Chapter 9

Biogeochemical modelling


CHAPTER
COORDINATORS

Elodie Gutknecht
CHAPTER
AUTHORS

Laurent Bertino, Pierre Brasseur, Stefano Ciavatta, Gianpiero Cossarini, Katja Fennel, David Ford, Marilaure Grégoire, Diane Lavoie, and Patrick Lehodey

9.1 General introduction to Biogeochemical models

Marine biogeochemistry is the study of essential chemical elements in the ocean (such as carbon, nitrogen, oxygen, and phosphorus), and of their interactions with marine organisms. Biogeochemical cycles are driven by physical transport, chemical reactions, absorption, and transformation by plankton and other organisms, which form the basis of the oceanic food web. 

In the last decades, the interest for this cross-disciplinary science has greatly increased due to the occurrence of significant changes in the marine environment closely linked to the alteration of the biogeochemical cycles in the ocean. These alterations include phenomena such as acidification, coral bleaching, eutrophication, deoxygenation, harmful algal blooms, regime shifts in plankton, invasive species, and other processes deteriorating water quality and impacting the whole marine ecosystem. 

Monitoring and forecasting the biogeochemical and ecosystem components of the ocean, also referred to as “Green Ocean”, are essential for a better understanding of the current status and changes in ocean health and ecosystem functioning. Such operational systems provide indicators useful to scientists, industry (e.g. fisheries and aquaculture), policy makers and environmental agencies for the prediction of events, the management of living marine resources, and can support the decision-making process to respond to environmental changes. 

This chapter gives an overview of the Green Ocean component of OOFS. The first section addresses the objectives, applications and beneficiaries of the Green Ocean and introduces the fundamental theoretical knowledge of marine biogeochemical modelling. The second section details and discusses each component of a biogeochemical OOFS to guide new forecasters in biogeochemistry. Modelling of higher trophic levels is introduced. Finally, several operational systems are mentioned as examples.

9.1.1 Objective, applications and beneficiaries

Human activities, primarily the combustion of fossil fuels, cement production, and the industrial production of nitrogen-based fertilisers, are leading to ocean warming, acidification, deoxygenation, and coastal eutrophication, thus putting ever-increasing and compounding pressures on marine ecosystems (Figure 9.1).

Figure 9.1. Threats on marine ecosystems. Changes and alterations in the marine environment observed in recent decades include acidification, coral bleaching, eutrophication, deoxygenation, harmful algal blooms, changes in planktonic regimes, invasive species, etc.

At the same time, the ocean is serving as a major sink of carbon dioxide (CO2), the most important anthropogenic greenhouse gas. This contributes to mitigating global warming, but the magnitude of this sink is likely to diminish. Our ability to quantify these phenomena and project their future course hinges on a mechanistic understanding of the biogeochemical cycles of carbon, oxygen, and nutrients in the ocean and how they are changing.

The Marine BGC, the study of elemental cycles and their interactions with the environment and living organisms, is a multidisciplinary science at the crossroads between ocean physics, chemistry, and biology, and intersects with atmospheric and terrestrial sciences as well as social science and environmental policy. As an example, Figure 9.2 illustrates the complex carbon cycle in the ocean and the interactions between biological, chemical, and physical processes.

Figure 9.2. Cycling of carbon in the marine food chain. Phytoplankton assimilate CO2 via photosynthesis in the euphotic zone and are consumed by zooplankton. Zooplankton are the initial prey for many small and large aquatic organisms. Carbon is thus transferred further up the food web to higher-level predators. Different mechanisms contribute to the export and storage of carbon into the deep ocean. The carbon cycle in the ocean is complex and influenced by biological, chemical, and physical processes - credit: Oak Ridge National Laboratory at https://www.ornl.gov/

Ocean BGC models describe the base of the marine food chain from bacteria to mesozooplankton and couple the cycles of carbon (C), nitrogen (N), oxygen (O2), phosphorus (P) and silicon (Si). They mostly focus on plankton, classifying the plankton diversity in accordance with their functional characteristics, the so-called Plankton Functional Types (PFTs). Species at higher trophic levels, such as fish and marine mammals play a lesser role in elemental cycling, they are thus generally not explicitly represented in BGC models, but they are very important for ecosystem models that focus on the ecology/biology of marine organisms. BGC and ecosystem models are sometimes referred to indistinctly because they can overlap in their representation of the lower trophic levels. Specific modelling approaches, like Lagrangian modelling, habitat modelling, or food web models, are used to connect BGC with the high trophic levels (e.g. fish).

The implementation of accurate OOFSs requires sustained, systematic, and NRT observation from (sub)mesoscale to large scale to initialise, parameterize, and validate ocean models. NRT information in operational oceanography means a description of the present situation with a delay of a few minutes to a few days.

The forecast of ocean physics has considerably improved in the last decades, reaching a high level of predictability (Chapter 5). The evolving equations governing the physical dynamics are based on physical laws, the model parameterizations are quite well-established, and the abundance of observations for temperature, salinity, and sea level height offers a way to improve model predictions through data assimilation. Forecasting of the Green Ocean has been developed more recently and it has not yet reached the same level of maturity, in most cases being incorporated into already existing physical OOFS. The formulation of ecosystem models is still empirical and the scarcity of in-situ biological and BGC data critically limits the capabilities to constrain their parameterization and to improve their performances through a robust data-model comparison exercise and data assimilation. The scarcity of data is even more critical in NRT, limiting data assimilation to surface chlorophyll-a (Chla) derived from satellite reflectance (Fennel et al., 2019).

The advent of in-situ robotic platforms combined with high resolution satellite products for the Green Ocean have the potential to palliate this deficiency. For instance, the advent of hyper-spectral satellites is promising in terms of delivery of surface information on PFTs, detection of harmful algal blooms, and benthic habitat mapping, while the boost in robotic platforms will offer huge opportunities to map the (deep) seafloor with an unprecedented level of details. The combination of marine robotics, image analysis, machine learning, new sensor development, and the coordination of robotic platforms and satellite sensors will constitute a significant breakthrough in our knowledge of marine ecosystems. All this information would need to be integrated in models for forecasting and producing high quality reanalysis of the Green Ocean to support the production of added value products and innovative services. Coordination of Ocean OSSEs can help to design the new observing biological and biogeochemical systems with maximal impact to users, yet their development is still insufficient and should be encouraged (Le Traon et al., 2019).

Ultimately, BGC OOFS systems serve major environmental and societal issues, including the Ocean's role in the global carbon cycle and the impacts of natural changes and anthropogenic stressors in the physical-chemical marine environment on ecosystems and human activities. Applications range from multi-decadal retrospective simulations (namely, “reanalyses”), operational analysis of the current conditions (“nowcasts”), short-term and seasonal predictions (“forecasts”), scenario simulations, and climate change projections. These integrated systems are essential not only for a better understanding of the current status of key biogeochemical and ecosystem processes in the ocean and how they are changing, but also to provide stakeholders, policy makers and environmental agencies with indicators of ocean health in order to take appropriate mitigation, adaptation, conservation, and protection measures for living marine organisms and their habitats but also for human health.

“A predicted ocean whereby society has the capacity to understand current and future ocean conditions, forecast change and impact on human wellbeing and livelihoods” is an expected outcome of the United Nations Decade of Ocean Science for Sustainable Development, 2021-2030 (Ryabinin et al., 2019), supported also by the Sustainable Development Goals 14 (Life below water), 8 (Decent work and economic growth), and 9 (Industry, innovation and infrastructure).

9.1.2 Fundamental theoretical background

9.1.2.1 Biogeochemical modelling

Plankton (including phytoplankton and zooplankton) are organisms which are carried by tides and currents, or do not swim well enough to move against them. They form the base of the marine ecosystem and are a central component of the BGC models that simulate the cycling of elements through seawater and plankton.

Most models take an “NPZD” approach, simulating:

  • Nutrients: substances which organisms require for growth.
  • Phytoplankton: microscopic algae which obtain energy from sunlight through photosynthesis.
  • Zooplankton: planktonic animals which obtain energy by eating other organisms.
  • Detritus: dead and excreted organic matter.

Each of these is represented by one or more state variables, depending on the complexity of the model. Rather than considering individual organisms, state variables represent concentrations of elements such as nitrogen or carbon. They are often called tracers because they are transported and diffused by ocean dynamics.

As with physical models, BGC models are discretized on a grid covering the region of interest and require suitable initial and boundary conditions for each state variable. At each grid point, the evolution of a state variable C is given by the equation:

where ∇ ∙ (CU) and DC are the advection and diffusion terms equivalent to those used for temperature and salinity in physical models (please refer to Chapter 5). ∇ is the generalised derivative vector operator, t is the time, U the vector velocity, and Dc is the parameterization of small-scale physics for the tracer. The SMS(C) stands for source-minus-sink terms for the tracer C and represents the BGC processes simulated by the model. Each 1D water column is normally treated independently, with lateral interactions limited to advection and diffusion. Most BGC models are formulated to conserve mass.

Figure 9.3. Schematic of a basic NPZD model considering four state variables, one for each compartment.

Unlike the Navier-Stokes equations for physical models (Chapter 5), there is no known set of laws defining biological behaviour. Instead, empirical relationships are used to describe observed processes such as growth and mortality.

The basic source-minus-sink terms usually modelled in a NPZD model (Figure 9.3) are:

  • Phytoplankton growth or Primary production: the creation of organic matter through photosynthesis. It is a function of phytoplankton concentration, nutrient availability, and light availability. It can also be regulated by temperature.
  • Grazing: zooplankton eating phytoplankton and detritus.
  • Mortality: death through natural causes, e.g. viruses, predation by higher trophic levels (fish and marine mammals), etc.
  • Messy feeding: zooplankton graze inefficiently, and a proportion of organic matter enters the nutrient or detritus pool rather than being ingested by zooplankton.
  • Remineralisation: bacteria break down the organic matter in detritus, which is converted back to nutrients.
  • Sinking: detritus sinks through the water column due to gravity.

In this case, the differential equations for phytoplankton (P), zooplankton (Z), detritus (D), and nutrients (N) are as follows:

where phytoplankton evolution depends on primary production, grazing and mortality;

where zooplankton evolution depends on grazing and mortality;

where detritus evolution depends on mortality, grazing, messy feeding, remineralisation and sinking;

where nutrients evolution depends on primary production, messy feeding, and remineralisation.

