abandoning sverdrup’s critical depth hypothesis on...

13
Ecology, 91(4), 2010, pp. 977–989 Ó 2010 by the Ecological Society of America Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms MICHAEL J. BEHRENFELD 1 Department of Botany and Plant Pathology, Cordley Hall 2082, Oregon State University, Corvallis, Oregon 97331-2902 USA Abstract. The Critical Depth Hypothesis formalized by Sverdrup in 1953 posits that vernal phytoplankton blooms occur when surface mixing shoals to a depth shallower than a critical depth horizon defining the point where phytoplankton growth exceeds losses. This hypothesis has since served as a cornerstone in plankton ecology and reflects the very common assumption that blooms are caused by enhanced growth rates in response to improved light, temperature, and stratification conditions, not simply correlated with them. Here, a nine-year satellite record of phytoplankton biomass in the subarctic Atlantic is used to reevaluate seasonal plankton dynamics. Results show that (1) bloom initiation occurs in the winter when mixed layer depths are maximum, not in the spring, (2) coupling between phytoplankton growth (l) and losses increases during spring stratification, rather than decreases, (3) maxima in net population growth rates (r) are as likely to occur in midwinter as in spring, and (4) r is generally inversely related to l. These results are incompatible with the Critical Depth Hypothesis as a functional framework for understanding bloom dynamics. In its place, a ‘‘Dilution–Recoupling Hypothesis’’ is described that focuses on the balance between phytoplankton growth and grazing, and the seasonally varying physical processes influencing this balance. This revised view derives from fundamental concepts applied during field dilution experiments, builds upon earlier modeling results, and is compatible with observed phytoplankton blooms in the absence of spring mixed layer shoaling. Key words: Critical Depth Hypothesis; grazing; growth rates; North Atlantic; phytoplankton blooms; remote sensing. INTRODUCTION The vernal, or spring, bloom is a renowned feature of many seasonal seas in the global ocean. Perhaps most famous of all is the annual greening event, clearly detectable from space (Fig. 1A), that occurs at middle and high latitudes of the North Atlantic. The physical– biological interactions responsible for this phenomenon have been studied since the 19th century (Banse 1992) and this interest continues unabated today. The emergence of plankton-rich waters in spring and early summer from the clear waters of winter, when light is low and storms frequent, naturally leads to the conclusion that the North Atlantic bloom is a conse- quence of improved upper ocean growth conditions in the spring that stimulate phytoplankton growth (anal- ogous to spring growth of terrestrial vegetation in temperate ecosystems, where increased cell division drives biomass accumulation). This relationship between growth conditions and phytoplankton biomass was encapsulated in 1935 by Gran and Braarud in the concept of a ‘‘critical depth,’’ which was later formalized by Sverdrup (1953) into the familiar ‘‘Critical Depth Hypothesis.’’ The fundamental tenet of the critical depth concept is that there exists, for any given date and location in the ocean, a surface mixing depth at which phytoplankton growth is precisely matched by losses of phytoplankton biomass within this depth interval. In Sverdrup’s hypothesis, the vernal bloom is initiated when the surface mixed layer shoals to a depth less than the critical depth, allowing the bulk phytoplankton specific growth rate (l) to exceed the bulk specific loss rate (l ) such that net population growth (r) is possible: r ¼ l l . 0: ð1Þ Importantly, Sverdrup assumed that l was a constant over time and thus independent of l. Accordingly, he envisioned the vernal bloom as resulting from, not simply being correlated with, an increase in phytoplank- ton growth rates. This view remains prevalent today, but will be challenged herein. While often misinterpreted, Sverdrup’s loss rate included sinking export, predator consumption (i.e., grazing), and respiration by the phytoplankton themselves, despite his use of the term ‘‘respiration’’ for l (Smetacek and Passow 1990). In more contemporary treatments (e.g., Crumpton and Wetzel 1982, Evans and Parslow 1985, Platt et al. 1991, Banse 1994), it is common practice to separate l into its component terms, such as r ¼ l g s p f ð2Þ Manuscript received 3 July 2009; accepted 20 July 2009. Corresponding Editor: O. Sarnelle. 1 E-mail: [email protected] 977

Upload: hadiep

Post on 14-Jul-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

Ecology, 91(4), 2010, pp. 977–989! 2010 by the Ecological Society of America

Abandoning Sverdrup’s Critical Depth Hypothesison phytoplankton blooms

MICHAEL J. BEHRENFELD1

Department of Botany and Plant Pathology, Cordley Hall 2082, Oregon State University, Corvallis, Oregon 97331-2902 USA

Abstract. The Critical Depth Hypothesis formalized by Sverdrup in 1953 posits thatvernal phytoplankton blooms occur when surface mixing shoals to a depth shallower than acritical depth horizon defining the point where phytoplankton growth exceeds losses. Thishypothesis has since served as a cornerstone in plankton ecology and reflects the very commonassumption that blooms are caused by enhanced growth rates in response to improved light,temperature, and stratification conditions, not simply correlated with them. Here, a nine-yearsatellite record of phytoplankton biomass in the subarctic Atlantic is used to reevaluateseasonal plankton dynamics. Results show that (1) bloom initiation occurs in the winter whenmixed layer depths are maximum, not in the spring, (2) coupling between phytoplanktongrowth (l) and losses increases during spring stratification, rather than decreases, (3) maximain net population growth rates (r) are as likely to occur in midwinter as in spring, and (4) r isgenerally inversely related to l. These results are incompatible with the Critical DepthHypothesis as a functional framework for understanding bloom dynamics. In its place, a‘‘Dilution–Recoupling Hypothesis’’ is described that focuses on the balance betweenphytoplankton growth and grazing, and the seasonally varying physical processes influencingthis balance. This revised view derives from fundamental concepts applied during field dilutionexperiments, builds upon earlier modeling results, and is compatible with observedphytoplankton blooms in the absence of spring mixed layer shoaling.

Key words: Critical Depth Hypothesis; grazing; growth rates; North Atlantic; phytoplankton blooms;remote sensing.

INTRODUCTION

The vernal, or spring, bloom is a renowned feature ofmany seasonal seas in the global ocean. Perhaps mostfamous of all is the annual greening event, clearlydetectable from space (Fig. 1A), that occurs at middleand high latitudes of the North Atlantic. The physical–biological interactions responsible for this phenomenonhave been studied since the 19th century (Banse 1992)and this interest continues unabated today. Theemergence of plankton-rich waters in spring and earlysummer from the clear waters of winter, when light islow and storms frequent, naturally leads to theconclusion that the North Atlantic bloom is a conse-quence of improved upper ocean growth conditions inthe spring that stimulate phytoplankton growth (anal-ogous to spring growth of terrestrial vegetation intemperate ecosystems, where increased cell divisiondrives biomass accumulation). This relationship betweengrowth conditions and phytoplankton biomass wasencapsulated in 1935 by Gran and Braarud in theconcept of a ‘‘critical depth,’’ which was later formalizedby Sverdrup (1953) into the familiar ‘‘Critical DepthHypothesis.’’

The fundamental tenet of the critical depth concept isthat there exists, for any given date and location in theocean, a surface mixing depth at which phytoplanktongrowth is precisely matched by losses of phytoplanktonbiomass within this depth interval. In Sverdrup’shypothesis, the vernal bloom is initiated when thesurface mixed layer shoals to a depth less than thecritical depth, allowing the bulk phytoplankton specificgrowth rate (l) to exceed the bulk specific loss rate (l )such that net population growth (r) is possible:

r ! l" l. 0: #1$

Importantly, Sverdrup assumed that l was a constantover time and thus independent of l. Accordingly, heenvisioned the vernal bloom as resulting from, notsimply being correlated with, an increase in phytoplank-ton growth rates. This view remains prevalent today, butwill be challenged herein. While often misinterpreted,Sverdrup’s loss rate included sinking export, predatorconsumption (i.e., grazing), and respiration by thephytoplankton themselves, despite his use of the term‘‘respiration’’ for l (Smetacek and Passow 1990). In morecontemporary treatments (e.g., Crumpton and Wetzel1982, Evans and Parslow 1985, Platt et al. 1991, Banse1994), it is common practice to separate l into itscomponent terms, such as

r ! l" g" s" p" f #2$

Manuscript received 3 July 2009; accepted 20 July 2009.Corresponding Editor: O. Sarnelle.

