater slows the equilibration rate of 2.1 Concept – overview dissolved CO2 with the atmosphere. In addition, the re- sponses to warming/cooling partially counteract the ef- In a classical atmospheric inversion, a spatio-temporal field fects of ocean-internal sources/sinks. A further problem of surface-to-atmosphere carbon fluxes is estimated such that of atmospheric inversions based on data from a discrete its corresponding mixing ratio field – as simulated by an at- set of measurement sites is that information is only pro- mospheric tracer transport model – matches as closely as vided on scales comparable to or larger than the distance possible a set of mixing ratio observations. The match is between the sites or the sampling frequency, while vari- gauged by a quadratic cost function to be minimized (Ap- ability exists also on smaller scales. pendix A2). A further data-based method to estimate sea–air CO2 fluxes Here we extend the inversion framework by not only con- uses ocean-interior carbon data in inverse calculations of sidering the process of atmospheric transport but also pro- oceanic transport (Gloor et al., 2003; Mikaloff Fletcher et al., cesses in the oceanic mixed layer: The atmospheric trans- 2006). This method is independent of parameterizations of port model is supplemented by a chain of parameterizations gas exchange as well. However, it only yields long-term sea– of gas exchange, carbonate chemistry, and a carbon bud- air CO2 fluxes over large spatial regions, not its temporal or get equation, which determine the sea-to-air CO2 exchange CO2 high-resolution spatial variability. The long-term global sea– (fma ) as a function of ocean-internal carbon sources and air CO2 flux has also been estimated from observed trends sinks (Fig. 1). Then the inversion is not directly adjusting the in atmospheric oxygen (Keeling and Shertz, 1992) as well sea-to-air fluxes any more, but only indirectly through adjust- 13 as from C isotopic ratios in atmospheric CO2 (Ciais et al., ing the ocean-internal fluxes instead (Fig. 2, from left to mid- 1995). dle). This offers two advantages: (1) the equations describing

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ 22 C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data

CO Atm. mixing ratio [ppm]: c 2 (i,t) K CO2 CO2 meteo., fnee , f Atm. Transport fos → CO Sea-air fluxes [mol/m2/s]: f 2 (x,y,t) ma K u, T , S, pbaro , XCO2 Solub., Gas exch. → CO Mixed-layer partial press. [µatm]: p 2 (x,y,t) K

CO2,Ref DIC,Ref T , S, A, γDIC, pm , C Carbon. chem. m → DIC Mixed-layer concentration [µmol/kg]: C (x,y,t) m K h C Budget →

2 DIC Ocean-internal sources/sinks [mol/m /s]: fint (x,y,t)

Fig. 1. Summary of the main quantities and process parameteriza- tions involved in the diagnostic ocean mixed-layer model. Thick boxes denote the process parameterizations given in Sect. 2.2, each expressing the quantity right above as a function of the quantity right below. Quantities at the arrows on the left represent driver data entering the parameterizations. See Table 2 for mathematical CO symbols, and Appendix A1 for the equations and further explana- C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 195 22 C. R¨odenbeck et al.: Sea-airtion. CO2 flux estimated from CO2 partial pressure data

CO Atm. mixing ratio [ppm]: c 2 (i,t) Atm. Inv. For ATM For SFC K CO2 Jc CO2 CO2 Jc CO2 CO2 CO2 c c c c meteo., fnee , f Atm. Transport obs obs fos → ⇐⇒ K ⇐⇒ K

2 CO2 AT AT Sea-air fluxes [mol/m /s]: fma (x,y,t) K CO2 CO2 CO2 fma fma fma CO K K u, T , S, pbaro , X 2 Solub., Gas exch. → → GE GE CO2 CO J Mixed-layer partial press. [µatm]: p (x,y,t) pCO2 p 2 p pCO2 K K obs ⇐⇒ K CO2,Ref DIC,Ref T , S, A, γDIC, pm , Cm Carbon. chem. CC CC → DIC DIC DIC Cm Cm Mixed-layer concentration [µmol/kg]: C (x,y,t) K K m K CB CB h C Budget → f DIC f DIC → int → int 2 DIC Ocean-internal sources/sinks [mol/m /s]: fint (x,y,t) Fig.Fig. 2. 2. IllustrationIllustration of of the the inverse inverse procedure procedure (three (three modi). modi). Boxes Boxes de- note the process parameterizations from Figure 1, causally linking Fig.Fig. 1. 1.SummarySummary of the main quantities quantities and and process process parameteriza- parameteriza- denote the process parameterizations from Fig. 1, causally linking tions involved in the diagnostic ocean mixed-layer model. Thick quantities from bottom to top. Double arrows symbolize the match- tions involved in the diagnostic ocean mixed-layer model. Thick quantities from bottom to top. Double arrows symbolize the match- boxes denote the process parameterizations given in Sect. 2.2, each ing between observed and modelled quantities, as gauged by a cost boxes denote the process parameterizations given in Sect. 2.2, each ing between observed and modelled quantities, as gauged by a cost expressing the quantity right above as a function of the quantity function J (Appendix A2). Thin arrows indicate the adjustments expressing the quantity right above as a function of the quantity function J (Appendix A2). Thin arrows indicate the adjustments right below. Quantities at the arrows on the left represent driver of the respective unknowns done to minimize the model-data mis- right below. Quantities at the arrows on the left represent driver data of the respective unknowns done to minimize the model–data mis- data entering the parameterizations. See Table 2 for mathematical match. Left: Pure atmospheric transport inversion not used here but entering the parameterizations. See Table 2 for mathematical sym- match.given Left: for reference: pure atmospheric The sea-air transport fluxes f inversionCO2 are adjusted not used to here match but symbols, and Appendix A1 for the equations and further explana- COma givenatmospheric for reference: mixing the ratios sea–air (terrestrial fluxes Netf Ecosystem2 are adjusted Exch toange match is bols,tion. and Appendix A1 for the equations and further explanation. ma atmosphericadjusted as mixing well but ratios omitted (terrestrial from this net schematic ecosystem for exchange clarity). Mid- is ad- DIC justeddle: asIn wellcase butATM omitted, ocean-internal from this sources/sinks schematic forfint clarity).are adjusted Middle: Atm. Inv. For ATM For SFC instead of sea-air fluxes to match atmospheric CODIC2 mixing ratios the individual processes impose spatial and temporal struc- in the case of ATM, ocean-internal sources/sinks fint are adjusted instead(again of Net sea–air Ecosystem fluxes Exchange to match is atmospheric further adjusted). CO mixing Right: ratiosIn tureCO2 toJ thec sea–airCO2 fluxes,CO2 basedJc onCO2 the spatio-temporal infor- 2 c c c c case SFC, ocean-internal sources/sinks are adjusted to match the mationobs ⇐⇒ fromK the driverobs data⇐⇒ (sea-surfaceK (SST), (again net ecosystem exchange is further adjusted). Right: in the casesurface-ocean of SFC, ocean-internalCO2 partial pressure sources/sinks observations. are adjusted Sea-air to fluxes match are the wind speed,AT etc., Fig. 1). (2) TheAT explicit representation of surface-oceanthen calculated CO frompartial the estimated pressure partial observations. pressure Sea–air field. fluxes are oceanic processesf CO2 involves furtherf CO2 quantities, notablyf CO2 sea- 2 ma maK maK then calculated from the estimated partial pressure field. surface→ CO partial pressure. Through cost function contri- 2 GE GE butions gauging the match of modelled and measured pCO2 , CO CO Jp CO these data can be used as an observationalp 2 constraintp 2 p replac-2 K obs ⇐⇒ K ysis (Kalnay et al., 1996). The model fields are sampled at ing the atmospheric data (Fig. 2, right). In this mode of op- CC CC the location and time of the individual mixing ratio measure- eration, atmospheric transport models and atmospheric data DIC DIC ments used. TM3 has performed well in intercomparisons of are actually no longer used in theCm inverse calculation,C butm the K K state-of-the-art global tracer transport models (e.g. Stephens Bayesian framework, including a-priori spatial and temporal CB CB et al., 2007; Law et al., 2008). correlations, is still applied in theDIC same way as in theDIC atmo- f f Solubility and gas exchange. Diffusive sea-to-air gas ex- spheric mode. → int → int change is proportional to the over-/undersaturation of CO in The calculation is global over the time period 1985–2011, 2 Fig. 2. Illustration of the inverse procedure (three modi). Boxes de- the surface ocean and to the piston velocity with the quadratic withnote thea temporal process parameterizations resolution of 1 from day Figureand a 1,horizontal causally linking resolu- wind speed dependence by Wanninkhof (1992), scaled to tionquantities of approximately from bottom to 4 top.◦ latitude Double× arrows5◦ longitude symbolize (TM3 the m modelatch- match the global average piston velocity of Naegler (2009) grid).ing between observed and modelled quantities, as gauged by a cost function J (Appendix A2). Thin arrows indicate the adjustments (Appendix A1.1). 2.2of the Process respective parameterizations unknowns done to minimize the model-data mis- Carbonate chemistry. The carbon species relevant for gas match. Left: Pure atmospheric transport inversion not used here but exchange (CO2) only account for a small part of the car- CO2 Allgiven processes for reference: considered The sea-air explicitly fluxes fma areare given adjusted in Fig. to match1 and bon relevant in the ocean-internal budget (dissolved inor- atmospheric mixing ratios (terrestrial Net Ecosystem Exchange is summarized in the following. Details, including the equa- ganic carbon, DIC). The link between CO2 abundance (ex- adjusted as well but omitted from this schematic for clarity). Mid- CO2 tionsdle: In used, case areATM found, ocean-internal in Appendix sources/sinksA. f DIC are adjusted pressed in terms of partial pressure pm ) and DIC abun- int DIC insteadAtmospheric of sea-air transport fluxes to match. Atmospheric atmospheric COCO2 mixingmixing ratios ratio dance (in terms of its concentration Cm ) is determined fields(again in Net response Ecosystem to surface-to-atmosphereExchange is further adjusted). fluxes Right: are simu-In by chemical equilibria, which we assume to be attained in- case SFC, ocean-internal sources/sinks are adjusted to match the CO2 DIC lated by the global off-line atmospheric transport model TM3 stantaneously. The non-linear dependence of pm on Cm (surface-oceanHeimann andCO K2orner¨ partial, 2003 pressure) with observations. a spatial resolution Sea-air fluxes of ≈ are4◦ has been linearized in standard ways (Sarmiento and Gru- lat.then× calculated5◦ long. from× 19 the vertical estimated levels. partial The pressure model field. is driven by ber, 2006). The chemistry parameterization also contains a 6-h interannual meteorological fields derived from the Na- temperature-dependent factor, and contributions from sea- tional Centers for Environmental Prediction (NCEP) reanal- sonal variations in alkalinity and (Appendix A1.2). www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 CO 196 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability

Mixed-layer carbon budget. Changes in the spatio- 3.1 Overview temporal field of dissolved inorganic carbon in the ocean mixed layer need to be balanced by the sum of fluxes (Ap- To illustrate the characteristics of the quantities and the tem- pendix A1.3). As the sea–air flux itself depends on carbon poral scales involved in the scheme, Fig. 3 shows time se- concentration, this balance can be expressed as a linear first- ries of key quantities (as of run SFC, blue) at the exam- ◦ order differential equation. Its solution gives DIC concen- ple pixel around Station M in the Norwegian Sea (66 N, DIC DIC 2◦ E). The ocean-internal sources/sinks (bottom panel) are tration Cm as a function of the ocean-internal sinks fint (representing the total effect of biological conversion and relatively smooth, as the a-priori temporal correlations (Ap- vertical/horizontal advection/diffusion on mixed-layer car- pendix A2.2) suppress fast variations. The mixed-layer DIC bon concentration). An additional important item in the bud- concentration (panel above) responds to these ocean-internal get is the re-entrainment during mixed-layer deepening of fluxes and the emerging sea–air exchange. Being their tem- carbon left behind previously during mixed-layer shoaling, poral integral, the DIC concentration is slightly shifted in which we represent as a “history flux” fhist. We further con- phase with respect to the ocean-internal fluxes. In addition, sider the influence of freshwater fluxes. the DIC concentration is rising in response to the atmo- spheric CO2 increase. The surface-ocean CO2 partial pres- 2.3 Main adjustable degrees of freedoms sure (3rd panel from bottom) shares the rise and the varia- tions with the DIC concentration, but the seasonality is again DIC The ocean-internal carbon sources and sinks (fint ) at the slightly shifted because of the temperature and alkalinity ef- end of the described chain of process parameterizations is fects. Finally, the sea–air CO2 flux (top panel) is dominated the basic unknown to be adjusted to match the surface-ocean by the pCO2 variability; in addition, it shows high-frequency CO p 2 or atmospheric CO2 data (next section). It is treated (daily) variability due to wind speed and solubility changes. similarly to the unknown sea–air flux in the pure atmospheric Figure 3 also illustrates the dampening effect of the carbon- transport inversion (Fig. 2), including Bayesian a-priori spa- ate chemistry on sea–air exchange, as the seasonal amplitude tial and temporal correlations. The detailed specification and of the sea–air flux (top) is much lower than that of the ocean- further explanations are found in Appendix A2.2. interior flux (bottom). The spatial resolution and domain of the calculation is il- 2.4 Data constraints – base runs lustrated by Fig. 4, showing the amplitude of the mean sea- sonal cycle of pCO2 for each pixel. We will present results of two cases that differ in the data set used as main constraint (Fig. 2): 3.2 Data and model constraints SFC. In the main case used to create the primary prod- CO uct of this study, the ocean part of the diagnostic scheme The estimated p 2 and sea–air CO2 flux fields combine in- is fit to surface pCO2 data points from the SOCAT v1.5 formation from the partial pressure data and from the pro- database (Pfeil et al., 2012, http://www.socat.info/). Data cess parameterizations. To illustrate this, Fig. 3 also shows DIC = pre-treatment and further details are given in Table 1. SOCAT the a-priori state (thin dashed grey) defined by fint, pri 0 DIC = data cover about 6 million pixels/time steps globally within (Appendix A2.2, see bottom panel). The choice fint, pri 0 the calculation period (until 2007, see Supplement Figs. S7.3 means that, without data knowledge, it is equally likely to and S7.4 for data distribution). have an internal source or an internal sink at any given lo- ATM. As a comparison case used to assess the consistency cation and time, and thus corresponds to a state of no infor- DIC DIC between surface-ocean and atmospheric data, the diagnostic mation on fint . The variations in the prior values of Cm , CO CO2 scheme is fit to atmospheric CO2 mixing ratios measured ap- p 2 , and the sea–air fluxes fma , which then follow from proximately weekly or hourly at a set of sites by various in- the process parameterizations, only comprise responses to stitutions. Details are given in Table 1. variations in sea surface temperature and the other driving variables (including rising atmospheric CO2), but miss vari- ations in the ocean-internal sources/sinks. As this a-priori 3 Results and discussion pCO2 field contradicts the data (black dots), the estimation DIC procedure now adjusts the internal flux fint in such a way The main product of this study is a data-based estimate of that the data are matched as closely as possible (blue). Most the global spatio-temporal CO2 partial pressure field and ul- prominently, this reduces or – as at the example pixel of CO timately the sea–air CO2 fluxes. Here we characterize and Fig. 3 – even reverses the seasonal variations of p 2 (re- evaluate its robustness, errors, and information content. In flecting that thermal, biological, and physical effects partially addition, we compare these results with independent data and oppose each other). The fit at further locations is given below previous estimates. The consideration of interannual vari- (Sect. 3.4). ations is done in the companion paper (Rodenbeck¨ et al., For most pixels, the SOCAT data set only sporadi- 2013). cally contains data points (in particular, periods of dense

