Ocean dinitrogen fixation and its potential effects on ocean primary ...

1 downloads 0 Views 6MB Size Report
Feb 15, 2018 - L. K. K. −. −. = = +. +. (3d). Kx, where x is either ammonium, nitrate, phosphate or iron, is also a ..... PP rates are available over the period 1989–2014 (Bates et al., 2001; Lomas ...... 855–872. John Wiley & Sons, Inc. DOI: https://.
Riche, OGJ and Christian, JR 2018 Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming. Elem Sci Anth, 6: 16. DOI: https://doi.org/10.1525/elementa.277

RESEARCH ARTICLE

Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming O. G. J. Riche*,‡ and J. R. Christian†,‡ Dinitrogen fixation (DNF) provides a large fraction of the ‘new’ nitrogen supporting upper ocean productivity, and is associated with environmental conditions likely to show substantial change under anthropogenic warming. For example, surface warming induces stronger stratification, weaker nutrient supply and more rapid nutrient depletion. Using six Earth System Models, we have examined spatial patterns and trends of DNF in the CMIP5 historical and RCP 8.5 experiments. Four models (CanESM2, CESM1-BGC, IPSL-CM5R-LR, and UVicESCM) show high DNF rates in warm, stratified waters mostly associated with the western parts of the ocean basins, while GFDL-ESM2M and MPI-ESM-LR show elevated rates near the eastern boundaries because of coupling of DNF and denitrification. Despite a growing body of data, the spatial pattern of DNF is still insufficiently resolved by available observations, and none of the models agrees well with the observations. Modelled and observed rates are mostly in the same general range except for UVicESCM, and frequency distributions are similar, but spatial pattern correlations are weak and in most cases not statistically significant. Only a few models show strong trends in DNF and primary production in a warming climate, and the signs of the trends are inconsistent. Observations of primary production at the benchmark subtropical station ALOHA (22.75°N, 158°W) and proxies for historical DNF from the same region appear to corroborate trends in CanESM2 that are not present in other models. However, the CanESM DNF parameterization does not include any limitation by P or Fe, so modelled future trends may not materialize due to nutrient limitation. Analysis of available models and observations suggests that our understanding of environmental controls on ocean DNF remains limited and future trends are highly uncertain. Long-term global simulations of DNF will only be meaningful if we maintain long-term observations and extend coverage to undersampled regions. Keywords: ocean dinitrogen fixation; Earth system modelling; anthropogenic climate change

1 Introduction The global ocean nitrogen inventory depends on the balance between dinitrogen fixation (DNF) and deni­ trification, and provides an important control on ocean uptake of atmospheric CO2 (Galbraith et al., 2008; M ­ atear et al., 2010). Rapid climate change associated with the anthropogenic enhanced greenhouse effect could change the balance between these two biological processes so as to either enhance or attenuate the ocean sink for CO2. Climate change in the coming century will directly affect ocean ecosystems and temperature-dependent rates of

* School of Earth and Ocean Sciences, University of Victoria, Victoria, BC, CA Fisheries and Oceans Canada, Institute of Ocean Sciences, Sidney, BC, CA



Canadian Centre for Climate Modelling and Analysis, Victoria, BC, CA



Corresponding author: O. G. J. Riche ([email protected])

biogeochemical processes, but what the net effect on ocean CO2 uptake will be is not known (Gruber, 2011). At present, the open ocean (bottom depth > 180 m) accounts for the largest share of global photosynthetic biomass production at 83% of the ocean fraction; the remaining 17% includes coastal zones and upwelling regions (Karl et al., 1996). Field et al. (1998) estimated about equal marine and terrestrial contributions, implying an open ocean contribution of >40% of the global (ocean + terrestrial) total. In low-latitude subtropical oceans, as much as half of primary production is supported by DNF as its principal nitrogen source (Karl et al., 1997). At higher latitudes, recent observations in the Canadian Arctic (Blais et al., 2012) suggest a small (~4%) contribution of DNF to new primary production. Enhanced stratification under anthropogenic warming can be expected to result in reduced primary production via reduced vertical nutrient supply (e.g., Steinacher et al., 2010; Gruber, 2011; Roxy et al., 2016; Yasunaka et al., 2016). Yet enhanced stratification can also potentially increase DNF if adequate sources

Art. 16, page 2 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

of phosphorus or iron are available, with opposing effects on primary production (Karl et al., 1995, 1997; Karl, 1999). As diazotrophic phytoplankton thrive in warm, stratified waters (Sohm et al., 2011), investigating the potentially enhanced role of diazotrophs in tropical and subtropical ocean ecosystems in a warming climate is important. Proxy δ15N measurements of the DNF contribution to the N inventory in the subtropical Pacific suggest that the recent apparent enhancement discussed by Karl (1999) is a long-term trend throughout the 20th century, possibly associated with anthropogenic warming (Sherwood et al., 2014). We have examined trends in DNF in a variety of Earth System Models associated with the CMIP5 project (Taylor et al., 2012) to determine which if any show future trends and what the controlling processes may be. The simulations considered include present day conditions and future warming scenarios such as RCP 8.5 (a ‘no-mitigation’ scenario in which atmospheric CO2 concentration approaches 1000 ppm by 2100). The available models present a diversity of approaches to modelling DNF and therefore encompass a range of possible responses to future climate changes that are consistent with present knowledge of ocean DNF. 2 Data and methods

2.1 CMIP5 models and experiments

The CMIP5 data repository contains five models that include ocean DNF as an output field: CanESM2 (Zahariev et al., 2008; Arora et al., 2011), CESM1-BGC (Moore et al., 2002, 2004, 2013), GFDL-ESM2M (Dunne et al., 2013), IPSLCM5A-LR (Dufresne et al., 2013; Aumont et al., 2015) and MPI-ESM-LR (Ilyina et al., 2013). In addition we included output from the intermediate-complexity UVic Earth System Climate Model (UVicESCM; Weaver et al., 2001), which includes a prognostic diazotroph group within its ocean biogeochemistry module (Schmittner et al., 2005, 2008), and has been run with radiative forcing from most of the CMIP5 core experiments (M. Eby, pers. comm.).

The experiments considered were the Historical and RCP 8.5 experiments. The Historical experiment (1850–2005) uses radiative forcing based on observations, which include concentrations or emissions of natural and anthropogenic aerosols, solar forcing, atmospheric composition and land use (Taylor et al., 2012), while the RCPs are hypothetical future scenarios over 2006–2100 (Moss et al., 2010; Taylor et al., 2012). We examined the differences between a “present day” climate taken as a monthly climatology of 1986–2005 of the Historical run, and the most comprehensive data base of DNF observations available (see Section 2.3). In examining the time series stations HOT (22.75°N, 158°W) and BATS (31.67°N, 64.17°W), we also considered the full 250 years of Historical + RCP8.5 (1850–2100). In the case of UVicESCM, only two periods of 20 years (1980–1999 and 2080–2099) were available to us. We refer to the individual models as CanESM, CESM, GFDL, IPSL, MPI and UVic to signify CanESM2, CESM1BGC, GFDL-ESM2M, IPSL-CM5A-LR, MPI-ESM-LR, and UVicESCM. Since understanding of ocean DNF is still very limited (Diez et al., 2008; Sohm et al., 2011; Luo et al., 2012, 2014; Voss et al., 2013), the models are quite diverse in their approaches to modelling DNF. Most models represent DNF as a function of light, temperature, and nutrient concentrations, or some combination of these. Table 1 summarizes the major characteristics of the models, in particular diazotroph requirements or parameterizations of dependence on nutrients and environmental conditions. Some models have a prognostic diazotroph group (e.g., CESM), while others do not. Most use parameterizations based on observations of and experiments with Trichodesmium spp., although many different pelagic diazotrophs are now known, and others may be regionally more important than Trichodesmium in terms of total contribution to DNF (Zehr and Ward, 2002; Karl and Church, 2014; Zehr and Bombar, 2015). A brief description of the various parameterizations follows.

Table 1: Controls on marine dinitrogen fixation in the different models. DOI: https://doi.org/10.1525/elementa.277.t1 Model

DNF dependencea

References

T

PAR

N

P

Fe

O2

Biomass

CanESM

+b

+

–c

ndd

nd

nd

nd

Zahariev et al., 2008

CESM

+

+



+

+

nd

+

Holl and Montoya, 2005; Moore et al., 2002, 2007, 2013

GFDL

+

+



+

+



+

Holl and Montoya, 2005; Dunne et al., 2013

IPSL

+

+



+

+

nd

nd

Séférian et al., 2013; Aumont et al., 2015

MPI

nd

nd

e



+

nd

nd

nd

Ilyina et al., 2013

UVic

+

+

nd

+

nd

nd

+

Schmittner et al., 2008

e

Except for MPI, all models depend on temperature and PAR (see Section 2.2). Indicates an increase of DNF with increase of the environmental factor (e.g., P limitation); under Biomass, + implies prognostic diazotrophs, which are also subject to the other limitations or inhibitions. c Indicates a decline in DNF with increase of the environmental factor (e.g., N inhibition). d No dependence on the environmental factor. e In MPI, N and P dependencies are linearly combined; i.e., DNF depends on –N* = 16P – N (see Section 2.2). a

b

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

2.2 Brief description of DNF models

The simplest parameterization, as exemplified by MPI (Equation 1), scales DNF to availability of nitrogen relative to phosphorus. This scaling is similar to the concept of the quasi-conservative tracer N* = [NO3− ] − R N :P [PO34− ], where RN:P = 16 is the N:P Redfield ratio, and N* is a proxy of DNF when N* > 0 (Gruber and Sarmiento, 1997). In MPI, DNF is assumed to occur, at rate λMPI, in the surface layer wherever there is an excess of P with respect to N.