1 E-mail: [email protected]

977

Page 2: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

where g is mortality due to grazing, s is losses due tosinking, p is losses due to viral infection and parasitism,and f includes physical flushing losses, including dilutionby vertical mixing (Crumpton and Wetzel 1982, Banse1994).Sverdrup evaluated the Critical Depth Hypothesis by

comparing temporal changes in phytoplankton abun-dance observed in 1949 at Weather Ship ‘‘M’’ (668 N, 28E) to coincident changes in mixing depth relative to thecritical depth horizon. He concluded that the observa-tions were consistent with the hypothesis, but not withoutsome astute (and often overlooked) caveats. First,Sverdrup noted that the ‘‘bloom’’ in phytoplanktonbiomass observed two days after ‘‘the depth of the mixedlayer was for the first time smaller than the critical depth’’(Sverdrup 1953:249) likely reflected the advection of highbiomass waters into the Weather Ship ‘‘M’’ site, ratherthan rapid growth of the local phytoplankton populationin response to the changed relationship between mixingdepth and critical depth. He also recognized that, in thepresence of grazers, the ‘‘phytoplankton population mayremain small in spite of heavy production’’; in otherwords, bloom formation is critically dependent on therelationship between phytoplankton growth and grazing.Finally and importantly, Sverdrup noted that, even in hisspring 1949 data set, the first increase in phytoplanktonbiomass occurred prior to vertical stratification of theupper water column. With these issues in mind, Sverdrupconcluded his classic paper with the following statement:‘‘It is, therefore, not advisable to place too great emphasison the agreement between theory and [the Weather Ship‘‘M’’] observations.’’For over five decades, the Critical Depth Hypothesis

has repeatedly reemerged in discussions of phytoplank-ton bloom dynamics and found its way into oceano-graphic and limnological textbooks. Nevertheless,criticisms have been raised (e.g., Smetacek and Passow1990, Backhaus et al. 2003), with one reoccurring issuebeing the observation of significant spring phytoplank-ton blooms in the apparent absence of water columnstratification (Heimdal 1974, Schei 1974, Townsend etal. 1992, Ellertsen 1993, Backhaus et al. 1999, Dale et al.1999, Kortzinger et al. 2008). Here, I take a fresh look atseasonal phytoplankton cycles in the North Atlantic byusing a synoptic tool unavailable to Sverdrup in 1953:satellite observations. With these data, it is clear that theCritical Depth Hypothesis is an inadequate frameworkfor understanding vernal blooms. As a potentialreplacement, a ‘‘Dilution–Recoupling Hypothesis’’ isproposed that focuses on annual cycles in physicalprocesses that alter the delicate balance betweenphytoplankton growth and losses.

THE NORTH ATLANTIC FROM SPACE

Seasonal cycles in North Atlantic phytoplankton wereinvestigated using eight-day resolution remote sensingdata from the Sea-viewing Wide Field-of-view Sensor

(SeaWiFS) for the period January 1998 to December2006. SeaWiFS-derived water leaving radiances emanatefrom the upper first optical depth of the water column(i.e., depth to which 10% of surface irradiance pene-trates), thus phytoplankton features deeper in the watercolumn are not detected. However, uniformity ofphytoplankton properties within an active mixing layerimplies that SeaWiFS products are generally represen-tative of this layer, even during the deep winter mixingof the North Atlantic (Townsend et al. 1992, Ward andWaniek 2007, Boss et al. 2008; note, during midwinter,vertical velocities from convective mixing are stillsufficient for mixed layer transit times of a few days orless [Backhaus et al. 2003, D’Asaro 2008]).For the current study, satellite products used as

indices of phytoplankton abundance were surfacechlorophyll concentration (Chlsat, mg/m3; Fig. 1A) andphytoplankton carbon concentration (Cphyt, mg/m3).Chlsat values were from the NASA standard OC4-V4algorithm (the most familiar product to SeaWiFS users).For comparison, OC4-V4 Chlsat estimates are, onaverage, 29% higher than those from the Garver-Siegel-Maritorena (GSM) semi-analytical algorithm(Garver and Siegel 1997, Maritorena et al. 2002, Siegelet al. 2002b). Despite this bias, the two algorithms yieldtemporal patterns that are highly correlated (r ! 0.97),so conclusions drawn here regarding phytoplanktonseasonal cycles are insensitive to this choice of Chlsatproduct. Phytoplankton carbon concentrations werederived from GSM particulate backscattering coeffi-cients (bbp) following Behrenfeld et al. (2005).The weakness of using Chlsat as an index of

phytoplankton standing stock is that chlorophyll is alsoinfluenced by light- and nutrient-driven changes inintracellular pigment levels. The weakness of Cphyt isthat it is derived from a scattering property that is notuniquely algal, so its relation to phytoplankton biomassis influenced by the stability of the particle sizedistribution. However, Cphyt is insensitive to physiolog-ical variability and appears to track phytoplanktonbiomass in the few studies conducted to date (Behrenfeldand Boss 2003, 2006, Behrenfeld et al. 2005, Siegel et al.2005, Huot et al. 2007, Westberry et al. 2008, Dall’Olmoet al. 2009).North Atlantic phytoplankton seasonal cycles were

evaluated for 12 58 latitude 3 108 longitude bins in thecentral open ocean basin and away from continentalmargins (Fig. 1A). The bins are roughly a factor of fourlarger than the high-latitude bins used by Banse andEnglish (1994) and were chosen to minimize theinfluence of advection between eight-day periods, whilemaintaining a sense of spatial variability in phytoplank-ton seasonal properties. The lowest latitude bins (NA-1–NA-3) lie between 408 N and 458 N (Fig. 1A) andcorrespond to the ‘‘transitional region’’ discussed byHenson et al. (2008). For additional orientation, the1989 North Atlantic Bloom Experiment (NABE) waslocated near the center of bin NA-4 (white star in Fig.

MICHAEL J. BEHRENFELD978 Ecology, Vol. 91, No. 4

Page 3: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

1A) and the recent North Atlantic Bloom (NAB) study(March–June 2008) coincides with the eastern edge ofbin NA-12 (pink star in Fig. 1A). For the northernmostbins (NA-7–NA-12), satellite data were unavailable for afew weeks each year during midwinter. For all remainingtimes and locations, bin values represent means for all

observations within a given eight-day period, averaging.1200 cloud-free pixels per period (maximum potential! 1800 pixels). Less than 4% of the bin values werederived from ,200 cloud-free pixels. Within-bin stan-dard deviations for each eight-day mean value areshown for Cphyt in Fig. 1B.

FIG. 1. (A) Typical late bloom surface chlorophyll concentrations (Chlsat; note log scale) as observed from the satellite Sea-viewing Wide Field-of-view Sensor (SeaWiFS) in June 2002. Also shown are the 12 study bins and their designations. Red box:Figs. 1B, C, 2A, C, 3, and 4A–C show results from bin NA-5. Heavy black box: bins used in Fig. 2B–F. White star: 1989 NorthAtlantic Bloom Experiment (NABE) location. Pink star: 2008 North Atlantic Bloom (NAB) study location. (B) Nine-year recordof phytoplankton biomass (Cphyt: black symbols, left axis) and Chlsat (green symbols, right axis) at eight-day resolution for bin NA-5. Gray bars indicate within-bin standard deviations for Cphyt. (C) Nine-year record of Cphyt (black symbols, left axis, same as panelB), photosynthetically active radiation (PAR: red line, lower right axis), and mixed layer depth (MLD: blue line, upper right axis)for bin NA-5. Vertical dashed lines in panels B and C indicate 1 January of each year.