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data 23

Sea-air CO2 flux (PgC/yr) 0.04

0.02

0.00

-0.02

-0.04 pCO2 (uatm) 500 475 450 425 400

375 350 325 300 275 250

CO2 C. Rodenbeck¨ et al.: Global surface-ocean p and sea–air CO2 flux variability DIC concentration (umol/kg) 197 CO2140 CO C. R¨odenbeck et al.: Sea-air 21202 flux estimated from 2 partial pressure data 23 2100 Sea-air CO2 flux (PgC/yr) 0.04 2080

0.02 2060 2040 0.00 2020 -0.02 2000 -0.04 1980 pCO2 (uatm) 500 Internal sources/sinks (PgC/yr) 475 450 0.04 425 400

375 350 0.02 325 300 275 250 0.00 DIC concentration (umol/kg) 2140 2120 -0.02 2100 2080 2060 -0.04 2040 2020 1990 1995 2000 2005 2010 2000 1980 year (A.D.) Internal sources/sinks (PgC/yr)

0.04 ◦ ◦ Fig.0.02 3. Example time series (full daily resolution) at the model pixel at around Station M (66 N, 2 E) of the main quantities of the diagnostic CO2 CO2 DIC model0.00 (Figure 1): Sea-air CO2 fluxes (fma ), surface-ocean CO2 partial pressure (pm ), mixed-layer DIC concentration (Cm ), as well as DIC -0.02 ocean-internal carbon addition to (removal from) the mixed layer (fint ). Estimates have been obtained by fitting the diagnostic scheme to -0.04 SOCAT surface-ocean1990 CO21995 partial pressure2000 data (SFC2005, blue). The SOCAT2010 data points that happen to fall in the example pixel are shown as year (A.D.) black dots in the pCO2 panel. The Bayesian prior (f DIC = 0 and corresponding CDIC, pCO2 , and f CO2 , Appendix A2.2, Sect. 3.2) is also given m ◦ int◦ m m ma Fig.Fig. 3. 3.ExampleExample time time series series (full (full daily daily resolution) resolution) at the model model pix pixelel at around around Station Station M M (66 (66◦ N,N, 2◦ E) of the main main quantities quantities of of the the diagnostic diagnostic (thin dashed gray).CO2 CO2 DIC model (Figure 1): Sea-air CO2 fluxesCO (f2ma ), surface-ocean CO2 partial pressure (pCOm2 ), mixed-layer DIC concentration (CmDIC), as well as model (Fig. 1): sea–air CO2 fluxes (fma ), surface-ocean CO2 partialDIC pressure (pm ), mixed-layer DIC concentration (Cm ), as well as ocean-internalocean-internal carbon carbon addition addition to to (removal (removal from) from) the the mixed mixed layerlayer ( (ffDICint ).). Estimates Estimates have have been been obtained obtained by by fitting fitting the the diagnostic diagnostic sch schemeeme to to CO int SOCATSOCAT surface-ocean surface-ocean CO 2partialpartial pressure pressure data data ( (SFCSFC,, blue). blue). The The SOCAT SOCAT data data points that that happen happen to to fall fall in in the the examp examplele pixel pixel are are shown shown as as black dots in the pCO2 panel.2 The Bayesian prior (f DIC = 0 and corresponding CDIC, pCO2 , and f CO2 , Appendix A2.2, Sect. 3.2) is also given COm 2 intDIC = m DICm CO2 ma CO2 black(thin dots dashed in the gray).pm panel. The Bayesian prior (fint 0 and corresponding Cm , pm , and fma , Appendix A2.2, Sect. 3.2) is also given (thin dashed grey lines).

Fig. 4. Amplitude of the seasonal cycle of surface-ocean CO2 par- tial pressure (µatm) estimated by fitting the diagnostic scheme to the SOCAT data (run SFC). The amplitude is given as temporal CO2 standard deviation of the 1997-2009 monthly mean pm at each pixel.

Fig. 4. Amplitude of the seasonal cycle of surface-ocean CO2 partial pressure (µatm) estimated by fitting the diagnostic scheme to the Fig. 4. Amplitude of the seasonal cycle of surface-ocean CO2 par-CO2 SOCAT data (run SFC). The amplitude is given as temporal standard deviation of the 1997–2009 monthly mean pm at each pixel. tial pressure (µatm) estimated by fitting the diagnostic scheme to the SOCAT data (run SFC). The amplitude is given as temporal www.ocean-sci.net/9/193/2013/CO2 Ocean Sci., 9, 193–216, 2013 standard deviation of the 1997-2009 monthly mean pm at each pixel. 24 C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data

CO 198 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability

Constrained Extrapolated from pseudo data sampled from this field at the times and lo- pixel pixel cations of the SOCAT data (Appendix B). Constraining inter- annual variations requires more even data distribution in time CO2 CO2 than constraining seasonal variability; this density of data is fma fma K K only available is some areas of the ocean (Rodenbeck¨ et al., GE GE 2013). CO J p 2 p pCO2 pCO2 The spatio-temporal a-priori correlations also act to obs ⇐⇒ K K DIC smooth fint on small scales (Fig. 3, bottom), as they damp CC CC small-scale and short-term variations. This has been done DIC DIC because the available data do not have the spatial or tem- Cm Cm K K poral density required to constrain these fine-scale features CB CB adequately. Nevertheless, despite the smooth f DIC the corre- correlated int DIC DIC pCO2 fint fint sponding and sea–air CO2 flux fields (Fig. 3, top) do → adjustments−→ comprise fast variations as represented in the process param- eterizations, including responses to variations in temperature Fig. 5. Illustration of the information flow in the spatio-temporal Fig. 5. Illustration of the information flow in the spatio-temporal (changes in solubility and ) or in wind extrapolation through a-priori correlations. Where data values ex- extrapolation through a-priori correlations. Where data values ex- speed (e.g. accelerated depletion of an oversaturated mixed ist, pixels are directly constrained (left column, which is identical ist, pixels are directly constrained (left column, which is identical to to Fig. 2). Due to the a-priori spatial correlations (Appendix A2.2), layer in a high wind speed event, Bates et al., 1997). These Figure 2). Due to the a-priori spatial correlations (Appendix A2.2), adjustments to f DIC at the constrained pixel also change f DIC at effects already account for a considerable part of short-term adjustments to intf DIC at the constrained pixel also changeint f DIC at int CO2 int variations (see Sect. 3.4 below). Only variations related to neighbouring unconstrained pixels. p COis2 then calculated pas- neighbouring unconstrainedDIC pixels. p is then calculated pas- sively from the extrapolated f DICby the process parameterizations, fast ocean-internal processes, such as the sub-weekly varia- sively from the extrapolatedintf by the process parameterizations, using the local values of the drivingint fields (right column). tions of upwelling events or algal blooms, will be missing. In using the local values of the driving fields (right column). In data-dense areas, local adjustments to f DIC (both for constrained any case, the pseudo-data tests (Appendix B) confirm that the In data-dense areas, local adjustments tointf DIC (both for constrained and for extrapolated pixels) will need to compromiseint between sev- degrees of freedom in the diagnostic scheme provide suffi- and for extrapolated pixels) will need to compromise between sev- eral data constraints within the correlation radius (de-weighted with cient flexibility to follow variability as contained in the pCO2 eral data constraints within the correlation radius (de-weighted with distance due to the decay of the correlation, Fig. A1). As with fields from a biogeochemical process model simulation (as distance due to the decay of the correlation, Figure 13). As with any smoothing, this may be beneficial in suppressing outliers, but the model has less day-to-day variability than the real pCO2 alsoany adverse smoothing, in partly this suppressing may be beneficial small-scale in suppressing signals (the structures outliers, but also adverse in partly suppressing small-scale signals (the structures field, however, this test may underestimate errors from alias- from the driving data of the parameterizations are never smoothed ing such high-frequency variations into variations on the re- however,from the as driving the smoothing data of the only parameterizations acts on f DIC). In are data-poor never smo areas,othed int DIC solved timescales). therehowever, may exist as the pixels smoothing that do notonly have acts any on constrainedfint ). In data-poor pixel within areas, thethere correlation may exist radius; pixels then thatf doDIC notwill have stay any close constrained to the prior. pixel within intDIC 3.3 Robustness Inthe time, correlation extrapolation radius; happens then fint notwill only stay due close to the to a-priori the prior. correla- tions,In time, but extrapolationadditionally due happens to the notmemory only effect due to resulting the a-priori from correla- the relaxationtions, but time additionally of the budget due to equation the memory Eq. (A21 effect). resulting from the Figures 6 and 7 show the seasonality of the sea–air CO2 relaxation time of the budget equation Eq. (A21). flux and the surface-ocean CO2 partial pressure as estimated from SOCAT data by run SFC. The fields have been inte- grated/averaged over a set of regions splitting the ocean into regular sampling, as at Station M since 2006, are a rare ex- basins and latitude bands. ception). Thus, only a small fraction of pixels/time steps is To test how robustly the results are determined from the constrained directly in the described way. However, most data and to assess the impact of errors in the parameteri- parts of the space–time domain are still constrained indi- zations and their driving fields, individual important model rectly through extrapolation via the spatial and temporal a- elements have been changed in sensitivity runs: (i) increase priori correlations defined in Appendix A2.2. The mecha- and decrease of the a-priori uncertainty by a factor 2, leav- nism is illustrated in Fig. 5. In particular, as seasonal vari- ing more/less freedom for inverse adjustments, (ii) decrease DIC DIC ations in fint are allowed to be adjusted much more freely of the a-priori uncertainty of non-seasonal variations in fint than non-seasonal variations (equivalent to a-priori correla- by a factor of 2, (iii) increase in the spatial correlation lengths DIC tions between adjustments to fint at any given day of year by a factor of 3, (iv) increase and decrease of the global mean in each year, Appendix A2.2), the mean seasonal cycle is ex- piston velocity by 3.2 cm (range given by Naegler, 2009), and trapolated throughout the time period of the calculation. This (v) increase and decrease mixed-layer depth h by a factor of also means that data points in any year can contribute to con- 2. The resulting range of results (grey bands in Figs. 6 and strain this mean seasonality. Thus, though there are regions 7) is generally very narrow compared to the seasonal ampli- or periods too far away from data points for extrapolation, tude, in particular for pCO2 . In most regions, the sensitivity of larger-scale seasonal variations are constrained almost every- the sea–air flux is slightly larger than that of pCO2 , almost en- where in the ocean. This has been confirmed by assessing tirely due to the uncertainty of gas exchange (variation in pis- how well a given pCO2 field can be retrieved by the scheme ton velocity). Larger spread in pCO2 than in the sea–air flux

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 199

Fig. 7. Analogously to Fig. 6, monthly mean seasonal cycle of Fig. 6. Monthly mean seasonal cycle of sea–air CO2 fluxes, esti- CO2 mated by fitting the diagnostic scheme to SOCAT data (run SFC, surface-ocean partial pressure pm , estimated by blue). The grey shading around the standard result comprises sen- fitting the diagnostic scheme to SOCAT data (run SFC, blue). The sitivity cases listed in Sect. 3.3. Fluxes have been integrated over grey shading comprises the same sensitivity cases. For comparison, CO2 latitude bands split into ocean basins (including open ocean, coastal the pm climatology by Takahashi et al. (2009) is given (violet, areas and marginal seas; panels in geographical arrangement). Sea- with a small additive adjustment to transfer the nominal year 2000 sonal cycles have been calculated over the period 1997–2009; the average into a 1997–2009 average according to the rising atmo- CO2 first half year is repeated for clarity. spheric CO2 partial pressure field pa , Eq. A3).

is found mostly in the North Pacific and the temperate North ◦ ◦ Atlantic, entirely due to a few marginal areas (including pix- from scale mismatch between the model pixel (≈ 4 lat. × 5 els in the far north of the Pacific, or the Black Sea and the long.) and the point measurement, but likely also from the Caspian Sea added to the Atlantic region, see Fig. 4) where need to compromise between the signals of neighbouring lo- pCO2 cations in the fit: the statistics of the misfit in the surrounding the seasonality is exceptionally high due to high SST ◦ ◦ seasonality, but where gas exchange is low such that their area (Pacific 45 –90 N) reveal very small seasonal biases contribution to regional sea–air flux is nevertheless small. If (well below 1 µatm, Fig. 9). The distribution of residuals de- pCO2 is averaged only over the open ocean, the pCO2 spread viates somewhat from the Gaussian assumed mathematically becomes narrow in all regions (see Supplement Fig. S7.8). in the Bayesian inference by having wider tails, but is sym- metric. Example locations STM and – to a lesser degree – 3.4 Fit to the data BTM show close seasonal fit (Fig. 8), and agreement also in some high-frequency features (see below). A less successful The match of the estimated pCO2 field and measured data fit is found at Stratus (South Pacific) in an area of low data points somewhat depends on the considered location (Fig. 8). density (note that the independent data at this location are At the North Pacific example pixel (i.e. within the region of mainly from after the end of the SOCAT v1.5 data period; largest spread among the sensitivity cases in Fig. 7), the sea- thus the agreement relies on the extrapolating effect of the sonality is underestimated by some µatm. This may result dominating seasonal degrees of freedom (Appendix A2.2) www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 26 C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data

SOCAT Data Independent Data (not in SOCAT v1.5) N Pacific (155W,55N) 475 450 425 400 BTM (64W,32N) 375 450 CO 200350 C. Rodenbeck¨ et al.: Global425 surface-ocean p 2 and sea–air CO2 flux variability 325 26 C. R¨odenbeck et al.: Sea-air400 CO2 flux estimated from CO2 partial pressure data

300 375 pCO2 (uatm) 275 SOCAT Data Independent Data (not in SOCAT v1.5) 350 N Pacific (155W,55N) 250 475 pCO2 (uatm) 225 450 325 200 425 300

400 BTM (64W,32N) 1996 3751998 2000 2002 2004 450 2006 2007 2008 350 year (A.D.) 425 year (A.D.)

325 400

300 STM (2E,66N) 375 Stratus (85.62W,19.70S) pCO2 (uatm) 275 400 350 450 250 pCO2 (uatm) 375 225 325 425 200 300

1996 1998 2000 2002 2004 2006 2007 2008 350 year (A.D.) 400 year (A.D.)