µP is the growth rate of phytoplankton due to photosynthesis; mP and mzare the mortality rates of phytoplankton and zooplankton; GP and GD are the grazing rates of zooplankton on phytoplankton and detritus; αDand αNrepresent the efficiency of the grazing; (1-αD) and (1-αN) the non-assimilated fractions of grazing by zooplankton that return to detritus and nutrients; remD is the remineralisation rate of detritus and wD is the sinking speed of detritus.

The exact equations used differ between models, the ones given above are common examples. Other processes are often considered as well, notably respiration, excretion, and egestion, which cause loss of organic matter. Of course, additional processes may be included in more complex models.

The processes can be modelled using different mathematical forms, often with parameter values which are uncertain and can be tuned. While sinking and mortality rates are usually single parameters (linear functions), phytoplankton growth rate requires multiple parameters. µP is usually a function of nutrients, light and temperature:

µPmax is the maximum growth rate, f(T) is the temperature effect, f(I) and f(N) are the limitation terms due to light and nutrients. Different formulations exist for each of these terms, but usually NPZD-type models characterise nutrient limitation of phytoplankton growth rate using Michaelis-Menten kinetics:

K is known as the half-saturation constant for nutrient uptake, and N is the nutrient concentration. If nutrient is plentifully available, then N/(K+N) ≈1 and phytoplankton growth is not limited by the nutrient.

The state variables of NPZD models represent concentrations of a given chemical element, often nitrogen, with other elements such as carbon derived using constant stoichiometry between carbon, nitrogen and phosphorus, i.e. the Redfield ratio of 106:16:1 (Redfield, 1934).

More complex models include additional variables for each compartment. Phytoplankton can be split into PFTs, grouping together species which perform a similar function within the ecosystem (Le Quéré et al., 2005). PFTs are often based on organism size. It is also common to separate out diatoms, which form silicate shells and play an important role in the sinking of carbon. In models, PFTs are distinguished by differing parameters for traits such as maximum phytoplankton growth rates, grazing, and nutrient affinity. Zooplankton can also be split into functional types, again often based on size, with different feeding preferences. Note that the current paradigm neglects the fact that many plankton are mixotrophs: they both photosynthesize and eat other organisms (Flynn et al., 2013; Glibert et al., 2019).

Variable stoichiometry (elemental ratios) can also be introduced. Each PFT is then described by separate state variables for each element, such as nitrogen, carbon, and phosphorus.

Chla is often included into BGC models as it is the main photosynthetic pigment found in phytoplankton, and measurement of its concentration in water is used as an indicator of the phytoplankton biomass. Chla can be represented as a constant ratio to the carbon biomass, or a variable ratio depending on nutrient, light levels, and temperature (Geider et al., 1997).

Most models incorporate dissolved inorganic nitrogen as a nutrient, which includes nitrate and ammonium. Phosphate and iron may be modelled too, and silicate if diatoms are a PFT. Nutrient inputs from rivers and the atmosphere can also be specified. Detritus may be split into different sizes, with different sinking rates, and into different elements. Some models explicitly simulate bacteria and viruses, rather than just parameterising their effects.

Besides NPZD variables, models can also include other related processes, such as the oxygen and carbon cycles. The carbon cycle is usually represented by the state variables DIC and total alkalinity, the latter being the capacity of seawater to neutralise an acid. From these and other variables, quantities such as pH and air-sea CO2 flux can be calculated (Zeebe and Wolf-Gladrow, 2001).

BGC models are closely related to higher trophic level models or ecosystem models. The latter require the underlying biogeochemistry, and BGC models require at least some parameterisation of the ecosystem, i.e. the explicit representation of part of the living component of the ocean (e.g. phytoplankton, zooplankton) with zooplankton mortality as a closure term, parameterising the predation of zooplankton by higher trophic levels such as fish and top predators (see Section 9.2.8).

Adding complexity to BGC models means that less important processes are neglected or amalgamated, but also increases the uncertainties associated with approximated formulations. There is no consensus on optimal structure and complexity, which will vary depending on the purpose (Fulton et al., 2003). Adding extra variables also increases computational cost, split between the computation of transport (advection and diffusion) for each state variable and the computation of the non-linear functions relating the state variables of the BGC model. In an operational context, the balance between model complexity and computational costs is critical and must be carefully evaluated. BGC models should be as simple as possible and as complex as necessary to answer specific questions.

9.1.2.2 Model calibration

As already mentioned, biogeochemical models are based on empirical relationships to describe the dynamics of biological processes. Observational data are then essential for tuning, adjusting or revising the formulations, i.e. making the model results match the observed distributions and fluxes of inorganic and organic quantities. Model calibration can be performed "by hand", i.e. by adjusting certain parameters of the biogeochemical models until the models show a "good" fit to the observed tracer fields, or by using objective optimisation methods (Kriest et al., 2020). The resulting set of biogeochemical parameters is often closely linked to the ocean circulation, mixing, and ventilation derived from the physical model used, with its specificities and defaults.

9.1.2.3 Physical-Biogeochemical coupling

Ocean physics advects and diffuses BGC model variables, thus redistributing inorganic and organic amounts. In addition, some BGC processes depend on physical conditions such as temperature or salinity, particularly crucial for the carbon cycle. Thus, there is a very strong link between the physical conditions and the BGC, which makes the BGC models closely dependent on the physical models.

Vertical motions are particularly critical to bring nutrients from nutrient-rich deep waters into the uppermost layer that receives the sunlight needed for photosynthesis and marine life. Two critical layers together regulate phytoplankton production:

  • The mixed layer is the upper layer of the ocean that interacts with the atmosphere. It is assumed to be mixed and homogeneous through convective/turbulent processes, generated by winds, surface heat fluxes, or processes modifying salinity. The deeper it is, the deeper phytoplankton are mixed, which will take them away from the light required for photosynthesis. Deep mixing also replenishes near-surface nutrient stocks.
  • The euphotic zone is the layer from the surface down to the depth at which irradiance is 1% of the surface irradiance. The deeper the euphotic depth, the deeper the layer in which photosynthesis and phytoplankton production can occur. It extends from a few metres in turbid estuaries to approximately two hundred metres in the open ocean.
Figure 9.4. Schematic representation of the interplay between mixed layer depth (yellow line) and upper-ocean euphotic zone (light blue area) on the initiation of phytoplankton bloom (modified from Dall'Olmo et al., 2016).

The mixed layer may develop within the euphotic layer (in stratified situations), or over a greater thickness of up to several hundred metres (in well-mixed situations). The interplay between these two critical layers controls the plankton exposure to sunlight and the coincident exposure to nutrients, thus regulating phytoplankton production (Figure 9.4). Exact mechanisms are still debated. Please refer to Ford et al. (2018) for more details.

In turn, phytoplankton abundance may feed back to physics, by absorbing radiation in the surface layers and therefore affecting heat penetration into the water column (Lengaigne et al., 2007).

9.1.2.4 From open ocean to coastal ecosystems

Different considerations are generally needed for open ocean and coastal ecosystems. In the open ocean, the seasonal cycle is quite well defined and recurring (Figure 9.5). Seasonal increases in temperature and solar radiation drive the phytoplankton spring bloom. The peak persists for a few weeks to months until nutrient limitation and grazing cause the bloom to collapse. A secondary biomass peak can develop in late summer or autumn.

Figure 9.5. Seasonal cycle of phytoplankton relative to variations in sunlight, nutrients, and zooplankton (Copyright: 2004 Pearson Prentice Hall, Inc).

In contrast, coastal ecosystems can be very complex, subject to a succession of blooms having different origins, thus requiring additional model complexity. Correct specification of river inputs also becomes more critical. Furthermore, the equations in Section 9.1.2.1 are for the pelagic (water column) ecosystem. In shallow waters, such as shelf seas, it becomes important to include the benthic (seafloor) ecosystem into the BGC models. This requires the addition of extra variables, though they do not need to be advected or diffused. Finally, coastal waters are often turbid, and the effect of sediments and coloured dissolved organic matter on light and therefore primary production should be included. Dedicated optical models are sometimes used for this purpose (Gregg and Rousseaux, 2016).

9.1.2.5 Potential predictability of ocean biogeochemistry

The potential predictability of ocean biogeochemistry varies considerably depending on the scales and quantities of interest. A lot of variability is driven by physics, with changes in mixing and stratification affecting light and nutrients and therefore primary production. When these physics changes can be predicted, e.g. changes in stratification with a warming climate and interannual variability related to phenomena such as the El Niño Southern Oscillation, associated large-scale changes to ocean biogeochemistry can also be predicted. Similarly, changes to the ocean carbon cycle and acidification with increasing atmospheric CO2 concentrations can be predicted. When considering local regions and/or shorter time scales, both physics and biogeochemistry become harder to be accurately predicted.

Furthermore, various biogeochemical quantities change at very different rates. Phytoplankton react quickly to changes in light and nutrient availability and can double in concentration over a day (Laws, 2013). Zooplankton will exhibit a slightly more lagged response to these changes. Meanwhile, nutrient concentrations will typically change more slowly, and the carbon cycle even more slowly, although surface concentrations (of nutrients and carbon) can change rapidly, for example during a storm. These different rates of change have implications for the scales of predictability.

For accurate predictions, it is important to initialise models using data assimilation (see Section 9.2.5). At seasonal-to-decadal time scales, predictability is dominated by physics, and this must be accurately initialised and simulated. Physics remains important at shorter time scales, but is essential to initialise nutrient concentrations correctly, as this will help to determine the primary productivity. For short-range predictions, phytoplankton concentrations should be initialised, though the memory of the phytoplankton variables may be as short as a few days, given that they react to changes in nutrients and mixing. Accurate model formulations and parameterisations are also required, otherwise the model will react incorrectly to the data assimilation.