April 2010 979ABANDONING SVERDRUP’S BLOOM HYPOTHESIS

Page 4: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

Throughout the SeaWiFS record, regular Chlsat andCphyt seasonal cycles of variable amplitude (Henson etal. 2008) are found for each bin (Fig. 1B; see also theAppendix: Fig. A1 for all 12 bins). Spring chlorophyllpeaks typically range from 0.5 to 2.0 mg/m3, withoccasional values exceeding 2.5 mg/m3 (Fig. 1B;Appendix: Fig. A1). On average, peak concentrationsof Chlsat and Cphyt are lowest in the southern-most bins(Chlsat ! 0.72 6 0.28 mg/m3 [mean 6 SD]; Cphyt ! 33.96 7.2 mg/m3) and increase progressively to the north toreach values of Chlsat ! 1.18 6 0.32 mg/m3 and Cphyt !81.9 6 45.1 mg/m3 in bins NA-10, NA-11, and NA-12.Timing of these seasonal peaks also changes withlatitude, occurring roughly two weeks later with every58 increase in latitude (Siegel et al. 2002a, Henson et al.2008). Annual minima in Chlsat and Cphyt are typicallyobserved in January (Fig. 1B; Appendix: Fig. A1), withChlsat ranging from ;0.1 to 0.2 mg/m3, in excellentagreement with field observations (Parsons and Lalli1988, Banse 2002, Backhaus et al. 2003, Ward andWaniek 2007, Boss et al. 2008). During each seasonalcycle and in each bin, changes in Chlsat and Cphyt arehighly correlated (Fig. 1B; Appendix: Fig. A1; mean r!0.84, range ! 0.76–0.93), with the largest deviationsobserved during summer in the lowest latitude bins(NA-1–NA-3) and consistent with light-driven photo-acclimation responses to elevated mixed layer light levels(Behrenfeld et al. 2005, Siegel et al. 2005). The strongcorrespondence between Chlsat and Cphyt reflects thelarge dynamic range in phytoplankton biomass in theNorth Atlantic overwhelming coincident physiologicalprocesses acting to distinguish these two properties.Accordingly, the same overall conclusions emergeregarding the North Atlantic bloom whether phyto-plankton abundance is assessed using Chlsat or Cphyt.For clarity, I have, therefore, restricted the remainingdiscussion and figures to seasonal cycles in Cphyt, butprovide illustrations of seasonal Chlsat cycles for eachbin in the Appendix.In Fig. 1C, the nine-year record of Cphyt for NA-5 (red

box in Fig. 1A) is compared to coincident changes inmixed layer depth (MLD) and SeaWiFS cloudiness-corrected incident photosynthetically active radiation(PAR: 400–700 nm). Here, MLD values are from theFleet Numerical Meteorology and Oceanography Cen-ter (FNMOC) model (Clancy and Sadler 1992) and theSimple Ocean Data Assimilation (SODA) model (avail-able online).2 Both are tuned to available in situ data(i.e., they are ‘‘data-assimilating models’’). Perhaps themost striking features of Fig. 1C are that (1) winterthrough spring changes in phytoplankton biomass arehighly correlated with changes in PAR, and (2) thelargest changes in Cphyt coincide with spring MLDshoaling. Similar relationships are found for all 12 binsand may at first appear to confirm the Critical Depth

Hypothesis, but in fact they do not. The key problem isthat correlation between phytoplankton standing stock(Cphyt), PAR, and MLD shoaling does not imply that asimilar relationship exists for net growth rates (r). This isa crucial point because Sverdrup’s Critical DepthHypothesis is a statement that ‘‘shoaling of the mixedlayer above the critical depth horizon initiates a springbloom because it demarcates the first time when rbecomes positive.’’ To test this notion, one musttherefore evaluate seasonal cycles in r. As demonstratedin the following sections, such an analysis tells a verydifferent story about the North Atlantic bloom than thatimplied by Fig. 1C.

ABANDONING SVERDRUP

The net specific growth rate of a phytoplanktonpopulation can be calculated from two measures ofbiomass (C0, C1) separated by a period of time (Dt! t1"t0) following

r ! ln#C1=C0$=Dt #3$

so long as the phytoplankton suspension is not beingdiluted by phytoplankton-free media over the period,Dt. Eq. 3 implies that temporal changes in r can bevisualized by plotting phytoplankton biomass on alogarithmic scale against time. In Fig. 2A, the Cphyt

record is plotted on a logarithmic scale for bin NA-5(red box in Fig. 1A). From this plot, it is immediatelyapparent that Cphyt begins to increase in the middle ofwinter, not in the spring. (Vertical dashed lines in Fig.2A indicate 1 January.) Moreover, the rate of increase(i.e., the slope of log(Cphyt) vs. time) is highlyconstrained throughout the winter-through-spring peri-od in most cases (Fig. 2A).In Fig. 2B–F, mean seasonal cycles in Cphyt (black

symbols) for bins NA-2, NA-5, NA-8, NA-11, and NA-12 are compared to coincident changes in MLD (solidblack line; note that these five bins correspond to theheavy black box in Fig. 1A). These five plots illustratelatitudinal similarities and differences in mean annualphytoplankton cycles. In each case, Cphyt is alreadyincreasing when MLDs are at their maximum. Similarresults are found for all 12 bins for both Cphyt and Chlsat(Appendix: Fig. A2). Indeed, in 104 out of the total 108annual cycles available for the 12 bins, the onset of netpositive growth in Cphyt is found to coincide with thecessation of mixed layer deepening. In no case is positivenet growth delayed until significant shoaling of themixed layer occurs. Furthermore, alternative sources ofMLD estimates all yield consistent relationships be-tween Cphyt and MLD (gray symbols and standard errorbars in Fig. 2B–F; see figure legend for details).The satellite record clearly indicates that North

Atlantic bloom initiation is not a springtime event, butrather occurs in midwinter, at the latest. This findingstrongly refutes the Critical Depth Hypothesis, butraises the new question: ‘‘Why does cessation of mixedlayer deepening initiate a net increase in phytoplankton2 hhttp://web.science.oregonstate.edu/ocean.productivityi

MICHAEL J. BEHRENFELD980 Ecology, Vol. 91, No. 4

Page 5: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

concentration?’’ To answer to this question, it isimportant to recognize (as alluded to previously) thatnet phytoplankton specific growth rates are not accu-rately reflected by changes in carbon concentrationwhen a population is being diluted. A quintessentialexample of this issue is a chemostat culture, wheregrowth rate is equivalent to dilution rate and indepen-dent of biomass (which is held constant). As explainedby Evans and Parslow (1985), dilution effects must beconsidered in the North Atlantic when mixed layerdeepening entrains phytoplankton-free water frombelow. If this dilution effect is sufficiently large, it ispossible for phytoplankton concentration to decrease(e.g., in late autumn and early winter) despite net in situpopulation growth. To achieve a more completeunderstanding of North Atlantic seasonal phytoplank-ton cycles, it is therefore necessary to first account fordilution effects and then evaluate variability in r directly

(as opposed to simply viewing log-transformed plots ofCphyt). To conduct such an analysis using satellite data,Eq. 3 can be used to estimate r from changes in Cphyt

during periods of mixed layer shoaling and when mixedlayer deepening entrains significant phytoplanktonbiomass from below. During periods when mixingentrains relatively phytoplankton-free water, r mustinstead be estimated from changes in mixed layerintegrated phytoplankton biomass (R Cphyt ! Cphyt 3MLD; expressed as mg C/m2), which accounts fordilution of the population over a larger volume.

For the current study, r was estimated from R Cphyt

whenever MLD exceeded the euphotic depth (Zeu !depth to which 1% of surface irradiance penetrates).This criterion was chosen because global field data sets(e.g., Behrenfeld and Falkowski 1997) show phyto-plankton biomass throughout the euphotic zone whenMLD , Zeu, while chlorophyll profiles in the North

FIG. 2. (A) Nine-year record of phytoplankton biomass (Cphyt) at eight-day resolution for bin NA-5 plotted with a log-transformed y-axis (same data as in Fig. 1B, C). Gray bars indicate within-bin standard deviations. Vertical dashed lines indicate 1January of each year. (B–F) Mean annual cycles in Cphyt at eight-day resolution (black symbols, log-transformed left axis) andmixed layer depth (MLD) from the Fleet Numerical Meteorology and Oceanography Center (FNMOC) and the Simple OceanData Assimilation (SODA) models (black line, right axis). Gray symbols (right axis) represent mean mixed layer depth for threemodels, showing close agreement in timing of mixed layer deepening and shoaling despite differences in magnitude of midwintermixing depth (gray bars indicate standard deviation between models). The three MLD estimates are from the FNMOC/SODAmodel (as in Fig. 1), the Miami Isopycnic Coordinate Ocean Model (MICOM; Bleck et al. 1992), and a higher resolution (20–40km) version of the MICOM (Hatun et al. 2005; see Milutinovi!c et al [2009] for additional details). Annual cycles begin in Januaryand end in December (x-axes).