λMPI  =  λ0  (−N ∗ )  for N ∗ < 0 (1)

where λ0 is a reference rate in d–1 (Table 2). An intermediate approach has a more complex parameterization of DNF, but still parameterizes DNF as introduction of ‘new’ inorganic N without having a prognostic diazotroph group. CanESM prescribes a vertical profile of diazotroph abundance, linear light and temperature dependence, and inhibition by reactive nitrogen (Equation 2). λCanESM λ0 (C 1  e  a Nf  z  e −aNf z + C 0    = ) max (T − T min , 0)   K N I 0 − k − k chl z      e ( w chl 0 )    T max − T min K N + N I max

(2)

Art. 16, page 3 of 18

Kx, where x is either ammonium, nitrate, phosphate or iron, is also a function of phytoplankton biomass for ammonium and nitrate (Aumont et al., 2015, their equations 7a–c). Aumont et al. (2015) explain that “half-saturation constants increase with the size of the phytoplankton cell as a consequence of a smaller surface-to-volume ratio”:

P1  = min (P , Pmin ) (3e)



= P2   max (0, P − Pmin ) (3f)



K X = K Xmin

P1 + S sat P2 (3g) P1 + P2

where P is the biomass, Pmin a minimum phytoplankton biomass, Kxmin the minimum half-saturation constant (for either NO3– or NH4+; Table 2) and Ssat the size ratio of phytoplankton (set to 3). UVic (Equation 4) exemplifies the general form of models with a prognostic diazotroph group. Diazotroph biomass (DN in mmol N m–3) is combined with a temperature-dependent growth rate and a nutrient limitation factor, in this case phosphate limitation (Schmittner et al., 2008). In the UVic model, the rate of DNF is determined by the most limiting factor, either nutrient or light at depth z (Schmittner et al., 2008).

where C0, C1 and aNf are coefficients determining the vertical profile of abundance of Trichodesmium spp., T is the sea-surface temperature in Kelvin, KN and Imax are coeffi⎛ ⎞⎟ 3− ⎟⎟ ⎜⎜⎜ ⎡⎣⎢PO 4 ⎤⎦⎥ α I (z ) cients determining the dependence of DNF on inorganic ⎟⎟(4a) , λUVic = λmax DN min ⎜⎜ 3 − ⎟⎟ 2 N concentration and irradiance, Io is surface PAR, and kx ⎜⎜ K P + ⎡⎢PO4 ⎤⎥ 2 + λ α I z ( ) ⎣ ⎦ ( ) ⎟⎟ ⎜ max ⎝ ⎠ are light attenuation coefficients for water and (surface) chlorophyll (chl0). CanESM has no limitation by other T min ⎞ T ⎛ nutrients such as P or Fe. Parameters for all of the models ⎟ ⎜⎜ T ref − e T ref ⎟ (4b) = λ μ max 0  , e ⎟⎟ ⎜ max 0 are defined in Table 2. ⎜⎝⎜ ⎟ ⎠ IPSL has the most complex diagnostic parameterization (Equations 3a–g). In particular, inhibition by reac- Diazotrophs can use nitrate where available, but are not tive nitrogen depends on both ammonium and nitrate. growth-limited by nitrogen. Equation A3 in Schmittner et Temperature dependence is exponential for T ≥ 20°C, and al. (2008) defined NO3– uptake as follows: factors for iron and phosphorus limitation are included ⎡NO− ⎤ (Equation 3d). Note that, in IPSL, the light dependence of NO3− ⎣⎢ 3 ⎥⎦ μUVic = λUVic (4c) DNF is an exponential function of PAR, represented as I(z) 16 K P + ⎡⎣⎢NO3− ⎤⎦⎥ here. CESM (Moore et al., 2013) is mostly based on the earlier BEC model by Moore et al. (2002). The CESM DNF imple= λIPSL   λ0 max ( μ (T ) − μ0 ,  0)   mentation differs in several ways from the original BEC  I ( z ) ⎞ ⎛ − ⎟ DNF implementation. For instance, DNF depends on the ⎜⎜ fix ⎟ K min (LFe ,  LPO 4 )  1 ⎜⎜ − e  I ⎟⎟ (3a) carbon biomass, growth rate, and a biomass-specific DNF ⎟ ⎝⎜ ⎠⎟ rate. The modelled diazotrophs can utilize both fixed dinitrogen and inorganic nitrogen as a nutrient source follow⎧⎪  L   K  =  ⎪⎨ (3b) ing Holl and Montoya (2005); nitrate and ammonium half⎪⎪⎩ 0.01 if L  ≤ 0.2 saturation constants have been set so that diazotrophs do not strongly compete with the other phytoplankton (0.1 μM, about 5 to 10 times larger than for the other groups; K NO−  K NH+ 3 4 L  = Moore et al., 2013). Below (Equation 5) we reproduce the − + (3c) K NO−  K NH+  +  K NH+  [NO3 ] +  K NO− [NH4 ] 3 4 4 3 BEC DNF (Moore et al., 2002), which depends on cell N quota (Q), diazotroph biomass and temperature. Follow3− ing Geider et al. (1998), as for the other phytoplankton PO [Fe] 4 = L Fe = , L PO4 groups, the modelled inorganic nitrogen uptake declines 3 − (3d) K Fe + [Fe] K P + [PO4 ] as Q approaches Qmax. Temperature-dependence is param-

Art. 16, page 4 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Table 2: Model parameters. DOI: https://doi.org/10.1525/elementa.277.t2 Model

Parameter

Symbol

MPI

reference DNF rate

λ0

Redfield nitrogen to phosphorus ratio

RN:P

reference DNF rate

λ0

7.2 × 10–5

nitrogen half-saturation constant

KN

0.1

maximum reference diazotroph concentration

C1

500

trichomes L–1

surface reference diazotroph concentration

C0

50

trichomes L–1

inverse depth of concentration maximum

aNf

0.1

m–1

minimum temperature

Tmin

20

°C

maximum temperature

Tmax

30

°C

maximum reference surface irradiance

Imax

350

W m–2

light attenuation coefficient for water

kw

0.04

m–1

light attenuation coefficient for chlorophyll

kchl

0.03

m–1 (mg chl m–3)–1

reference DNF rate

λ0

0.013

temperature-dependent growth coefficient

µ(T)

minimum growth rate

CanESM

IPSL

CESM

GFDL

5 × 10

16

d–1 mol N (mol P)–1 µmol trichome–1d–1 µmol N L–1

µmol L–1 d–1

µ0

2.15

d–1

photosynthetic parameter

Ifix

50

W m–2

half-saturation constant for P limitation

KP

0.8

µmol P L–1

half-saturation constant for Fe limitation

KFe

0.1

nmol Fe L–1

KminNO3–

0.13

µmol N L–1

KminNH4+

0.013

µmol N L–1

temperature growth coefficient

µ0

0.055

d–1

reference temperature

Tref

15.65

°C

minimum temperature

Tmin

15

°C

half-saturation constant for P limitation

KP

initial slope of P-I curve

α

~0.044 0.1

reference DNF growth coefficient

λ0

0.4

Arrhenius coefficient

AT

4000

K

reference temperature

Tref

303.15

K

minimum and maximum N:C cell quotas

Qmin, Qmax

reference growth coefficient

µ0

0.65

d–1

half-saturation constant for N inhibition

KN

µmol N L–1

dissolved oxygen reference concentration

O2ref

~7 ~300

temperature sensitivity coefficient

α

0.063

°C–1

maximum P:N cell quota

[QP:N]max

0.167

mol P mol N–1

maximum Fe:N cell quota

[QFe:N]max

4.4 × 10–3

mol Fe mol N–1

Fe:N half saturation constant

KFe:N

2.4 × 10–4

mol Fe mol N–1

eterized according to the Arrhenius equation, with a minimum temperature for diazotroph growth of 16°C.

Q   1  − f λCESM  = λ0   max    DN  e Q 1.015 − f

⎡1 ⎢ − 1 − AT    ⎢T T ref ⎣

⎤ ⎥ ⎥ ⎦

(5a)

0.034, 0.17

D DC

µmol P L–1 (W m–2)–1 d–1 d–1

mol N mol C–1

µmol O2 L–1

Q  −  Q min (5b) Q max − Q min

N = Q    = ,  f    



Unit –3

0.6 × 1.066T

minimum half-saturation constants for N limitation UVic

Value

DN (DC) is diazotroph biomass in moles N (C) m–3; Q, Qmin, and Qmax are the N:C cell quota and minimum and

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

maximum quotas, respectively; and T is the temperature expressed in Kelvin. Diazotroph biomass is modelled ­prognostically according to modified Geider et al. (1998) iron- and phosphorus-limited growth (5 nM and 90 pM, phosphate and iron half-saturation constants, respectively, in Moore et al., 2007). GFDL (Equation 6) is the most complex and complete prognostic model, with micro- and macronutrient limitations based on iron and phosphorus cell quotas, an inorganic nitrogen inhibition factor, and an oxygen inhibition factor. Temperature dependence is exponential and not limited to the tropical range as in the four previous models (no temperature dependence in the case of MPI). ⎛ ⎞⎟ 4 ⎜ [O2 ] KN ⎟⎟ = λGFDL μ DN ⎜⎜⎜1 − 4⎟ ⎟ 4 ref ⎡ ⎜ [ ] + ⎡O ⎤ ⎟⎟ K N + NO3− ⎤ + ⎡NH+4 ⎤ (6a) ⎢⎣ ⎥⎦ ⎢⎣ ⎥⎦ 2 ⎢⎣ 2 ⎦⎥ ⎠ ⎝⎜  O



fP =

μ  = μ0  min (f P ,  f Fe )  e α T  e 1−f (PAR) (6b)

Q [Q P :N ]max

P :N , f Fe =

⎛D

Q Fe2 :N , and K Fe :N + Q Fe2 :N 2

⎡D , ⎢ x ⎢ ⎝⎜ DN ⎣ D N

Q X :N = min ⎜⎜⎜

x

⎤ ⎞⎟ ⎥ ⎟⎟ ⎥ ⎟⎟ ⎦ max ⎠

(6c)

The term e1–f(PAR) represents the light limitation following Geider et al. (1997), where f(PAR) depends on the initial slope of the PI curve, the carbon to chlorophyll ratio and temperature. DN (DP, DFe) is the diazotroph biomass in moles N (P, Fe). Q and Qmax are the phosphorus or iron cell quota (relative to nitrogen) and maximum quota, respectively. O2 and O2ref are the dissolved oxygen concentration and inhibition concentration (Dunne et al., 2013); KN is a half-saturation constant for reactive nitrogen inhibition. The modelled diazotroph biomass follows Geider et al. (1997) and can assimilate dissolved inorganic nitrogen following Holl and Montoya (2005). 2.3 Observations