325 STM (2E,66N) 375 Stratus (85.62W,19.70S) 400 450

pCO2 (uatm) 300 375 425pCO2 (uatm) 350 275 350 400 325 2006 325 2007 2008 375 2007 2008 2009 2010

pCO2 (uatm) 300 year (A.D.) pCO2 (uatm) 350 year (A.D.) 275 325 2006 2007 2008 2007 2008 2009 2010 CO2 Fig. 8. Comparison between the estimated yearp (A.D.)field (run SFC, blue) and measurements (blackyear (A.D.) dots) at test locations. Left: Data contained in SOCAT v1.5 (largerFig. 8. Comparison dots) at a between pixel the in estimated the NorthpCO2 field Pacific, (run SFC and, blue)the and pixel measurements around (black Station dots) at M test in locations. the North Left: Data Atlantic. contained Right: Two mooring ◦ CO◦ 2 ◦ ◦ sitesFig. in 8. theComparison North Atlanticin SOCAT between (BTM,v1.5 the (larger estimated (64 dots)W,32 at ap pixelN)) infield the and North (run South Pacific,SFC Pacific, and blue)the and pixel (Stratus measurements around (85.62 Station MW,19.70 in (black the North dots) Atlantic.S)) at by test Right:Sabine locations. Two et m al.ooring Left: (2010) data contained that are not in part ◦ ◦ ◦ ◦ SOCAT v1.5 (largersites dots) in the Northat a pixel Atlantic in (BTM,CO the2 North (64 W,32 Pacific,N)) and and South the Pacificpixel (Stratusaround (85.62 StationW,19.70 M inS)) the by North Sabine Atlantic. et al. (2010) Right: that are two not part mooring sites in the of SOCAT v1.5 (smaller dots). The p fieldCO has been picked at the times where the measurements exist (connected by straight lines for North Atlantic (BTM,of SOCAT 32◦ v1.5N,64 (smaller◦ W) dots). and TheSouthp 2 Pacificfield has (Stratus, been picked 19.70 at the◦ timesS, 85.62 where◦ theW) measurements by Sabine exis ett al. (connected(2010) by that straight are linesnot part for of SOCAT v1.5 clarity); the time axesclarity); haveCO the time been axes limited have been to limited respective to respective year yearss with with data. data. The The monthly monthly climatology climatology by Takahashi byet al. Takahashi (2009) (regridded et al. and (2009) (regridded and (smaller dots). The p 2 field has been picked at theCO times where the measurements exist (connected by straight lines for clarity); the time with added trend parallelwith added to trend the parallelatmospheric to the atmosphericCO2 increase,2 increase, violet) violet) is is also also given. given. See Supplementary See Supplementary material for comparisons material to for SOCAT comparisons data to SOCAT data axes have been limited(Figure S7.5) to respective and independent years data with (Figure data. S7.6) The at further monthlylocations. climatology by Takahashi et al. (2009) (regridded and with added trend (Figure S7.5) and independent data (Figure S7.6) at further locations. parallel to the atmospheric CO2 increase, violet) is also given. See Supplement for comparisons to SOCAT data (Fig. S7.5) and independent data (Fig. S7.6) at further locations.

0.05 Winter (JFM) Histogram Gaussian Summer (JAS) Histogram 0.05 0.04 Gaussian CAT v1.5 data set (Fig. 10, upper panels). In the extra-tropics Winter (JFM) Histogram Gaussian (example sites BSO or MOSEAN+WHOTS), many short- 0.03 Summer (JAS) Histogram term features are reproduced in run SFC, though mostly with Gaussian 0.04 reduced amplitude. Discrepancies (e.g. beginning of 2007 at 0.02 BSO) can largely be explained by the deviation of sea surface Rel. frequency (1/uatm) ◦ ◦ 0.03 0.01 temperature in our driving field – regridded to the ≈ 4 × 5 pixels around the stations – from the local temperature (lower 0 0.02 -100 -50 0 50 100 panels). Note that this discrepancy does not preclude that the pCO2 difference (uatm) pixel average of pCO2 is actually reproduced more accurately

Rel. frequency (1/uatm) than the local value, to the extent that our SST field repre- 0.01 Fig. 9. Histograms of the residuals of the fit to SOCAT data (i.e., CO sents the pixel’s average temperature; however, this cannot p 2 estimated by run SFC minus data, µatm, horizontal axis). Thin lines show the relative abundance (within bins of 5µatm) of be validated or falsified based on the point data. ◦ 0 residuals over all pixels from the North Pacific (North of 45 N) and In contrast to the extra-tropics, the high-frequency varia- -100 all time-50 steps with data in winter0 (Jan-Feb-Mar, 50 blue) or in s100ummer (Jul-Aug-Sep,pCO2 red). difference Respective (uatm) thick lines give a Gaussian of the tions in our estimate are unrelated or even opposite to those same mean and standard deviation. See Supplementary material, of the measurements in the tropics (example site TAO170W), Figure S5.2 for global maps of seasonal residuals. Fig. 9. Histograms of the residuals of the fit to SOCAT data (i.e. even though the SST field performs similarly as at the Fig.pCO 9. 2 Histogramsestimated by of run theSFC residualsminus of data, the fit µatm, to SOCAT horizontal data axis). (i.e., extra-tropical sites. While in the extra-tropics warmer SST CO2 CO2 p Thinestimated lines show by the run relativeSFC minus abundance data, (within µatm bins, horizontal of 5 µatm) axis). of leads to higher pm as reproduced by the solubility and ◦ residuals over all pixels from the North Pacific (north of 45 N) and CO2 Thin lines show the relative abundance (within bins of 5µatm) of chemistry parameterizations, tropical pm increases during ◦ residualsall time over steps all with pixels data from in winter the North (January–February–March, Pacific (North of 45 N) blue) and cold events, as in the upwelling of cold and carbon-rich or in summer (July–August–September, red). Respective thick lines all time steps with data in winter (Jan-Feb-Mar, blue) or in summer deep water. Such events would have to be reproduced as (Jul-Aug-Sep,give a Gaussian red). of the Respective same mean thick and lines standard give deviation. a Gaussian Seeof Sup- the plement Fig. S5.2 for global maps of seasonal residuals. high-frequency variations in the ocean-internal sources/sinks same mean and standard deviation. See Supplementary material, DIC fint , which however cannot be constrained from the SO- Figure S5.2 for global maps of seasonal residuals. CAT data set (potentially except at very few locations in pe- riods of high-frequency sampling). but misses any interannual anomalies). See the Supplement As similar conditions are found at other extra-tropical and (Sect. S5, Figs. S7.5 and S7.6) for more comparisons and tropical sites (Supplement Figs. S7.6, S7.7), we conclude statistics. that our results reproduce at least part of the real short-term In order to validate the high-frequency variations, we com- variations in the extra-tropics dominated by solubility and pare them to mooring observations independent of the SO-

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 201

Table 1. Data sets used as constraints or as driver data in the process parameterizations (Sect. A1).

Quantity Data set Reference Pre-treatment/resolution/remarks Used for Surface-ocean data:

CO2 pm SOCATv1.5 Pfeil et al. (2012), Data are used having WOCE-flag = 2 and valid fields for SFC http://www.socat.info/ fugacity, temperature, and salinity. Values below 200 µatm or above 600 µatm have been excluded as being local compared to the grid cells. Values have been transferred from fugacity to partial pressure by dividing by 0.996. Atmospheric data: CO CO c 2 Various Conway et al. (1994) CO2 mixing ratios c 2 . Data points are individual ATM measurement and many others flask pair averages (approximately weekly) or hourly averages, programs taken at 57 sites globally. Driver data: T OAFlux Yu and Weller (2007) Gridded, daily All ε OAFlux Yu and Weller (2007) Gridded, daily All S WOA 2001 Conkright et al. (2002) Gridded monthly climatology (interpolated, All values taken from A data set) A Lee et al. (2006) Gridded monthly climatology (interpolated) All γ DIC Egleston et al. (2010) Mean spatial pattern All DIC,Ref Cm GLODAP Key et al. (2004) Mean spatial pattern All h LOCEAN de Boyer Montegut´ et al. (2004) Temperature criterion; All monthly climatology (interpolated) u NCEP reanalysis Kalnay et al. (1996) All Meteo. NCEP reanalysis Kalnay et al. (1996) Driver for TM3 atmospheric transport model All Comparison data:

CO2 pm Takahashi et al. (2009) Gridded monthly climatology Sect. 3.5 PO4 Cm WOA 2005 Garcia et al. (2006) Gridded monthly climatology Sect. 3.7

Glossary: GLODAP = Global Ocean Data Analysis Project; LOCEAN = Laboratoire d’oceanographie´ et du climat: experimentations´ et approches numeriques;´ Meteo. = Meteorological variables as listed in Heimann and Korner¨ (2003); NCEP = National Centers for Environmental Prediction; OAFlux = Objectively Analysed air–sea Fluxes; SOCAT = Surface Ocean CO2 Atlas; WOCE = World Ocean Circulation Experiment; WOA = World Ocean Atlas; Math symbols see Table 2. chemistry. In the tropics, more realistic short-term variations tion1. In data-dense areas, results should not depend much on could potentially be obtained by parameterization of fast pro- the extrapolation method. DIC cesses (upwelling events) in fint . In most regions, the seasonal cycle estimated from SO- CAT by the diagnostic scheme is similar in phase to the pCO2 climatology by Takahashi et al. (2009) (Fig. 7), though the seasonal amplitude is somewhat larger in many regions. Even better agreement is found when only considering the 3.5 Comparison to the Takahashi climatology open ocean: the open-ocean seasonalities in the temperate North Pacific and Atlantic almost fully coincide (Supplement Fig. S7.8). Open-ocean agreement is also confirmed at sev- eral of our test locations (Fig. 8). The monthly mean seasonal cycle calculated from our spatio- Differences between our seasonalities and the Takahashi CO temporal p 2 field has been compared to the widely et al. (2009) climatology are somewhat bigger in areas of low used observation-based climatology by Takahashi et al. data density (e.g. the Southern Ocean), as expected. Com- CO (2009), which differs from our method in the p 2 data set parison to data at typical test locations tends to confirm the used (Lamont-Doherty Earth Observatory (LDEO) database, slightly larger seasonal amplitudes in our estimate (Fig. 8 Takahashi et al., 2010) and in the way of spatio-temporal ex- trapolation of the point data into pixels without data: while 1 As a further methodological difference, the diagnostic scheme the diagnostic scheme extrapolates the internal flux field uses each data point at its original sampling time; thus no transfer to DIC CO2 fint and then infers p from it via the process represen- a nominal year as in Takahashi et al. (2009) is needed and interan- CO2 tations (Fig. 5), Takahashi et al. (2009) extrapolates the p nual variations in atmospheric CO2 above the ocean are fully taken field itself, using a 2-dimensional advection–diffusion equa- into account. www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 CO 202 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability

Table 2. Mathematical symbols (also see Table 3). Some quantities are indexed by species or other distinctions explained in the text.

Quantity Unita Meaning A ppm (µmol m−2 s−1)−1 Atmospheric transport matrix, Eq. (A25) DIC −1 Cm µmol kg Mixed-layer DIC concentration DIC −1 Ca µmol kg DIC concentration in equilibrium with atmosphere cobs ppm Set of observed atmospheric mixing ratios cmod ppm Modelled mixing ratios, Eq. (A25) c0 ppm Initial mixing ratio −2 −1 fma µmol m s Sea-to-air tracer flux, Eq. (A17) −2 −1 fhist µmol m s History (re-entrainment) flux, Eq. (A20) −2 −1 fint µmol m s Ocean-internal tracer sources/sinks, Eq. (A18) F µmol m−2 s−1 Flux model matrix, Eq. (A26) f µmol m−2 s−1 Vector of flux discretized at pixels/time steps h m Mixed layer depth J 1 Cost function, Eq. (A27) Jc 1 Cost function, atmospheric data part, Eq. (A24) Jp 1 Cost function, surface-ocean data part, Eq. (A30) k m s−1 Piston velocity, Eq. (A2) L mol kg−1 atm−1 Solubility (including fugacity correction), Eqs. (A1) and (S1.1) pm µatm Partial pressure in mixed layer, Sect. A1 pa µatm Atmospheric partial pressure, Eq. (A3) pbaro atm Barometric pressure at ocean surface pdry atm Dry-air pressure at ocean surface pH2O µatm Saturation water vapor pressure above ocean, Eq. (S1.3) S ‰ Mixed-layer salinityb Sc 1 Schmidt number, Appendix S1.1 T ◦C Mixed-layer temperature (sea surface temperature, SST) t s Time coordinate ti, te s Start time, end time of inversion period u m s−1 Wind speed at 10 m above surface x 1 Vector of adjustable parameters, Eq. (A26) X ppm Dry-air molar mixing ratio in the atmosphere above the ocean z m Vertical coordinate β 1 Temperature factor of CO2 partial pressure, Eqs. (A5) and (A6) 0 1 Scaling of piston velocities, Eq. (A2) γ DIC µmol kg−1 Buffer factor (response factor), Eq. (A11) ε 1 Ice-free fraction of ocean surface (0 = ice-covered, 1 = ice-free)

a ppm = µmol mol−1; b for the present purposes, salinity in ‰ or on the Practical Salinity Scale are considered interchangeable. and Supplement Figs. S7.5 and S7.6). In particular, in the tology using its open-ocean values. For example, this con- North Pacific where differences between the two methods tributes to the differences in the temperate North Atlantic due are biggest, larger seasonalities are confirmed both at the in- to the Mediterranean (see Supplement, Fig. S7.5). dividual pixels (see Fig. 8 for a typical example) and in the Partly, the differences in Fig. 7 can be traced to the fact larger seasonal bias of about 8.5 µatm in the differences be- that the two estimates are based on different pCO2 data sets, tween Takahashi et al. (2009) and SOCAT data points in that though there is considerable overlap between the SOCAT and area (however, the largest contribution to the difference in the LDEO databases (this influence has explicitly been tested in pCO2 average of Fig. 7 comes from the northernmost pixels the Supplement Sect. S6). Sensitivity to data selection par- where data are too sparse for a conclusive validation). Data tially arises from interannual variations; in particular, differ- at Station M (Fig. 8) also confirm larger seasonality in the ences in the tropics may be related to the exclusion of El Nino˜ North Atlantic. data in Takahashi et al. (2009). Moreover, the synthetic data Differences between the two estimates are also bigger to- test (Appendix B) reveals slightly too high seasonality in our wards the coast. Note that part of these coastal differences results, though most of this mismatch is confined to certain come from the fact that, in order to cover the entire ocean marginal areas (that do not contribute much to the sea–air surface, we extrapolated the Takahashi et al. (2009) clima- flux due to low gas exchange).

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data 27

CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 203

BSO (19.5E,72.5N) MOSEAN+WHOTS (157.92W,22.77N) TAO170W (170W,2S) 375 425 500

350 400 475 325 450 375 300 425 350 pCO2 (uatm) 275 pCO2 (uatm) pCO2 (uatm) 400 250 325 375 2007 2008 2009 2006 year (A.D.) year (A.D.) year (A.D.) BSO (19.5E,72.5N) MOSEAN+WHOTS (157.92W,22.77N) TAO170W (170W,2S) 12 28 30 10 26 28 8 6 24 26

SST (degrC) 4 SST (degrC) SST (degrC) 2 22 24 2007 2008 2009 2006 year (A.D.) year (A.D.) year (A.D.)