April 2010 981ABANDONING SVERDRUP’S BLOOM HYPOTHESIS

Page 6: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

Atlantic indicate sharp decreases below the mixed layerduring periods of deep mixing when MLD . Zeu (Wardand Waniek 2007). Thus, r was calculated from eight-day resolution satellite data for each North Atlantic binfollowing

r ! lnX

Cphyt-1

.XCphyt-0

! ".Dt;

if MLD is deepening and . Zeu #4a$

r ! ln#Cphyt-1=Cphyt-0$=Dt;

if MLD is shoaling or , Zeu #4b$

where Cphyt-0 and R Cphyt-0 are initial biomass levels andCphyt-1 and R Cphyt-1 are biomass levels after the timeinterval, Dt! eight days. As a noteworthy aside, Eq. 4aprovides an accurate assessment of r if mixed layerdeepening entrains phytoplankton-free water fromdepth and dilution is complete (which seems areasonable assumption given mixed layer transit timesof a few days or less; Backhaus et al. 2003, D’Asaro2008). Eq. 4b, on the other hand, would provide anaccurate estimate of r if the deep water entrained bymixing had a phytoplankton concentration equivalentto the initial surface population. Thus, the true value ofr is bounded by Eqs. 4a and 4b if entrained deep wateris not entirely devoid of phytoplankton. This uncer-tainty influences the confidence that can be assigned toretrieved values of r for any given eight-day period, buthas little impact on the overall pattern in r across themixed layer deepening period where MLDs increase byhundreds of meters and the clear water assumption fordeep water is a reasonable approximation (Ward andWaniek 2007).In Fig. 3, the nine-year record of r based on Eqs. 4a

and 4b is shown for bin NA-5 (open symbols). Fig. 3

also shows net specific rates of change in biomass basedon Cphyt alone (r0 ! black symbols). In other words, r0

represents the net specific growth rates that would resultif dilution effects of mixed layer deepening were ignoredand Eq. 4b was applied to all the data in Fig. 2A. Thiscomparison illustrates a few important points. First,seasonal cycles and variability in r and r0 are highlycorrelated (r! 0.90), indicating that overall patterns andvariability across the nine-year record are robust to theapproach for calculating r. Second, the dilution correc-tion (Eq. 4a) is generally modest and only impacts rduring a short period of late autumn and early winter(upper panel of Fig. 3), implying that the uncertainties indeep water phytoplankton concentrations describedabove do not compromise basic findings describedherein. Finally, the primary influence of the dilutioncorrection is that it advances the period of positive netpopulation growth to slightly earlier in the year, butinitiation of this positive growth phase is found inmidwinter whether based on r or r 0. Again, thisconclusion is illustrated in Fig. 2 by the increases inCphyt observed from January onward.In Fig. 4A, the nine-year record of r for bin NA-5

(same data as in Fig. 3, open symbols) is plotted alongwith coincident changes in PAR (gray line, lower rightaxis) and MLD (black line, upper right axis; note thatvalues increase upward). Here we see that during eachseasonal cycle r is negative over most of the summer andearly autumn, then increases in parallel with increasingMLD and decreasing PAR to become positive by lateautumn (horizontal dashed line indicates r ! 0), andthereafter remains positive until early summer (Fig. 4A).For all eight complete seasonal cycles at NA-5, thispositive growth phase begins while PAR is decreasingand, on average, precedes the post-solstice increase inPAR by 45 days (range 16–80 days). Clearly, the

FIG. 3. Nine-year record of phytoplankton net specific growth rate (r; open symbols) for bin NA-5. Also shown are net growthrates that would result if dilution effects of mixed layer deepening were not accounted for (r0, solid symbols). This r0 is equivalent toapplying Eq. 4b to all the data in Fig. 2A. A three-bin running boxcar averaging has been applied to dampen high frequencyvariability. Upper panel: dots indicate where Eqs. 4a and 4b were applied for calculating r for each eight-day period.

MICHAEL J. BEHRENFELD982 Ecology, Vol. 91, No. 4

Page 7: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

positive net growth phase is not initiated by increasingPAR or mixed layer shoaling. The NA-5 record alsoillustrates the high temporal variability in r and showsthat the peak in r is not consistently found in spring. Forexample, a spring peak in r is found during 2005–2006,but the peak during 2000–2001 and 2002–2003 occursduring midwinter and prior to the MLD shoaling (Fig.4A). During some years, there is no clear peak in rthroughout the positive growth phase (e.g., 1999–2000),while in other years similar magnitude peaks in r arefound in midwinter and spring (e.g., 1998–1999 and2001–2002; Fig. 4A). Similar variability is found for theother North Atlantic bins, but the overall pattern thatemerges is consistent: the spring maximum in phyto-plankton concentration represents the culmination of a

positive population growth phase that begins during lateautumn/winter. This positive growth phase is initiatedduring mixed layer deepening and is first expressed as anincrease in Cphyt (and Chlsat) when deep-water entrain-ment into the mixed layer stops (i.e., dilution; Fig. 2;Appendix: Fig. A2).

In Fig. 4B, the mean annual cycle for r at NA-5 isshown (open symbols), along with mean values forMLD and Zeu (note that in this figure the x-axis beginswith July and MLD increases downward on the rightaxis). Here we see that the positive growth phase(December to June) begins roughly when the mixedlayer penetrates below the euphotic layer (i.e., MLD .Zeu; Fig. 3B). This correspondence is also found in theother bins with complete satellite coverage over the

FIG. 4. (A) Nine-year record of phytoplankton net specific growth rate (r; open symbols, left axis! same data as open symbolsin Fig. 3), mixed layer depth (MLD; heavy black line, upper right axis), and photosynthetically active radiation (PAR: light grayline, lower right axis). Note that MLD increases upward in this figure to illustrate correlation with r. All data are at eight-dayresolution. A three-bin running boxcar averaging has been applied to r to dampen high frequency variability. (B) Annual meancycles from July to following July of r (open symbols, left axis), MLD (heavy black line, right axis), and euphotic depth (Zeu: heavydotted line, right axis) for bin NA-5. Note that in this figure MLD increases downward. Gray bars indicate standard deviation in rfor each eight-day period over the nine-year satellite record (panel A). ‘‘Positive net growth phase’’ is indicated at top. Horizontaldashed line indicates r ! 0. (C) Annual mean cycles in phytoplankton specific growth rate (l) for NA-5 (left axis) based on netprimary production (NPP) estimates from (solid line) the standard Vertically Generalized Production Model (VGPM); (dashedline) the VGPM with an exponential description of chlorophyll-specific light-saturated photosynthesis (Pb

opt); and (dotted line) theVGPM with a regionally tuned description of Pb

opt (see the Abandoning Sverdrup section for details). Note again that the x-axisbegins in July and ends in July a year later. Also shown is the mean annual cycle in r from panel B (open symbols, right axis,with same scale as in panel B). Gray hatched box shows where the entire annual range in r is found when plotted on the same leftaxis as l, emphasizing how small r is relative to l throughout the year.

April 2010 983ABANDONING SVERDRUP’S BLOOM HYPOTHESIS

Page 8: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

annual cycle and does not simply reflect the criteriadescribed above for estimating r, as it is also found if r isestimated from R Cphyt whenever MLD is increasing(i.e., if the MLD . Zeu criteria is dropped in Eq. 4a).Another characteristic of the mean annual cycles in r is aslight midwinter depression (e.g., January throughMarch in Fig. 4B), which becomes more prominent athigher latitudes and gives the positive growth phase a‘‘dual peaked’’ appearance (Fig. 4B). The mean annualcycle also shows the small and brief positive growthperiod associated with the September–October ‘‘fallbloom’’ (Fig. 4B).With the satellite record of net population growth

rates in the North Atlantic (Fig. 4A, B), we can nowdismiss the Critical Depth Hypothesis as a validexplanation for bloom initiation. To better understandhow a positive growth phase can persist during winterwhen PAR is low and MLDs are maximal, it isinstructive to compare absolute values of r withcoincident variations in phytoplankton specific growthrates (l). To estimate l, net primary production (NPP)was calculated from satellite Chlsat data using threeapproaches: (1) the standard parameterization of theVertically Generalized Production Model (VGPM;Behrenfeld and Falkowski 1997), (2) the VGPM withan exponential function describing chlorophyll-specificlight-saturated photosynthesis (Pb

opt) following Morel(1991), and (3) the VGPM with a regionally tunedestimate of Pb

opt. The first two of these approaches aredescribed in Campbell et al. (2002) and were found inthe round-robin algorithm testing study of Carr et al.(2006) to be representative of the two major categoriesof satellite NPP estimates. In the third approach,phytoplankton Chl:C ratios were described as a functionof mixed layer light level following Behrenfeld et al.(2005) and Pb

opt described as an inverse function of Chl:Cratio and tuned to give values consistent with measuredPbopt during the North Atlantic Bloom Experiment

(available online).3 All three models give NPP estimatesper square meter, which were then divided by activephytoplankton biomass (i.e., integrated carbon from thesurface to the greater of Zeu and MLD) to estimate l.In Fig. 4C, seasonal cycles in l are shown for NA-5.