Luo et al. (2012) provide the largest up-to-date dataset of observed DNF rates. We have used a subset of the Luo et al. (2012) dataset: 15N2 assimilation measurements over the whole euphotic zone. These are believed to be the most accurate estimates of water column DNF rate, due to consistent use of a modern and accurate method (Montoya et al., 1996). A preliminary examination of IPSL and UVic rates suggested that they have extremely high DNF rates (total global 657.2 and 315.9 Tg N y–1, respectively), and their maximum areal rates are outside the range of the other CMIP5 models. In IPSL the long-term nitrogen budget was balanced by scaling DNF to denitrification (Séférian et al., 2013), and denitrification is high because volume of the oxygen minimum zone (OMZ) in the Pacific is overestimated (Séférian et al., 2013, Cabré et al., 2015). For illustrative purposes, IPSL and UVic DNF rates have been rescaled in Figure 1 by 0.17 and 0.40, respectively. UVic was scaled to the global total DNF rate estimate of Gruber and Sarmiento (1997); IPSL was scaled to its value in an

Art. 16, page 5 of 18

ocean-only simulation with the same biogeochemistry model (Aumont et al., 2015). For IPSL only, this rescaling is retained in all of the figures and tables, and IPSL is excluded from any discussion of absolute magnitude. Most of the available observations in Luo et al. (2012) are from subtropical regions. We limited our analyses to latitudes between 45°S and 45°N. Since the observations are mostly in the open ocean and the models have coarse resolution, we also limited our analysis to the open ocean, excluding coastal waters and some of the marginal seas. In the Atlantic, we included the Gulf of Mexico and the Caribbean Sea, as these are large areas of warm, stratified ocean known from historical observations to have active diazotroph populations. In the Indian Ocean, we included the Arabian Sea but excluded the Bay of Bengal (north of 5°N), because coarse resolution models create a nonphysical “nutrient trap” there, and because observations suggest that DNF is much greater in the Arabian Sea (Naqvi, 2008). All Pacific marginal seas were excluded, as they are poorly resolved by coarse-resolution global models; DNF is known to occur in the South China Sea, but at much lower rates than in open ocean waters of the low-latitude Pacific (Chen et al., 2003, 2014). Modelled DNF rates were regridded to a common regular 2° × 2° grid. Observed DNF was projected onto the same grid and available observations pooled and averaged within each grid cell. Total DNF was calculated for the models and the observations using the regridded and averaged maps. Individual-basin totals are estimated for 0–45°N or S. The Indian Ocean total includes both the northern and southern hemispheres (north of 45°S), excluding the Bay of Bengal north of 5°N. Estimates of the total observed DNF rates are based on Luo et al. (2012) within the same areas. Mean observed rates are calculated (weighted by grid cell area) within each region and multiplied by the area of the region. We also compared the modelled DNF and primary production (PP) at the benchmark time series stations HOT (22.75°N, 158°W) and BATS (31.67°N, 64.17°W). DNF data at these locations were taken from Luo et al. (2012) while PP rates are available over the period 1989–2014 (Bates et al., 2001; Lomas et al., 2002; Brix et al., 2006; Karl and Church, 2014) represented by 247 and 320 approximately monthly profiles at HOT and BATS, respectively. For DNF and PP time series comparisons, a single model grid cell containing the stations was chosen, and data were averaged to annual means. At BATS only a very limited number of observed DNF rates was available (see Section 3). At BATS we also considered a geochemical estimate of DNF based on long-term nutrient concentration time series (Singh et al., 2013) and modelled DNF estimates by Hood et al. (2001). For modelled DNF and PP at these locations, a single 2° × 2° grid cell containing the stations was chosen, and model outputs were averaged to annual means. Linear trends were calculated for modelled DNF and PP at HOT and BATS for the period 1986–2100, except in the case of UVic where only the 1980–1999 and 2080–2099 time periods were available. In this case, estimating the trend might be more appropriately done as simply the difference between means for these two periods, but

Art. 16, page 6 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Figure 1: Global depictions of mean depth-integrated DNF rate from the six models examined. Values as estimated by a) CanESM, b) CESM, c) IPSL, d) UVic, e) GFDL, and f) MPI, for 1986–2005 of the Historical experiment, except UVic which is based on the period 1980–1999. All models have been regridded to a common 2° × 2° grid. A common colour scale is used to indicate DNF rate; black areas correspond to values >400 μmol m–2 d–1. DOI: https:// doi.org/10.1525/elementa.277.f1 calculating regression coefficients for all available data gives essentially the same answer. PP trends were calculated for HOT and BATS for the time period of data availability. Not enough data were available to estimate trends in observed DNF. Regression coefficients were tested for significance using a standard t-test, assuming a decorrelation time scale of 3 years for the models and 3 months for the observations. This approach is conservative in all but a few cases, but it makes no difference to the results. All significant trends were significant regardless of what assumptions were made regarding autocorrelation, with most p values at 0.01 or 0.001. 3 Results Table 3 shows the range of total DNF rates and their distribution among ocean basins. In the Pacific, most of the estimated basin totals show good agreement with observations. UVic systematically overestimates DNF, but the

distribution among basins is similar to other models. In the Atlantic, the total DNF rates tend to be larger than observations, especially in the South Atlantic. South Atlantic rates are low in the Luo et al. (2012) data base, but similar rates have been reported elsewhere (Fernandez et al., 2010). CanESM, CESM, and GFDL represent the lower end of the range of modelled rates of DNF. Global total DNF for CanESM, CESM, and GFDL fall within 100–200 TgN y–1, with MPI just slightly higher; observation-based estimates range from 103 to 177 TgN y–1 (Großkopf et al., 2012; Luo et al., 2012). As mentioned earlier, IPSL CMIP5 data were rescaled to the rate of denitrification, resulting in a global total about five times larger than the observed range; ocean-only simulations with the same biogeochemistry model produce rates within this range (Aumont et al., 2015). UVic represents the higher end showing areal values as high as 1135 μmol m–2 d–1 (see Table S1 for statistics of observed and modelled distributions) and a global

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

total of 316 TgN y–1. Latitudes less than 45° account for the vast majority of the global totals (>90%, except for GFDL with 80% and MPI with 73%; Table 3). GFDL has substantial DNF in the mid-latitudes, particularly in the Southern Hemisphere (Figure 1), along the southern boundaries of the ocean basins where there is potential for episodic equatorward transport of surface nutrients from the HNLC region. The Pacific Ocean represents the largest share across all models with 40–57% of the global total (Table 3), which is mainly due to the larger surface area of the Pacific Ocean as areal DNF rates are usually within the same range across ocean basins. Totals for latitudes less than 45° tend to be higher than the observations. The Atlantic and Indian Ocean fractions (Table 3) are of similar magnitude, particularly in CanESM, CESM, and UVic. The modelled total for the Indian Ocean ranges between 9 and 27% (mean of 18.0%) of the global total; Atlantic Ocean fractions are slightly higher (13–26%, mean of 19.9%). In all models, the totals for the northern and southern hemispheres in the Pacific and Atlantic Oceans are similar. Figure 1 shows the geographic distribution of DNF for all six models. Note that UVic has been rescaled for illustrative purposes here (see Section 2.3). Models can be grouped according to spatial distribution. CanESM, CESM, IPSL and UVic all have high DNF in oligotrophic regions on the western boundaries, i.e., maximal outside of the coastal and equatorial zones with high concentrations of upwelled inorganic nitrogen. This distribution is not surprising as the models are parameterized to have inhibition of DNF by inorganic nitrogen (see Section 2.2). In these models, DNF is also dependent on warm temperatures, with minimum temperatures from 15°C (UVic) to 20°C (CanESM/IPSL). GFDL and MPI exhibit very different distributions from the other four models, with the largest DNF in the coastal and equatorial upwelling zones (Figure 1). In these models, DNF occurs in waters with excess P relative to N. In MPI, it is a simple linear relationship of DNF to –N* (Equation 1), and DNF is not inhibited by low temperature. In GFDL, nitrogen inhibition is combined with oxygen inhibition (Equation 6) which occurs at surface concentrations of

Art. 16, page 7 of 18

oxygen (~300 μM; Table 2). GFDL ­nitrogen inhibition has a high half-saturation constant (7 μM) compared to the other models, and is restricted mostly to subsurface waters or upwelling regions. In these two models, there is also high DNF at mid-latitudes where temperature limits DNF activity in the other models: GFDL in the Southern Ocean and MPI in the North Pacific. Spatial pattern correlations among models also suggest two groups (Table S2). MPI and GFDL are weakly correlated with the other models (r < 0.18), and are relatively weakly correlated with each other (r = 0.32). The other four models tend to have stronger spatial correlation with each other (r > 0.5), with the exception of CESM and UVic (r = 0.31). In the Indian Ocean, most of the models agree south of 15–20°S. Most models show a steep gradient from high to very low DNF rates, except for GFDL and MPI which have enhanced rates at higher latitudes (Figure 1). Rates are low near Western Australia, except in CESM (along most of the coast) and GFDL. In the interior of the Indian Ocean, CanESM and CESM have uniformly high DNF rates, while MPI is uniformly low and the other models are spatially variable. In CanESM, CESM, IPSL and UVic, high DNF areas spread across the tropical and subtropical oceans with a large area of low rates in the eastern tropical Pacific Ocean. IPSL tends to have high DNF on the eastern side of the Indian Ocean and low DNF in the southeast Pacific Ocean. Figure 2 shows comparisons of modelled DNF rates to observations compiled by Luo et al. (2012, 2014). Figure 2a shows the spatial distribution of measurements; almost all of the data are located in the Pacific and Atlantic Oceans and during the 2000–2010 time period (Luo et al., 2012). Within the same region, the data can range over several orders of magnitude. For example, in the North Atlantic the range is 1 to 248 μmol m–2 d–1. The models overestimate DNF rates, especially in the south Atlantic Ocean (Figure 2b–g). The small number of rates exceeding 200 μmol m–2 d–1 are not consistently reproduced by any model. Counting the points above and below the 1:1 line, all of the models overestimate DNF, with IPSL (rescaled by 0.17) having the highest fraction of locations overestimated (95 of 133) and GFDL the