References

Abecassis, M., Senina, I., Lehodey, P., Gaspar, P., Parker, D., Balazs, G., Polovina, J. (2013). A Model of Loggerhead Sea Turtle (Caretta caretta) Habitat and Movement in the Oceanic North Pacific. PLoS ONE, 8(9), e73274, https://doi.org/10.1371/journal.pone.0073274

Allen, J. I., M. Eknes, and G. Evensen. (2003). An Ensemble Kalman Filter with a complex marine ecosystem model: hindcasting phytoplankton in the Cretan Sea. Annales Geophysicae, 21(1), 399-411, https://doi.org/10.5194/angeo-21-399-2003

Almroth-Rosell, E., Eilola, K., Kuznetsov, Hall, I.P.O.J., and Meier, H.E.M. (2015). A new approach to model oxygen dependent benthic phosphate fluxes in the Baltic Sea. Journal of Marine Systems, 144, 127-141, https://doi.org/10.1016/j.jmarsys.2014.11.007

Anderson, L.A., Robinson, A.R., Lozano, C.J. (2000). Physical and biological modeling in the Gulf Stream region: I. Data assimilation methodology. Deep Sea Research Part I: Oceanographic Research Papers, 47, 1787-1827, https://doi.org/10.1016/S0967-0637(00)00019-4

Audzijonyte, A., Pethybridge, H., Porobic, J., Gorton, R., Kaplan, I., Fulton, E.A. (2019). Atlantis: A spatially explicit end-to-end marine ecosystem model with dynamically integrated physics, ecology and socio-economic modules. Methods in Ecology and Evolution, 10, 1814-1819, https://doi.org/10.1111/2041-210X.13272

Aumont, O., Ethé, C., Tagliabue, A., Bopp, L., and Gehlen, M. (2015). PISCES-v2: an ocean biogeochemical model for carbon and ecosystem studies. Geoscientific Model Development, 8, 2465-2513, https://doi.org/10.5194/gmd-8-2465-2015

Aumont O., Maury O., Lefort S., Bopp L. (2018). Evaluating the Potential Impacts of the Diurnal Vertical Migration by Marine Organisms on Marine Biogeochemistry. Global Biogeochemical cycles, 32 (11), 1622-1643, https://doi.org/10.1029/2018GB005886

Baretta, J. W., Ebenhöh, W., and Ruardij, P. (1995). The European regional seas ecosystem model, a complex marine ecosystem model. Netherlands Journal of Sea Research, 33, 233-246, https://doi.org/10.1016/0077-7579(95)90047-0

Bell, J.D., Ganachaud, A., Gehrke, P.C., Griffiths, S.P., Hobday, A.J., Hoegh-Guldberg, O., Johnson, J.E., Le Borgne, R., Lehodey, P., Lough, J.M., Matear, R.J., Pickering, T.D., Pratchett, M.S., Sen Gupta, A., Senina, I. and Waycott, M. (2013). Mixed response of tropical Pacific fisheries and aquaculture will respond differently to climate change. Nature Climate Change, 3, 591-599, https://doi.org/10.1038/nclimate1838

Bell, J.D., Senina, I., Adams, T., Aumont, O., Calmettes, B., Clark, S., Dessert, M., Hampton, J., Hanich, Q., Harden-Davies, H., Gehlen, M., Gorgues, T., Holmes, G., Lehodey, P., Lengaigne, M., Mansfield, B., Menkes, C., Nicol, S., Pasisi, C., Pilling, G., Ota, Y., Reid, C., Ronneberg, E., Sen Gupta, A., Seto, K., Smith, N., Taei, S., Tsamenyi, M., Williams, P. (2021). Pathways to sustaining tuna-dependent Pacific Island economies during climate change. Nature Sustainability, 4, 900-910, https://doi.org/10.1038/s41893-021-00745-z

Berline, L., Brankart, J.M., Brasseur, P., Ourmières, Y., and Verron, J. (2007). Improving the physics of a coupled physical–biogeochemical model of the North Atlantic through data assimilation: Impact on the ecosystem. Journal of Marine Systems, 64(1-4), 153-172, https://doi.org/10.1016/j.jmarsys.2006.03.007

Berthet, S., Séférian, R., Bricaud, C.,Chevallier, M., Voldoire, A., and Ethé, C. (2019). Evaluation of an online grid-coarsening algorithm in a global eddy-admitting ocean biogeochemical model. Journal of Advances in Modeling Earth Systems, 11, 1759-1783, https://doi.org/10.1029/2019MS001644

Bertino, L., Evensen, G., and Wackernagel, H. (2003). Sequential data assimilation techniques in oceanography. International Statistical Review, 71(2), 223-241, http://www.jstor.org/stable/1403885. 

Biogeochemical-Argo Planning Group (2016). The Scientific Rationale, Design andImplementation Plan for a Biogeochemical-Argo Float Array. https://doi.org/10.13155/46601

Blanchard, J.L., Heneghan, R.F., Everett, J.D., Trebilco, R., Richardson, A.J. (2017). From bacteria to whales: Using function size spectra to model marine ecosystems. Trends in Ecology and Evolution, 32(3), 174-186, https://doi.org/10.1016/j.tree.2016.12.003

Boyer, T.P., Baranova, O.K., Coleman, C., Garcia, H.E., Grodsky, A., Locarnini, R.A., Mishonov, A.V., Paver, C.R., Reagan, J.R., Seidov, D., Smolyar, I.V., Weathers, K., Zweng, M.M. (2018). World Ocean Database 2018. A.V. Mishonov, Technical Ed., NOAA Atlas NESDIS 87, https://www.ncei.noaa.gov/products/world-ocean-database

Brankart, J.-M., Testut, C.-E., Béal, D., Doron, M., Fontana, C., Meinvielle, M., and Brasseur, P. (2012). Towards an improved description of oceanographic uncertainties : effect of local anamorphic transformations on spatial correlations. Ocean Science, 8, 121-142, https://doi.org/10.5194/os-8-121-2012

Brasseur, P., Gruber, N., Barciela, R., Brander, K., Doron, M., El Moussaoui, A., Hobday, M. Huret, A.J., Kremeur, A.-S., Lehodey, P., and others. (2009). Integrating biogeochemistry and ecology into ocean data assimilation systems. Oceanography, 22(3), 206-215, https://doi.org/10.5670/oceanog.2009.80

Bricaud, C., Le Sommer, J., Madec, G., Calone, C., Deshayes, J., Éthé, C., Chanut, J., Lévy, M. (2020). Multigrid algorithm for passive tracer transport in the NEMO ocean circulation model: a case study with the NEMO OGCM (version 3.6). Geoscientific Model Development, 13(11), 5465-5483, https://doi.org/10.5194/gmd-13-5465-2020

Brown, C. D., and Davis, H. T. (2006). Receiver operating characteristics curves and related decision measures: A tutorial. Chemometrics and Intelligent Laboratory Systems, 80(1), 24-38 https://doi.org/10.1016/j.chemolab.2005.05.004

Butenschön, M., Clark, J., Aldridge, J. N., Allen, J. I., Artioli, Y., Blackford, J., Bruggeman, J., Cazenave, P., Ciavatta, S., Kay, S., Lessin, G., van Leeuwen, S., van der Molen, J., de Mora, L., Polimene, L., Sailley, S., Stephens, N., and Torres, R. (2016). ERSEM 15.06: a generic model for marine biogeochemistry and the ecosystem dynamics of the lower trophic levels. Geoscientific Model Development, 9, 1293-1339, https://doi.org/10.5194/gmd-9-1293-2016

Campbell, J. (1995). The lognormal distribution as a model for bio-optical variability in the sea. Journal of Geophysical Research: Oceans, 100, C7, 13237-13254, https://doi.org/10.1029/95JC00458

Capet, A., Meysman, F. J. R., Akoumianaki, I., Soetaert, K., and Grégoire, M. (2016). Integrating sediment biogeochemistry into 3-D oceanic models: A study of benthic-pelagic coupling in the Black Sea. Ocean Modelling, 101, 83-100, https://doi.org/10.1016/j.ocemod.2016.03.006

Carlotti, F., Wolf, K.U. (1998). A Lagrangian ensemble model of Calanus finmarchicus coupled with a 1-D ecosystem model. Fisheries Oceanography, 7,191-204, https://doi.org/10.1046/j.1365-2419.1998.00085.x

Carrassi, A., Bocquet, M., Bertino, L., and Evensen, G. (2018). Data assimilation in the geosciences: An overview of methods, issues, and perspectives. Wiley Interdisciplinary Reviews: Climate Change, e535, https://doi.org/10.1002/wcc.535

Carroll, D., Menemenlis, D., Adkins, J.F., Bowman, K.W., Brix, H., Dutkiewicz, S., Fenty, I., Gierach, M.M., Hill, C., Jahn, O., and Landschützer, P. (2020). The ECCO-Darwin data-assimilative global ocean biogeochemistry model: Estimates of seasonal to multidecadal surface ocean pCO2 and air-sea CO2 flux. Journal of Advances in Modeling Earth Systems, 7(3-4), 191-204, https://doi.org/10.1029/2019MS001888

Chai, F., Johnson, K., Claustre, H., Xing, X., Wang, Y., Boss, E., Riser, S., Fennel, K., Schofield, O., Sutton, A. (2020). Monitoring ocean biogeochemistry with autonomous platforms. Nature Reviews Earth & Environment, 1, 315-326, https://doi.org/10.1038/s43017-020-0053-y

Chambault P., de Thoisy B., Heerah K., Conchon A., Barrioz S., Dos Reis V., Berzins R., Kelle L., Picard B., Roquet F., Le Maho Y., Chevallier D. (2016). The influence of oceanographic features on the foraging behavior of the olive ridley sea turtle Lepidochelys olivacea along the Guiana coast. Progress in Oceanography, 142: 58-71, https://doi.org/10.1016/j.pocean.2016.01.006

Christensen, V., and Walters, C. (2004). Ecopath With Ecosim: Methods, Capabilities and Limitations. Ecological Modelling, 172, 109-139, https://doi.org/10.1016/j.ecolmodel.2003.09.003

Ciavatta, S., Torres, R., Saux-Picart, S. and Allen, J.I. (2011). Can ocean color assimilation improve biogeochemical hindcasts in shelf seas? Journal of Geophysical Research: Oceans, 116(C12), https://doi.org/10.1029/2011JC007219