Despite differing magnitudes, all three models givesimilar seasonal cycles because changes in NPP (thusl) are dominated by changes in Chlsat and PAR. Alsoshown in Fig. 4C and plotted on the right hand axis isthe seasonal cycle in r from Fig. 4B (open symbols). Thiscomparison illustrates the general inverse relationshipbetween l and r and strongly implies that the positivegrowth phase leading to the spring peak in phytoplank-ton biomass is less a consequence of springtime increasesin l as it is a reflection of seasonal changes in loss rates(l ). In other words, the only way for r to increase while lis decreasing is for the fraction of l escaping losses to

increases as l decreases. To clearly understand thisfinding, it is important to recognize in Fig. 4C the verydifferent scales for l (left axis) and r (right axis). Ifinstead both properties were plotted on the left axis, thefull seasonal cycle in r would fall within the hatched boxat the bottom of the figure. This comparison illustrateswhy the positive growth phase can persist throughoutthe winter: r is so small that it requires very little NPP tosupport it (Fig. 4C). Indeed, the mean value of r frommid-December to mid-April at NA-5 is only 0.018/d,corresponding to a population doubling time of morethan a month. Again, the essential requirement toachieve comparable values of r from winter throughspring (Fig. 4B) is that phytoplankton loss rates (l ) mustcovary closely with l, and herein lies the crucial flaw inthe Critical Depth Hypothesis: Sverdrup assumed l to beconstant over time.

THE DILUTION–RECOUPLING HYPOTHESIS

The annual North Atlantic phytoplankton bloom isthe consequence of a decoupling between phytoplanktongrowth and losses (Eqs. 1, 2) and it terminates witheither the exhaustion of surface nutrients or overgrazingby heterotrophs (Banse 1992, 2002). The key question iswhether this decoupling is due to increased phytoplank-ton growth rates or decreased losses (Eq. 1). Asdescribed above, the positive net growth phase of theannual cycle leading to the late spring climax in biomassstarts in late winter (Fig. 4), which dismisses the CriticalDepth Hypothesis as a valid explanation for bloominitiation. Onset of the positive phase also coincides withdecreasing incident irradiance and, in general, r is foundto be inversely related to PAR, NPP, and l. Theseobservations deemphasize the importance of the springstimulation in phytoplankton photosynthesis andgrowth, and emphasize the importance of seasonalperturbations in the loss terms for phytoplanktonbiomass (Banse 1992). In this section, I propose aDilution–Recoupling Hypothesis for the North Atlanticbloom that focuses on physical processes impactingphytoplankton growth–loss relationships, incorporatesseasonal changes in l, and builds upon the earlier workof Evans and Parslow (1985), Banse (1992, 2002), andMarra and Barber (2005). While the hypothesis providesa framework for understanding the North Atlanticseasonal phytoplankton cycle and is consistent with thesatellite observations (Figs. 2–4), it is only a hypothesisand other interpretations and complexities should beconsidered in the future. It is also important to recognizethat satellite data provide limited information fordistinguishing different loss terms for phytoplanktonbiomass. Thus, in the words of Sverdrup, ‘‘It is notadvisable to place too great emphasis on the agreementbetween theory and observation.’’Eq. 2 captures the primary loss terms influencing r. Of

these, the ‘‘flushing’’ term ( f ) can be dismissed as asignificant factor influencing retrieved values of rbecause the large bin sizes used in the current analysis3 hhttp://usjgofs.whoi.edu/jg/dir/jgofs/i

MICHAEL J. BEHRENFELD984 Ecology, Vol. 91, No. 4

Page 9: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

minimize the horizontal advection effects between anytwo eight-day sampling periods and the impact ofdilution from mixed layer deepening has been accountedfor in calculations of r. Seasonal changes in sinkinglosses (s), on the other hand, can contribute to theannual phytoplankton cycle. During mixed layer deep-ening, vertically exported phytoplankton can be en-trained back into the surface layer (Ward and Waniek2007). Seasonal changes in taxonomy also occur, withsmaller, slow sinking (often flagellated) species domi-nating during winter and more rapidly sinking diatomsbecoming prevalent later in the bloom (Halldal 1953,Paasche 1960, Ramsfjell 1960, Paasche and Rom 1962,Parsons and Lalli 1988, Dale et al. 1999, Backhaus et al.2003, Ward and Waniek 2007). Aggregate formationcan also significantly increase s during later stages of thebloom, as coagulation increases as a squared function ofparticle concentration, and may even set a limit tomaximum achievable bloom concentrations (Jackson2005). Much less is known about seasonal changes inviral infection and parasitism in the North Atlantic.However, while many factors clearly contribute tophytoplankton losses, grazing (g) is by far the dominantterm at most times (e.g., Steemann Nielsen 1957, Banse1992, Banse and English 1994) and, importantly, g canrespond rapidly to changes in phytoplankton biomassand l. Accordingly, the Dilution–Recoupling Hypoth-esis focuses on phytoplankton–grazer interactions andphysical processes perturbing the balance between l andg. To begin, a brief digression is helpful regarding effectsof dilution.In planktonic ecosystems, the impact of dilution on

predator–prey interactions is commonly exploited forassessing phytoplankton specific growth rates (l). Thetechnique is referred to as a ‘‘dilution experiment’’(Landry and Hassett 1982, Landry 1993, Landry et al.1995). A dilution experiment is conducted by dispensingdifferent ratios of whole and filtered seawater intovessels of equal size, measuring initial concentrations ofchlorophyll, incubating the samples for some timeperiod, and then assessing final chlorophyll concentra-tions at the end of the incubation (Landry and Hassett1982, Landry et al. 1995). Phytoplankton net specificgrowth rates (r) are then calculated (Eq. 3) for eachtreatment and plotted against dilution ratio. Theresultant relationship typically shows r increasinglinearly with dilution (e.g., Landry et al. 1995), althoughother nonlinear relationships are possible (Evans andParanjape 1992). Extrapolation of this relationship tothe intercept provides the estimate of l. The dilutionexperiment works because addition of filtered seawaterto whole seawater decreases l by diluting predators andprey (i.e., changes predator–prey encounter rates), butideally does not alter l (Landry and Hassett 1982).Importantly, relatively short incubations (;1 day) areemployed to prevent predator and prey populationgrowth from negating the impact of the dilutionprocedure. In other words, the dilution effect is

transient. To sustain this effect over a much longerperiod, new filtered seawater would need to be added ata rate comparable to population growth rates (thusmaintaining constant predator–prey interactions). Therelevance of this small volume experimental technique toplankton dynamics of the subarctic North Atlantic basinis that it provides a tangible example of a simple processaltering the balance between l and l that occursnaturally over vast areas during mixed layer deepening.

The underlying concept of the Dilution–RecouplingHypothesis is that the complex heterotrophic food webof planktonic ecosystems allows a constant tightcoupling (rapid response time) between phytoplanktongrowth and losses, but seasonal mixed layer deepeninghas the potential to slightly decouple l and g byimpacting predator–prey interactions through dilutionof both phytoplankton and grazers (as in a dilutionexperiment). A strong tendency for g to balance limplies that this decoupling effect will only be main-tained through continued dilution. Thus, cessation ofmixed layer deepening should advocate a strengtheningin predator–prey coupling. As described in the modelingstudy of Evans and Parslow (1985), this recoupling canbe further enhanced by mixed layer shoaling, which bothsevers nonmobile phytoplankton at depth from thephotic zone (Backhaus et al. 2003) and concentrates themobile grazing community into a shrinking ‘‘photosyn-thetically active’’ volume (i.e., the surface mixed layer oreuphotic zone). The following paragraph applies thesebasic concepts to narrate one potential interpretation ofthe repeated seasonal cycles in r resolved from theSeaWiFS data and illustrated in Fig. 4.