Table 3: Total DNF rates (Tg N y–1) by region. DOI: https://doi.org/10.1525/elementa.277.t3 Model

Global total

Regional totalsa N Pacific

S Pacific

N Atlantic

Latitude < 45°

144.4

40.2

30.4

19.1

11.5

31

132.3

CESM

175.1

38.5

32.1

26.4

19.2

47.4

163.6

b

IPSL

111.8

34.1

29.0

10.2

11.8

15.4

100.6

UVic

315.9

87.3

92.7

34.5

18.1

65.2

297.8

GFDL

158.8

29.4

38.2

15.7

20.8

23.6

127.7

203.7

48.1

56.6

11.8

13.9

18.3

148.7

34.3

49.5

10.9

2.3

na

97

Observed

c

na

d

Sum of individual basin/hemisphere totals to latitude < 45°. IPSL rescaled by 0.17 (see Section 2.3). c Values estimated from data compiled by Luo et al. (2012). d Indicates not available. b

Indian Ocean

CanESM

MPI

a

S Atlantic

Art. 16, page 8 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Figure 2: Observed depth-integrated DNF and scatter plots of modelled vs. observed values. a) Observed depthintegrated DNF rate (monthly climatology) from the MAREMIP database, and (b–g) comparison between mean modelled (y-axis) and observed (MAREMIP, x-axis) rates: b) CanESM, c) CESM, d) IPSL (rescaled by 0.17), e) UVic, f) GFDL, and g) MPI. The black dashed line indicates 1:1. An outlier observed value (1189 μmol m–2 d–1) has been omitted. Note different time period for UVic. DOI: https://doi.org/10.1525/elementa.277.f2 lowest (79 of 131). The relative overestimation is usually larger in the lower part of the observed range, e.g., 0–100 μmol m–2 d–1; 77% of the observations fall in this range. CanESM, CESM, GFDL and MPI (excluding an outlier in the latter) have an overall range lower than the observed rates: observed values go up to ~600 μmol m–2 d–1 (one outlier at 1189 μmol m–2 d–1 has been excluded for visualization purposes) while modelled values are less than 300 μmol m–2 d–1. GFDL and MPI mostly cluster close to the 1:1 line at the lower end of the range. The spatial correlation between models and observations is not statistically significant in most cases. Correlations are significant only for CESM (r = 0.21) and IPSL (r = 0.23) for untransformed data, and for IPSL (r = 0.32) and UVic (r = 0.21) for logtransformed data (Table S2). Figure 3 shows frequency histograms of DNF rates. Observed rates vary from 0 to 600 μmol m–2 d–1 (excluding

one outlier as above). The distribution is strongly positively skewed with a large proportion of low rates and a tail of high rates. Histograms of modelled DNF rates at the locations of the observations show that IPSL, UVic, and GFDL agree on the shape of the distribution. MPI mostly agrees with the observations, although it has a larger frequency of high DNF rates and lower frequency of low rates. In the North and South Pacific and South Atlantic, MPI exhibits a mode around 75 μmol m–2 d–1. CanESM and CESM differ from the observations in having a more Gaussian distribution with a strong mode at intermediate DNF rates. Figure 4 shows frequency histograms similar to Figure 3 but considering all locations, not just those where observations exist. In some cases, the shapes of the histograms change substantially. Overall the shapes of the histograms are closer to that of the observations, particularly for CanESM and CESM. IPSL and UVic change

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Art. 16, page 9 of 18

Figure 3: Histograms of DNF rates (monthly climatology) at locations of observations. a) Mean observed depth-integrated DNF rate from the MAREMIP database, and (b–g) modelled DNF rates: b) CanESM, c) CESM, d) IPSL (rescaled by 0.17), e) UVic, f) GFDL, and g) MPI. Modelled values are monthly, for a climatology calculated over the 20-year period indicated. An outlier observed value (1189 μmol m–2 d–1) has been omitted. Note different scales and time period for UVic. DOI: https://doi.org/10.1525/elementa.277.f3 the least, and overall have a shape closest to the observations (rescaling does not affect this result). All of the models approximate the observed pattern of a positively skewed distribution (with varying degrees of skew) but several (CanESM, GFDL, MPI) have modes at intermediate rates that are not present in the observations. Statistics of the distributions are provided in Table S1. GFDL is the closest to the observed distribution in terms of maximum, standard deviation and skewness. All exhibit positive skew ranging from 0.2 (CanESM) to 4.8 (MPI); even the lowest values are significantly different from 0 (P < 0.001). Figure 5 compares observations and models at the benchmark subtropical stations HOT and BATS. For DNF, observations at HOT cover a longer time period than at BATS and thus capture more of the variance in actual rates. At HOT, observed primary production (PP) shows an increasing trend (Figure 5a; Table 4). For the models, trends are calculated over a much longer period (1986– 2100), with only CanESM and UVic showing positive trends in PP. While the absolute rate of PP is very low in

CanESM, the relative magnitude of the trend is similar to that in the observations, suggesting approximately a doubling of PP over a century. At BATS (Figure 5b), CanESM shows a similar trend, but the observations do not show a statistically significant trend. Based on 1980–1999 and 2080–2099, UVic PP increases slightly over time at both stations, but declines at HOT after 2080. All of the other models show a decrease of PP with climate warming. In CanESM, DNF supports about ~45% of PP at HOT and 3–30% at BATS, assuming a Redfield scaling. Modelled DNF rates (Figure 5c) all fall within the observed range at the HOT station (16–628 μmol m–2 d–1), although CESM is clearly at the high end. At BATS (Figure 5c), only two models (UVic, MPI) fall within the lower observed range of 15–39 μmol m–2 d–1 (which is based on only 6 data points) in the present climate. The other three models fall within the higher range of 111–129 μmol m–2 d–1 estimated from long-term time series of nutrient concentrations at BATS (Singh et al., 2013). When comparing modelled DNF to earlier estimates based on

Art. 16, page 10 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

°

Figure 4: Histograms of DNF rates for all locations. a) Mean observed depth-integrated DNF rate from the MAREMIP database (same as Figure 3a, reproduced here for ease of comparison), and (b–g) modelled DNF rates for all locations: b) CanESM, c) CESM, d) IPSL (rescaled by 0.17), e) UVic, f) GFDL, and g) MPI. An outlier observed value (1189 μmol m–2 d–1) has been omitted. Note different scales and time period for UVic. Histogram construction and colour codes are identical to Figure 3. DOI: https://doi.org/10.1525/elementa.277.f4 models constrained by observations of DIC and N export (Hood et al., 2001), most of the CMIP5 models fall within the lower half of this range. CESM has higher rates in the present climate but declines into the observed range after about 2040. IPSL (rescaled by 0.17) and UVic rates are the lowest at this location. Rates in the tropical-subtropical Atlantic decline with latitude in CESM and CanESM, but this gradient is much more abrupt in IPSL and UVic (Figure 1a–d). DNF rates at HOT show a range of trends (Figure 5c), with half of the models projecting a long-term increase (CanESM, CESM and GFDL) while the others show a decline. At BATS (Figure 5d), there is similar divergence of trends with two models (CanESM and IPSL) showing a long-term increase and the other four declining. Only half of the models show trends of the same sign at both stations, and all four possible combinations (+ +, + –, – +, and – –) are represented among the six models (Tables 4 and 5). The physical processes underlying the positive and negative trends vary. Trends in environmental factors are

summarized in Table 5. Figure S1 shows the corresponding time series of the environmental controls for CESM, IPSL, UVic, and GFDL. Table S3 shows the temporal correlation between DNF rates and environmental controls across all the models. In CanESM, DNF increases at both stations despite an increase in the concentration of dissolved inorganic N (DIN) at BATS (Table 5). The concentration increase is not enough to inhibit DNF, and DNF appears to be driving DIN rather than the reverse. In MPI, DNF declines at both stations, as the supply of P declines relative to that of N (Table 5). Note that in CESM and IPSL, inhibition of DNF by ammonium is stronger than by nitrate (Table 2). In CESM, GFDL, and IPSL, processes leading to nutrient limitation differ at BATS and HOT, but P limitation generally prevails over iron limitation. Dissolved iron concentration is high; e.g., the minimum iron concentration at HOT in IPSL is ~150 pM (Figure S1e), whereas the halfsaturation constant is 100 pM (Table 2). P decreases over time in all but one case: GFDL at BATS (Table 5). In IPSL, P concentration at BATS is higher than at HOT, but the rate of decline is greater at HOT (Figure S1e–f); P is completely

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Art. 16, page 11 of 18

Figure 5: Time series of modelled and observed DNF and primary production at stations HOT and BATS. Model outputs represent the 2° × 2° grid point nearest the stations. a) Primary production at HOT, b) primary production at BATS, c) DNF rate at HOT, d) DNF rate at BATS. Observed DNF rates are depth-integrated 15N2 assimilation measurements (Luo et al., 2012); primary production is from 14C uptake as per HOT and BATS protocols. The upper bound for DNF is 222 μmol N m–2 d–1 for strong mixing and 518 μmol N m–2 d–1 for weaker mixing (Hood et al., 2001). The time period of the observations is 1989–2014. The horizontal positions from Hood et al. (2001) for strong mixing, weaker mixing, and from Singh et al. (2013) are located in the centre of the associated time periods, 1989–1994, 1995–1996, 1988–2009, respectively (grey-scaled bars). IPSL DNF rate has been rescaled by 0.17. DOI: https://doi.org/10.1525/ elementa.277.f5 Table 4: Linear trends in DNF and primary production (PP) over 100 years, expressed as percentage of the mean value in 1986a, at the time-series stations HOT and BATS. DOI: https://doi.org/10.1525/elementa.277.t4 Observations or models

HOT (22.75°N, 158°W) PP

Observations

DNF

PP

b

DNF

108.3

b

na

22.2 (NS)

na

CanESM

79.0

78.7

201.2

93.2

CESM

–8.6

21.2

–11.4

–63.9

IPSL

–25.4

–103.0

–12.5 (NS )