Ciavatta, S., Torres, R., Martinez-Vicente, V., Smyth, T., Dall’Olmo, G., Polimene, L., and Allen, J. I. (2014). Assimilation of remotely-sensed optical properties to improve marine biogeochemistry modelling. Progress in Oceanography, 127, 74-95, https://doi.org/10.1016/j.pocean.2014.06.002

Ciavatta, S., Kay, S., Saux-Picart, S., Butenschön, M. and Allen, J.I. (2016). Decadal reanalysis of biogeochemical indicators and fluxes in the North West European shelf-sea ecosystem. Journal of Geophysical Research: Oceans, 121(3), 1824-1845, https://doi.org/10.1002/2015JC011496

Ciavatta, S., Brewin, R. J. W., Skákala, J., Polimene, L., de Mora, L., Artioli, Y., and Allen, J. I. (2018). Assimilation of ocean-color plankton functional types to improve marine ecosystem simulations. Journal of Geophysical Research: Oceans, 123, 834-854, https://doi.org/10.1002/2017JC013490

Ciavatta, S., Kay, S., Brewin, R.J., Cox, R., Di Cicco, A., Nencioli, F., Polimene, L., Sammartino, M., Santoleri, R., Skakala, J. and Tsapakis, M. (2019). Ecoregions in the Mediterranean Sea through the reanalysis of phytoplankton functional types and carbon fluxes. Journal of Geophysical Research: Oceans, 124(10), 6737-6759, https://doi.org/10.1029/2019JC015128

Cossarini, G., Mariotti, L., Feudale, L., Mignot, A., Salon, S., Taillandier, V., Teruzzi, A., D'Ortenzio, F. (2019). Towards operational 3D-Var assimilation of chlorophyll Biogeochemical-Argo float data into a biogeochemical model of the Mediterranean Sea. Ocean Modelling, 133, 112-128, https://doi.org/10.1016/j.ocemod.2018.11.005

Daewel, U., and Schrum, C. (2013). Simulating long-term dynamics of the coupled North Sea and Baltic Sea ecosystem with ECOSMO II: Model description and validation. Journal of Marine Systems, 119-120, 30-49, https://doi.org/10.1016/j.jmarsys.2013.03.008

Dall'Olmo, G., Dingle, J., Polimene, L., Brewin, R.J.W., Claustre, H. (2016). Substantial energy input to the mesopelagic ecosystem from the seasonal mixed-layer pump. Nature Geoscience, 9, 820-823, https://doi.org/10.1038/ngeo2818

De Roos, A.M., Schellekens, T., Van Kooten, T., Van De Wolfshaar, K., Claessen, D., Persson, L. (2008).  Simplifying a physiologically structured population model to a stage-structured biomass model. Theoretical Population Biology, 73, 47-62, https://doi.org/10.1016/j.tpb.2007.09.004

DeAngelis, D.L., Gross, L.J. (1992). Individual-based Models and Approaches in Ecology. Chapman & Hall, London.

Diaz, F., D Bǎnaru, P. Verley and Y.-J. Shin (2019). Implementation of an end-to-end model of the Gulf of Lions ecosystem (NW Mediterranean Sea). II. Investigating the effects of high trophic levels on nutrients and plankton dynamics and associated feedbacks. Ecological Modelling, 405, 51-68, https://doi.org/10.1016/j.ecolmodel.2019.05.004

Dickey, T.D. (1991). The emergence of concurrent high-resolution physical and bio-optical measurements in the upper ocean and their applications. Reviews on Geophysics, 29(3), 383- 413, https://doi.org/10.1029/91RG00578

Dornan, T., Fielding, S., Saunders, R.A., Genner, M.J. (2019). Swimbladder morphology masks Southern Ocean mesopelagic fish biomass. Proceedings of the Royal Society B Biological Science, 286(1903), http://dx.doi.org/10.1098/rspb.2019.0353

Doron, M., Brasseur, P., Brankart, J.M. (2011). Stochastic estimation of biogeochemical parameters of a 3d ocean coupled physical-biogeochemical model: twin experiments. Journal of Marine Systems, 87, 194-207, https://doi.org/10.1016/j.jmarsys.2013.02.007

Dragon A-C., Senina, Hintzen N.T., Lehodey P. (2018). Modelling South Pacific Jack Mackerel spatial population dynamics and fisheries. Fisheries Oceanography, 27(2), 97-113, https://doi.org/10.1111/fog.12234

Dueri S., Faugeras B., Maury O. (2012). Modelling the skipjack tuna dynamics in the Indian Ocean with APECOSM-E: Part 1. Model formulation. Ecological Modelling, 245, 41-54, https://doi.org/10.1016/j.ecolmodel.2012.02.008

Dueri, S., Bopp, L., Maury, O. (2014). Projecting the impacts of climate change on skipjack tuna abundance and spatial distribution. Global Change Biology, 20(3): 742-753, https://doi.org/10.1111/gcb.12460

Dupont, F. (2012). Impact of sea-ice biology on overall primary production in a biophysical model of the pan-Arctic Ocean. Journal of Geophysical Research: Oceans,117(C8), https://doi.org/10.1029/2011JC006983

Eilola, K., Meier, H.E.M., Almroth, E. (2009). On the dynamics of oxygen, phosphorus and cyanobacteria in the baltic sea: A model study. Journal of Marine Systems, 75, 163-184, https://doi.org/10.1016/j.jmarsys.2008.08.009

Escobar-Flores, P.C., O’Driscoll, R.L., Montgomery, J.C. (2018). Spatial and temporal distribution patterns of acoustic backscatter in the New Zealand sector of the Southern Ocean. Marine Ecology Progress Series, 592, 19-35, https://doi.org/10.3354/meps12489

Evensen, G. (1994). Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics. Journal of Geophysical Research: Oceans, 99(C5), 10143-10162, https://doi.org/10.1029/94JC00572

Fennel, K., Gehlen, M., Brasseur, P., Brown, C. W., Ciavatta, S., Cossarini, G., Crise, A., Edwards, C. A., Ford, D., Friedrichs, M. A. M., Grégoire, M., Jones, E., Kim, H.-C., Lamouroux, J., Murtugudde, R., Perruche, C. and the GODAE OceanView Marine Ecosystem Analysis and Prediction Task Team (2019). Advancing marine biogeochemical and ecosystem reanalyses and forecasts as tools for monitoring and managing ecosystem health. Frontiers in Marine Science, 6, 89, https://doi.org/10.3389/fmars.2019.00089

Feudale, L., Teruzzi, A., Salon, S., Bolzon, G., Lazzari, P., Coidessa, G., Di Biagio, V., Cossarini, G. (2021). Product Quality Document For the Mediterranean Sea Production Centre, MEDSEA_ANALYSISFORECAST_BGC_006_014, https://doi.org/10.25423/cmcc/medsea_analysisforecast_bgc_006_014_medbfm3

Flynn, K. J., Stoecker, D. K., Mitra, A., Raven, J. A., Glibert, P. M. Hansen, P. J., Granéli, E., and Burkholder, J. M. (2013). Misuse of the phytoplankton-zooplankton dichotomy: the need to assign organisms as mixotrophs within plankton functional types. Journal of Plankton Research, 35, 3-11, https://doi.org/10.1093/plankt/fbs062

Follows, M. J., Dutkiewicz, S., Grant, S., and Chisholm, S. W. (2007). Emergent biogeography of microbial communities in a model ocean. Science, 315(5820), 1843-1846. https://doi.org/10.1126/science.1138544

Fontana, C., Brasseur, P., and Brankart, J. M. (2013). Toward a multivariate reanalysis of the North Atlantic Ocean biogeochemistry during 1998–2006 based on the assimilation of SeaWiFS chlorophyll data. Ocean Science, 9(1), 37-56, https://doi.org/10.5194/os-9-37-2013

Ford, D., and Barciela, R. (2017). Global marine biogeochemical reanalyses assimilating two different sets of merged ocean colour products. Remote Sensing of Environment, 203, 40-54, https://doi.org/10.1016/j.rse.2017.03.040

Ford, D., Key, S., McEwan, R., Totterdell, I. and Gehlen, M. (2018). Marine biogeochemical modelling and data assimilation for operational forecasting, reanalysis, and climate research. New Frontiers in Operational Oceanography, 625-652, https://doi.org/10.17125/gov2018.ch22

Ford, D. (2021). Assimilating synthetic Biogeochemical-Argo and ocean colour observations into a global ocean model to inform observing system design. Biogeosciences, 18(2), 509-534, https://doi.org/10.5194/bg-18-509-2021

Fulton, E.A. (2010). Approaches to end-to-end ecosystem models. Journal of Marine Systems 81, 171- 183, https://doi.org/10.1016/j.jmarsys.2009.12.012

Fulton, E., Smith, A., Johnson, C. (2003). Effect of Complexity of Marine Ecosystem Models. Marine Ecology Progress Series, 253:1-16, doi:10.3354/meps253001

Garcia, H.E., Weathers, K., Paver, C.R., Smolyar, I., Boyer, T.P., Locarnini, R.A., Zweng, M.M, Mishonov, A.V., Baranova, O.K., Seidov, D., and Reagan, J.R. (2018a). World Ocean Atlas 2018, Volume 3: Dissolved Oxygen, Apparent Oxygen Utilization, and Oxygen Saturation. A. Mishonov Technical Ed.; NOAA Atlas NESDIS 83, 38pp.

Garcia, H.E., Weathers, K., Paver, C.R., Smolyar, I., Boyer, T.P., Locarnini, R.A., Zweng, M.M, Mishonov, A.V., Baranova, O.K., Seidov, D., and Reagan, J.R. (2018b). World Ocean Atlas 2018, Volume 4: Dissolved Inorganic Nutrients (phosphate, nitrate and nitrate+nitrite, silicate). A. Mishonov Technical Ed.; NOAA Atlas NESDIS 84, 35pp.