CO2 Fig. 10. Top: Time series of surface-oceanCO2 p (µatm) at independent mooring sites (not part of SOCATv1.5): Barents Sea Opening◦ Fig. 10. Top: time◦ series◦ of surface-ocean p (µatm) at independent mooring sites (not part of SOCATv1.5):◦ ◦ Barents Sea Opening (72.5 N, ◦ (72.5 N,19.5 E, Arthun et al. (2012)), combined MOSEAN and WHOTS moorings◦ (22.77◦ N,157.92 W, Sabine et al. (2010)), and from 19.5 E, Arthun et al., 2012◦), combined◦ MOSEAN and WHOTS moorings (22.77 N, 157.92 W, Sabine et al., 2010), and from the TAO the TAO array (2 S,170 W, Sabine et al. (2010)). Comparison between the measurements (black dots) and the pCO2 field estimated by array (2◦ S, 170◦ W, Sabine et al., 2010). Comparison between the measurements (black dots) and the pCO2 field estimated by fitting to fitting to the SOCAT data (blue). The estimated field has been picked at the times where measurements exist; one year as been selected for the SOCAT data (blue). The estimated field has been picked at the times where measurementsCO2 exist; one year has been selected for each each site. Bottom: Sea surface temperature measured simultaneously with p (black dots) compared to our SST driving field at this pixel CO2 site. Bottom:(light sea grey). surface temperature measured simultaneously with p (black dots) compared to our SST driving field at this pixel (light grey).

Beyond seasonality, agreement is also found in the im- ment Sect. S4), this is largely related to the additional degrees plied long-term average global sea–air flux. The 1990–2009 of freedom in adjusting land-to-atmosphere fluxes: signals in average from run SFC is −1.2 PgC yr−1 (range between the atmospheric data are partially wrongly attributed to land −0.95 PgC yr−1 and −1.45 PgC yr−1 due to sensitivity cases or ocean. Consistently, best agreement between ATM and with lower/higher gas exchange rate, Sect. 3.3). Within un- SFC occurs in the southern regions away from land influ- certainties, this agrees well with the value of (−1.36 ± ence, while the large disagreement in the temperate North 0.6) PgC yr−1 recomputed from Takahashi et al. (2009) by Pacific is due to the short but strong sink period during sum- Wanninkhof et al. (2012, Table 2, “Net flux” and “Coastal mer, more typical for terrestrial boreal ecosystems. area”) using CCMP winds (Atlas et al., 2011). The results establish that oceanic and atmospheric data In summary, we take the similarity between the seasonali- are consistent in regions away from land influences, but that ties from the two methods based on similar data as an indica- estimates based on pCO2 data provide much stronger con- tion that the estimated pCO2 field in most regions is informed straints on internal ocean processes, in particular in ocean ar- by the data and not primarily method-dependent. eas closer to land. Combining the oceanic and atmospheric constraints is therefore both possible and beneficial (com- 3.6 Comparison of pCO2 -based and atmospheric pare Jacobson et al., 2007). It can be implemented by adding CO results together the cost function contributions of p 2 data (from SFC) and of atmospheric CO2 (from ATM). The estimated fields in this combined case (not shown) are almost identi- Figure 11 compares the SOCAT-based fit from Fig. 6 (run cal to the SFC run, as expected from the weakness of the SFC) with the fit of the diagnostic scheme to atmospheric atmospheric constraint on oceanic fluxes. However, as the at- CO data (run ATM). In the Southern Hemisphere, there is 2 mospheric data constrain the sum of ocean and land fluxes, broad agreement in the phasing and amplitude of the sea- improvements of the estimated land fluxes can be expected. sonal cycle. In particular, the difference between the two runs Of course, the combined run is essentially equivalent to us- is much smaller than their differences from the (common) ing the sea–air fluxes of the SFC run as priors in a subsequent prior, i.e. than the adjustments due to the respective data con- “classical” atmospheric inversion (similar as in the joint in- straints. Less agreement is found in the tropics and the North- version by Jacobson et al., 2007). Details will be discussed ern Hemisphere, in particular in the temperate latitudes. further in the context of interannual variations in the com- The ocean fluxes estimated from the atmospheric data panion paper (Rodenbeck¨ et al., 2013). are much less robust (more sensitive to changes in uncer- tain model details, not shown) than those estimated from the pCO2 data. As investigated by synthetic data testing (Supple- www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 28 C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data

28 C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data

CO 204 C. Rodenbeck¨ et al.: Global surface-oceanPacific 45N-90N p 2Atlanticand 45N-90N sea–air CO2 flux variability 0.5 0.2 0.4

0.3 0.2 0.1

Pacific 45N-90N Atlantic 45N-90N 0.1 Pacific 45N-90N Atlantic 45N-90N PO4 climatology 1.5 1.5 0.00.5 0.00.2 -0.10.4 1.0 1.0 Fit to SOCAT -0.20.3 -0.10.1 -0.30.2 0.5 0.5 Pacific 45N-90N Atlantic 45N-90N Prior -0.40.1 PO4 climatology 0.01.5 0.01.5 -0.50.0 -0.20.0 Fit to SOCAT -0.1 -0.51.0 -0.51.0 Pacific 15N-45N Atlantic 15N-45N Fit to SOCAT -0.2 Fit to atmospheric CO2 0.2 0.2-0.1 0.5 0.5 -0.3 -1.0 -1.0 Prior -0.4 -1.50.0 -1.50.0 0.1-0.5 0.1-0.2 Fit to SOCAT -0.5 Pacific 15N-45N -0.5 Atlantic 15N-45N Pacific 15N-45N Atlantic 15N-45N 1.5 1.5 Fit to atmospheric CO2 0.00.2 0.00.2 -1.0 -1.0 1.0 1.0 -1.5 -1.5 -0.10.1 -0.10.1 0.5 0.5 Pacific 15N-45N Atlantic 15N-45N 0.01.5 0.01.5 -0.20.0 -0.20.0

-0.51.0 -0.51.0 Pacific 15S-15N Atlantic 15S-15N Indian 15S-30N 0.2-0.1 0.2-0.1 0.2 -1.00.5 -1.00.5 -1.50.0 -1.50.0 0.1-0.2 0.1-0.2 0.1

-0.5 Pacific 15S-15N -0.5 Atlantic 15S-15N Indian 15S-30N Pacific 15S-15N Atlantic 15S-15N Indian 15S-30N 1.5 1.5 1.5 0.00.2 0.00.2 0.00.2 -1.0 -1.0 1.0 1.0 1.0 -1.5 -1.5 -0.10.1 -0.10.1 -0.10.1 0.5 0.5 0.5 Pacific 15S-15N Atlantic 15S-15N Indian 15S-30N 0.01.5 0.01.5 0.01.5 -0.20.0 -0.20.0 -0.20.0

1.0 1.0 1.0 Pacific 45S-15S Atlantic 45S-15S Indian 45S-15S -0.5 -0.5 -0.5 PO4 concentration (umol/kg) 0.2-0.1 0.2-0.1 0.2-0.1 -1.00.5 -1.00.5 -1.00.5 -1.50.0 -1.50.0 -1.50.0 0.1-0.2 0.1-0.2 0.1-0.2

Sea-air CO2 flux (PgC/yr) Pacific 45S-15S Atlantic 45S-15S Indian 45S-15S Pacific 45S-15S Atlantic 45S-15S Indian 45S-15S -0.5 -0.5 -0.5 PO4 concentration (umol/kg) 1.5 1.5 1.5 0.00.2 0.00.2 0.00.2 -1.0 -1.0 -1.0 1.0 1.0 1.0 -1.5 -1.5 -1.5 -0.10.1 -0.10.1 -0.10.1 0.5 0.5 0.5 Sea-air CO2 flux (PgC/yr) Pacific 45S-15S Atlantic 45S-15S Indian 45S-15S 0.01.5 0.01.5 0.01.5 -0.20.0 -0.20.0 -0.20.0

-0.51.0 -0.51.0 -0.51.0 Pacific 90S-45S Atlantic 90S-45S Indian 90S-45S 0.2-0.1 0.2-0.1 0.2-0.1 -1.00.5 -1.00.5 -1.00.5 -1.50.0 -1.50.0 -1.50.0 0.1-0.2 0.1-0.2 0.1-0.2

-0.5 Pacific 90S-45S -0.5 Atlantic 90S-45S -0.5 Indian 90S-45S Pacific 90S-45S Atlantic 90S-45S Indian 90S-45S 1.5 1.5 1.5 0.00.2 0.00.2 0.00.2 -1.0 -1.0 -1.0 1.0 1.0 1.0 -1.5 -1.5 -1.5 -0.10.1 -0.10.1 -0.10.1 0.5 0.5 0.5 Pacific 90S-45S Atlantic 90S-45S Indian 90S-45S 0.01.5 0.01.5 0.01.5 -0.20.0 -0.20.0 -0.20.0 Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul -0.51.0 -0.51.0 -0.51.0 -0.1 -0.1 -0.1 -1.00.5 -1.00.5 -1.00.5 -1.50.0 -1.50.0 -1.50.0 -0.2 -0.2 -0.2 Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul -0.5 -0.5 -0.5

-1.0 -1.0 -1.0 Fig. 12. Mixed-layer phosphate concentration (monthly climatol- Fig. 11.-1.5 Monthly mean seasonal-1.5 cycle of sea–air-1.5 CO2 fluxes as Fig. 12. Mixed-layer phosphate concentration (monthly climatol- Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul ogy, mean subtracted). Seasonality inferred from the SOCAT fit Fig. 6. Estimates by fitting the diagnostic scheme to SOCAT data (blue)ogy, mean assumes subtracted). all ocean-internal Seasonality carbon inferred sources from and the sinks SOCATare re- fit Fig. 11. Monthly mean seasonal cycle of sea-air CO2 fluxes as Fig- Fig. 12. Mixed-layer phosphate concentration (monthly climatol- (run SFC, blue) are compared to the fit of the scheme to atmospheric lated(blue) to assumes phosphate all sources ocean-internal and sinks carbon in Redfield sources proportions and sinks. are For re- ure 6. Estimates by fitting the diagnostic scheme to SOCAT data ogy, mean subtracted). Seasonality inferred from the SOCAT fit CO2 mixing ratio data (run ATM, magenta), as well as to the prior comparison,lated to phosphate the PO sources4 climatology and sinks from inthe Redfield World proportions. Ocean Atlas For (run SFC, blue) are compared to the fit of the scheme to atmospheric (blue) assumes all ocean-internal carbon sources and sinks are re- (noFig. data 11. constraint,Monthly thin mean dashed seasonal grey) cycle of sea-air CO2 fluxes as Fig- (WOA,comparison, Garcia the et al., PO 2006)4 climatology is given (green). from the The World grey uncerta Oceaninty Atlas CO2 mixing ratio data (run ATM, magenta), as well as to the prior lated to phosphate sources and sinks in Redfield proportions. For ure 6. Estimates by fitting the diagnostic scheme to SOCAT data band(WOA, aroundGarcia the et SOCAT-based al., 2006) is givenestimates (green). again The gives grey the uncertainty range of (no data constraint, thin dashed gray). comparison, the PO4 climatology from the World Ocean Atlas (run SFC, blue) are compared to the fit of the scheme to atmospheric sensitivityband around cases the as SOCAT-based in Figure 7; note estimates however again that calculat gives theed PO range4 is of (WOA, Garcia et al., 2006) is given (green). The grey uncertainty CO2 mixing ratio data (run ATM, magenta), as well as to the prior alsosensitivity sensitive cases to several as in Fig. further7; modelnote however elements that not calculated included in PO this is band around the SOCAT-based estimates again gives the range4 of 3.7(no Plausibility data constraint, of thin the dashed ocean-internal gray). carbon range.also sensitive to several further model elements not included in this sensitivity cases as in Figure 7; note however that calculated PO4 is sources and sinks range. also sensitive to several further model elements not included in this range. As the diagnostic scheme (in run SFC) is constructed to match the CO2 partial pressure data, any errors in the pa- rameterizations linking pCO2 and f DIC (namely carbonate long as we are only interested in the pCO2 (and sea–air flux) int DIC chemistry (including temperature or alkalinity dependence), fields, fint could be just considered a mathematical device or mixed-layer carbon budget (including freshwater and his- in the mapping of pCO2 . tory fluxes), see Fig. 1) will be compensated by spurious ad- On the other hand, if the ocean-internal fluxes are of inter- ditional adjustments to the ocean-internal carbon sources and est themselves (or in the envisaged extensions of the scheme sinks. On the one hand, this means that the estimated pCO2 using further data streams like oxygen coupled to carbon via DIC (and sea–air flux) fields will be only little affected by im- fint ), correctness of the parameterizations is needed. On perfections in these parameterizations. Even at pixels not di- the seasonal timescale considered here, the plausibility of CO2 DIC rectly constrained by a p data value but only indirectly the estimated fint field can be tested in the light of mixed- DIC via extrapolation of the fint field through its a-priori spatio- layer phosphate concentrations. As explained in Appendix C, temporal correlations (Fig. 5), any such bias will largely can- a modelled mixed-layer phosphate field can approximately DIC cel out (to the extent that it is the same at the directly con- be calculated from our fint estimate, and can be compared strained and the extrapolated pixels). This leads to the robust- to an observation-based monthly climatology (World Ocean CO ness of the p 2 estimate shown in Sect. 3.3. Therefore, as Atlas, Garcia et al., 2006). The inferred PO4 concentration

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 205 seasonality is broadly similar in phase and amplitude to An advantage of explicit parameterizations, as in the pro- the observations (Fig. 12), including characteristic region-to- posed mixed-layer scheme, is that the unknowns (here: in- region differences, though overestimation is seen especially ternal sources/sinks) are physical quantities which are poten- in the temperate latitudes. Given that part of the mismatch tially open for interpretation (compare Sect. 3.7). They also arises because the modelled PO4 concentration neglects non- enable us to use measurements of further quantities of the Redfieldian sources/sinks (as justified in Appendix C) and scheme (such as mixed-layer DIC concentration, nutrients, because of errors in the observed PO4 climatology which are etc.) as data constraints as well. These further data streams CO2 likely significant (and amplified by the Redfield ratio), we can either be used instead of the present ones (pm or at- take this agreement as broad confirmation of the employed mospheric data) to create alternative estimates for compari- parameterizations. son, or in conjunction with the present ones to potentially in- The carbonate chemistry parameterization (including its troduce further unknowns and so distinguish more processes driving fields, in particular alkalinity) can specifically be (e.g. carbonate chemistry). checked using DIC data. This is done in the companion paper All carbon sources/sinks in the ocean and on land are (Rodenbeck¨ et al., 2013). linked to oxygen sinks/sources, most of them in rather well- defined stoichiometric ratios. The diagnostic scheme can 3.8 The diagnostic mixed-layer scheme in the context therefore be extended to also make use of atmospheric or CO2 of other pm mapping methods oceanic measurements of oxygen (Rodenbeck¨ et al., 2013). Beyond the scheme as presented here, multi-linear re- CO2 The suitability of individual pm mapping methods (Sect. 1) gression could be brought in by expressing the unknown depends on the intended application. The primary motivation ocean-internal fluxes as a linear combination of (further) of the diagnostic mixed-layer scheme presented here was to driving variables and then match the data by adjusting time- be able to process different data streams (so far pCO2 and independent weights. This would make the estimation more atmospheric CO2 mixing ratio) in a consistent way, and to similar to data assimilation, with the benefits and risks men- combine them. The intention was to achieve this in a data- tioned above. Further, instead of linear combinations, pa- dominated way, with the least complex model possible. rameterizations of the ocean-internal processes involving ad- CO2 Statistical methods like the reference pm climatology by justable parameters could be used if available. This would Takahashi et al. (2009) or the interpolation by Jones et al. also open up the way to use even further data streams, such (2012b) have the advantage of being data-dominated, quite as ocean color from satellite observation. independent of driver data sets and process parameteriza- tions. Regression methods (multi-linear regression, neural networks) use driver data sets to bring in information about 4 Summary and conclusions modes of variability not resolved by the data, thus addressing the problem of data sparsity; still, they leave flexibility in the We presented a temporally and spatially resolved estimate CO2 of the global surface-ocean CO partial pressure field and way of how pm and these driving fields are related. Finally, 2 data assimilation into process models then also prescribes the sea–air CO2 flux, obtained on the basis of CO2 partial this relation analytically, which brings in process knowledge pressure data from the SOCAT database. In order to inter- as a further constraint, with the risk of suppressing real vari- polate the point data into a spatio-temporal field and to add ability not captured in the chosen parameterizations. short-term variations not constrained from the data, a simple The diagnostic scheme is similar to the regression meth- model of surface-ocean biogeochemistry, including represen- ods, thereby also reproducing some of the short-term vari- tations of sea–air gas exchange, solubility, carbonate chem- ability (Sect. 3.4). It is more rigid, like data assimilation, for istry, mixed-layer DIC budget, and seasonal re-entrainment some processes that may be considered relatively well known (“history”), was fitted to the observations. The inverse proce- (sea–air gas exchange, carbonate chemistry, and mixed-layer dure was similar to atmospheric inversion calculations, using mass conservation). It is more free, on the other hand, for the spatial and temporal a-priori correlations to extrapolate the ocean-internal sources and sinks (it is independent of any as- data information. sumptions except for smoothness as needed to regularize the Focusing first on seasonal variations, the following con- mathematics), i.e. it is purely statistical here. The smoothness clusions were drawn: assumption represents the mechanism of spatio-temporal in- terpolation2. – The diagnostic scheme can be robustly fit to SOCAT pCO2 data, to estimate the spatio-temporal pCO2 field and sea–air CO flux (run SFC). 2 This results in different notions of output resolution: Formally, 2 the resolution is daily and ≈ 4◦ × 5◦ pixels, but part of this fine- CO scale variability is only coming from the driving data, while only – In terms of seasonality, the resulting p 2 field agrees the larger scales (> monthly and > 1500 km areas) are actually in- well with the climatology by Takahashi et al. (2009), formed by the SOCAT data. confirming that the estimates are informed by the data. www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 CO 206 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability

Differences mainly occur in coastal areas and marginal A1.1 Solubility and gas exchange seas, as well as in regions where data density is low. Sea–air flux is expressed in the usual way in terms of a partial – The atmospheric CO2 constraint on surface-ocean bio- pressure difference, geochemistry (run ATM) is not robust in general, but   CO2 CO2 CO2 CO2 CO2 consistent with the oceanic constraint (run SFC) in ar- fma = k %L pm − pa (A1) eas away from land influences. This confirms that it is appropriate and advisable to use the oceanic constraint (atmospheric sign convention: positive = source to atmo- CO2 as oceanic prior in an atmospheric inversion, to improve sphere), where pa is the CO2 partial pressure in the at- CO2 the inferred land fluxes. mosphere, and pm the CO2 partial pressure in the ho- mogeneous mixed layer3. The temperature-dependent solu- – The diagnostic scheme involves interpretable quanti- bility LCO2 relates the partial pressure difference to a dis- ties, as indicated by the comparison of predicted and solved CO2 concentration difference. The seawater density % observed seasonal phosphate variations. The scheme is needed to relate volume-based and mass-based quantities. can be extended by parameterizations of ocean-internal The diffusive resistance to sea–air gas exchange is described sources and sinks, that allow us to use further data by the piston velocity kCO2 . We use the formulation accord- streams to constrain individual biogeochemical pro- ing to Wanninkhof (1992): cesses. −0.5 CO 2  CO Ref – Using synthetic data tests (as in Appendix B), the diag- k 2 = 0u Sc 2 /Sc , (A2) nostic mixed-layer model can also be applied to B as- calculated from 6-hourly NCEP wind speed u (Kalnay et al., sess possible impacts of additional data on constraining 1996) and Schmidt number Sc given in the Supplement ocean biogeochemistry, helping to design future obser- Sect. S1.1. The global scaling factor 0 is chosen such that vation strategies. CO2 the global mean CO2 piston velocity kGlob matches the value The results also include interannual variations, which will from Table 3 given by Naegler (2009) (average of values by be considered in a companion paper (Rodenbeck¨ et al., Naegler and Levin (2006), Krakauer et al. (2006), Sweeney 2013). et al. (2007), and Muller¨ et al. (2008) inferred from radiocar- CO The gridded p 2 and sea–air CO2 flux estimates will be bon data). made available for download in digital form from the Jena in- The atmospheric partial pressure version website, www.bgc-jena.mpg.de. Regular updates are pCO2 = XCO2 pdry planned. a (A3) inherits variability not only from atmospheric dry-air pres- sure pdry (Supplement Sect. S1.1) but also from the atmo- Appendix A CO spheric CO2 molar mixing ratio X 2 just above the ocean surface, which has considerable increasing trend, seasonality, Model documentation and – even over the ocean – synoptic variability. It is taken For reference, all mathematical symbols are listed in Tables 2 from the atmospheric CO2 mixing ratio fields provided by the and 3. Jena inversion s81 v3.4 (see http://www.bgc-jena.mpg.de/ ∼christian.roedenbeck/download-CO2/). This XCO2 field has A1 Ocean process parameterizations been obtained from a forward run of the atmospheric trans- port model (TM3, resolution ≈ 4◦ × 5◦ × 19 layers), with This section details the equations used to express sea–air flux surface fluxes from a classical atmospheric transport inver- CO as a function of partial pressure (Sect. A1.1), partial pres- sion based on atmospheric CO2 measurements. The X 2 sure as a function of mixed-layer concentration (Sect. A1.2), field constructed that way is compatible with measured CO2 and mixed-layer concentration as a function of ocean-interior mixing ratios at the sites used, and therefore realistic at least source/sinks (Sect. A1.3), going down the chain of parame- in the rising trend and large-scale seasonality; the sensitivity terizations in Fig. 1. of the results presented here to the XCO2 field is small (see The variables of the following parameterizations are Supplement Sect. S2).4 spatio-temporal fields, with a temporal resolution of 1 day 3 More precisely, it is the CO partial pressure in air in equilib- and a horizontal resolution of ≈ 4◦ × 5◦, i.e. the equations 2 rium with sea water of the present chemical composition. are applied to every pixel of the atmospheric transport model. 4 Of course, the use of inversion-based XCO2 already creates Vertically, we consider an oceanic mixed layer assumed to be some dependence of our prior from atmospheric data. In run ATM, perfectly homogeneous in temperature and chemical compo- this formally violates assumptions of Bayesian inference, which sition, exchanging inorganic carbon with the overlying atmo- however could only affect calculations of a-posteriori covariances. sphere, the underlying ocean interior, neighbouring pixels, Even there, we would not expect problems due to the small sensi- and the organic carbon pool. tivity on XCO2 .

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 207

Table 3. Physical constants, and geophysical quantities assumed constant.

Quantity Value Meaning −1 −1 ◦ cp 3993 J kg K Specific heat capacity of sea water (value at 20 C) CO2 −1 = · −5 −1 kGlob 16.5 cm h ( 4.58 10 m s ) Global mean piston velocity of CO2 (Naegler, 2009) −1 −1 Rgas 8.3144 J mol K Gas constant ◦ T0 273.15 K Absolute temperature at 0 C 3 −1 Videal 0.0224136 m mol Molar volume of an ideal gas % 1025 kg m−3 Density of sea water

Table 4. Adjustable degrees of freedom.

Quantity A-priori Spatial Temporal value resolution resolution Carbon cycle components: DIC + fint Ocean-internal DIC sources/sinks Zero corr. pix. lt seas CO2 fnee Land NEE (run ATM only) corr. pix. lt, seas, var Technical degrees of freedom: DIC,ini 1Cm Initial condition of budget equation Zero corr. pix. – CO2 fini Flux pulse for atmospheric initial condition (run ATM only) Zero 3 lat. bands –

Glossary: lt, – = constant (long-term); seas = seasonal; var = interannual and short-term variability; corr. pix. = correlated pixels; math symbols see Table 2.

A1.2 Carbonate chemistry Even though this relation has been fitted to data in the tem- perature range of about 2 to 25 ◦C, it is used here for all tem- The CO2 partial pressure in the mixed layer is a function of peratures. DIC CO2 mixed-layer DIC concentration Cm , alkalinity A, tempera- The remaining dependences of pm are non-linear, but ture T , and salinity S: monotonic and can be linearized to good approximation   (Sarmiento and Gruber, 2006). Linearization of the depen- CO2 CO2 DIC DIC pm = pm Cm ,A,T,S . (A4) dence on Cm is important here in order to be able to use the fast minimization algorithm of Rodenbeck¨ (2005) In the diagnostic scheme, the relevant dependence is that on (Sect. A2.1). We consider deviations (1) of the driver vari- DIC ables from temporally constant reference values (superscript Cm , while the influences of A, T , and S are calculated a- priori from driver data (Table 1). While the exact function is “Ref”, see Sect. A1.4 below): a 5th order polynomial (Zeebe and Wolf-Gladrow, 2001), we CDIC = 1CDIC +CDIC,Ref (A7) use existing approximations, as described in the following. m m m Ref At first, we single out the temperature dependence into a A = 1A +A (A8) CO2 Ref temperature-dependent factor and pm at a reference tem- S = 1S +S (A9) perature T Ref: Linearization of Eq. (A5) around these references gives     CO2 = Ref · CO2 DIC Ref CO2 Ref pm β T,T pm (Cm ,A,T , S). (A5) pm = β T,T × (A10)

" CO2 CO2 CO2 # ∂pm ∂pm ∂pm For iso-chemical sea water (i.e. constant CDIC, A, and S), . pCO2,Ref + 1CDIC + 1A + 1S m m ∂CDIC m ∂A ∂S Takahashi et al. (2009) experimentally determined the ratio m CO2 of pm of samples at different to be The DIC sensitivity has been calculated from the response factors γ DIC by Egleston et al. (2010): n −1 β(T1,T2) = exp 0.0433K (T1 − T2) (A6) CO2 CO2,Ref ∂pm pm −5 −2  2 2o = . −4.35 × 10 K T − T . DIC DIC (A11) 1 2 ∂Cm γ www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 CO 208 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability

The alkalinity sensitivity has been calculated using the ap- consider that changes in the spatio-temporal field of dis- proximation solved carbon in the mixed layer need to be balanced by the sum of fluxes: CO2 CO2,Ref ∂pm ∂pm = × (A12) d 1   ∂A ARef DIC CO2 DIC DIC DIC Cm = −fma + fhist + ffrw + fint . (A18) −(ARef)2 dt h%  DIC,Ref Ref Ref DIC,Ref 2Cm − A A − Cm Fluxes (carbon exchange per unit horizontal area and unit CO2 time) comprise loss through sea–air exchange fma , seasonal (Eq. (8.3.17) of Sarmiento and Gruber, 2006, note missing re-entrainment f DIC and freshwater effects f DIC explained minus sign). The salinity sensitivity (direct effect on chemi- hist frw below, and fluxes f DIC due to all other ocean-internal pro- cal equilibrium only) has been set to int cesses (biological conversion, vertical advection/diffusion, CO2,Ref CO2,Ref or horizontal advection/diffusion). This formulation assumes pm pm = (A13) that fluxes are immediately diluted vertically throughout the ∂S SRef mixed layer. Mixed-layer depth h is prescribed as a climatol- (Eq. (8.3.5) of Sarmiento and Gruber, 2006). Any sea- ogy by de Boyer Montegut´ et al. (2004) (downloaded from sonal and interannual variations in the proportionalities http://www.locean-ipsl.upmc.fr/∼cdblod/mld.html using the CO2 DIC ∂pm /∂Cm , etc., are neglected. The variations in alkalin- MLD climatology mldT02 sk.nc according to a temperature ity, A, and salinity, S, are fixed according to input data sets criterion; chosen because the MLD climatology based on a CO2 (Table 1). Their effect on pm is illustrated in the Supple- density criterion has incomplete coverage, and there is no ment Fig. S7.2. Note that the effect of the alkalinity vari- MLD product directly based on data that also represents in- ations, which are strongly related to freshwater fluxes, is terannual variations). The density % of surface water (as- DIC mainly offset by the freshwater effects on DIC considered sumed constant, Table 3) is needed as Cm is mass based − below (Eq. A19). (given in µmol kg 1), while fluxes are volume/area based. Equation (A10) provides the desired linear link between Two specific ocean-internal processes are not considered DIC 5 CO2 DIC part of f but are parameterized explicitly in the budget changes in pm and Cm according to carbonate chemistry. int For simplification in the further use, we combine the gas ex- (Eq. A18): change and chemistry formulas: we first define an apparent – The freshwater effect f DIC describes the dilution or DIC concentration in equilibrium with the atmosphere, frw enhancement of the mixed-layer DIC concentration by −1 CO2 ! " CO2 precipitation, evaporation, river freshwater input, or sea- ∂pm pa 1CDIC = − pCO2,Ref (A14) ice formation/melting. Analogously to the often used a DIC Ref m ∂Cm β T,T salinity normalization (Sarmiento and Gruber, 2006), CO2 CO2 # we assume that freshwater does not contain either car- ∂pm ∂pm . − 1A − 1S bon or salt (and that all salinity variations are related ∂A ∂S to freshwater fluxes), and calculate the freshwater effect Using this definition, Eq. (A10) becomes on DIC from salinity changes:

CO2 DIC,Ref   ∂pm   C dS .pCO2 = pCO2 + β T,T Ref 1CDIC − 1C (A15) f DIC = h% m . m a DIC m a frw Ref (A19) ∂Cm S dt Now also defining an “apparent piston velocity” of DIC as This will tend to overestimate the effect to the extent that freshwater input does carry carbon. CO2 DIC CO CO  Ref ∂pm k = k 2 · L 2 β T,T , (A16) DIC DIC – The “history” flux f is part of the carbon exchange ∂Cm hist between the mixed layer and the underlying ocean. Con- Eqs. (A1) and (A15) can be combined and simplified into sider the hypothetical case that no sources or sinks are   acting below the mixed layer. Then, as soon as the CO2 DIC DIC fma = k % 1Cm − 1Ca , (A17) mixed layer is deepening, it re-entrains water that was left behind previously during mixed-layer shoaling. As expressing the sea–air CO2 flux as a linear function of the mixed-layer concentrations will generally have changed surface DIC concentration. 5 The explicit treatment of these processes is not strictly needed A1.3 Mixed-layer DIC budget if the only purpose of the scheme was mapping of pCO2 , because DIC they would otherwise be absorbed into fint during the fit to the To finally establish the link between the mixed-layer DIC data. However, splitting them off allows easier interpretation of DIC concentration and the ocean-interior sources and sinks, we fint (Sect. 3.7), and also prepares for the envisaged link to oxygen.