224.8

UVic

4.8

–73.6

4.0

–89.9

GFDL

–15.5

35.1

–1.9 (NS)

–12.7

MPI

–14.7

–16.4

–32.1

–45.5

Except for UVic (1980) and observed PP (1988). Not available due to insufficient data. c Not significant; all other trends are statistically significant (α = 0.05 or less). a

BATS (31.67°N, 64.17°W)

c

Art. 16, page 12 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Table 5: Signsa of the modelled trends of DNF and environmental factors at the time-series stations HOT and BATS. DOI: https://doi.org/10.1525/elementa.277.t5 Model

DNF

SSTb

NO3

NH4

Phosphorus

Iron

Biomassc

HOT

BATS

HOT

BATS

HOT

BATS

HOT

BATS

HOT

BATS

HOT

BATS

HOT

BATS

CanESM

+

+

+

+

+d

+

nd

nd

nd

nd

nd

nd

nd

nd

CESM

+



+

+



–d



+d





–d

+

+



IPSL



+

+

+

+

+

+

+





+

+

nd

nd

UVic





+

+

nd

nd

nd

nd





nd

nd





GFDL

+



+

+







+

+



+

+

+



MPI





nd

nd





nd

nd





nd

nd

nd

nd

d

d

d

+ and – signs indicate non-zero trends; nd, no dependence on the environmental factor. Sea surface temperature. c When DNF depends on biomass (CESM, GFDL, and UVic), the other dependencies (N, P, Fe) constrain a biomass-specific DNF rate. d Indicates trend is not statistically significant. a

b

depleted by 2016. In UVic, enhanced stratification limits the resupply of inorganic nutrients; in particular, surface PO4 steadily declines (Table 5, Figure S1g–h). As a result, both diazotroph biomass and per-cell DNF decline at both stations (Table 5; Figure S1g–h). In Table S4, we compare the trends of PP with the same environmental factors. As noted above, because iron concentration (CESM, IPSL, GFDL) is mostly above the halfsaturation constant, the PP decline is likely to depend on N or P. Except for CanESM (and UVic at BATS), enhanced thermal surface stratification reduces PP despite higher metabolic rates in warming waters. In UVic the long-term trend in PP at HOT is small (Figure 5). In CanESM, PP correlates significantly with surface nitrate at BATS; i.e., increasing PP is driven by increasing DNF. In CESM and GFDL, declining PP (Figure 5) follows declining N and P concentrations (Table S4) at both HOT and BATS; i.e., the trends that emerge from the simple parameterizations in CanESM are not present in the more complex models with prognostic diazotrophs and multiple nutrient limitations. 4 Discussion

4.1 How well do the models compare with the observations? Does one model emerge as best representing the observed spatial pattern? How do the models compare with each other?

The spatial pattern correlation between models and ­observations at the locations of the observations (Figure 2) suggests that to meaningfully distinguish among these models using these observations is not possible. The highest spatial correlation (log-transformed) was 0.32, and four of the six models showed no significant correlation. Using the observations to distinguish between the two model groups was also difficult (Figure 2; Table S2), as observations are mostly lacking in eastern boundary waters (Figure 2a). This finding indicates a need for future measurement campaigns in undersampled regions (Figure 2a). Figure 2b–g further show that models systematically overestimate (underestimate) in the low (high) range of the observations. In particular, the observed DNF rate in the South Atlantic is low (0–10 μmol m–2 d–1), while the modelled range is much broader (up to 300 μmol m–2 d–1).

In the South Atlantic, MPI has the least spread and the least positive bias with values mostly within 30–100 μmol m–2 d–1; these rates are the closest to the observations in this particular region (Figure 2a). Waters upwelling along eastern boundaries contain an excess of dissolved inorganic P with respect to N requirements (high negative N*). High P:N ratios tend to favour diazotrophs over other phytoplankton groups, which may result in strong spatial coupling between surface DNF and subsurface denitrification (Tyrrell et al., 1999; Deutsch et al., 2007; Landolfi et al., 2013). Denitrification removes DIN and potentially enhances DNF over OMZs in five of the six models (CESM, GFDL, IPSL, MPI, and UVic), all of which have P dependence of DNF. The spatial coupling is obvious in GFDL and MPI in equatorial and coastal upwelled waters (Figure 1). Model biases in ocean circulation and thermocline ventilation may result in expanded OMZs (Dunne et al., 2013; Cabré et al., 2015), so that the models may overestimate denitrification and upwelling of low N/P waters. In GFDL this bias is compounded by O2 inhibition of DNF, particularly in the Pacific. At HOT and BATS, only CanESM shows a positive trend in PP (mainly attributable to increasing DNF). These trends are consistent with previous observations (Karl, 1999; Sherwood et al., 2014). The other models encompass all possible combinations of PP and DNF trends. Increasing P limitation appears to be the most common environmental control among the models (Table 5; Figure S1). 4.2 Is the distribution of DNF Gaussian? Are the available observations sufficient to know? Do the models reproduce the observed distribution?

Figure 3a suggests a strongly positively skewed distribution of observed DNF rates (statistics are summarized in Table S1). This result is based only on 15N2 assimilation, but the distribution is similar to what was found using both 15N2 assimilation and C2H2-reduction assay measurements (Luo et al., 2012). Whether the positive skew of modelled DNF is due to the same processes as in the observations is not clear, but a positive skew of even 0.2 does not occur by chance (P < 0.0005 for N > 2000). Using all model grid points (Figure 4) gives a d ­ istribution

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

of ­modelled DNF rates that more closely resembles the observed distribution than one considering only locations where observations exist (Figure 3). CanESM, CESM, GFDL, and MPI show a much more positively skewed distribution, while IPSL and UVic distributions change little and fit the observed distribution fairly well. Luo et al. (2012) argued that the currently available observations did not cover the eastern equatorial regions for historical reasons (e.g., effort was focused on nutrientdepleted oligotrophic regions), even though these regions are important to understand the connection of DNF to processes like denitrification, upwelling, and O2 deficiency. The diversity of parameterizations in the CMIP5 models makes this data set potentially useful for understanding these processes as new observations emerge, although model error in processes such as ventilation and denitrification may limit their usefulness in some cases. 4.3 What are the important dependencies on environmental conditions? Are the models likely to respond to future changes in environmental conditions in a realistic way?

Besides temperature and PAR dependence, DNF parameterizations may assume iron and/or phosphorus limitation, and inhibition by DIN and O2. DIN inhibition is a dependence suggested by studies on Trichodesmium spp. (Capone et al., 1997; Holl and Montoya, 2005; Zehr, 2011); many of the models considered here do not have complete inhibition at high DIN concentrations. Recent laboratory experiments have qualified nitrogen inhibition by showing that DNF can occur when diazotrophs are acclimated to chronic high DIN concentrations, although at a lower rate than at low DIN concentration (Knapp, 2012; Knapp et al., 2012). DIN inhibition may be offset by phosphorus supply in iron-replete conditions as observed in laboratory experiments (Knapp et al., 2012) and field studies (Fernandez et al., 2011, 2015). Knapp et al. (2012) suggest that enhanced abundance of diazotrophs in P- and Fe-rich waters offsets reduced DNF due to nitrogen inhibition, resulting in DNF rates similar to low DIN conditions. In CESM, however, iron deficiency limits DNF in the southeastern tropical Pacific so that spatial coupling between DNF and denitrification is weak (Moore et al., 2007). Additional supply of dissolved organic phosphorus in oligotrophic regions is a mechanism that could potentially explain some discrepancies between observations and models (Anderson et al., 2015). However, there is no particular reason for the supply of dissolved organic phosphorus to increase in a warming climate, and even in the present climate there are net sources only in specific regions fed by transport from more productive regions (Anderson et al., 2015). It is now accepted that a large diversity of diazotrophs exists, including heterotrophic species (Sohm et al., 2011; Zehr et al., 2011). In particular, some recent studies strongly suggest that DNF occurs in nutrient-rich but low N/P waters (e.g., the Peru and Chile upwelling regions) with rates similar to those observed in oligotrophic surface waters, and may be mainly heterotrophic (Fernandez

Art. 16, page 13 of 18

et al., 2011, 2015). High DNF in eastern boundary current environments is consistent with the parameterizations ­ used by GFDL and MPI, although photoautotrophy is implied because of light (GFDL) or depth (MPI) dependence. Diazotroph adaptation to variable N/P environments can be represented by a dependence of diazotroph abundance on P concentration and DNF rate on N concentration or N/P ratio (Knapp et al., 2012). A prognostic approach (CESM, GFDL, UVic) may more realistically represent these effects than a diagnostic approach (CanESM, IPSL, MPI). In addition, a prognostic approach takes into account competition between diazotrophs and other phytoplankton groups for common resources such as P or Fe. Another potentially important control on DNF that has barely been explored in models is CO2. Experiments on Trichodesmium sp. by Levitan et al. (2007) showed that DNF, growth, biomass and cell C content were enhanced when Trichodesmium sp. cultures were acclimated to high pCO2 (900 ppm). Fowler et al. (2015) estimated trends of DNF under higher pCO2 (750–1000 ppm) based on similar experiments with Trichodesmium sp. and Crocosphaera watsonii. Hutchins et al. (2013) estimated DNF dependence on CO2 in isolates from the Atlantic and Pacific Oceans; these results suggest an increased DNF contribution to global new N supply with high CO2, an increase that is potentially irreversible (Hutchins et al., 2013, 2015). Gradoville et al. (2014), by contrast, found that DNF activity in Trichodesmium was insensitive to pCO2 enrichment in natural seawater conditions. These authors suggested that long-term experiments or field studies would help determine whether or not high-pCO2 species will eventually dominate the diazotroph community. It is important to resolve this question and develop adequate parameterizations for models, because CO2 will probably change at least as much as any of the environmental factors considered here. 4.4 What are the implications of CMIP5 results for DNF and primary production in an enhanced greenhouse climate?