Garnesson, P., Mangin, A., and Bretagnon M. (2021). Quality Information Document, Ocean Colour Production Centre, Satellite Observation Copernicus-GlobColour Products, https://catalogue.marine.copernicus.eu/documents/QUID/CMEMS-OC-QUID-009-030-032-033-037-081-082-083-085-086-098.pdf

Garnier, F., Brankart, J. M., Brasseur, P. and Cosme, E. (2016). Stochastic parameterizations of biogeochemical uncertainties in a 1/4° NEMO/PISCES model for probabilistic comparisons with ocean color data. Journal of Marine Systems, 155, 59-72, https://doi.org/10.1016/j.jmarsys.2015.10.012

Gehlen, M., Barciela, R., Bertino, L., Brasseur, P., Butenschön, M., Chai, F., Crise, A., Drillet, Y., Ford, D., Lavoie, D., Lehodey, P., Perruche, C., Samuelsen, A., and Simon, E. (2015). Building the capacity for forecasting marine biogeochemistry and ecosystems: recent advances and future developments. Journal of Operational Oceanography, 8:sup1, s168-s187, https://doi.org/10.1080/1755876X.2015.1022350

Geider, R.J., MacIntyre, H.L., and Kana, T.M. (1997). 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. Marine Ecology Progress Series, 148, 187-200, https://doi.org/10.3354/meps148187

Germineaud, C., Brankart, J.M., and Brasseur, P. (2019). An ensemble-based probabilistic score approach to compare observation scenarios: an application to biogeochemical-Argo deployments. Journal of Atmospheric and Oceanic Technology, 36(12), 2307-2326, https://doi.org/10.1175/JTECH-D-19-0002.1

Gharamti, M.E., Tjiputra, J., Bethke, I., Samuelsen, A., Skjelvan, I., Bentsen, M., Bertino, L. (2017). Ensemble data assimilation for ocean biogeochemical state and parameter estimation at different sites. Ocean Modelling, 112, 65-89, https://doi.org/10.1016/j.ocemod.2017.02.006

Glibert, P., Mitra, A., Flynn, K., Hansen, P., Jeong, H., and Stoecker, D. (2019). Plants Are Not Animals and Animals Are Not Plants, Right? Wrong! Tiny Creatures in the Ocean Can Be Both at Once! Frontiers for Young Minds, 7:48, https://doi.org/10.3389/frym.2019.00048

Gradinger, R. (2009). Sea-ice algae: Major contributors to primary production and algal biomass in the Chukchi and Beaufort Seas during May/June 2002. Deep Sea Research Part II: Topical Studies in Oceanography, 56, 1201-1212, https://doi.org/10.1016/j.dsr2.2008.10.016

Green, D.B., Bestley, S., Corney, S.P., Trebilco, R., Lehodey, P., and Hindell, M.A. (2021). Modeling Antarctic krill circumpolar spawning habitat quality to identify regions with potential to support high larval production. Geophysical Research Letters, 48, e2020GL091206, https://doi.org/10.1029/2020GL091206

Green, D.B., Bestley, S., Trebilco, R., Corney, S.P., Lehodey, P., McMahon, C.R., Guinet, C., and Hindell, M.A. (2020). Modelled mid-trophic pelagic prey fields improve understanding of marine predator foraging behaviour. Ecography, https://doi.org/10.1111/ecog.04939

Gregg, W.W., and Rousseaux, C.S. (2016). Directional and spectral irradiance in ocean models: effects on simulated global phytoplankton, nutrients, and primary production. Frontiers in Marine Science, 3, 240, http://dx.doi.org/10.1088/1748-9326/ab4667

Gregg, W.W., and Rousseaux, C.S. (2019). Global ocean primary production trends in the modern ocean color satellite record (1998-2015). Environmental Research Letters,14(12),124011,  https://doi:10.1088/1748-9326/ab4667

Grégoire, M., Raick, C., and Soetaert, K. (2008). Numerical modelling of the central black sea ecosystem functioning during the eutrophication phase. Progress in Oceanography, 76, 286 333, https://doi.org/10.1016/j.pocean.2008.01.002

Grégoire, M., Soetaert, K. (2010). Carbon, nitrogen, oxygen and sulfide budgets in the Black Sea: a biogeochemical model of the whole water column coupling the oxic and anoxic parts. Ecological Modelling, 221(19), 2287-2301.

Hazen, E. L., Scales, K.L., Maxwell, S.M., Briscoe, D.K., Welch, H., Bograd, S.J., Bailey, H., Benson, S.R., Eguchi, T., Dewar, H., Kohin, S., Costa, D.P., Crowder, L.B., Lewison R.L. (2018). A dynamic ocean management tool to reduce bycatch and support sustainable fisheries.Science Advances, 4(5), DOI:10.1126/sciadv.aar3001

Hernandez, F., Bertino, L., Brassington, G., Chassignet, E., Cummings, J., Davidson, F., Drevillon, M., Garric, G., Kamachi, M., Lellouche, J.M., Mahdon, R., Martin, M.J,, Ratsimandresy, A., and Regnier, C. (2009). Validation and Inter-comparison studies within GODAE. Oceanography, 22(3): 128-143, https://doi.org/10.5670/oceanog.2009.71

Hernandez, F., Smith, G., Baetens, K., Cossarini, G., Garcia-Hermosa, I., Drevillon, M., von Schuckman, K. (2018). Measuring performances, skill and accuracy in operational oceanography: New challenges and approaches. In: "New Frontiers in Operational Oceanography", E. Chassignet, A. Pascual, J. Tintoré, and J. Verron, Eds., GODAE OceanView, 759-796, DOI:10.17125/gov2018.ch29

Hernandez, O., Lehodey, P., Senina, I., Echevin, V., Ayon, P., Bertrand, A., Gaspar, P. (2014). Understanding mechanisms that control fish spawning and larval recruitment: Parameter optimization of an Eulerian model (SEAPODYM-SP) with Peruvian anchovy and sardine eggs and larvae data. Progress in Oceanography, 123, 105-122, http://dx.doi.org/10.1016/j.pocean.2014.03.001

Hipsey, M.R., Gal, G., Arhonditsis, G.B., Carey, C.C., Elliott, J.A., Frassl, M. A., ... & Robson, B.J. (2020). A system of metrics for the assessment and improvement of aquatic ecosystem models. Environmental Modelling & Software, 128, 104697, https://doi.org/10.1016/j.envsoft.2020.104697

Hobday, A.J., Hartog, J.R., Timmiss, T., Fielding, J. (2010). Dynamic spatial zoning to manage southern bluefin tuna (Thunnus maccoyii) capture in a multi-species longline fishery. Fisheries Oceanography, 19(3), 243-253, https://doi.org/10.1111/j.1365-2419.2010.00540.x

Houtekamer, P. L., and Mitchell, H. L. (1998). Data Assimilation Using an Ensemble Kalman Filter Technique. Monthly Weather Review,126(3), 796-811, https://doi.org/10.1175/1520-0493(1998)126<0796:DAUAEK>2.0.CO;2

Howell, E.A., Kobayashi, D.R., Parker, D.M., Balazs, G.H., Polovina, J.J. (2008). TurtleWatch: a tool to aid in the bycatch reduction of loggerhead turtles Caretta caretta in the Hawaii-based pelagic longline fishery. Endangered Species Research, 5:267-278, https://doi.org/10.3354/esr00096

Hu, J., Fennel, K., Mattern, J.P. and Wilkin, J. (2012). Data assimilation with a local Ensemble Kalman Filter applied to a three-dimensional biological model of the Middle Atlantic Bight. Journal of Marine Systems, 94, 145-156, https://doi.org/10.1016/j.jmarsys.2011.11.016

Huse, G., Melle, W., Skogen, M.D., Hjøllo, S.S., Svendsen, E., and Budgell, W.P. (2018). Modeling Emergent Life Histories of Copepods. Frontiers in Ecology and Evolution, 6(23), https://doi.org/10.3389/fevo.2018.00023

IOCCG. (2020). Synergy between Ocean Colour and Biogeochemical/Ecosystem Models. Dutkiewicz, S. (ed.), IOCCG Report Series, No. 19, International Ocean Colour Coordinating Group, Dartmouth, Canada, http://dx.doi.org/10.25607/OBP-711

Ishizaka, J. (1990). Coupling of coastal zone color scanner data to a physical-biological model of the southeastern US continental shelf ecosystem: 2. An Eulerian model. Journal of Geophysical Research: Oceans, 95(C11), 20183-20199, https://doi.org/10.1029/JC095iC11p20183

Jaccard,P.,Hjermann,D.Ø.,Ruohola,J.,Marty, S.,Kristiansen, T., Sørensen,K.,Kaitala, S.,Mangin,A.,Pouliquen, S., et al. (2021). Quality Information Document For Global Ocean Reprocessed in-situ Observations of Biogeochemical Products INSITU_GLO_BGC_REP_OBSERVATIONS_013_046, http://doi.org/10.13155/54846, https://catalogue.marine.copernicus.eu/documents/QUID/CMEMS-INS-QUID-013-046.pdf

Jech, J.M., Horne, J.K., Chu, D., Demer, D.A., Francis, D.T.I., Gorska, N., Jones, B., Lavery, A.C., Stanton, T.K., Macaulay, G.J., Reeder, D.B., Sawada, K. (2015). Comparisons among ten models of acoustic backscattering used in aquatic ecosystems. The Journal of the Acoustical Society of America, 138: 3742, https://doi.org/10.1121/1.4937607

Jennings, S., and Collingridge, K. (2015). Predicting Consumer Biomass, Size-Structure, Production, Catch Potential, Responses to Fishing and Associated Uncertainties in the World’s Marine Ecosystems. PLoS ONE, 10(7), e0133794. https://doi.org/10.1371/journal.pone.0133794

Jolliff, J.K., Kindle, J.C., Shulman, I., Penta, B., Friedrichs, M.A.M., Helber, R.W., and Arnone, R. (2009). Summary diagrams for coupled hydrodynamic-ecosystem model skill assessment. Journal of Marine Systems, 76, 64-82, https://doi.org/10.1016/j.jmarsys.2008.05.014

Jones, E.M., Baird, M.E., Mongin, M., Parslow, J., Skerratt, J., Lovell, J., Margvelashvili, N., Matear, R.J., Wild-Allen, K., Robson, B. and Rizwi, F. (2016). Use of remote-sensing reflectance to constrain a data assimilating marine biogeochemical model of the Great Barrier Reef. Biogeosciences, 13(23), 6441, https://doi.org/10.5194/bg-13-6441-2016