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 209

meanwhile, this re-entrainment is equivalent to a car- Solving this equation would be numerically involved but for- DIC,Ref bon flux which is proportional to the difference between tunately is not needed in practice: Firstly, Cm only de- DIC the current concentration Cm (t) and that at time tprev termines the absolute levels of DIC (Eq. A7) not considered CO2 when the mixed layer was as deep for the last time. We here, but does not directly affect our target quantities pm or call this the “history” flux. It can be written as sea–air fluxes (as it cancels out both from Eqs. (A15), (A17),   dh and (A21); we nevertheless use a data-based field because f DIC(t) = % · CDIC t  − CDIC(t) · 2 (A20) DIC,Ref hist m prev m dt a minor influence of Cm arises through Eq. (A12) (al- kalinity sensitivity) and Eq. (A19) (freshwater effect)). Sec- CO2,Ref The Heaviside step function 2 ensures that fhist only ondly, the partial pressure reference value pm has no CO2,Ref acts during mixed-layer deepening (dh/dt > 0). The effect either: Though changes in pm shift the long-term history flux is non-zero in the long-term mean: The mean of 1Ca through Eq. (A14), the same shift then happens deeper waters that get disconnected from the shoaling DIC to 1Cm through Eq. (A21) with Eq. (A22), such that it mixed layer will preserve the high DIC concentrations again cancels out from Eqs. (A15) and (A17). The choice of of the winter season. A budget equation that does not in- CO2,Ref CO2 pm is thus arbitrary (we use pa (ti) as for numerical clude the re-entrainment of this amount of DIC during CO reasons a value on the order of the actual p 2 is desirable). mixed-layer deepening would imply a systematic down- m ward “leak” of DIC. The history flux is thus essential to A2 Inversion DIC balance the long-term mean internal sources/sinks fint CO2 with the mean sea–air exchange fma (plus the mean The estimation procedure of this paper is derived from the DIC accumulation in the mixed layer itself). This has linear Bayesian atmospheric inversion (Newsam and Enting, been verified numerically, and Sect. S3 proves this bal- 1988). The implementation is based specifically on the atmo- ance analytically. spheric transport inversion (“Jena CO2 inversion”) described Any changes of the DIC concentration below the mixed in Rodenbeck¨ (2005) and reviewed in Sect. A2.1. This atmo- layer will lead to additional entrainment fluxes, forming spheric transport inversion is extended by representing sea– an important part of f DIC. air fluxes by the ocean parameterizations (Sect. A2.2) and int by using the pCO2 constraint instead of or in addition to the Substituting Eq. (A17) into Eq. (A18) gives atmospheric constraint (Sect. A2.3). d kDIC kDIC 1CDIC = − 1CDIC + 1CDIC (A21) A2.1 The classical atmospheric transport inversion dt m h m h a 1 DIC DIC 1 DIC 1 DIC The atmospheric inversion is based on a set of observed mix- + fhist {1Cm } + ffrw + fint , h% h% h% ing ratios cobs (time series at observation sites, see Table 1). where we further exploited that the temporally constant ref- The inversion calculation seeks fluxes f that lead to the best DIC,Ref erence Cm disappears from the left-hand side. The ref- match between cobs and corresponding modelled mixing ra- erence also cancels out in the history flux (Eq. A20), which tios cmod(f ), in the sense that the value of the cost function DIC is thus a linear functional of 1Cm (t). Equation (A21) thus st DIC represents a 1 order differential equation in 1Cm that can DIC 1 T −1 be solved (numerically) for given “forcing” f . At the be- Jc = (cobs − cmod) Q (cobs − cmod) (A24) int 2 c ginning ti of the time period, we use the initial condition DIC DIC,ini is minimal (superscript T means vector transpose). The (di- 1Cm (ti) = 1Ca(ti) + 1Cm . (A22) agonal) matrix Qc introduces a weighting among the con- DIC,Ref centration values, involving assumed measurement uncer- Up to a constant offset (determined by Cm ) and an DIC,ini tainty, location-dependent model uncertainty, and a data den- initial transient (determined by 1Cm , see Supplement Sect. S1.2), Eq. (A21) thus provides a linear dependence of sity weighting, as described in Rodenbeck¨ (2005). The mod- CDIC on f DIC for every pixel. elled mixing ratios, taken at the same time and location as the m int observations, are simulated by a numerical transport model A1.4 Reference values (here, the TM3 model, Heimann and Korner¨ , 2003), which can formally be written as DIC,Ref Ref Ref Ref For the reference values Cm , A , T , and S , we use the long-term averages of the respective driver data sets cmod = Af + c0 (A25) (Table 1) for each pixel. The partial pressure reference value CO2,Ref with the transport matrix A and the initial concentration c0. pm in Eq. (A10) formally follows from the functional relation Since the atmospheric data at the discrete set of sites can-   not fully constrain the flux field at all pixels and time steps, CO2,Ref CO2 DIC,Ref Ref Ref Ref pm = pm Cm ,A ,T ,S . (A23) Bayesian a-priori constraints are introduced to regularize the www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 CO 210 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability estimation. They are implemented here by writing the fluxes (superscript “lt”), mean seasonality (“seas”), and as a function of a set of parameters x: non-seasonal variations (“var”, interannual and high- frequency variations). The specification of the priors, f (x) = f pri + Fx. (A26) a-priori sigmas, and correlation structure is taken from the Jena inversion v3.2 (as documented in Rodenbeck¨ In this linear flux model (Rodenbeck¨ , 2005), each parameter (2005)√ except for a-priori uncertainties tightened by the in x acts as a multiplier for one of the columns of the ma- factor 8, and length scales of a-priori spatial correla- trix F to be explained below. A-priori, the parameters x are tions increased by a factor 3 in longitude direction and assumed to have zero mean and unit variance and to be un- by 1.5 in latitude direction). correlated, expressed by adding the Bayesian cost function contribution – f fos: burning. Only a fixed (prior) term, 1 T as in the Jena inversion v3.2 (yearly emissions from J = Jc + x x. (A27) 2 the EDGAR v4.0 database, update of Olivier and Berdowski, 2001). This is equivalent to assuming fluxes with a-priori mean f pri and a-priori covariance matrix Q = FFT. f ,pri – f : purely technical component. Flux pulse at begin- The specification of the flux model elements f and F is ini pri ning of inversion period to adjust initial atmospheric detailed in Rodenbeck¨ (2005). The flux is first written as a mixing ratio field (see Rodenbeck¨ , 2005). sum of contributions from several components: The numerical minimization of the cost function is done f = f + f + f + f (A28) ma nee fos ini by Conjugate Gradient descent with re-orthonormalization (Rodenbeck¨ , 2005). (f ma = sea–air flux, f nee = terrestrial net ecosystem ex- change (NEE), f fos = fossil fuel burning emissions, f ini = flux pulse at beginning of inversion period to adjust initial A2.2 Extension: using ocean parameterization atmospheric mixing ratio field). Each component i is repre- sented by a structure as Eq. (A26), i.e. The inversion set-up used in this study is identical to the Jena inversion system as of Sect. A2.1 in most respects, except X  f (x) = f pri,i + Fixi , (A29) that the sea–air flux f ma is not estimated directly but im- i plemented as a function of the ocean-internal flux f int using the parameterizations of Sect. A1. Then, instead of f ma, the where the vectors x denote subsets of parameters in x, and i ocean-internal flux f int is adjusted to match the data con- the matrices F denote the corresponding groups of columns i straint. The detailed flux-model specifications for f int are of F. The columns of Fi represent spatio-temporal base func- similar to those for the sea–air flux in the atmospheric in- tions out of which the spatially and temporally varying ad- version: justments to the flux field f i are composed. In space, the base functions represent flux elements localized at each of the grid DIC = – The a-priori state has been defined by fint,pri 0 (no in- cells of the transport model. These elements overlap each ternal sources or sinks). The “error” of this prior model other by exponential tails, acting to smooth the flux field. In DIC = fint,pri 0 is then identical with the process fluxes time, the base functions represent Fourier modes. Smoothing themselves. Therefore, the a-priori uncertainty around is achieved by down-weighting the higher-frequency modes. this prior should reflect the expected amplitude of vari- In addition to the a-priori correlations implemented that ations. In the standard set-up, we make no assump- way, all columns of Fi are proportional to a spatio-temporal tions about spatial structure of the amplitude, and take a weighting allowing flux adjustments in areas of activity of constant per-ocean-area uncertainty over all the ocean. the component (e.g. over the ocean for f ma) while suppress- No adjustments are allowed in ice-covered regions (see ing flux adjustments elsewhere. Supplement S1.3). The a-priori uncertainties are scaled The detailed flux model settings for the components f i of such that the implied a-priori uncertainty of the globally − the surface-to-atmosphere CO2 flux in the “Jena inversion” integrated sea–air flux is 1200 Tmol yr 1 (with respect are to 3-monthly anomalies).

– f ma: sea–air flux, being of central interest here. In the – Bayesian a-priori spatial and temporal correlations have pure transport inversion as in Rodenbeck¨ (2005), this DIC component is estimated directly by decomposition into been implemented that dampen variability in fint on flux elements as just described. scales smaller than around 640 km (longitude), 320 km (latitude), and about 2 weeks (time), Fig. A1. The cho- – f nee, lt + f nee, seas + f nee, var: terrestrial net ecosystem sen e-folding lengths are one-third of the ocean flux cor- exchange (NEE), further split into long-term flux relations in the standard Jena Inversion v3.4 (update of

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data 29

Pacific 45N-90N Atlantic 45N-90N 425 425 400 400 Prior 375 375 Fit to SOCAT-sampled PlankTOM5

350 350 Fit to PlankTOM5 (all pixels) CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability325 211 325 PlankTOM5 ("Known truth") 300 300 Pacific 15N-45N Atlantic 15N-45N 425 425 400 400 375 375 350 350 325 325 300 300 Pacific 15S-15N Atlantic 15S-15N Indian 15S-30N 425 425 425 400 400 400 375 375 375 350 350 350

pCO2 (uatm) 325 325 325 300 300 300 Pacific 45S-15S Atlantic 45S-15S Indian 45S-15S 425 425 425 400 400 400 375 375 375 350 350 350 325 325 325 300 300 300 Pacific 90S-45S Atlantic 90S-45S Indian 90S-45S 425 425 425 400 400 400 375 375 375 350 350 350 325 325 325 300 300 300 ◦ ◦ Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Fig. A1. Top: map of a-priori correlation coefficients with respect to an example pixel at (70 E, 49 S), illustrating how inverse adjustments DIC to fint de-correlate withFig. distance. 13. Bottom:Top: Map a-priori of correlation a-priori coefficients correlation with coefficients respect to an with example respect time step in the middle of the inversion period; in additionto to an the example de-correlation pixel away at (70 from◦ theE,49 example◦ S),illustrating time step, correlations how inverse rise again adjust- each year due to the dominance of the seasonal cycle. DIC ments to fint de-correlate with distance. Bottom: A-priori correla- CO2 tion coefficients with respect to an example time step in the middle Fig. 14. Testing the capacity to retrieve the p field from the SOCAT data set: Fit of the diagnostic scheme to the pCO2 field Rodenbeck¨ , 2005)6,of and the similar inversion to those period; of pCO in2 found addition to thethe de-correlation regional average away (test from not shown). Shorter correla- from correlation analysisthe example along cruise time tracks step, by correlationsJones risetions again also each prevent year implausible due to the leakagesimulated of interannual by the PlankTOM5 biogeochemical model (Buitenhuis et al. (2012a). Thedominance correlations of are the seen seasonal as a mathe- cycle. variability of the tropical Pacific northwardet al., and 2010) south- (combined with Takahashi et al. (2009)’s seasonality) matical device to stabilize the estimation by suppress- ward (see Rodenbeck¨ et al., 2013). Onsub-sampled the other hand, at locations/times of the SOCAT measurements (blue), ing noise which results, for example, from unequal sam- longer correlations may have been beneficialcompared in theto data- the simulated pCO2 field itself (“known truth”, violet). pling (they represent the main mechanism by which the sparse Southern Ocean as they leadFor to slightly further better comparison, a fit of the scheme to the simulation at ev- information from the discrete atmospheric or oceanic performance in synthetic-data tests (analogously to Ap- data points is extrapolated into space and time, see pendix B, test not shown). In any case,ery simple grid distance- cell and time step (“complete information”, green) and the Fig. 5 for illustration). The correlation scales need to based de-correlation necessarily representsBayesian a crude prior ap- (“no data information”, thin dashed gray, partially CO be large enough to accomplish this extrapolation, but proximation to the unknown real correlationoff-scale) structure is given. The p 2 fields have been averaged over the small enough not to unduly dampen actual signals. Cor- of the ocean-internal sources/sinks determinedregions byas com-in Figure 7. relations shorter than in the standard atmospheric in- plex processes. To ensure that this approximation does version have been chosen to reflect the higher spatial not critically determine the result, runs with different detail in the pCO2 data. For example, this is impor- de-correlation lengths are included in the set of sensi- tant in the North Atlantic where shorter correlations are tivity cases in our robustness test (Sect. 3.3). needed to prevent a large influence of coastal data on In the time domain, we additionally dampen all non- 6 As an additional minor technical change, the basic spatial grid seasonal variability by a factor 16 with respect to the of the adjustable degrees of freedom is every pixel of the TM3 trans- seasonal variability (this is accomplished by implement- DIC port model grid, rather than aggregates of 4 neighbouring pixels, ing the time dependence of fint as a Fourier series and reflecting the higher spatial resolution of the data. reducing the a-priori uncertainty of the non-seasonal www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 CO 212 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability

modes, see Rodenbeck¨ , 2005). Such tighter a-priori un- For the case combining atmospheric and oceanic con- certainties for the interannual variations are consistent straints (Sect. 3.6), both cost function contributions are added = + + 1 T with the expectation that their amplitudes are smaller (J Jc Jp 2 x x). than those of the seasonal variations. Moreover, they dampen spurious interannual variations that may other- wise arise from the very different coverage of the SO- Appendix B CAT data in the individual years, and extrapolate the DIC mean seasonal cycle of fint throughout the inversion Test of retrieval capability period. As the chosen dampening factor is largely arbi- trary, it has also been varied in sensitivity tests. As a prerequisite to interpreting the results of fitting the CO2 DIC,ini scheme to surface-ocean p data, we need to test whether A further degree of freedom is 1Cm , needed only for DIC,ini (1) the scheme contains sufficient degrees of freedom to re- technical reasons: Through adjusting 1Cm , the initial produce the spatial and temporal variability of the pCO2 field, conditions of the carbon budget equation can become consis- and (2) the information available in the SOCAT data set is tent with the data (see Supplement Sect. S1.2). Adjustments sufficient to constrain it. This test is done by a synthetic run: DIC,ini to 1Cm are spatially resolved with the same spatial cor- (1) We create a spatio-temporal pCO2 field to be used as syn- relations as fint, but constant in time; the a-priori value is thetic “truth” on the basis of a biogeochemical process model zero. simulation (NEMOv2.3 with PlankTOM5, Buitenhuis et al., In run ATM, there are further degrees of freedom not 2010), mapped to the grid of our scheme. The seasonal cy- needed in run SFC: as in the atmospheric inversion, land cle of this field is similar to the observations at most loca- fluxes are adjustable as specified in Sect. A2.1, as well as tions except for the high latitudes, which is why we subtract the adjustable atmospheric initial condition. a mean seasonal cycle and replace it by that of Takahashi Table 4 summarizes the variables that are adjusted in the et al. (2009). This combined field then contains variations cost function minimization. similar to reality on all relevant timescales (though variations on the fastest (weekly to daily) timescales are smaller than CO2 A2.3 Extension: the p constraint observed). (2) This synthetic pCO2 field is sampled at the lo- cations and times of the SOCAT measurements, in the same Analogously to the observed atmospheric mixing ratios c , obs way as the modelled pCO2 field (Sect. A2.3). This gives a set all selected partial pressure values from the SOCAT database of pseudo data representing the same amount of information (Table 1) are collected into a vector p . A corresponding obs (in terms of data density available to the scheme) as the actual vector p of modelled partial pressures is formed by sam- mod data (note that SOCAT also contains along-track variability pCO2 pling the gridded m field of the model at the locations on scales much shorter than the size of our grid cells, which and times of the measurements (i.e. the modelled value is the scheme cannot make use of in any case). (3) We then fit that found in the gridcell/timestep which encloses the loca- the diagnostic scheme to the pseudo data, and compare the tion/time of the corresponding observation). If several SO- result with the synthetic truth as the known correct answer. CAT data points fall into the same gridcell and timestep, they In terms of mean seasonality considered here, the retrieval p are averaged together and only form a single element in obs (blue) fits the “known truth” (violet) closely (Fig. B1). Most p and mod. region-to-region differences are reproduced. The largest dif- p The vector mod is a function of the adjustable variable ferences are found in the North Pacific, and in parts of the f , as is p (except that p is also a function of fur- int mod mod temperate Southern Hemisphere. If pCO2 is averaged over f f ther adjustable variables, nee and ini, which are irrelevant open ocean only (not shown), Northern Hemisphere and trop- now). We can define an analogous cost function contribution: ical mismatches further reduce, in particular with the essen- tially perfect match in the temperate North Atlantic. 1 T −1  To be able to judge the match, we set it into perspective Jp = p − p Q p − p . (A30) 2 obs mod p obs mod with two further cases representing situations with less or with more available information, respectively: The covariance matrix Qp is chosen diagonal as well; the un- certainty for every individual pixel with two or more SOCAT – The black line shows the Bayesian prior, which does not data points√ is set to 10 µatm (for pixels with a single data yet use any SOCAT or atmospheric data (Sect. 2.3). The point to 2 × 10 µatm). The fit to the SOCAT data is then pCO2 seasonality thus just represents a-priori knowl- done in the same way as to the atmospheric data, just with edge on the response to temperature-induced changes Jc replaced by Jp (those degrees of freedom that do not de- in the chemistry (and to lesser extend in solubility, CO2 pend on pm – land fluxes and atmospheric initial condition Sect. A2.2). It is opposite in phase compared to the syn- – are unconstrained in this inversion and just remain at their thetic truth in high latitudes. The pseudo data are able a-priori values). to correct this large difference almost completely.

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ C. R¨odenbeck et al.: Sea-air CO2 flux estimated from CO2 partial pressure data 29

CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 213

Pacific 45N-90N Atlantic 45N-90N ity but not contained in our synthetic “truth” potentially 425 425 deteriorate the performance of our fit to real data). 400 400 Prior 375 375 Fit to SOCAT-sampled PlankTOM5 The test confirms that the differences between our results 350 350 Fit to PlankTOM5 (all pixels) and Takahashi et al. (2009) in the northernmost part of the 325 325 PlankTOM5 ("Known truth") North Pacific are due to the lack of data (see Supplement 300 300 Pacific 15N-45N Atlantic 15N-45N Fig. S7.4) in conjunction with spatial variability: In Taka-

425 425 hashi et al. (2009), seasonal amplitudes are large in a region 400 400 ◦ 375 375 at around 50–60 N and strongly drop going northward (see 350 350 Supplement Fig. S7.1). Though the diagnostic model would 325 325 be able follow the spatial pattern in the climatology if data 300 300 would say so (green line), in the absence of data it extra- Pacific 15S-15N Atlantic 15S-15N Indian 15S-30N 425 425 425 polates the field northward keeping the high seasonal cycle 400 400 400 from the pixels south of this area. 375 375 375 Seasonal data coverage is also limiting in the tropi- 350 350 350 cal Indian, the other problematic region (see Supplement

pCO2 (uatm) 325 325 325 300 300 300 Fig. S7.4). Pacific 45S-15S Atlantic 45S-15S Indian 45S-15S 425 425 425 400 400 400 Appendix C 375 375 375 350 350 350 Calculating phosphate concentrations from f DIC 325 325 325 int 300 300 300 Pacific 90S-45S Atlantic 90S-45S Indian 90S-45S Considering the ocean-internal processes affecting mixed-

425 425 425 layer concentrations of biogeochemical tracers, carbon 400 400 400 375 375 375 addition/removal occurs both through biological respira- 350 350 350 tion/photosynthesis in the mixed layer and through mixing-in 325 325 325 of water masses with higher/lower carbon concentration. In 300 300 300 the second case, in turn, there is a contribution from the pre- Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul Jan Jul formed carbon concentration originating from the last con- Fig. 13. Top: Map of a-priori correlation coefficients with respect to an example pixel at (70◦ E,49◦ S), illustrating how inverse adjust- Fig. B1. Testing the capacity to retrieve the pCO2 field from the tact of the arriving water parcel with the atmosphere, and a DIC SOCAT data set: fit of the diagnostic scheme to the pCO2 field sim- contribution from carbon added to the water parcel along its ments to fint de-correlate with distance. Bottom: A-priori correla- CO2 ulatedFig. 14. by theTesting PlankTOM5 the capacity biogeochemical to retrieve model the p (Buitenhuisfield from et theal., way through the ocean interior by remineralization. tion coefficients with respect to an example time step in the middle CO2 of the inversion period; in addition to the de-correlation away from 2010SOCAT) (combined data set: with Fit of seasonality the diagnostic from schemeTakahashi to the et al.p (2009field)) The biological contribution to the sources and sinks of car- the example time step, correlations rise again each year due to the sub-sampledsimulated by at the locations/times PlankTOM5 of biogeochemical the SOCAT measurements model (Buitenhu (blue),is bon (both in the mixed layer and the ocean interior) is directly et al., 2010) (combined with Takahashi et al. (2009)’s seasonality) dominance of the seasonal cycle. compared to the simulated pCO2 field itself (“known truth”, violet). linked to sources and sinks of phosphate, and assumed to pro- Forsub-sampled further comparison, at locations/times a fit of of the the scheme SOCAT to measurements the simulation (blue), at ev- compared to the simulated CO2 field itself (“known truth”, violet). ceed in Redfield proportions rC:P = 106. The remainder is ery grid cell and time step (“completep information”, green) and the For further comparison, a fit of the scheme to the simulation at ev- proportional to the differences between the pre-formed car- Bayesian prior (“no data information”, thin dashed grey, partially ery grid cell and time step (“complete information”, green) and the bon (or nutrient) concentrations of arriving water parcels and off-scale) is given. The pCO2 fields have been averaged over the Bayesian prior (“no data information”, thin dashed gray, partially those of the destination mixed-layer. Considering only their regions as in Fig. 7. CO off-scale) is given. The p 2 fields have been averaged over the variability as relevant here, these differences are assumed to regions as in Figure 7. be either small or close to Redfield proportions as well (this is equivalent to assuming that the internal sources/sinks of the DIC* DIC PO4 tracer Cm = Cm −rC:PCm (Gloor et al., 2003) have no – We also did the retrieval based on maximum informa- variability). Then the variations of ocean-internal phosphate CO tion about the p 2 field available, by constraining the PO4 sources and sinks (fint ) are proportional to the variations of scheme at every grid point and time step, not only where f DIC. We thus calculated the potential seasonal changes in actual SOCAT data exist. The result (green) fits the int the mixed-layer PO4 concentration using a PO4 budget equa- known truth almost perfectly, showing that there are suf- tion as Eq. (A18) (including history and freshwater fluxes), ficient degrees of freedom in the scheme to reproduce except that (1) the ocean-internal flux was reduced by the the mean seasonality. However, as the performance of PO4 DIC Redfield ratio (f = f /r : ), (2) the long-term mean the SOCAT-sampled pseudo data is only slightly worse, int int C P PO4 we conclude that the SOCAT data density is essentially was subtracted from fint , and (3) there was no sea–air flux PO4 sufficient to constrain the seasonality of surface-ocean of phosphate (fma = 0). As the initial condition for the bud- biogeochemistry on the scale of the considered regions get equation is unknown and phosphate does not relax to- PO4 (note that strong modes of variability existing in real- wards an atmospheric value, calculated Cm contains an ar-

www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 CO 214 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability bitrary offset, such that we only compare the deviation from Chen, L., Xu, S., Gao, Z., Chen, H., Zhang, Y., Zhan, J., and Li, W.: the temporal mean. Estimation of monthly air-sea CO2 flux in the southern Atlantic and Indian Ocean using in-situ and remotely sensed data, Remote Sens. Environ., 115, 1935–1941, 2011. Supplementary material related to this article is Ciais, P., Tans, P. P., Trolier, M., White, J. W. C., and Francey, R. J.: available online at: http://www.ocean-sci.net/9/193/2013/ A Large Northern Hemisphere Terrestrial CO2 Sink Indicated 13 12 os-9-193-2013-supplement.pdf. by the C/ C Ratio of Atmospheric CO2, Science, 269, 1098– 1102, doi:10.1126/science.269.5227.1098, 1995. Conkright, M., Locarnini, R. A., Garcia, H., O’Brien, T., Boyer, T., Stephens, C., and Antonov, J.: World Ocean Atlas 2001: Objec- tive Analyses, Data Statistics, and Figures, CD-ROM Documen- Acknowledgements. We would like to thank all contributors to tation, National Oceanographic Data Center, Silver Spring, MD, the SOCAT database, as well as all providers of atmospheric CO2 17 pp., 2002. data, which are the basis of this work. We gratefully acknowl- Conway, T., Tans, P., Waterman, L., Thoning, K., Kitzis, D., edge helpful discussions with Nicolas Cassar, Andrew Dickson, Masarie, K., and Zhang, N.: Evidence for interannual variabil- Niki Gruber, Roberta Hamme, Steve Jones, Bob Key, Armin Kohl,¨ ity of the carbon cycle from the national oceanic and atmo- Wolfgang Koeve, Iris Kriest, Corinne Le Quer´ e,´ Miguel Mahecha, spheric administration climate monitoring and diagnostics labo- Bill Munger, Andreas Oschlies, Taro Takahashi, Yasunori Tohjima, ratory global air sampling network, J. Geophys. Res., 99, 22831– and Lisan Yu. Sara Mikaloff Fletcher and Rik Wanninkhof made 22855, 1994. very helpful comments as referees. This study was also supported Corbiere,` A., Metzl, N., Reverdin, G., Brunet, C., and Takahashi, T.: through EU FP7 project CARBOCHANGE (264879) and is a Interannual and decadal variability of the oceanic carbon sink in contribution to the international IMBER/SOLAS projects. the North Atlantic subpolar gyre, Tellus B, 59, 168–178, 2007. The service charges for this open access publication have been de Boyer Montegut,´ C., Madec, G., Fischer, A. S., Lazar, A., and covered by the Max Planck Society. Iudicone, D.: Mixed layer depth over the global ocean: an exam- ination of profile data and a profile-based climatology, JGR, 109, Edited by: M. Hoppema C12003, doi:10.1029/2004JC002378, 2004. Egleston, E. S., Sabine, C. L., and Morel, F. M. M.: Revelle revis- ited: Buffer factors that quantify the response of ocean chemistry References to changes in DIC and alkalinity, Global Biogeochem. Cy., 24, GB1002, doi:10.1029/2008GB003407, 2010. Archer, D., Kheshgi, H., and E, M.-R.: Multiple timescales for neu- Garcia, H. E., Locarnini, R. A., Boyer, T. P., and Antonov, J. I.: tralization of fossil fuel CO2, Geophys. Res. Lett., 24, 405–408, World Ocean Atlas 2005, Volume 4: Nutrients (phosphate, ni- 1997. trate, silicate)., in: NOAA Atlas NESDIS 64, edited by Levitus, Arthun, M., Bellerby, R. G. J., Omar, A. M., and Schrum, S., US Government Printing Office, Washington, DC, 2006. C.: Spatiotemporal variability of air-sea CO2 fluxes in Gloor, M., Gruber, N., Sarmiento, J., Sabine, C. L., Feely, R. A., the Barents Sea, as determined from empirical relation- and Rodenbeck,¨ C.: A first estimate of present and preindus- ships and modeled hydrography, J. Mar. Sys., 98–99, 40–50, trial air-sea CO2 flux patterns based on ocean interior car- doi:10.1016/j.jmarsys.2012.03.005, 2012. bon measurements and models, Geophys. Res. Lett., 30, 1010, Atlas, R., Hoffman, R. N., Ardizzone, J., Leidner, S. M., Jusem, doi:10.1029/2002GL015594, 2003. J. C., Smith, D. K., and Gombos, D.: A cross-calibrated, multi- Heimann, M. and Korner,¨ S.: The global atmospheric tracer model platform ocean surface wind velocity product for meteorological TM3, Tech. Rep. 5, Max Planck Institute for Biogeochemistry, and oceanographic applications, B. Am. Meteorol. Soc., 92, 157– Jena, Germany, 2003. 174, 2011. Jacobson, A. R., Mikaloff Fletcher, S. E., Gruber, N., Sarmiento, Baker, D. F., Law, R. M., Gurney, K. R., Rayner, P., Peylin, P., J. L., and Gloor, M.: A joint atmosphere-ocean inver- Denning, A. S., Bousquet, P., Bruhwiler, L., Chen, Y.-H., Ciais, sion for surface fluxes of carbon dioxide: 1. Methods and P., Fung, I. Y., Heimann, M., John, J., Maki, T., Maksyutov, global-scale fluxes, Global Biogeochem. Cy., 21, GB1019, S., Masarie, K., Prather, M., Pak, B., Taguchi, S., and Zhu, doi:10.1029/2005GB002556, 2007. Z.: TransCom 3 inversion intercomparison: Impact of trans- Jahne,¨ B., Heinz, G., and Dietrich, W.: Measurement of the diffu- port model errors on the interannual variability of regional sion coefficients of sparingly soluble gases in water with a mod- CO2 fluxes, 1988–2003, Global Biogeochem. Cy., 20, GB1002, ified Barrer method, J. Geophys. Res., 92, 10767–10776, 1987. doi:10.1029/2004GB002439, 2006. Jones, S. D., Le Quer´ e,´ C., and Rodenbeck,¨ C.: Autocor- Bates, N. R., Knap, A. H., and Michaels, A. F.: Contribution of relation characteristics of surface ocean pCO2 and air- hurricanes to local and global estimates of air-sea exchange of sea CO2 fluxes, Glob. Biogeochem. Cy., 26, GB2042, CO2, Nature, 395, 58–61, 1997. doi:10.1029/2010GB004017, 2012a. Bousquet, P., Peylin, P., Ciais, P., Le Quer´ e,´ C., Friedlingstein, P., Jones, S. D., Le Quer´ e,´ C., Rodenbeck,¨ C., and Manning, A.: A and Tans, P. P.: Regional changes in carbon dioxide fluxes of land Global Interpolation of Surface Ocean pCO2 , J. Adv. Model. and oceans since 1980, Science, 290, 1342–1346, 2000. Earth Sys., (in review), 2012b. Buitenhuis, E., Rivkin, R., Sailley, S., and Le Quer´ e,´ C.: Biogeo- Joos, F., Plattner, G.-K., Stocker, T. F., Marchal, O., and Schmittner, chemical fluxes through microzooplankton, Global Biogeochem. A.: Global warming and marine carbon cycle feedbacks an future Cy., 24, GB4015, doi:10.1029/2009GB003601, 2010.