Starting in late spring, surface nutrient depletion orovergrazing terminates the annual North Atlantic bloom(Banse 1992, 2002). From this point through theremaining stratified summer months, losses (grazing,sinking, others) exceed l, with biomass decreasingroughly exponentially (Fig. 2) at a specific rate betweenr! 0 and r!"0.01 per day (Fig. 4). In late summer andearly autumn, the mixed layer begins to penetrate deeperreaches of the euphotic zone, entraining nutrients as wellas phytoplankton and grazers from below. During thisperiod, the ‘‘dilution effect’’ is minimal (i.e., predator–prey encounter rates are not strongly altered) and onlyminor changes in r are observed (Fig. 4B, September–October) that largely reflect nutrient-driven changes inl. Eventually, however, mixed layer deepening begins toentrain relatively plankton-free water from below. Thisevent roughly coincides with MLD . Zeu (Fig. 4B,November–December) and, according to the scenariodescribed above, perturbs the balance between l and g(as in a dilution experiment) and initiates the positive netgrowth phase and the annual North Atlantic bloom(Fig. 4B). The start of this positive r phase correspondsto an increase in areal biomass (R Cphyt) but notnecessarily phytoplankton concentration (Cphyt; m"3)because gains in biomass are being distributed over agrowing volume of water (i.e., MLD is still increasing).

April 2010 985ABANDONING SVERDRUP’S BLOOM HYPOTHESIS

Page 10: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

Once the mixed layer stops deepening, though, thepositive population growth rate becomes apparent inCphyt, resulting in the observed coincidence betweeninitial increases in Cphyt and cessation of mixed layerdeepening (Fig. 2). From this point forward to the endof bloom, grazing pressure increases (because decou-pling by dilution has ended) but is countered by light-driven increases in l, allowing r to remain high (evenincrease) well into the spring (Fig. 4). Importantly, thisscenario does not require stratification for a bloom tooccur (Evans and Parslow 1985), only cessation ofmixed layer deepening, making this new view compatiblewith observations of bloom development in the absenceof shoaling (Heimdal 1974, Schei 1974, Townsend et al.1992, Eilertsen 1993, Backhaus et al. 1999, Dale et al.1999). Nevertheless, stratification does have two signif-icant impacts on the North Atlantic bloom under theDilution–Recoupling Hypothesis: it accelerates thespring increase in grazing pressure within the mixedlayer by concentrating mobile heterotrophs (Evans andParslow 1985), while simultaneously increasing lthrough improvements in median mixed layer lightlevels (Reynolds 2006). Thus, springtime increases in lretain an important role by prolonging the positivegrowth phase despite stratification-driven increases in g,as well as biomass-driven increases in s (Jackson 2005).The Dilution–Recoupling Hypothesis narrated above

largely focuses on only two terms in Eq. 2: l and g.Despite this shortcoming, it does appear to account forfundamental features in the seasonal cycle in r and italso provides a useful context for interpreting additionalfeatures of the satellite data. For example, as anextension of field dilution experiments, the Dilution–Recoupling Hypothesis assumes that predator–preydecoupling from mixed layer deepening results fromchanges in predator and prey concentrations, not fromchanges in l. Accordingly, when r is expressed as afraction of l using data in Fig. 4C (i.e., r/l), this fractionis found to increase and decrease in parallel with MLDdeepening and shoaling (r ! 0.91, data not shown),without the midwinter depression observed in r (Fig.4B, C). The same result is found for all 12 bins (mean r!0.87). Thus the typical ‘‘dual peaked’’ appearance inmean seasonal cycles of r (Fig. 4B, C) can be interpretedas reflecting changes in the balance between mixed layereffects on l and light-driven effects on l. Similarly,latitudinal changes in this balance resulting fromdifferent seasonal cycles in MLD and incident lightcan be related to the more pronounced midwinterdepressions in r noted earlier for higher latitude bins.Finally, comparison of satellite-derived r and l for the12 bins consistently indicates that the decoupling of land l from mixed layer deepening is roughly twice aslarge as the recoupling effect from an equivalentshoaling of the mixed layer. This finding is consistentwith the ‘‘asymmetric effect’’ envisioned by Evans andParslow (1985) to result from mixed layer deepening

diluting both predators and prey, while mixed layershoaling primarily concentrates predators alone.

CONCLUSION

The North Atlantic phytoplankton bloom is amongthe most conspicuous features in the global satelliterecord and has been a topic of ecological interest forover a century (see Banse 1992). The appearance ofelevated phytoplankton concentrations coinciding withspringtime improvements in upper ocean growth condi-tions (light, temperature, stratification) led early on tothe assumption that the vernal bloom was a consequenceof increased phytoplankton growth rates. Sverdrup’sCritical Depth Hypothesis subsequently provided aframework for understanding the environmental condi-tions necessary to initiate a bloom, and this hypothesishas served the ecological science community for overhalf a century. Today, satellite observations providecoverage of the North Atlantic bloom in a mannerunimaginable in Sverdrup’s time and, with these data,basic assumptions of the Critical Depth Hypothesis arefound wanting. Results presented here demonstrate thatbloom initiation occurs in the winter when mixed layerdepths are maximum (Figs. 2, 4), that the couplingbetween phytoplankton growth and losses increasesrather than decreases during spring stratification (Fig.4C), that annual maxima in net population growth ratesare as likely to occur in midwinter as in spring (Fig. 4A),and that phytoplankton loss rates are not constant (asassumed by Sverdrup), but proportional to, andperpetually coupled tightly with, phytoplankton growthrates. The current analysis also shows that the positivenet growth phase in the North Atlantic begins prior toan increase in light and generally occurs while mixingdepths are still increasing. With these findings it isnecessary to abandon the notion that the bloom simplyreflects an increase in phytoplankton division rate.Indeed, the satellite record indicates an overall inverserelationship between r and l (Fig. 4C). That net growthrate can exhibit the opposite sign of cell division rate wasearlier acknowledged, for example, by Riley (1946),Margalef et al. (1957), and Banse (2002).As a replacement for the Critical Depth Hypothesis, a

Dilution–Recoupling Hypothesis is outlined here thatfocuses on the balance between phytoplankton growthand grazing, and seasonally varying physical processesinfluencing this balance. The central role of grazing onphytoplankton population dynamics has been recog-nized (yet often overlooked) since the mid-20th century(see discussion in Banse 1992:428–430), but a completeunderstanding of the annual cycle in phytoplankton–grazer interactions has been hindered by infrequentsampling and a bias in field campaigns toward latestages of the bloom. Fortunately, satellite observationsalready provide the necessary measurement frequency tobegin resolving temporal aspects of plankton dynamicsin the seasonal seas. The Dilution–Recoupling Hypoth-esis described here adopts concepts from Evans and

MICHAEL J. BEHRENFELD986 Ecology, Vol. 91, No. 4

Page 11: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

Parslow (1985) and distinguishes two primary impactsof MLD changes on the balance between l and l. In the‘‘dilution phase,’’ mixed layer deepening lowers predatorand prey concentrations and thus weakens the linkbetween l and g by reducing encounter rates. In the‘‘recoupling phase,’’ the tendency of planktonic ecosys-tems to balance l and l is favored. During this phase, gincreases relative to l because the decoupling period ofmixed layer deepening has ended. Importantly, mixedlayer shoaling is not requisite for the bloom, but insteadintensifies g by concentrating mobile predators into ashrinking volume, which is offset by concurrent increas-es in l (Reynolds 2006). With or without stratification,springtime increases in PAR are critical to maintain labove a growing grazing pressure. At very highlatitudes, spring increases in PAR are particularlyimportant because they also mark the end of polarnight and thus permit growth.The broader applicability of the Dilution–Recoupling

Hypothesis to other environments remains to be tested.However, the impact of mixed layer deepening on rthrough dilution of phytoplankton and grazers asdescribed by Evans and Parslow (1985) was recentlyapplied to studies on monsoon-driven blooms in theArabian Sea by Marra and Barber (2005). In this case,the authors adopt the Critical Depth Hypothesis, butargue that seasonal mixing is inadequate for lightlimitation of phytoplankton growth (Barber et al.2001, Marra and Barber 2005). Nitrogen is likewisedismissed as a limiting factor for l in the Arabian Sea.The appearance of elevated biomass following monsoon-driven erosion of the mixed layer was therefore proposedto result from dilution effects on grazing (Marra andBarber 2005), as per Evans and Parslow (1985). J. Marraand T. S. Moore, II, unpublished manuscript, have sinceinsightfully stated that ‘‘monsoon-induced vertical mix-ing is more a factor in resource limitation for the micro-grazers than the phytoplankton,’’ but they also make adistinction that ‘‘In the Arabian Sea, unlike the NorthAtlantic, . . . mixing never exceeds the critical depth.’’While this later statement is clearly challenged herein,the modeling work of Evans and Parslow (1985), theArabian Sea study of Marra and Barber (2005), and theNorth Atlantic analysis of the current study are allconverging, at a fundamental level, toward a consistentview of seasonal plankton dynamics dominated byphysical disruption and ecological recoupling of phyto-plankton–grazer interactions.The primary motivation for this effort has been to