Kaufman, D.E. (2017). Using High-Resolution Glider Data and Biogeochemical Modeling to Investigate Phytoplankton Variability in the Ross Sea. Dissertations, Theses, and Masters Projects, William & Mary, Paper 1499449869, http://dx.doi.org/10.21220/M2BK8V

Kriest, I., Kähler, P., Koeve, W., Kvale, K., Sauerland, V., and Oschlies, A. (2020). One size fits all? Calibrating an ocean biogeochemistry model for different circulations. Biogeosciences, 17, 3057-3082, https://doi.org/10.5194/bg-17-3057-2020

Krishnamurthy, A., Moore, J. K., Mahowald, N., Luo, C., and Zender, C.S. (2010). Impacts of atmospheric nutrient inputs on marine biogeochemistry. Journal of Geophysical Research: Biogeosciences, 115(G1), https://doi.org/10.1029/2009JG001115

Lambert, C., Mannocci, L., Lehodey, P., Ridoux, V. (2014). Predicting Cetacean Habitats from Their Energetic Needs and the distribution of Their Prey in Two Contrasted Tropical Regions. PLoS ONE, 9(8), e105958, https://doi.org/10.1371/journal.pone.0105958

Lamouroux, J., Perruche, C., Mignot, A., Paul, J., Szczypta, C. (2019). Quality Information Document For Global Biogeochemical Analysis and Forecast Product, https://catalogue.marine.copernicus.eu/documents/QUID/CMEMS-GLO-QUID-001-028.pdf

Lamouroux, J., Perruche, C., Mignot, A., Gutknecht, E., Ruggiero, G., Evaluation of the CMEMS global biogeochemical simulation, with assimilation of satellite Chla concentrations, in prep.

Laws, E.A. (2013). Evaluation of In Situ Phytoplankton Growth Rates: A Synthesis of Data from Varied Approaches. Annual Review of Marine Science, 5(1), 247-268, https://doi.org/10.1146/annurev-marine-121211-172258

Le Quéré, C., Harrison, S. P., Prentice, I. C., Buitenhuis, E. T., Aumont, O., Bopp, L., Claustre, H., Cotrim Da Cunha, L., Geider, R., Giraud, X., Klaas, C., Kohfeld, K. E., Legendre, L., Manizza, M., Platt, T., Rivkin, R. B., Sathyendranath, S., Uitz, J., Watson, A. J., and Wolf‐Gladrow, D. (2005). Ecosystem dynamics based on plankton functional types for global ocean biogeochemistry models. Global Change Biology, 11, 2016-2040, https://doi.org/10.1111/j.1365-2486.2005.1004.x

Le Traon et al. (2017). The Copernicus marine environmental monitoring service: main scientific achievements and future prospects. Special Issue Mercator Océan International #56. Available at: https://marine.copernicus.eu/news/mercator-ocean-journal-56-cmems-special-issue

Le Traon, P. Y., Reppucci, A., Fanjul, E. A., Aouf, L., Behrens, A., Belmonte, M., Bentamy, A., Bertino, L., Brando, V. E., Kreiner, M. B., Benkiran, M., Carval, T., Ciliberti, S. A., Claustre, H., Clementi, E., Coppini, G., Cossarini, G., De Alfonso Alonso- Muñoyerro, M., Delamarche, A., Dibarboure, G., Dinessen, F., Drevillon, M., Drillet, Y., Faugere, Y., Fernández, V., Fleming, A., Garcia-Hermosa, M. I., Sotillo, M. G., Garric, G., Gasparin, F., Giordan, C., Gehlen, M., Grégoire, M., Guinehut, S., Hamon, M., Harris, C., Hernandez, F., Hinkler, J. B., Hoyer, J., Karvonen, J., Kay, S., King, R., Lavergne, T., Lemieux-Dudon, B., Lima, L., Mao, C., Martin, M. J., Masina, S., Melet, A., Nardelli, B. B., Nolan, G., Pascual, A., Pistoia, J., Palazov, A., Piolle, J. F., Pujol, M. I., Pequignet, A. C., Peneva, E., Gómez, B. P., de la Villeon, L. P., Pinardi, N., Pisano, A., Pouliquen, S., Reid, R., Remy, E., Santoleri, R., Siddorn, J., She, J., Staneva, J., Stoffelen, A., Tonani, M., Vandenbulcke, L., von Schuckmann, K., Volpe, G., Wettre, C., and Zacharioudaki, A. (2019). From observation to information and users: The Copernicus Marine Service Perspective. Frontiers in Marine Science, https://doi.org/10.3389/fmars.2019.00234

Lehodey, P., Senina, I., Murtugudde, R. (2008). A Spatial Ecosystem And Populations Dynamics Model (SEAPODYM) - Modelling of tuna and tuna-like populations. Progress in Oceanography, 78, 304-318, https://doi.org/10.1016/j.pocean.2008.06.004

Lehodey, P., Senina, I., Wibawa, T.A., Titaud, O., Calmettes, B., Tranchant, B., and Gaspar, P. (2017). Operational modelling of bigeye tuna (Thunnus obesus) spatial dynamics in the Indonesian region. Marine Pollution Bulletin, 131, 19-32, https://doi.org/10.1016/j.marpolbul.2017.08.020

Lehodey, P., Conchon, A., Senina, I., Domokos, R., Calmettes, B., Jouanno, J., Hernandez, O., and Kloser, R. (2015). Optimization of a micronekton model with acoustic data. ICES Journal of Marine Science, 72(5), 1399-1412, https://doi.org/10.1093/icesjms/fsu233

Lehodey, P., Murtugudde, R., and Senina, I. (2010). Bridging the gap from ocean models to population dynamics of large marine predators: A model of mid-trophic functional groups. Progress in Oceanography, 84, 69-84.

Lehodey, P., Senina, I., Calmettes, B, Hampton, J, Nicol S. (2013). Modelling the impact of climate change on Pacific skipjack tuna population and fisheries. Climatic Change, 119 (1): 95-109, DOI 10.1007/s10584-012-0595-1

Lengaigne, M., Menkes, C., Aumont, O., Gorgues, T., Bopp, L., André, J. M., and Madec, G. (2007). Influence of the oceanic biology on the tropical Pacific climate in a coupled general circulation model. Climate Dynamics, 28(5), 503-516. https://doi.org/10.1007/s00382-006-0200-2

Libralato, S., and Solidoro, C. (2009). Bridging biogeochemical and food web models for an Endto-End representation of marine ecosystem dynamics: The Venice lagoon case study. Ecological Modelling, 220(21), 2960-2971, https://doi.org/10.1016/j.ecolmodel.2009.08.017

Longhurst, A. (1998). Ecological geography in the sea. Academic Press.

Mattern, J. P., Fennel, K., and Dowd, M. (2012). Estimating time-dependent parameters for a biological ocean model using an emulator approach. Journal of Marine Systems, 96, 32-47, https://doi.org/10.1016/j.jmarsys.2012.01.015

Mattern, J.P., Dowd, M. and Fennel, K. (2013). Particle filter-based data assimilation for a three-dimensional biological ocean model and satellite observations. Journal of Geophysical Research: Oceans, 118(5), 2746-2760, https://doi.org/10.1002/jgrc.20213

Maunder, M.N., Punt, A.E. (2013). A review of integrated analysis in fisheries stock assessment. Fisheries Research, 142, 61-74, https://doi.org/10.1016/j.fishres.2012.07.025

Maury, O., Faugeras, B., Shin, Y.-J., Poggiale, J.-C., Ben Ari, T., and Marsac, F. (2007). Modeling environmental effects on the size-structured energy flow through marine ecosystems. Part 1: The model. Progress in Oceanography, 74(4), 479-499, https://doi.org/10.1016/j.pocean.2007.05.002

McEwan, R., S. Kay, D. Ford (2021). Quality Information Document For the Atlantic - European North West Shelf Production Centre, https://catalogue.marine.copernicus.eu/documents/QUID/CMEMS-NWS-QUID-004-002.pdf

Melsom, A. and Ç. Yumruktepe (2021), Quality Information Document For the Arctic Production Centre, https://catalogue.marine.copernicus.eu/documents/QUID/CMEMS-ARC-QUID-002-004.pdf, https://doi.org/10.48670/moi-00003

Mignot, A., Claustre, H., Cossarini, G., D’Ortenzio, F., Gutknecht, E., Lamouroux, J., Lazzari, P., Perruche, C., Salon, S., Sauzède, R., Taillandier, V., Teruzzi, A. (2021). Defining BGC-Argo-based metrics of ocean health and biogeochemical functioning for the evaluation of global ocean models. Biogeosciences, https://doi.org/10.5194/bg-2021-2

Miller, C.B., Lynch, D.R., Carlotti, F., Gentleman, W., Lewis, C.V.W. (1998). Coupling of an individual-based population dynamic model of Calanus finmarchicus to a circulation model for theGeorges Bank region. Fisheries Oceanography, 7(3-4), 219-234, https://doi.org/10.1046/j.1365-2419.1998.00072.x

Mogensen, K.S., Balmaseda, M.A., Weaver, A., Martin, M.J., Vidard, A. (2009). NEMOVAR: A variational data assimilation system for the NEMO ocean model. ECMWF Newsl., 120, 17-21.

