Subsurface maxima of phytoplankton and chlorophyll: Steady‐state solutions from a simple model

In oligotrophic lakes and oceans, the deep chlorophyll maximum may form independently of a maximum of phytoplankton biomass, because the ratio of chlorophyll to phytoplankton biomass (in units of carbon) increases with acclimation to reduced light and increased nutrient supply at depth. Optical data (beam attenuation as proxy for phytoplankton biomass and chlorophyll fluorescence and absorption as proxies for chlorophyll concentration) and conventional measurements of biovolume, particulate organic carbon, and chlorophyll from two oligotrophic systems (Crater Lake, Oregon, and Sta. ALOHA in the subtropical North Pacific Ocean) are presented and show a vertical separation of the maxima of biomass and chlorophyll by 50‐80 m during stratified conditions. We use a simple mathematical framework to describe the vertical structure of phytoplankton biomass, nutrients, and chlorophyll and to explore what processes contribute to the generation of vertical maxima. Consistent with the observations, the model suggests that biomass and chlorophyll maxima in stable environments are generated by fundamentally different mechanisms. Maxima in phytoplankton biomass occur where the growth rate is balanced by losses (respiration and grazing) and the divergence in sinking velocity, whereas the vertical distribution of chlorophyll is strongly determined by photoacclimation. A deep chlorophyll maximum is predicted well below the particle maximum by the model. As an interpretation of these results, we suggest a quantitative criterion for the observed coexistence of vertically distinct phytoplankton assemblages in oligotrophic systems: the vertical position at which a species occurs in highest abundance in the water column is determined by the ‘‘general compensation depth’’— that is, the depth at which specific growth and all loss rates, including the divergence of sinking/swimming and vertical mixing, balance. This prediction can be tested in the environment when the divergence of sinking and swimming is negligible.

[1]  Richard J. Geider,et al.  Seasonal and latitudinal dependencies of phytoplankton carbon-to-chlorophyll a ratios: results of a modelling study , 1997 .

[2]  J.,et al.  Number 3 A Study of Production in the Gulf of Mexico ' , 2022 .

[3]  Hugh L. MacIntyre,et al.  A dynamic model of photoadaptation in phytoplankton , 1996 .

[4]  Ricardo M Letelier,et al.  Seasonal and interannual variations in photosynthetic carbon assimilation at Station , 1996 .

[5]  D. Kiefer,et al.  A steady state description ofgrowth and light absorption in the marine planktonic diatom Skeletonema costatum , 1989 .

[6]  David M. Karl,et al.  Seasonal and interannual variability in primary production and particle flux at Station ALOHA , 1996 .

[7]  James K. B. Bishop,et al.  Transmissometer measurement of POC , 1999 .

[8]  R. Eppley,et al.  Chlorophyll maximum layers of the southern-california bight and possible mechanisms of their formation and maintenance , 1981 .

[9]  J. Raven,et al.  LIGHT DEPENDENCE OF GROWTH AND PHOTOSYNTHESIS IN PHAEODACTYLUM TRICORNUTUM (BACILLARIOPHYCEAE) 1 , 1985 .

[10]  J. Zaneveld,et al.  Intermediate Nepheloid Layers Observed off Oregon and Washington , 1980 .

[11]  K. Mann,et al.  Dynamics of Marine Ecosystems , 1991 .

[12]  E. D’Asaro,et al.  Observations of Persistent Mixing and Near-Inertial Internal Waves , 1986 .

[13]  J. Cullen,et al.  Changes in buoyancy and chemical composition during growth of a coastal marine diatom: ecological and biogeochemical consequences , 1995 .

[14]  P. Falkowski Light-Shade Adaptation in Marine Phytoplankton , 1980 .

[15]  Sallie W. Chisholm,et al.  Phytoplankton population dynamics at the Bermuda Atlantic Time-series station in the Sargasso Sea , 2001 .

[16]  M. Perry,et al.  Phytoplankton rate processes in the oligotrophic waters of the central North Pacific Ocean , 1980 .

[17]  T. Kana,et al.  Dynamic model of phytoplankton growth and acclimation: responses of the balanced growth rate and the chlorophyll a:carbon ratio to light, nutrient-limitation and temperature , 1997 .

[18]  Tommy D. Dickey,et al.  Optical and physical variability on timescales from minutes to the seasonal cycle on the New England shelf: July 1996 to June 1997 , 2001 .

[19]  P. Franks Thin layers of phytoplankton: a model of formation by near-inertial wave shear , 1995 .

[20]  Lisa R. Moore,et al.  Photophysiology of the marine cyanobacterium Prochlorococcus: Ecotypic differences among cultured isolates , 1999 .

[21]  E. Venrick Phytoplankton in an Oligotrophic Ocean: Observations and Questions , 1982 .

[22]  R. Lande,et al.  Suspension times of particles in the upper ocean , 1987 .

[23]  J. Steele,et al.  The vertical distribution of chlorophyll , 1960, Journal of the Marine Biological Association of the United Kingdom.

[24]  James C. Kitchen,et al.  On the noncorrelation of the vertical structure of light scattering and chlorophyll α in case I waters , 1990 .

[25]  Ann E. Gargett,et al.  Vertical eddy diffusivity in the ocean interior , 1984 .