None of the CMIP5 models has shown that a particular set of environmental controls on DNF satisfactorily explains the spatial pattern in the observations. Furthermore, the models divide into two distinct groups, one of which associates DNF with relatively nutrient-rich regions (GFDL, MPI), while the other favours stratified, oligotrophic regions (Figure 1; Table S2). The lack of observations in the eastern boundary current waters and the generally weak spatial correlation between models and observations prevents any firm statement that one or the other group of models is more realistic. However, the limited metrics available suggest that GFDL and MPI are as skillful as any of the others, despite the bias in the distribution of observations towards habitats that the other group assumes are most important. Warmer, more stratified oceans and a poleward expansion of such regions are robust consequences of increased greenhouse gas forcing (e.g., Xu et al., 2012). Four of the six models have DNF occurring primarily in this type of environments (Figure 1). Among these models, only CanESM

Art. 16, page 14 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

shows a strong increasing trend in DNF over the 20th and 21st centuries at HOT and BATS (Figure 5c–d and Table S3). If we assume that the trend is real (Sherwood et al., 2014), we must then ask whether the differences between CanESM and the other models indicate that CanESM has captured something essentially correct in the mechanism, or whether the model having reproduced this trend is coincidental. In CanESM, the secular trend is a result of DNF being primarily a function of temperature (Bissett et al., 1999), without limitation by other environmental factors except for inhibition at high concentrations of DIN. In the other models other factors, especially phosphorus, are limiting as the area of habitat that appears generally favourable expands. The Bissett et al. (1999) model was based on the idea that other phytoplankton groups would outcompete diazotrophs in nutrient-replete waters and that diazotrophs would outcompete other phytoplankton in oligotrophic regions (Capone et al., 1997). However, it was not intended to represent the sort of long-term trend seen in CanESM under anthropogenic global warming, which would require either great metabolic flexibility with regard to N/P ratios or additional sources of P or both. CanESM is the only model with no nutrient (P, Fe) limitation. The trends at HOT suggest that either diazotrophs are accessing new sources of P or depleting existing stocks. The ability of diazotrophs to efficiently use dissolved organic phosphorus represents a competitive advantage over non-diazotrophic phytoplankton (Dyhrman et al., 2006; Landolfi et al., 2015). However, if the dissolved organic pool is a source of ‘new’ phosphorus in a novel climate that favours diazotrophy, that pool could be depleted over time. The changes to date have been small relative to those expected in the future, and whether the kind of trends seen in CanESM can be sustained is not known. The negative correlation between PP and sea-surface temperature (Table S4) highlights the effect of enhanced thermal stratification. Enhanced stratification combined with declining DNF in most models suggests that PP (and in some cases DNF) is increasingly limited by the supply of nutrients. In several models the trend is toward greater limitation of DNF by P, especially at BATS (Table S3). At HOT, PP shows an upward trend over fully 25 years of observations (Figure 5; Table 4). Although this trend could result from interdecadal climate variability, its time scale approaches the range of scales for which this explanation is unlikely (e.g., Henson et al., 2010; Christian, 2014; Cummins and Masson, 2014; Rodgers et al., 2015). The PP trend has been associated with DNF, presumed to result from climate variability (Karl et al., 1995; Karl, 1999); more recently, Sherwood et al. (2014) identified enhanced DNF in this region as a long term trend. There are many weaknesses in the CanESM solution and no a priori reason to believe that it is qualitatively ‘better’ than the other models, but the increasing trend in DNF in CanESM represents a potentially large negative feedback on atmospheric CO2 growth and deserves further investigation. It is also important to consider that none of the models address the effect of CO2 on diazotrophy. In a world with much higher CO2,

direct effects of CO2 on DNF may be as large as climate change effects on temperature and nutrient supply. 5 Summary and conclusions Comparisons among models, and between models and observations, show that models are qualitatively consistent with the observations in terms of relative spatial distribution among the Pacific, Atlantic, and Indian Oceans (Table 3). However, the magnitude of DNF and the spatial distribution vary widely among models, and the spatial pattern correlation between modelled and observed DNF rates is weak, and in most cases not statistically significant. The frequency distribution of DNF rates (Figures 3 and 4) shows a positively skewed distribution which the models reproduce to varying degrees. The model distributions are generally narrower and less skewed than appears to be the case in the observations; i.e., the modelled space/time distribution of DNF is less variable and episodic than in the real world. The available observational data base is an impressive achievement, as prior to about 2000 there were almost no such data, yet spatial and temporal coverage is still very incomplete. As a result, these data are insufficient to distinguish among the examined models in terms of their relative skill. In the present generation of climate models, the first order question is the global aggregate rate of DNF, and the spatial distribution is secondary. If the process is as episodic as it appears in this compilation of observations, then modelling the underlying mechanisms correctly is likely to be extremely difficult. Continuing collection of high quality observations is necessary, and should expand further into areas not traditionally considered primary areas of diazotrophy; models can continue to be tested against these observations in a variety of local, regional and global frameworks. Data Accessibility Statement UVic output products are available through Michael Eby at the School of Earth and Ocean Sciences, University of ­Victoria. A subset of these data used for the paper is ­available at ftp://ftp.cccma.ec.gc.ca/pub/jchristian/DNF/. The CMIP5 data are publically available through one of the CMIP web nodes (for instance, the Lawrence Livermore National Laboratory node at http://pcmdi9.llnl.gov), part of the Earth System Grid Federation peer-to-peer enterprise system. The Luo et al. (2012) dataset is publically available through the PANGAEA portal at https://doi.pangaea.de/10.1594/PANGAEA.774851. Supplemental Files The supplemental files for this article can be found as ­follows: • Figure S1. Trends in environmental conditions at HOT and BATS. DOI: https://doi.org/10.1525/elementa.277.s1 • Table S1. Statistics of observed and modelled frequency distributions. DOI: https://doi.org/10.1525/ elementa.277.s2 • Table S2. Spatial correlation among CMIP5 models,

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

and between observations and models. DOI: https:// doi.org/10.1525/elementa.277.s3 • Table S3. Temporal correlation of DNFa rate with ­environmental factors at BATS and HOT in each model. DOI: https://doi.org/10.1525/elementa.277.s4 • Table S4. Temporal correlation of primary ­productiona with environmental factors at BATS and HOT in each model. DOI: https://doi.org/10.1525/ elementa.277.s5 Acknowledgements We thank Jody Deming, Elementa Ocean Science Editorin-Chief, and two anonymous reviewers for their useful input. We acknowledge the World Climate Research Program’s Working Group on Coupled Modelling, which is responsible for CMIP5, and we thank the data originators for producing and making available their model output. We thank Michael Eby for providing output from the UVic Earth System Climate model for our analysis and for authorizing the release of our working subset. We also thank Luo and collaborators for the collection and compilation of observed nitrogen fixation, and the data originators and the PANGAEA Team for making this dataset ­publically available. Funding information This project was supported by the Marine Environmental Observation Prediction and Response (MEOPAR) network, funded through the Network of Centres of Excellence of Canada. Competing interests The authors have no competing interests to declare. Author contributions • Contributed to conception and design: OGJR, JRC • Contributed to acquisition of data: OGJR, JRC • Contributed to analysis and interpretation of data: OGJR, JRC • Drafted and/or revised the article: OGJR, JRC • Approved the submitted version for publication: OGJR, JRC References Anderson, TR, Christian, JR and Flynn, KJ 2015 Modelling DOM biogeochemistry. Biogeochemistry of marine dissolved organic matter. Chapter 15, Hansell, D and Carlson, C (eds.), 2nd ed., 635–667. DOI: https://doi. org/10.1016/B978-0-12-405940-5.00015-7 Arora, VK, Scinocca, JF, Boer, GJ, Christian, JR, Denman, KL, Flato, GM, Kharin, VV, Lee, WG and Merryfield, WJ 2011 Carbon emission limits required to satisfy future representative concentration pathways of greenhouse gases. Geophys Res Lett 38(5). DOI: https://doi. org/10.1029/2010GL046270 Aumont, O, Ethé, C, Tagliabue, A, Bopp, L and Gehlen, M 2015 PISCES-v2: an ocean biogeochemical model for carbon and ecosystem studies. Geosci Model Dev

Art. 16, page 15 of 18

8(2): 1375–1509. DOI: https://doi.org/10.5194/ gmd-8-2465-2015 Bates, NR, Samuels, L and Merlivat, L 2001 Biogeochemical and physical factors influencing ­ seawater ƒCO2 and air-sea CO2 exchange on the Bermuda coral reef. Limnol Oceanogr 46(4): 833–846. DOI: https://doi.org/10.4319/lo.2001.46.4.0833 Bissett, WP, Walsh, JJ, Dieterle, DA and Carder, KL 1999 Carbon cycling in the upper waters of the Sargasso Sea: I. Numerical simulation of differential carbon and nitrogen. Deep-Sea Res PT I 46(2): 205–269. DOI: https://doi.org/10.1016/ S0967-0637(98)00062-4 Blais, M, Tremblay, J-É, Jungblut, AD, Gagnon, J, ­Martin, J, Thaler, M and Lovejoy, C 2012 Nitrogen fixation and identification of potential diazotrophs in the Canadian Arctic. Global Biogeochem Cy 26(3). DOI: https://doi.org/10.1029/2011GB004096 Brix, H, Gruber, N, Karl, DM and Bates, NR 2006 On the relationships between primary, net community, and export production in subtropical gyres. DeepSea Res PT II 53(5–7): 698–717. DOI: https://doi. org/10.1016/j.dsr2.2006.01.024 Cabré, A, Marinov, I, Bernardello, R and Bianchi, D 2015 Oxygen minimum zones in the tropical Pacific across CMIP5 models: mean state differences and climate change trends. Biogeosciences 12: 5429–5454. DOI: https://doi.org/10.5194/ bg-12-5429-2015 Capone, DG, Zehr, JP, Paerl, HW, Bergman, B and ­Carpenter, EJ 1997 Trichodesmium, a globally significant marine cyanobacterium. Science 276(5316): 1221–1229. DOI: https://doi.org/10.1126/ science.276.5316.1221 Chen, YL, Chen, HY and Lin, YH 2003 Distribution and downward flux of Trichodesmium in the South China Sea as influenced by the transport from the ­Kuroshio Current. Mar Ecol Prog Ser 259: 47–57. DOI: https://doi.org/10.3354/meps259047 Chen, YL, Chen, HY, Lin, YH, et al. 2014 The relative contributions of unicellular and filamentous diazotrophs to N2 fixation in the South China Sea and the upstream Kuroshio. Deep-Sea Res PT I 85(2): 56–71. DOI: https://doi.org/10.1016/j. dsr.2013.11.006 Christian, JR 2014 Timing of the departure of ocean biogeochemical cycles from the preindustrial state. PLoS One 9(11). DOI: https://doi.org/10.1371/journal.pone.0109820 Cummins, PF and Masson, D 2014 Climatic variability and trends in the surface waters of coastal British Columbia. Prog Oceanogr 120: 279–290. DOI: https://doi.org/10.1016/j.pocean.2013.10.002 Deutsch, C, Sarmiento, JL, Sigman, DM, Gruber, N and Dunne, JP 2007 Spatial coupling of nitrogen inputs and losses in the ocean. Nature 445(7124): 163–167. DOI: https://doi.org/ 10.1038/nature05392 Diez, B, Bergman, B and El-Shehawy, R 2008 Marine diazotrophic cyanobacteria: out of the blue. Plant