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/ CO C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability 215

atmospheric CO2, Science, 284, 464–467, 1999. doi:10.1029/2009GB003658, 2010. Kalnay, E., Kanamitsu, M., Kistler, R., Collins, W., Deaven, D., Mikaloff Fletcher, S. E., Gruber, N., Jacobson, A. R., Doney, S. C., Gandin, L., Iredell, M., Saha, S., White, G., Woollen, J., Zhu, Dutkiewicz, S., Gerber, M., Follows, M., Joos, F., Lindsay, K., Y., Chelliah, M., Ebisuzaki, W., Higgins, W., Janowiak, J., Mo, Menemenlis, D., Mouchet, A., Muller,¨ S. A., and Sarmiento, K. C., Ropelewski, C., Wang, J., Leetmaa, A., Reynolds, R., J. L.: Inverse estimates of anthropogenic CO2 uptake, transport, Jenne, R., and Joseph, D.: The NCEP/NCAR 40-year reanalysis and storage by the ocean, Global Biogeochem. Cy., 20, GB2002, project, Bull. Am. Met. Soc., 77, 437–471, 1996. doi:10.1029/2005GB002530, 2006. Keeling, R. F. and Shertz, S. R.: Seasonal and interannual variations Muller,¨ S. A., Joos, F., Plattner, G.-K., Edwards, N. R., and in atmospheric oxygen and implications for the global carbon Stocker, T. F.: Modelled natural and excess radiocarbon cycle, Nature, 358, 723–727, 1992. - sensitivities to the gas exchange formulation and ocean Key, R. M., Kozyr, A., Sabine, C. L., Lee, K., Wanninkhof, R., transport strength, Global Biogeochem. Cy., 22, GB3011, Bullister, J. L., Feely, R. A., Millero, F. J., Mordy, C., and Peng, doi:10.1029/2007GB003065, 2008. 14 T.-H.: A global ocean carbon climatology: Results from Global Naegler, T.: Reconciliation of excess C-constrained global CO2 Data Analysis Project (GLODAP), Global Biogeochem. Cy., 18, piston velocity estimates, Tellus B, 61, 372–384, 2009. GB4031, doi:10.1029/2004GB002247, 2004. Naegler, T. and Levin, I.: Closing the global radiocar- Krakauer, N. Y., Randerson, J. T., Primeau, F. W., Gruber, N., and bon budget 1945–2005, J. Geophys. Res., 111, D12311, Menemenlis, D.: Carbon isotope evidence for the latitudinal dis- doi:10.1029/2005JD006758, 2006. tribution and wind speed dependence of the air-sea gas transfer Newsam, G. N. and Enting, I. G.: Inverse problems in atmospheric velocity, Tellus B, 58, 390–417, 2006. constituent studies: I. Determination of surface sources under a Law, R. M., Peters, W., Rodenbeck,¨ C., Aulagnier, C., Baker, diffusive transport approximation, Inverse Problems, 4, 1037– I., Bergmann, D. J., Bousquet, P., Brandt, J., Bruhwiler, L., 1054, 1988. Cameron-Smith, P. J., Christensen, J. H., Delage, F., Denning, Olivier, J. G. J. and Berdowski, J. J. M.: Global emissions sources A. S., Fan, S., Geels, C., Houweling, S., Imasu, R., Karstens, U., and sinks, in: The Climate System,Berdowski, J., Guicherit, R., Kawa, S. R., Kleist, J., Krol, M. C., Lin, S.-J., Lokupitiya, R., and Heij, B. J., A. A. Balkema Publishers/Swets and Zeitlinger Maki, T., Maksyutov, S., Niwa, Y., Onishi, R., Parazoo, N., Pa- Publishers, Lisse, The Netherlands, ISBN 9058092550, 33–78, tra, P. K., Pieterse, G., Rivier, L., Satoh, M., Serrar, S., Taguchi, 2001. S., Takigawa, M., Vautard, R., Vermeulen, A. T., and Zhu, Z.: Park, G.-H., Wanninkhof, R., Doney, S. C., Takahashi, T., Lee, K., TransCom model simulations of hourly atmospheric CO2: Ex- Feely, R. A., Sabine, C. L., Trinanes,˜ J., and Lima, I. D.: Variabil- perimental overview and diurnal cycle results for 2002, Global ity of global net sea-air CO2 fluxes over the last three decades Biogeochem. Cy., 22, GB3009, doi:10.1029/2007GB003050, using empirical relationships, Tellus B, 62, 352–368, 2010. 2008. Pfeil, B., Olsen, A., Bakker, D. C. E., Hankin, S., Koyuk, H., Le Quer´ e,´ C., Rodenbeck,¨ C., Buitenhuis, E. T., Conway, T. J., Kozyr, A., Malczyk, J., Manke, A., Metzl, N., Sabine, C. L., Langenfelds, R., Gomez, A., Labuschagne, C., Ramonet, M., Akl, J., Alin, S. R., Bellerby, R. G. J., Borges, A., Boutin, J., Nakazawa, T., Metzl, N., Gillett, N., and Heimann, M.: Satu- Brown, P. J., Cai, W.-J., Chavez, F. P., Chen, A., Cosca, C., ration of the Southern Ocean CO2 sink due to recent climate Fassbender, A. J., Feely, R. A., Gonzalez-D´ avila,´ M., Goyet, change, Science, 316, 1735–1738, 2007. C., Hardman-Mountford, N., Heinze, C., Hood, M., Hoppema, Le Quer´ e,´ C., Takahashi, T., Buitenhuis, E. T., Rodenbeck,¨ C., and M., Hunt, C. W., Hydes, D., Ishii, M., Johannessen, T., Jones, Sutherland, S. C.: Impact of climate change and variability on S. D., Key, R. M., Kortzinger,¨ A., Landschutzer,¨ P., Lauvset, the global oceanic sink of CO2, Global Biogeochem. Cy., 24, S. K., Lefevre,` N., Lenton, A., Lourantou, A., Merlivat, L., Mi- GB4007, doi:10.1029/2009GB003599, 2010. dorikawa, T., Mintrop, L., Miyazaki, C., Murata, A., Nakadate, Lee, K., Tong, L. T., Millero, F. J., Sabine, C. L., Dickson, A. G., A., Nakano, Y., Nakaoka, S., Nojiri, Y., Omar, A. M., Padin, Goyet, C., Park, G.-H., Wanninkhof, R., Feely, R. A., and Key, X. A., Park, G.-H., Paterson, K., Perez, F. F., Pierrot, D., Pois- R. M.: Global relationships of total alkalinity with salinity and son, A., R´ıos, A. F., Santana-Casiano, J. M., Salisbury, J., Sarma, temperature in surface waters of the world’s oceans, Geophys. V. V. S. S., Schlitzer, R., Schneider, B., Schuster, U., Sieger, R., Res. Lett., 33, L19605, doi:10.1029/2006GL027207, 2006. Skjelvan, I., Steinhoff, T., Suzuki, T., Takahashi, T., Tedesco, K., Lefevre,` N., Watson, A. J., and Watson, A. R.: A comparison of Telszewski, M., Thomas, H., Tilbrook, B., Tjiputra, J., Vande- multiple regression and neural network techniques for mapping mark, D., Veness, T., Wanninkhof, R., Watson, A. J., Weiss, R., in situ pCO2 data, Tellus B, 75, 375–384, 2005. Wong, C. S., and Yoshikawa-Inoue, H.: A uniform, quality con- Lourantou, A. and Metzl, N.: Decadal evolution of carbon sink trolled Surface Ocean CO2 Atlas (SOCAT), Earth Syst. Sci. Data within a strong bloom area in the subantarctic zone, Geophys. Discuss., 5, 735–780, doi:10.5194/essdd-5-735-2012, 2012. Res. Lett., 38, L23608, doi:10.1029/2011GL049614, 2011. Rayner, P., Enting, I., Francey, R., and Langenfelds, R.: Recon- 13 Matear, R. J. and Hirst, A. C.: Climate change feedback on the fu- structing the recent carbon cycle from atmospheric CO2, δ CO2 ture oceanic CO2 uptake, Tellus B, 51, 722–733, 1999. and O2/N2 observations, Tellus B, 51, 213–232, 1999. Metzl, N., Corbiere,` A., Reverdin, G., Lenton, A., Takahashi, Rodenbeck,¨ C.: Estimating CO2 sources and sinks from atmo- T., Olsen, A., Johannessen, T., Pierrot, D., Wanninkhof, spheric mixing ratio measurements using a global inversion of R., Olafsd´ ottir,´ S. R., Olafsson, J., and Ramonet, M.: Re- atmospheric transport, Tech. Rep. 6, Max Planck Institute for cent acceleration of the sea surface fCO2 growth rate in Biogeochemistry, Jena, Germany, 2005. the North Atlantic subpolar gyre (1993–2008) revealed by Rodenbeck,¨ C., Houweling, S., Gloor, M., and Heimann, M.: CO2 winter observations, Global Biogeochem. Cy., 24, GB4004, flux history 1982–2001 inferred from atmospheric data using a

www.ocean-sci.net/9/193/2013/ Ocean Sci., 9, 193–216, 2013 CO 216 C. Rodenbeck¨ et al.: Global surface-ocean p 2 and sea–air CO2 flux variability

global inversion of atmospheric transport, Atmos. Chem. Phys., Takahashi, T., Sutherland, S. C., and Kozyr, A.: Global Ocean Sur- 3, 1919–1964, doi:10.5194/acp-3-1919-2003, 2003. face Water Partial Pressure of CO2 Database: Measurements Per- Sabine, C., Maenner, S. and Sutton, A.: High-resolution ocean formed during 1968–2010 (Version 2010), Tech. Rep. NDP-088, and atmosphere pCO2 time-series measurements from moor- ORNL/CDIAC-152, 2010. ing TAO140W. http://cdiac.esd.ornl.gov/ftp/oceans/Moorings/ Telszewski, M., Chazottes, A., Schuster, U., Watson, A., Moulin, C., TAO140W/. Carbon Dioxide Information Analysis Center, Oak Bakker, D., Gonzalez-D´ avila,´ M., Johannessen, T., Kortzinger,¨ Ridge National Laboratory, US Department of Energy, Oak A., Luger,¨ H., Olsen, A., Omar, A., Padin, X., R´ıos, A., Steinhoff, Ridge, Tennessee, 2010. T., Santana-Casiano, M., Wallace, D. W. R., and Wanninkhof, Sarmiento, J. L. and Gruber, N.: Ocean Biogeochemical Dynamics, R. H.: Estimating the monthly pCO2 distribution in the North Princeton Univ. Press, Princeton, USA, 2006. Atlantic using a self-organizing neural network, Biogeosciences, Sarmiento, J. L. and Le Quer´ e,´ C.: Oceanic carbon dioxide uptake 6, 1405–1421, doi:10.5194/bg-6-1405-2009, 2009. in a model of century-scale global warming, Science, 274, 1346– Valsala, K. V. and Maksyutov, S.: Simulation and assimilation of CO 1350, 1996. global ocean p 2 and air-sea CO2 fluxes using ship obser- Sarmiento, J. L., Gloor, E., Gruber, N., Beaulieu, C., Jacobson, vations of surface ocean pCO2 in a simplified Biogeochemi- A. R., Fletcher, S. E. M., Pacala, S., and Rodgers, K.: Trends and cal offline model, Tellus B, 62, 821–840, doi:10.1111/j.1600- Regional Distributions of Land and Ocean Carbon Sinks, Bio- 0889.2010.00495.x, 2010. geosciences, 7, 2351–2367, doi:10.5194/bg-7-2351-2010, 2010. Wanninkhof, R.: Relationship between wind speed and gas ex- Sigman, D. M. and Boyle, E. A.: Glacial/interglacial variations in change over the ocean, J. Geophys. Res.-Oceans, 97, 7373–7382, atmospheric carbon dioxide, Nature, 407, 859–869, 2000. 1992. Stephens, B. B., Gurney, K. R., Tans, P. P., Sweeney, C., Peters, W., Wanninkhof, R., Park, G.-H., Takahashi, T., Sweeney, C., Feely, R., Bruhwiler, L., Ciais, P., Ramonet, M., Bousquet, P., Nakazawa, Nojiri, Y., Gruber, N., Doney, S. C., McKinley, G. A., Lenton, T., Aoki, S., Machida, T., Inoue, G., Vinnichenko, N., Lloyd, A., Le Quer´ e,´ C., Heinze, C., Schwinger, J., Graven, H., and J., Jordan, A., Heimann, M., Shibistova, O., Langenfelds, R. L., Khatiwala, S.: Global ocean carbon uptake: magnitude, vari- Steele, L. P., Francey, R. J., and Denning, A. S.: Weak northern ability and trends, Biogeosciences Discuss., 9, 10961–11012, and strong tropical land carbon uptake from vertical profiles of doi:10.5194/bgd-9-10961-2012, 2012. atmospheric CO2, Science, 316, 1732–1735, 2007. Weiss, R.: Carbon dioxide in water and seawater: the solubility of a Sweeney, C., Gloor, E., Jacobson, A. R., Key, R. M., McKinley, G., non-ideal gas., Mar. Chem., 2, 203–205, 1974. and co authors: Constraining global air-sea gas exchange for CO2 Weiss, R. F.: The solubility of nitrogen, oxygen and argon in water with recent Bomb 14C measurements, Global Biogeoch. Cycl., and seawater, Deep-Sea Res., 17, 721–735, 1970. 21, GB2015, doi:10.1029/2006GB002784, 2007. Weiss, R. F. and Price, B. A.: Nitrous oxide solubility in water and Takahashi, T., Sutherland, S. C., Wanninkhof, R., Sweeney, C., seawater, Mar. Chem., 8, 347–359, 1980. Feely, R. A., Chipman, D. W., Hales, B., Friederich, G., Chavez, While, J., Totterdell, I., and Martin, M.: Assimilation of pCO2 Data F., Sabine, C., Watson, A., Bakker, D. C. E., Schuster, U., Metzl, Into a Global Coupled Physical-Biogeochemical Ocean Model, N., Yoshikawa-Inoue, H., Ishii, M., Midorikawa, T., Nojiri, Y., J. Geophys. Res., 117, C03037, doi:10.1029/2010JC006815, Kortzinger, A., Steinhoff, T., Hoppema, M., Olafsson, J., Arnar- 2012. son, T. S., Tillbrook, B., Johannessen, T.and Olsen, A., Bellerby, Yu, L. and Weller, R. A.: Objectively Analyzed air-sea heat Fluxes R., Wong, C. S., Delille, B., Bates, N. R., and de Baar, H. J. W.: (OAFlux) for the global oceans, Bull. Am. Met. Soc., 88, 527– Climatological mean and decadal change in surface ocean pCO2 539, 2007. and net sea-air CO2 flux over the global oceans, Deep-Sea Res. Zeebe and Wolf-Gladrow: CO2 in Seawater: Equilibrium, Kinetics, II, 56, 554–577, 2009. Isotopes, Elsevier, Amsterdam, 2001.

Ocean Sci., 9, 193–216, 2013 www.ocean-sci.net/9/193/2013/