Moore, A. M., Martin, M. J., Akella, S., Arango, H. G., Balmaseda, M., Bertino, L., ... and Weaver, A. T. (2019). Synthesis of ocean observations using data assimilation for operational, real time and reanalysis systems: A more complete picture of the state of the ocean. Frontiers in Marine Science, 6:90, https://doi.org/10.3389/fmars.2019.00090

Moriarty, R., and O’Brien, T. (2013). Distribution of mesozooplankton biomass in the global ocean. Earth System Science Data, 5(1), 45-55, https://doi.org/10.5194/essd-5-45-2013

Natvik, L., and Evensen, G. (2003). Assimilation of ocean colour data into a biochemical model of the North Atlantic. Part 1: Data assimilation experiments. Journal of Marine Systems, 41, 127-153, https://doi.org/10.1016/S0924-7963(03)00016-2

Nerger, L. andGregg,W.W.(2007).AssimilationofSeaWiFSdata intoa globalocean-biogeochemicalmodelusing a local SEIK filter. Journal of Marine Systems, 68, 237-254, https://doi.org/10.1016/j.jmarsys.2006.11.009

Nerger, L., and Gregg, W.W. (2008). Improving assimilation of SeaWiFS data by the application of bias correction with a local SEIK filter. Journal of Marine Systems, 73, 87-102, https://doi.org/10.1016/j.jmarsys.2007.09.007

Nerger, L., Janjić, T., Schröter, J., and Hiller, W. (2012). A Unification of Ensemble Square Root Kalman Filters. Monthly Weather Review, 140(7), 2335-2345, https://doi.org/10.1175/MWR-D-11-00102.1

Neumann, T. (2000). Towards a 3D-ecosystem model of the Baltic Sea. Journal of Marine Systems, 25(3), 405-419, https://doi.org/10.1016/S0924-7963(00)00030-0

Olsen, A., Lange, N., Key, R. M., Tanhua, T., Bittig, H. C., Kozyr, A., Àlvarez, M., Azetsu-Scott, K., Becker, S., Brown, P. J., Carter, B. R., Cotrim da Cunha, L., Feely, R. A., van Heuven, S., Hoppema, M., Ishii, M., Jeansson, E., Jutterström, S., Landa, C. S., Lauvset, S. K., Michaelis, P., Murata, A., Pérez, F. F., Pfeil, B., Schirnick, C., Steinfeldt, R., Suzuki, T., Tilbrook, B., Velo, A., Wanninkhof, R. and Woosley, R. J. (2020). GLODAPv2.2020 - the second update of GLODAPv2. Earth System Science Data, https://doi.org/10.5194/essd-12-3653-2020

Ourmières, Y., Brasseur, P., Lévy, M., Brankart, J.M. and Verron, J. (2009). On the key role of nutrient data to constrain a coupled physical-biogeochemical assimilative model of the North Atlantic Ocean. Journal of Marine Systems, 75(1-2), 100-115, https://doi.org/10.1016/j.jmarsys.2008.08.003

Palmer, J. R., and Totterdell, I. J. (2001). Production and export in a global ocean ecosystem model. Deep Sea Research Part I: Oceanographic Research Papers, 48(5), 1169-1198, https://doi.org/10.1016/S0967-0637(00)00080-7

Park, J.-Y., Stock, C. A., Yang, X., Dunne, J. P., Rosati, A., John, J., et al. (2018). Modeling global ocean biogeochemistry with physical data assimilation: A pragmatic solution to the equatorial instability. Journal of Advances in Modeling Earth Systems, 10, 891- 906, https://doi.org/10.1002/2017MS001223

Pérez-Jorge, S., Tobeña, M., Prieto, R., Vandeperre, F., Calmettes, B., Lehodey, P., Silva, M.A. (2020). Environmental drivers of large-scale movements of baleen whales in the mid-North Atlantic Ocean. Diversity and Distributions, 26(6), 683-698, https://doi.org/10.1111/ddi.13038

Petrik, C.M., Stock, C.A., Andersen, K.H., van Denderen, P.D., and Watson, J.R. (2019). Bottom-up drivers of global patterns of demersal, forage, and pelagic fishes. Progress in Oceanography, 176:102124, https://doi.org/10.1016/j.pocean.2019.102124

Petrik, C.M., Stock, C.A., Andersen, K.H., van Denderen, P.D., and Watson, J.R. (2020). Large pelagic fish are most sensitive to climate change despite pelagification of ocean foodwebs. Frontiers in Marine Science, 7, 588482, https://doi.org/10.3389/fmars.2020.588482

Pham, D.T., Verron, J., Roubaud, M.C. (1998). A singular evolutive extended Kalman filter for data assimilation in oceanography. Journal of Marine Systems, 16(3-4): 323-340, https://doi.org/10.1016/S0924-7963(97)00109-7

Pradhan, H.K., Völker, C., Losa, S.N., Bracher, A., and Nerger, L. (2020). Global assimilation of oceancolor data of phytoplankton functional types: Impact of different data sets. Journal of Geophysical Research: Oceans, 125, e2019JC015586, https://doi.org/10.1029/2019JC015586

Proud, R., Handegard, N. O., Kloser, R. J., Cox, M. J., and Brierley, A. S. (2018). From siphonophores to deep scattering layers:Uncertainty ranges for the estimation of global mesopelagic fish biomass. ICES Journal of Marine Science, 76, 718-733, https://doi.org/10.1093/icesjms/fsy037

Redfield, A.C. (1934). On the proportions of organic derivatives in sea water and their relation to the composition of plankton. James Johnson Memorial Volume, R. J. Daniel, Ed., University Press of Liverpool, 177-192.

Roberts, J. J., Best, B. D., Mannocci, L., Fujioka, E., Halpin, P. N., Palka, D. L., … Lockhart, G. G. (2016). Habitat-based cetacean density models for the U.S. Atlantic and Gulf of Mexico. Nature Scientific Reports, 6, 22615, https://doi.org/10.1038/srep22615

Romagosa M., Lucas C., Pérez-Jorge S., Tobeña M., Lehodey P., Reis J., Cascão I., Lammers M. O., Caldeira R. M. A., Silva M. A. (2020). Differences in regional oceanography and prey biomass influence  the presence of foraging odontocetes at two Atlantic seamounts. Marine Mammal Science, 36: 158-179, https://doi.org/10.1111/mms.12626

Romagosa, M., Pérez-Jorge, S., Cascão, I., Mouriño, H., Lehodey, P., Marques, T.A, Silva, M.A. (2021). Food talk: 40-Hz fin whale calls are associated with prey availability. Proceedings of the Royal Society B Biological Sciences, 288(1954), https://doi.org/10.1098/rspb.2021.1156

Rose, K.A., Fiechter, J., Curchitser, E.N., Hedstrom, K., Bernal, M., Creekmore, S., Haynie, A.C., Ito, S., Lluch-Cota, S.E., Megrey, B.A., Edwards, C.A., Checkley, D.M., Koslow, T., Mcclatchie, S., Werner, F.E., Maccall, A.D., and Agostini, V.N. (2015). Demonstration of a fully-coupled end-to-end model for small pelagic fish using sardine and anchovy in the California Current. Progress in Oceanography, 138, 348-380, https://doi.org/10.1016/j.pocean.2015.01.012

Russell, J., Sarmiento, J., Cullen, H., Hotinski, R., Johnson, K.S., Riser, S.C., et al. (2014). The Southern Ocean Carbon and Climate Observations and Modeling Program (SOCCOM). Ocean Carbon Biogeochem. News. Available at https://web.whoi.edu/ocb/wp-content/uploads/sites/43/2016/12/OCB_NEWS_FALL14.pdf

Ryabinin, V., Barbière, J., Haugan, P., Kullenberg, G., Smith, N., McLean, C., Troisi, A., Fischer, A., Aricò, S., Aarup, T., Pissierssens, P., Visbeck, M., Enevoldsen, H.O., Rigaud, J. (2019). The UN Decade of Ocean Science for Sustainable Development, Frontiers in Marine Science, 6,470, https://doi.org/10.3389/fmars.2019.00470

Sakov, P., and Oke, P.R. (2008). A deterministic formulation of the ensemble Kalman filter: an alternative to ensemble square root filters. Tellus A, 60, 361-371, https://doi.org/10.1111/j.1600-0870.2007.00299.x

Salon, S., Cossarini, G., Bolzon, G., Feudale, L., Lazzari, P., Teruzzi, A., Solidoro, C., Crise, A. (2019). Novel metrics based on Biogeochemical Argo data to improve the model uncertainty evaluation of the CMEMS Mediterranean marine ecosystem forecasts. Ocean Science, 15, 997-1022, https://doi.org/10.5194/os-15-997-2019

Santana-Falcon, Y., Brasseur, P., Brankart, J.M., and Garnier, F. (2020). Assimilation of chlorophyll data into a stochastic ensemble simulation for the North Atlantic ocean, Ocean Science, 16, 1297-1315, https://doi.org/10.5194/os-16-1297-2020

Schartau, M., Wallhead, P., Hemmings, J., Löptien, U., Kriest, I., Krishna, S., Ward, B. A., Slawig, T., and Oschlies, A. (2017). Reviews and syntheses: parameter identification in marine planktonic ecosystem modelling. Biogeosciences, 14, 1647-1701, https://doi.org/10.5194/bg-14-1647-2017

Scutt Phillips, J., Sen Gupta, A., Senina, I., van Sebille, E., Lange, M., Lehodey, P., Hampton, J., Nicol, S. (2018). An individual-based model of skipjack tuna (Katsuwonus pelamis) movement in the tropical Pacific Ocean. Progress in Oceanography, 164, 63-74, https://doi.org/10.1016/j.pocean.2018.04.007

Senina, I., Lehodey, P., Hampton, J., Sibert, J. (2019). Quantitative modelling of the spatial dynamics of South Pacific and Atlantic albacore tuna populations. Deep Sea Research Part II: Topical Studies in Oceanography, 175, 104667, https://doi.org/10.1016/j.dsr2.2019.104667

Senina, I., Lehodey, P., Sibert, J., Hampton, J. (2020). Improving predictions of a spatially explicit fish population dynamics model using tagging data. Canadian Journal of Aquatic and Fisheries Sciences, 77(3), 576-593, https://doi.org/10.1139/cjfas-2018-0470

Senina, I., Sibert, J., and Lehodey, P. (2008). Parameter estimation for basin-scale ecosystem-linked population models of large pelagic predators: Application to skipjack tuna. Progress in Oceanography, 78, 319-335, https://doi.org/10.1016/j.pocean.2008.06.003

Sibert, J., Senina, I., Lehodey, P., Hampton, J. (2012). Shifting from marine reserves to maritime zoning for conservation of Pacific bigeye tuna (Thunnus obesus). Proceedings of the National Academy of Sciences, 109(44): 18221-18225.