Art. 16, page 16 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Biotechnol 25(3): 221–225. DOI: https://doi. org/10.5511/plantbiotechnology.25.221 Dufresne, J-L, Foujols, M-A, Denvil, S, Caubel, A, Marti, O, Aumont, O, Balkanski, Y, Bekki, S, Bellenger, H, Benshila, R, et al. 2013 Climate change projections using the IPSL-CM5 Earth System Model: from CMIP3 to CMIP5. Clim Dynam 40(9– 10): 2123–2165. DOI: https://doi.org/10.1007/ s00382-012-1636-1 Dunne, JP, John, JG, Shevliakova, E, Stouffer, RJ, Krasting, JP, Malyshev, SL, Milly, PCD, Sentman, LT, Adcroft, AJ, Cooke, W, et al. 2013 GFDL’s ESM2 global coupled climate-carbon earth system models. Part II: carbon system formulation and baseline simulation characteristics. J Climate 26(7): 2247–2267. DOI: https://doi.org/10.1175/JCLI-D-12-00150.1 Dyhrman, ST, Chappell, PD, Haley, ST, Moffett, JW, Orchard, ED, Waterbury, JB and Webb, EA 2006 Phosphonate utilization by the globally important marine diazotroph Trichodesmium. Nature 439(7072): 68–71. DOI: https://doi.org/10.1038/ nature04203 Fernandez, A, Mourino-Carballido, B, Bode, A, Varela, M and Maranon, E 2010 Latitudinal distribution of Trichodesmium spp. and N2 fixation in the Atlantic Ocean. Biogeosciences 7(10): 3167–3176. DOI: https://doi.org/10.5194/bg-7-3167-2010 Fernandez, C, Farías, L and Ulloa, O 2011 Nitrogen fixation in denitrified marine waters. PLoS One 6(6). DOI: https://doi.org/10.1371/journal. pone.0020539 Fernandez, C, González, ML, Muñoz, C, Molina, V and Farias, L 2015 Temporal and spatial variability of biological nitrogen fixation off the upwelling system of central Chile (35–38.5°S). J Geophys Res Oceans 120(5): 3330–3349. DOI: https://doi. org/10.1002/2014JC010410 Field, CB, Behrenfeld, MJ, Randerson, JT and Falkowski, P 1998 Primary production of the biosphere: integrating terrestrial and oceanic components. Science 281(5374): 237–240. DOI: https:// doi.org/10.1126/science.281.5374.237 Fowler, D, Steadman, CE, Stevenson, D, Coyle, M, Rees, RM, Skiba, UM, Sutton, MA, Cape, JN, Dore, AJ, Vieno, M, et al. 2015 Effects of global change during the 21st century on the nitrogen cycle. Atmos Chem Phys 15(24): 13849–13893. DOI: https://doi. org/10.5194/acp-15-13849-2015 Galbraith, ED, Sigman, DM, Robinson, RS and ­Pedersen, TF 2008 Nitrogen in past marine environments. Nitrogen in the marine environment. Chapter 34, Capone, D (ed.), 2nd ed., 1497–1535. DOI: https://doi.org/10.1016/ B978-0-12-372522-6.00034-7 Geider, RJ, Maclntyre, HL and Kana, TM 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. Mar Ecol Prog Ser 148: 187–200. DOI: https://doi.org/10.3354/meps148187

Geider, RJ, Maclntyre, HL and Kana, TM 1998 A dynamic regulatory model of phytoplanktonic acclimation to light, nutrients, and temperature. Limnol Oceanogr 43(4): 679–694. DOI: https://doi.org/10.4319/ lo.1998.43.4.0679 Gradoville, MR, White, AE, Böttjer, D, Church, MJ and Letelier, RM 2014 Diversity trumps acidification: Lack of evidence for carbon dioxide enhancement of Trichodesmium community nitrogen or carbon fixation at Station ALOHA. Limnol Oceanogr 59(3): 645–659. DOI: https://doi.org/10.4319/ lo.2014.59.3.0645 Großkopf, T, Mohr, W, Baustian, T, Schunck, H, Gill, D, Kuypers, MMM, Lavik, G, Schmitz, RA, Wallace, DWR and LaRoche, J 2012 Doubling of marine dinitrogen-fixation rates based on direct measurements. Nature 488(7411): 361–364. DOI: https:// doi.org/10.1038/nature11338 Gruber, N 2011 Warming up, turning sour, losing breath: ocean biogeochemistry under global change. Philos T R Soc B 369(1943): 1980–1996. DOI: https://doi. org/10.1098/rsta.2011.0003 Gruber, N and Sarmiento, JL 1997 Global patterns of marine nitrogen fixation and denitrification. Global Biogeochem Cy 11(2): 235–266. DOI: https://doi. org/10.1029/97GB00077 Henson, SA, Sarmiento, JL, Dunne, JP, Bopp, L, Lima, I, Doney, SC, John, J and Beaulieu, C 2010 Detection of anthropogenic climate change in satellite records of ocean chlorophyll and productivity. ­Biogeosciences 7: 621–640. DOI: https://doi. org/10.5194/bg-7-621-2010 Holl, CM and Montoya, JP 2005 Interactions between nitrate uptake and nitrogen fixation in continuous cultures of the marine diazotroph Trichodesmium (Cyanobacteria). J Phycol 41(6): 1178–1183. DOI: https://doi.org/10.1111/j.1529-8817.2005.00146.x Hood, RR, Bates, NR, Capone, DG and Olson, DB 2001 Modeling the effect of nitrogen fixation on carbon and nitrogen fluxes at BATS. Deep-Sea Res PT II 48(8–9): 1609–1648. DOI: https://doi. org/10.1016/S0967-0645(00)00160-0 Hutchins, DA, Fu, F-X, Webb, EA, Walworth, N and ­Tagliabue, A 2013 Taxon-specific response of marine nitrogen fixers to elevated carbon dioxide concentrations. Nat Geosci 6(9): 790–795. DOI: https://doi.org/10.1038/ngeo1858 Hutchins, DA, Walworth, NG, Webb, EA, Saito, MA, Moran, D, McIlvin, MR, Gale, J and Fu, F-X 2015 Irreversibly increased nitrogen fixation in Trichodesmium experimentally adapted to elevated carbon dioxide. Nat Commun 6. DOI: https://doi. org/10.1038/ncomms9155 Ilyina, T, Six, KD, Segschneider, J, Maier-Reimer, E, Li, H and Núñez-Riboni, I 2013 Global ocean biogeochemistry model HAMOCC: model architecture and performance as component of the MPI-Earth system model in different CMIP5 experimental realizations. J Adv Model Earth Sy 5(2): 287–315. DOI: https:// doi.org/10.1029/2012MS000178

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Karl, D, Letelier, R, Tupas, L, Dore, J, Christian, J and Hebel, D 1997 The role of nitrogen fixation in biogeochemical cycling in the subtropical North Pacific Ocean. Nature 388(6642): 533–538. DOI: https:// doi.org/10.1038/41474 Karl, DM 1999 A sea of change: biogeochemical variability in the North Pacific Subtropical Gyre. Ecosystems 2(3): 181–214. DOI: https://doi.org/10.1007/ s100219900068 Karl, DM, Christian, JR, Dore, JE, Hebel, DV, Letelier, RM, Tupas, LM and Winn, CD 1996 Seasonal and interannual variability in primary production and particle flux at Station ALOHA. ­Deep-Sea Res PT II 43(2): 539–568. DOI: https://doi. org/10.1016/0967-0645(96)00002-1 Karl, DM and Church, MJ 2014 Microbial oceanography and the Hawaii Ocean Time-series programme. Nat Rev Microbiol 12(10): 699–713. DOI: https://doi. org/10.1038/nrmicro3333 Karl, DM, Letelier, R, Hebel, D, Tupas, L, Dore, J, ­Christian, J and Winn, C 1995 Ecosystem changes in the North Pacific subtropical gyre attributed to the 1991–92 El Niño. Nature 373(6511): 230–234. DOI: https://doi.org/10.1038/373230a0 Knapp, AN 2012 The sensitivity of marine N2 fixation to dissolved inorganic nitrogen. Frontiers Microbiol 3. DOI: https://doi.org/10.3389/fmicb.2012.00374 Knapp, AN, Dekaezemacker, J, Bonnet, S, Sohm, JA and Capone, DG 2012 Sensitivity of Trichodesmium erythraeum and Crocosphaera watsonii abundance and N2 fixation rates to varying NO3– and PO43– concentrations in batch cultures. Aquat Microb Ecol 66(3): 223–236. DOI: https://doi.org/10.3354/ame01577 Landolfi, A, Dietze, H, Koeve, W and Oschlies, A 2013 Overlooked runaway feedback in the marine nitrogen cycle: the vicious cycle. Biogeosciences 10: 1351–1363. DOI: https://doi.org/10.5194/bg-10-1351-2013 Landolfi, A, Koeve, W, Dietze, H, Kähler, P and Oschlies, A 2015 A new perspective on environmental controls of marine nitrogen fixation. Geophys Res Lett 42(11). DOI: https://doi. org/10.1002/2015GL063756 Levitan, O, Rosenberg, G, Setlik, I, Setlikova, E, Grigel, J, Klepetar, J, Prasil, O and Berman-Frank, I 2007 Elevated CO2 enhances nitrogen fixation and growth in the marine cyanobacterium Trichodesmium. Glob Change Biol 13(2): 531–538. DOI: https://doi. org/10.1111/j.1365-2486.2006.01314.x Lomas, MW, Bates, NR, Knap, AH, Karl, DM, Lukas, R, Landry, MR, Bidigare, RR, Steinberg, DK and Carlson, CA 2002 Refining our understanding of oceanic biogeochemistry and ecosystem functioning. Eos Trans Am Geophys Union 83(48): 559–567. DOI: https://doi.org/10.1029/2002EO000386 Luo, Y-W, Doney, SC, Anderson, LA, Benavides, M, Berman-Frank, I, Bode, A, Bonnet, S, Boström, KH, Böttjer, D, Capone, DG, et al. 2012 Database of diazotrophs in global ocean: abundance, biomass and nitrogen fixation rates. Earth Syst Sci Data 4(1): 47–73. DOI: https://doi.org/10.5194/essd-4-47-2012