[26]  J. Kremer,et al.  Origins of vertical patterns of phytoplankton and nutrients in the temperate, open ocean: a stratigraphic hypothesis , 1981 .

[27]  T. G. Owens,et al.  Light: shade adaptation , 1980 .

[28]  D. Redalje,et al.  Subsurface chlorophyll maximum in August-September 1985 in the CLIMAX area of the North Pacific , 1988 .

[29]  E. Venrick,et al.  Deep maxima of photosynthetic chlorophyll in the Pacific Ocean , 1973 .

[30]  P. Falkowski,et al.  Adaptation of the Cyanobacterium Microcystis aeruginosa to Light Intensity. , 1983, Plant physiology.

[31]  David A. Link,et al.  The Relationship Between Sphere Size And Settling Velocity , 1971 .

[32]  Dale A. Kiefer,et al.  Meridional variations in the concentration of chlorophyll and microparticles in the North Pacific Ocean , 1988 .

[33]  A. Morel Optical modeling of the upper ocean in relation to its biogenous matter content (case I waters) , 1988 .

[34]  M. Dacey,et al.  Stokes’ Settling and Chemical Reactivity of Suspended Particles in Natural Waters , 1974 .

[35]  C. Gibbs Chlorophyll b Interference in the Fluorometric Determination of Chlorophyll a and 'Phaeo-Pigments' , 1979 .

[36]  Nicholas R. Bates,et al.  Building the Long-Term Picture: The U.S. JGOFS Time-Series Programs , 2001 .

[37]  Richard J. Geider,et al.  LIGHT AND TEMPERATURE DEPENDENCE OF THE CARBON TO CHLOROPHYLL a RATIO IN MICROALGAE AND CYANOBACTERIA: IMPLICATIONS FOR PHYSIOLOGY AND GROWTH OF PHYTOPLANKTON , 1987 .

[38]  Dale A. Kiefer,et al.  A two‐component description of spectral absorption by marine particles , 1989 .

[39]  G. Evensen,et al.  Assimilation of ocean colour data into a biochemical model of the North Atlantic: Part 1. Data assimilation experiments , 2003 .

[40]  R. Lande,et al.  Phytoplankton growth rates estimated from depth profiles of cell concentration and turbulent diffusion , 1989 .

[41]  W. Gardner,et al.  Effects of monsoons on the seasonal and spatial distributions of POC and chlorophyll in the Arabian Sea , 1998 .

[42]  R. Olson,et al.  Another look at the nitrite and chlorophyll maxima in the central North Pacific , 1976 .

[43]  M. D. Hurley,et al.  Temperature, Water Chemistry, and Optical Properties of Crater Lake , 1996 .

[44]  M. Friedrichs Assimilation of JGOFS EqPac and SeaWiFS data into a marine ecosystem model of the Central Equatorial Pacific Ocean , 2001 .

[45]  Andrew H. Barnard,et al.  Spectral particulate attenuation and particle size distribution in the bottom boundary layer of a continental shelf , 2001 .

[46]  James K. B. Bishop,et al.  Spatial and temporal variability of POC in the northeast Subarctic Pacific , 1999 .

[47]  Ricardo M Letelier,et al.  Seasonal and interannual variations in photosynthetic carbon assimilation at Station ALOHA , 2003 .

[48]  John J. Cullen,et al.  The deep chlorophyll maximum comparing vertical profiles of chlorophyll a , 1982 .

[49]  E. Laws,et al.  Nutrient‐ and light‐limited growth of Thalassiosira fluviatilis in continuous culture, with implications for phytoplankton growth in the ocean , 1980 .

[50]  P. Glibert,et al.  Effect of irradiances up to 2000 μE m−2 s−1 on marine Synechococcus WH7803—I. Growth, pigmentation, and cell composition , 1987 .

[51]  M. Perry,et al.  Modeling in situ phytoplankton absorption from total absorption spectra in productive inland marine waters , 1989 .

[52]  P. Harrison,et al.  Sinking-rate response of natural assemblages of temperate and subtropical phytoplankton to nutrient depletion , 1984 .

[53]  Paul G. Falkowski,et al.  A consumer's guide to phytoplankton primary productivity models , 1997 .

[54]  M. Altabet,et al.  Nitrogen transport by vertically migrating diatom mats in the North Pacific Ocean , 1993, Nature.

[55]  E. Laws,et al.  of a marine diatom: implications to dynamics of chlorophyll maximum layers , 1983 .

[56]  Richard F. Davis,et al.  Reducing the effects of fouling on chlorophyll estimates derived from long‐term deployments of optical instruments , 1997 .

[57]  André Morel,et al.  Light scattering and chlorophyll concentration in case 1 waters: A reexamination , 1998 .

[58]  Walker O. Smith,et al.  Seasonal patterns of water column particulate organic carbon and fluxes in the Ross Sea, Antarctica , 2000 .

[59]  C. D. Mcintire,et al.  Taxonomic structure and productivity of phytoplankton assemblages in Crater Lake, Oregon , 1996 .

[60]  Pak,et al.  Meridional variations in the concentration of chlorophyll and microparticles in the North Pacific Ocean cc : , 2022 .