Simon, E. and Bertino, L. (2009). Application of the Gaussian anamorphosis to assimilation in a 3-D coupled physical-ecosystem model of the North Atlantic with the EnKF: a twin experiment. Ocean Science, 5(4), 495-510, https://doi.org/10.5194/os-5-495-2009

Simon, E. Bertino, L. (2012). Gaussian anamorphosis extension of the DEnKF for combined state parameter estimation: Application to a 1D ocean ecosystem model. Journal of Marine Systems, 89(1), 1-18, https://doi.org/10.1016/j.jmarsys.2011.07.007

Simon, E., Samuelsen, A., Bertino, L. and Mouysset, S. (2015). Experiences in multiyear combined state-parameter estimation with an ecosystem model of the North Atlantic and Arctic Oceans using the Ensemble Kalman Filter. Journal of Marine Systems, 152, 1–17, https://doi.org/10.1016/j.jmarsys.2015.07.004

Skákala, J., Ford, D., Brewin, R. J. W., McEwan, R., Kay, S., Taylor, B., Mora, L., and Ciavatta, S. (2018). The assimilation of phytoplankton functional types for operational forecasting in the northwest European shelf. Journal of Geophysical Research: Oceans, 123, 5230-5247, https://doi.org/10.1029/2018JC014153

Skakala, J., Bruggeman, J., Brewin, R.J., Ford, D.A. and Ciavatta, S. (2020). Improved representation of underwater light field and its impact on ecosystem dynamics: A study in the North Sea. Journal of Geophysical Research: Oceans, 125(7), e2020JC016122, https://doi.org/10.1029/2020JC016122

Skákala, J., Ford, D., Bruggeman, J., Hull, T., Kaiser, J., King, R.R., Loveday, B., Palmer, M.R., Smyth, T., Williams, C.A. and Ciavatta, S. (2021a). Towards a multi-platform assimilative system for North Sea biogeochemistry. Journal of Geophysical Research: Oceans, 126(4), e2020JC016649, https://doi.org/10.1029/2020JC016649

Skakala, J., Bruggeman, J., Ford, D.A., Wakelin, S.L., Akpınar, A., Hull, T., Kaiser, J., Loveday, B.R., Williams, C.A.J. and Ciavatta, S. (2021b). Improved consistency between the modelling of ocean optics, biogeochemistry and physics, and its impact on the North-West European Shelf seas. Earth andSpace Science Open Archive ESSOAr, https://doi.org/10.1002/essoar.10506737.2

Skogen, M.D. (1993). A User's guide to NORWECOM (the NORWegian ECOlogical Model system). Technical report 6, Inst.of Marine Research, Division of Marine Env., Pb 1870, N-5024 Bergen, Norway.

Skogen, M.D. and Søiland, H. (1998). A User's guide to NORWECOM v2.0 (the NORWegian ECOlogical Model system). Tech.rep. Fisken og Havet 18, Inst. of Marine Research, Pb.1870, N-5024 Bergen. Norway.

Song, H., Edwards, C.A., Moore, A.M., and Fiechter, J. (2016). Data assimilation in a coupled physical-biogeochemical model of the California current system using an incremental lognormal 4-dimensional variational approach: part 3 - Assimilation in a realistic context using satellite and in situ observations. Ocean Modelling, 106, 1531-145, https://doi.org/10.1016/j.ocemod.2016.04.001

Spindler, M. (1994). Notes on the biology of sea ice in the Arctic and Antarctic. Polar Biology, 14, 319-324, https://doi.org/10.1007/BF00238447

Storto, A., Oddo, P., Cipollone, A., Mirouze, I. and Lemieux-Dudon, B. (2018). Extending an oceanographic variational scheme to allow for affordable hybrid and four-dimensional data assimilation. Ocean Modelling, 128,.67-86, https://doi.org/10.1016/j.ocemod.2018.06.005

Stow, C. A., Jolliff, J., McGillicuddy Jr, D. J., Doney, S. C., Allen, J. I., Friedrichs, M. A., Wallhead, P., 2009. Skill assessment for coupled biological/physical models of marine systems. Journal of Marine Systems, 76(1-2), 4-15, https://doi.org/10.1016/j.jmarsys.2008.03.011

Taylor, K.E. (2001). Summarizing multiple aspects of model performance in a single diagram. Journal of Geophysical Research: Atmospheres, 106(D7), 7183-7192, https://doi.org/10.1029/2000JD900719

Teruzzi, A, Di Cerbo, P, Cossarini, G, Pascolo, E, Salon, S. (2019). Parallel implementation of a data assimilation scheme for operational oceanography: The case of the MedBFM model system. Computers & Geosciences, 124, 103-114, https://doi.org/10.1016/j.cageo.2019.01.003

Teruzzi, A., Dobricic, S., Solidoro, C., and Cossarini, G. (2014). A 3-D variational assimilation scheme in coupled transport-biogeochemical models: Forecast of Mediterranean biogeochemical properties. Journal of Geophysical Research: Oceans, 119, 200-217, https://doi.org/10.1002/2013JC009277

Tissier, A.-S., Brankart, J.M., Testut, C.E., Ruggiero, G., Cosme, E., and Brasseur, P. (2019). A multiscale ocean data assimilation approach combining spatial and spectral localization. Ocean Science, 15, 443-457, https://doi.org/10.5194/os-15-443-2019

Torres, R., Allen, J.I. and Figueiras, F.G. (2006). Sequential data assimilation in an upwelling influenced estuary. Journal of Marine Systems, 60(3-4), 317-329, https://doi.org/10.1016/j.jmarsys.2006.02.001

Uitz, J., Claustre, H., Morel, A., and Hooker, S. B. (2006). Vertical distribution of phytoplankton communities in open ocean: An assessment based on surface chlorophyll. Journal of Geophysical Research: Oceans, 111(C8), https://doi.org/10.1029/2005JC003207

van Leeuwen, P.J. (2010). Nonlinear data assimilation in geosciences: an extremely efficient particle filter. Quarterly Journal of the Royal Meteorological Society, 136(653), 1991-1999, https://doi.org/10.1002/qj.699

Vandenbulcke, L., and Barth, A. (2015). A stochastic operational system of the Black Sea: Tech. and validation. Ocean Modelling, 93, 7-21.

Verdy, A., and Mazloff, M.R. (2017). A data assimilating model for estimating Southern Ocean biogeochemistry. Journal of Geophysical Research: Oceans, 122(9), 6968-6988, https://doi.org/10.1016/j.ocemod.2015.07.010

Vetra-Carvalho, S., Van Leeuwen, P.J., Nerger, L., Barth, A., Altaf, M.U., Brasseur, P., Kirchgessner, P. and Beckers, J.M. (2018). State-of-the-art stochastic data assimilation methods for high-dimensional non-Gaussian problems. Tellus A: Dynamic Meteorology and Oceanography, 70(1), 1-43, https://doi.org/10.1080/16000870.2018.1445364

Vichi, M., Cossarini, G., Gutierrez Mlot, E., Lazzari, P., Lovato, T., Mattia, G., Masina, S., McKiver, W., Pinardi, N., Solidoro, C., Zavatarelli, M. (2015). The Biogeochemical Flux Model (BFM): Equation Description and User Manual. BFM version 5.1. BFM Report series N.1, March 2015, Bologna, Italy, pp. 89.

Wang, B., Fennel, K., Yu, L. and Gordon, C. (2020). Assessing the value of biogeochemical Argo profiles versus ocean color observations for biogeochemical model optimization in the Gulf of Mexico. Biogeosciences, 17(15), 4059-4074, https://doi.org/10.5194/bg-17-4059-2020

Warner, J.C., Armstrong, B., He, R., Zambon, J.B. (2010). Development of a Coupled Ocean–Atmosphere–Wave–Sediment Transport (COAWST) Modeling System. Ocean Modelling, 35(3), 230-244, https://doi.org/10.1016/j.ocemod.2010.07.010

Waters, J., Lea, D.J., Martin, M.J., Mirouze, I., Weaver, A. and While, J. (2015). Implementing a variational data assimilation system in an operational 1/4 degree global ocean model. Quarterly Journal of the Royal Meteorological Society, 141(687), 333-349, https://doi.org/10.1002/qj.2388

Waters, J., Bell, M. J., Martin, M. J., and Lea, D. J. (2017). Reducing ocean model imbalances in the equatorial region caused by data assimilation. Quarterly Journal of the Royal Meteorological Society, 143(702), 195-208.

While, J., Totterdell, I. and Martin, M. (2012). Assimilation of pCO2 data into a global coupled physical-biogeochemical ocean model. Journal of Geophysical Research: Oceans, 117(C3), https://doi.org/10.1029/2010JC006815

Wright, R.M., Le Quéré, C., Buitenhui, E., Pitois, S., Gibbons, M. (2021). Role of jellyfish in the plankton ecosystem revealed using a global ocean biogeochemical model. Biogeosciences, 18, 1291-1320, https://doi.org/10.5194/bg-18-1291-2021

Xiao, Y., and M.A.M. Friedrichs (2014). Using biogeochemical data assimilation to assess the relative skill of multiple ecosystem models: effects of increasing the complexity of the planktonic food web. Biogeosciences, 11(11), 3015-3030, https://doi.org/10.5194/bg-11-3015-2014

Yool, A., Popova, E. E., and Anderson, T. R. (2013). MEDUSA-2.0: an intermediate complexity biogeochemical model of the marine carbon cycle for climate change and ocean acidification studies. Geoscientific Model Development, 6, 1767-1811, https://doi.org/10.5194/gmd-6-1767-2013

Yu, L., Fennel, K., Bertino, L., El Gharamti, M., and Thompson, K. R. (2018). Insights on multivariate updates of physical and biogeochemical ocean variables using an Ensemble Kalman Filter and an idealized model of upwelling. Ocean Modelling, 126, 13-28, doi:10.1016/j.ocemod.2018.04.005

Zeebe, R., and Wolf-Gladrow, D. (2001). CO2 in Seawater: Equilibrium, Kinetics, Isotopes. Elsevier Oceanography Book Series, 65, 346 pp, Amsterdam.

Zhou M., Carlotti F., Zhu Y. (2010). A size-spectrum zooplankton closure model for ecosystem modelling. Journal of Plankton Research, 32(8), 1147-1165, https://doi.org/10.1093/plankt/fbq054

Chapter 9

Biogeochemical modelling

To start contributing, sharing knowledge and editing the WIKI, please login