Art. 16, page 17 of 18

Luo, Y-W, Lima, ID, Karl, DM, Deutsch, CA and Doney, SC 2014 Data-based assessment of ­environmental controls on global marine nitrogen fixation. ­Biogeosciences 11(3): 691–708. DOI: https://doi. org/10.5194/bg-11-691-2014 Matear, RJ, Wang, Y-P and Lenton, A 2010 Land and ocean nutrient and carbon cycle interactions. Curr Opin Environ Sustain 2(4): 258–263. DOI: https:// doi.org/10.1016/j.cosust.2010.05.009 Montoya, JP, Voss, M, Kahler, P and Capone, DG 1996 A Simple, high-Precision, high-sensitivity tracer assay for N2 fixation. Appl Environ Microb 62(3): 986–993. Moore, JK and Doney, SC 2007 Iron availability limits the ocean nitrogen inventory stabilizing feedbacks between marine denitrification and nitrogen fixation. Global Biogeochem Cy 21(2). DOI: https://doi. org/10.1029/2006GB002762 Moore, JK, Doney, SC, Kleypas, JA, Glover, DM and Fung, IY 2002 An intermediate complexity marine ecosystem model for the global domain. ­Deep-Sea Res PT II 49(1–3): 403–462. DOI: https://doi. org/10.1016/S0967-0645(01)00108-4 Moore, JK, Doney, SC and Lindsay, K 2004 Upper ocean ecosystem dynamics and iron cycling in a global three-dimensional model: Global ecosystem-biogeochemical model. Global Biogeochem Cy 18(4). DOI: https://doi.org/10.1029/2004GB002220 Moore, JK, Lindsay, K, Doney, SC, Long, MC and Misumi, K 2013 Marine ecosystem dynamics and biogeochemical cycling in the Community Earth System Model [CESM1(BGC)]: Comparison of the 1990s with the 2090s under the RCP4.5 and RCP8.5 ­Scenarios. J Climate 26(23): 9291–9312. DOI: https://doi.org/10.1175/JCLI-D-12-00566.1 Moss, RH, Edmonds, JA, Hibbard, KA, Manning, MR, Rose, SK, van Vuuren, DP, Carter, TR, Emori, S, Kainuma, M, Kram, T, et al. 2010 The next generation of scenarios for climate change research and assessment. Nature 463(7282): 747–756. DOI: https://doi.org/10.1038/nature08823 Naqvi, SAW 2008 The Indian Ocean. Nitrogen in the marine environment. Chapter 14, Capone, D (ed.), 2nd ed., 631–681. DOI: https://doi.org/10.1016/ B978-0-12-372522-6.00014.1 Rodgers, KB, Lin, J and Frölicher, TL 2015 Emergence of multiple ocean ecosystem drivers in a large ensemble suite with an Earth system model. Biogeosciences 12(11): 3301–3320. DOI: https://doi.org/10.5194/ bg-12-3301-2015 Roxy, MK, Modi, A, Murtugudde, R, Valsala, V, ­Panickal, S, Prasanna Kumar, S, Ravichandran, M, Vichi, M and Lévy, M 2016 A reduction in marine primary productivity driven by rapid warming over the tropical Indian Ocean. Geophys Res Lett 43(2). DOI: https://doi.org/10.1002/2015GL066979 Schmittner, A, Oschlies, A, Giraud, X, Eby, M and Simmons, HL 2005 A global model of the marine ecosystem for long-term simulations: ­Sensitivity to ocean mixing, buoyancy forcing, particle sinking, and dissolved organic matter cycling.

Art. 16, page 18 of 18

Riche and Christian: Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming

Global Biogeochem Cy 19(3). DOI: https://doi. org/10.1029/2004GB002283 Schmittner, A, Oschlies, A, Matthews, HD and ­Galbraith, ED 2008 Future changes in ­climate, ocean circulation, ecosystems, and biogeochemical cycling simulated for a business-asusual CO2 emission scenario until year 4000 AD. Global Biogeochem Cy 22. DOI: https://doi. org/10.1029/2007GB002953 Séférian, R, Bopp, L, Gehlen, M, Orr, JC, Ethe, C, ­Cadule, P, Aumont, O, Melia, DSY, Voldoire, A and Madec, G 2013 Skill assessment of three earth ­system m ­ odels with common marine biogeochemistry. Clim Dynam 40(9–10): 2549–2573. DOI: https://doi.org/10.1007/s00382-012-1362-8 Sherwood, OA, Guilderson, TP, Batista, FC, Schiff, JT and McCarthy, MD 2014 Increasing subtropical North Pacific Ocean nitrogen fixation since the Little Ice Age. Nature 505(7481): 78–81. DOI: https:// doi.org/10.1038/nature12784 Singh, A, Lomas, MW and Bates, NR 2013 Revisiting N2 fixation in the North Atlantic Ocean: significance of deviations from the Redfield ratio, atmospheric deposition and climate variability. Deep-Sea Res PT II 93: 148–158. DOI: https://doi.org/10.1016/j. dsr2.2013.04.008 Sohm, JA, Webb, EA and Capone, DG 2011 Emerging patterns of marine nitrogen fixation. Nat Rev Microbiol 9(7): 499–508. DOI: https://doi.org/10.1038/ nrmicro2594 Steinacher, M, Joos, F, Frölicher, TL, Bopp, L, C ­ adule, P, Cocco, V, Doney, SC, Gehlen, M, Lindsay, K, Moore, JK, et al. 2010 Projected 21st century decrease in marine productivity: a multi-model analysis. Biogeosciences 7(3): 979–1005. DOI: https:// doi.org/10.5194/bg-7-979-2010 Taylor, KE, Stouffer, RJ and Meehl, GA 2012 An Overview of CMIP5 and the experiment design. B Am Meteorol Soc 93(4): 485–498. DOI: https://doi.org/10.1175/ BAMS-D-11-00094.1 Tyrrell, T 1999 The relative influences of nitrogen and phosphorus on oceanic primary production. Nature 400(6744): 525–531. DOI: https://doi. org/10.1038/22941

Voss, M, Bange, HW, Dippner, JW, Middelburg, JJ, Montoya, JP and Ward, B 2013 The marine ­nitrogen cycle: recent discoveries, uncertainties and the potential relevance of climate change. Philos T R Soc B 368(1621). DOI: https://doi.org/10.1098/ rstb.2013.0121 Weaver, AJ, Eby, M, Wiebe, EC, Bitz, CM, Duffy, PB, Ewen, TL, Fanning, AF, Holland, MM, MacFadyen, A, Matthews, HD, et al. 2001 The UVic Earth S­ ystem Climate Model: model description, climatology, and applications to past, present and future climates. Atmos Ocean 39(4): 361–428. DOI: https://doi.org/ 10.1080/07055900.2001.9649686 Xu, L, Xie, S-P and Liu, Q 2012 Mode water ventilation and subtropical countercurrent over the North Pacific in CMIP5 simulations and future projections. J Geophys Res Oceans 117(C12). DOI: https://doi. org/10.1029/2012JC008377 Yasunaka, S, Ono, T, Nojiri, Y, Whitney, FA, Wada, C, Murata, A, Nakaoka, S and Hosoda, S 2016 ­ Long-term variability of surface n ­utrient concentrations in the North Pacific. Geophys Res Lett 43(7): 3389–3397. DOI: https://doi. org/10.1002/2016GL068097 Zahariev, K, Christian, JR and Denman, KL 2008 Preindustrial, historical, and fertilization simulations using a global ocean carbon model with new p ­ arameterizations of iron limitation, calcification, and N2 fixation. Progr Oceanogr 77(1): 56–82. DOI: https://doi.org/10.1016/j. pocean.2008.01.007 Zehr, JP 2011 Nitrogen fixation by marine cyanobacteria. Trends Microbiol 19(4): 162–173. DOI: https://doi. org/10.1016/j.tim.2010.12.004 Zehr, JP and Bombar, D 2015 Marine nitrogen fixation: organisms, significance, enigmas, and future directions. In: de Bruijn, FJ (ed.), Biological Nitrogen Fixation, 855–872. John Wiley & Sons, Inc. DOI: https:// doi.org/10.1002/9781119053095.ch84 Zehr, JP and Ward, BB 2002 Nitrogen cycling in the ocean: new perspectives on processes and paradigms. Appl Environ Microb 68(3): 1015–1024. DOI: https://doi.org/10. 1128/AEM.68.3.1015-1024.2002

How to cite this article: Riche, OGJ and Christian, JR 2018 Ocean dinitrogen fixation and its potential effects on ocean primary production in Earth system model simulations of anthropogenic warming. Elem Sci Anth, 6: 16. DOI: https://doi.org/10.1525/ elementa.277 Domain Editor-in-Chief: Jody W. Deming, University of Washington, US Knowledge Domain: Ocean Science Submitted: 15 February 2017

Accepted: 07 January 2018

Published: 15 February 2018

Copyright: © 2018 The Author(s). This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International License (CC-BY 4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. See http://creativecommons.org/licenses/by/4.0/. Elem Sci Anth is a peer-reviewed open access journal published by University of California Press.

OPEN ACCESS