invigorate discussion on seasonal phytoplankton bloomsand, at the very least, give the reader some pause toaccepting traditional interpretations. The Dilution–Recoupling Hypothesis, at best, provides a basicframework regarding first-order properties of the NorthAtlantic bloom. Clearly, the current treatment fails toaddress many of the ecological complexities regardingNorth Atlantic plankton, including species successions,short-term variability in r (Fig. 4A), and the basis for

temporal stability in r from winter through spring (Fig.2). As work continues along these lines, a shift in focus isstrongly recommended that emphasizes variability in rand seasonal changes in r as a fraction of l, rather thansimply focusing on standing stock. This recommenda-tion is intended to assist in understanding seasonalplankton dynamics, not to discount the significance ofbiomass accumulation. Indeed, higher trophic-levelfertility of the North Atlantic is linked to the vernalperiod of the bloom, as is seasonal pCO2 draw downassociated with organic carbon export. Nevertheless,population changes preceding the vernal climax andimprinted in the record of r play a crucial role in settingthe stage for the event. Resolving these processes willrequire a break from traditional late-bloom field studies,development of a balanced research program coveringthe full annual cycle, and the employment of new high-temporal-resolution technologies capable of prolongeddeployment (e.g., Boss et al. 2008). The significance ofthis revised philosophy was elegantly stated by Evansand Parslow (1985): ‘‘Observations that phytoplankton[are] abundant for a short period almost every spring . . .implies that the ecosystem not only produces a bloom inthe spring but also recreates, over the fall and winter, theconditions necessary for a bloom. A spring bloom is notan isolated event, but one feature of a roughly repeatingannual cycle.’’

A correct understanding of plankton dynamicsthroughout the annual cycle and the mechanismsunderlying this succession of events is particularlyessential today for interpreting current observed trendsand forecasting future change. For example, if oneadopts the notion that a vernal bloom emerges fromelevated phytoplankton growth rates (l) resulting frommixed layer shoaling above a critical depth, thenincreased stratification from climate warming may beanticipated to enhance North Atlantic blooms (or atleast advance their timing to earlier spring). In contrast,the Dilution–Recoupling Hypothesis described hereemphasizes the crucial role of mixing on the seasonaldecoupling of l and g. According to this view, deepmixing is essential for bloom formation, with deepermixing causing greater perturbations to predator–preyinteractions than shallower mixing. Consistent with thisnotion, higher latitudes of the North Atlantic are foundto have deeper winter mixing (Fig. 2), a greater winterdecoupling of l and l, higher winter-to-spring values of r(Appendix: Fig. A3, panel A), and higher peakchlorophyll concentrations (Appendix: Fig. A3, panelB). Extending this Dilution–Recoupling Hypothesis toclimate warming suggests that suppression of wintermixing may lead to reduced net phytoplankton growthrates and vernal biomass (i.e., an opposite conclusionfrom one based on the Critical Depth Hypothesis).Interestingly, the satellite record to date shows an inverserelationship between sea surface temperature changesand Chlsat at high latitudes (Behrenfeld et al. 2008,2009).

April 2010 987ABANDONING SVERDRUP’S BLOOM HYPOTHESIS

Page 12: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

ACKNOWLEDGMENTS

I thank Robert O’Malley for analysis of SeaWiFS andFNMOC data; Svetlana Milutinovi!c for MICOM1 andMICOM2 data; Emmanuel Boss, Karl Banse, Jerry Wiggert,Paula Bontempi, Eric D’Asaro, Michael Bender, David Seigel,Giorgio Dall’Olmo, Kimberly Halsey, Allen Milligan, PatrickSchultz, Paul Schrader, Toby Westberry, and Ronald Zaneveldfor helpful discussions and manuscript comments; SebastianDiehl and an anonymous reviewer for exceptional, thorough,and challenging reviews; and the NASA Goddard Ocean Colorresearch team for their continual dedication to achievinghighest quality satellite ocean color data, without which thecurrent work would not have been possible. This study wassupported by the National Aeronautics and Space Administra-tion Ocean Biology and Biogeochemistry Program under grantsNNX08AK70G, NNG05GD16G, and NNG05GR50G.

LITERATURE CITED

Backhaus, J. O., E. Hegseth, H. Wehde, X. Irigoien, K. Hatten,and K. Logemann. 2003. Convection and primary produc-tion in winter. Marine Ecology Progress Series 215:1–14.

Backhaus, J., H. Wehde, E. N. Hegseth, and J. Kampf. 1999.‘‘Phytoconvection’’—on the role of oceanic convection inprimary production. Marine Ecology Progress Series 189:77–92.

Banse, K. 1992. Grazing, temporal changes of phytoplanktonconcentrations, and the microbial loop in the open sea. Pages409–440 in P. G. Falkowski and A. D. Woodhead, editors.Primary productivity and biogeochemical cycles in the sea.Plenum Press, New York, New York, USA.

Banse, K. 1994. Grazing and zooplankton production as keycontrols of phytoplankton production in the open ocean.Oceanography 7:13–20.

Banse, K. 2002. Steeman Nielsen and the zooplankton.Hydrobiology 480:15–28.

Banse, K., and D. C. English. 1994. Seasonality of coastal zonecolor scanner phytoplankton pigment in the offshore oceans.Journal of Geophysical Research 99:7323–7345.

Barber, R. T., J. Marra, R. R. Bidigare, L. A. Codispoti, D.Halpern, Z. Johnson, M. Latasa, R. Goericke, and S. L.Smith. 2001. Primary productivity and its regulation in theArabian Sea during 1995. Deep-Sea Research II 48:1127–1172.

Behrenfeld, M. J., and E. Boss. 2003. The beam attenuation tochlorophyll ratio: an optical index of phytoplankton photo-acclimation in the surface ocean? Deep Sea Research 50:1537–1549.

Behrenfeld, M. J., and E. Boss. 2006. Beam attenuation andchlorophyll concentration as alternative optical indices ofphytoplankton biomass. Journal of Marine Research 64:431–451.

Behrenfeld, M. J., E. Boss, D. A. Siegel, and D. M. Shea. 2005.Carbon-based ocean productivity and phytoplankton phys-iology from space. Global Biogeochemical Cycles 19:GB1006.

Behrenfeld, M. J., and P. G. Falkowski. 1997. Photosyntheticrates derived from satellite-based chlorophyll concentration.Limnology and Oceanography 42:1–20.

Behrenfeld, M. J., D. A. Siegel, and R. T. O’Malley. 2008.Global ocean phytoplankton and productivity. Pages S56–S61 in J. M. Levy, editor. State of the climate in 2007.Bulletin of the American Meteorlogical Society, July 2008.

Behrenfeld, M. J., D. A. Siegel, R. T. O’Malley, and S.Maritorena. 2009. Global ocean phytoplankton. Pages 568–573 in T. C. Peterson and M. O. Baringer, editors. State ofthe climate in 2008. Bulletin of the American MeteorlogicalSociety, August 2009.

Bleck, R., C. Rooth, D. Hu, and L. T. Smith. 1992. Salinity-driven thermocline transients in a wind- and thermohaline-

forced isopycnic coordinate model of the North Atlantic.Journal of Physiological Oceanography 22:1486–1505.

Boss, E., D. Swift, L. Taylor, P. Brickley, R. Zaneveld, S. Riser,M. J. Perry, and P. G. Strutton. 2008. Observations ofpigment and particle distributions in the Western NorthAtlantic from an autonomous float and ocean color satellite.Limnology and Oceanography 53:2112–2122.

Campbell, J., et al. 2002. Comparison of algorithms forestimating primary productivity from surface chlorophyll,temperature and irradiance. Global Biogeochemical Cycles16(3):10.1029.

Carr, M.-E., et al. 2006. A comparison of global estimates ofmarine primary production from ocean color. Deep SeaResearch 53:741–770.

Clancy, R. M., and W. D. Sadler. 1992. The fleet numericaloceanography center suite of oceanographic models andproducts. Weather and Forecasting 7:307–327.

Crumpton, W. G., and R. G. Wetzel. 1982. Effects ofdifferential growth and mortality in the seasonal successionof phytoplankton populations in Lawrence Lake, Michigan.Ecology 63:1729–1739.

Dale, T., F. Rey, and B. R. Heimdal. 1999. Seasonaldevelopment of phytoplankton at a high latitude ocean site.Sarsia 84:1–17.

Dall’Olmo, G., T. K. Westberry, M. J. Behrenfeld, E. Boss, andW. Slade. 2009. Significant contribution of large particles tooptical backscattering in the open ocean. Biogeosciences 6:947–967.

D’Asaro, E. A. 2008. Convection and the seeding of the NorthAtlantic bloom. Journal of Marine Systematics 69:233–237.

Ellertsen, H. C. 1993. Spring blooms and stratification. Nature363:24.

Evans, G. T., and M. A. Paranjape. 1992. Precision of estimatesof phytoplankton growth and microzooplankton grazingwhen the functional response of grazers may be nonlinear.Marine Ecology Progress Series 80:285–290.

Evans, G. T., and J. S. Parslow. 1985. A model of annualplankton cycles. Biological Oceanography 3:327–347.

Garver, S. A., and D. A. Siegel. 1997. Inherent optical propertyinversion of ocean color spectra and its biogeochemicalinterpretation: I. Time series from the Sargasso Sea. Journalof Geophysical Research 102:18607–18625.

Gran, H. H., and T. Braarud. 1935. A quantitative study on thephytoplankton of the Bay of Fundy and the Gulf of Maine(including observations on hydrography, chemistry andmorbidity). Journal Biology Board Canada 1:219–467.

Halldal, P. 1953. Phytoplankton investigations from theweather ship M in the Norwegian Sea. Hvalrad Skr 38:1–91.

Hatun, H., A. B. Sandø, H. Drange, B. Hansen, and H.Valdimarsson. 2005. Influence of the Atlantic subpolar gyreon the thermohaline circulation. Science 309:1841–1844.

Heimdal, B. R. 1974. Composition and abundance of phyto-plankton in the Ullsfjord area, North Norway. Astarte 7:17–42.

Henson, S. A., J. P. Dunne, and J. L. Sarmiento. 2008. Decadalvariability in North Atlantic phytoplankton blooms. Journalof Geophysical Research 114(c4):C04013.

Huot, Y., M. Babin, F. Bruyant, C. Grob, M. S. Twardowski,and H. Claustre. 2007. Relationship between photosyntheticparameters and different proxies of phytoplankton biomassin the subtropical ocean. Biogeoscience 4:853–868.

Jackson, G. A. 2005. Coagulation theory and models of oceanicplankton aggregation. Pages 271–292 in I. Droppo, G.Leppard, S. Liss, and T. Milligan, editors. Flocculation innatural and engineered environments. CRC Press, BocaRaton, Florida, USA.

Kortzinger, A., U. Send, R. S. Lampitt, S. Hartman, D. W. R.Wallace, J. Karstensen, M. G. Villagarcia, O. Llinas, andM. D. DeGrandpre. 2008. The seasonal pCO2 cycle at 498 N/16.58 W in the northeastern Atlantic Ocean and what it tells

MICHAEL J. BEHRENFELD988 Ecology, Vol. 91, No. 4

Page 13: Abandoning Sverdrup’s Critical Depth Hypothesis on ...kudela/migrated/OS130/Readings/Behrenfeld...Abandoning Sverdrup’s Critical Depth Hypothesis on phytoplankton blooms ... many

us about biological productivity. Journal of GeophysicalResearch 113:C04020.

Landry, M. R. 1993. Estimating rates of growth and grazingmortality of phytoplankton by the dilution method. Pages715–722 in P. F. Kemp, B. F. Sherr, E. B. Sherr, and J. J.Cole, editors. Handbook of methods in aquatic microbialecology. Lewis Publishers, Boca Raton, Florida, USA.

Landry, M. R., and R. P. Hassett. 1982. Estimating the grazingimpact of marine micro-zooplankton. Marine Biology 67:283–288.

Landry, M. R., J. Kirshtein, and J. Constantinou. 1995. Arefined dilution technique for measuring the communitygrazing impact of microzooplankton, with experimental testsin the central equatorial Pacific. Marine Ecology ProgressSeries 120:53–63.

Margalef, R., F. Munoz, and J. Herrera. 1957. Fitoplancton delas costas de Castellon de enero de 1955 a junio de 1956.Investagaciones Pesqueras 7:3–31.

Maritorena, S., D. A. Siegel, and A. R. Peterson. 2002.Optimization of a semianalytical ocean color model forglobal-scale applications. Applied Optics 41:2705–2714.

Marra, J., and R. T. Barber. 2005. Primary productivity in theArabian Sea: a synthesis of JGOFS data. Progress inOceanography 65:159–175.

Milutinovi!c, S., M. J. Behrenfeld, J. A. Johannessen, and T.Johannessen. 2009. Sensitivity of remote sensing-derivedphytoplankton productivity to mixed layer depth: lessonsfrom the carbon-based productivity model. Global Biogeo-chemistry Cycles 23:GB4005.

Morel, A. 1991. Light and marine photosynthesis: a spectralmodel with geochemical and climatological implications.Progress in Oceanography 26:263–306.

Paasche, E. 1960. Phytoplankton distribution in the NorwegianSea in June, 1954, related to hydrography and compared withprimary production data. Fiskeridir Skr Ser Havunders 12:1–77.

Paasche, E., and A. M. Rom. 1962. On the phytoplanktonvegetation of the Norwegian Sea in May 1958. Nytt Magasinfor Botanik 9:33–60.

Parsons, T. R., and C. M. Lalli. 1988. Comparative oceanicecology of the plankton communities of the subarcticAtlantic and Pacific oceans. Oceanography and MarineBiology Annual Review 26:317–359.

Platt, T., D. F. Bird, and S. Sathyendranath. 1991. Criticaldepth and marine primary production. Proceedings of theRoyal Society B 246:205–217.

Ramsfjell, E. 1960. Phytoplankton distribution in the Norwe-gian Sea in June, 1952 and 1953. Fiskeridir Skr SerHavunders 12:1–112.

Reynolds, C. S. 2006. The ecology of phytoplankton. Cam-bridge University Press, Cambridge, UK.

Riley, G. A. 1946. Factors controlling phytoplankton popula-tions on Georges Bank. Journal of Marine Research 6:54–72.

Schei, B. 1974. Phytoplankton investigations in Skjomen, afjord in North Norway, 1970–1971. Astarte 7:43–59.

Siegel, D. A., S. C. Doney, and J. A. Yoder. 2002a. The NorthAtlantic spring bloom and Sverdrup’s critical depth hypoth-esis. Science 296:730–733.

Siegel, D. A., et al. 2002b. Global distribution and dynamicsof colored dissolved and detrital organic materials, Journalof Geophysical Research 107:3228.

Siegel, D. A., S. Maritorena, N. B. Nelson, and M. J.Behrenfeld. 2005. Independence and interdependences amongglobal ocean color properties: reassessing the bio-opticalassumption. Journal of Geophysical Research 110:C07011.

Smetacek, V., and U. Passow. 1990. Spring bloom initiationand Sverdrup’s critical-depth model. Limnology and Ocean-ography 35:228–234.

Steemann Nielsen, E. 1957. The general background of oceanicproductivity. Pages 91–120 in E. Steemann Nielsen and E. A.Jensen, editors. Primary oceanic production. The autotrophicproduction of organic matter in the ocean. Galathea Report1:43–136.

Sverdrup, H. U. 1953. On conditions for the vernal blooming ofphytoplankton. Journal du Conseil International pourl’Exploration de la Mer 18:287–295.

Townsend, D. W., M. D. Keller, M. E. Sieracki, and S. G.Ackleson. 1992. Spring phytoplankton blooms in the absenceof vertical water column stratification. Nature 360:59–62.

Ward, B. A., and J. J. Waniek. 2007. Phytoplankton growthconditions during autumn and winter in the Irminger Sea,North Atlantic. Marine Ecology Progress Series 334:47–61.

Westberry, T., M. J. Behrenfeld, D. A. Siegel, and E. Boss.2008. Carbon-based primary productivity modeling withvertically resolved photoacclimation. Global BiogeochemicalCycles 22:GB2024.

APPENDIX

Figures showing nine-year records of phytoplankton biomass and Chlsat at eight-day resolution for the 12 North Atlantic bins,mean annual cycles in phytoplankton carbon and Chlsat at eight-day resolution for the 12 North Atlantic bins, and 58 latitude zonalmean values of phytoplankton net specific growth rate during the positive phase of the annual cycle and spring maximum in surfacechlorophyll concentration (Ecological Archives E091-069-A1).

April 2010 989ABANDONING SVERDRUP’S BLOOM HYPOTHESIS