Skip to main content
Philosophical Transactions of the Royal Society B: Biological Sciences logoLink to Philosophical Transactions of the Royal Society B: Biological Sciences
. 2012 May 5;367(1593):1245–1255. doi: 10.1098/rstb.2011.0360

Global oceanic production of nitrous oxide

Alina Freing 1,, Douglas W R Wallace 1,, Hermann W Bange 1,*
PMCID: PMC3306629  PMID: 22451110

Abstract

We use transient time distributions calculated from tracer data together with in situ measurements of nitrous oxide (N2O) to estimate the concentration of biologically produced N2O and N2O production rates in the ocean on a global scale. Our approach to estimate the N2O production rates integrates the effects of potentially varying production and decomposition mechanisms along the transport path of a water mass. We estimate that the oceanic N2O production is dominated by nitrification with a contribution of only approximately 7 per cent by denitrification. This indicates that previously used approaches have overestimated the contribution by denitrification. Shelf areas may account for only a negligible fraction of the global production; however, estuarine sources and coastal upwelling of N2O are not taken into account in our study. The largest amount of subsurface N2O is produced in the upper 500 m of the water column. The estimated global annual subsurface N2O production ranges from 3.1 ± 0.9 to 3.4 ± 0.9 Tg N yr−1. This is in agreement with estimates of the global N2O emissions to the atmosphere and indicates that a N2O source in the mixed layer is unlikely. The potential future development of the oceanic N2O source in view of the ongoing changes of the ocean environment (deoxygenation, warming, eutrophication and acidification) is discussed.

Keywords: nitrous oxide, oceanic production, nitrification, denitrification

1. Introduction

Nitrous oxide (N2O) is an atmospheric trace gas which influences the Earth's climate both directly and indirectly [1,2]: (i) In the troposphere, it acts as a strong greenhouse gas and (ii) owning to a relatively long atmospheric lifetime, N2O can reach up to the stratosphere, where it acts as the major source for ozone-depleting nitric oxide radicals. Since the industrial revolution, the concentration of N2O in the atmosphere has increased rapidly by about 18 per cent [3]. The dominant natural sources are believed to be soils and oceans, whereas anthropogenic sources mostly result from agricultural and industrial activities. N2O is biologically produced in the ocean, and the resulting N2O emissions play a major role for the atmospheric N2O budget [4,5]. According to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, open ocean and coastal areas make up approximately 21 per cent and approximately 10 per cent of the total source of atmospheric N2O of 17.7 Tg N yr–1, respectively [5].

N2O is microbially produced via nitrification (NH4+ → NO2 → NO3) and denitrification (NO3 → NO2 → N2O → N2). They are widely accepted to be the main production mechanisms of N2O in the ocean [6]. N2O can occur as a by-product during nitrification and as an intermediate during denitrification. However, the exact metabolism used for N2O production during nitrification, the net behaviour of denitrification and their respective contributions to the global N2O inventory in the ocean remain unclear [69].

Here, we present an estimate of the N2O subsurface production rates by using the transient time distribution (TTD) approach. The discrepancy between the predicted global N2O production and ocean–atmosphere flux estimates and its potential reasons as well as implications are discussed.

2. Methods

In Freing et al. [10], TTDs calculated from tracer data together with in situ measurements of N2O were used to estimate the concentration of biologically produced N2O and N2O production rates in the central North Atlantic Ocean. This approach to estimate N2O production rates integrates the effects of potentially varying production/decomposition mechanisms along the transport path of a water mass. A new parametrization of N2O production during nitrification depending linearly on the apparent oxygen utilization (AOU) and exponentially on temperature/depth was developed by Freing et al. [10], which is used here together with global gridded World Ocean Atlas (WOA) temperature [11], salinity [12] and oxygen [13] data and global gridded CFC-12 (i.e. dichlorodifluoromethane, CF2Cl2) data from the Global Data Analysis Project (GLODAP) [14] to calculate the global annual subsurface N2O yield. We calculated a TTD age distribution and a mean age for every grid point. Because CFC-12 concentrations in the deep Pacific are too low to allow calculation of TTDs and the highest calculated mean age is 731 years, the mean age for every grid point with such a low CFC-12 concentration was set to 800 years. This choice results in a realistically smooth age distribution and is in line with the results of Matsumoto [15], who, using 14C, found the deep ocean to be characterized by centennial rather than millennial timescales.

In order to estimate the background N2O signal (N2Oeq) in the ocean from the TTD, we need to know the time-dependent history of atmospheric concentrations. We used a synthesis of ice-core and firn data [16,17] merged with more recent air measurements. The data are available at http://daac.ornl.gov [18]. We assume a constant atmospheric mixing ratio of N2O of 275 ppb prior to year 1800 and used a polynomial fit of the data from 1800 to the present day. For details on the atmospheric history of N2O used and the use of TTDs to calculate [N2O]eq, see Freing et al. [10].

(a). Open ocean apparent nitrous oxide production rate

The bottom depth at each grid point was determined by using the 60 s resolution Earth Topography Digital Dataset (ETOPO) global elevation map distributed with the Ferret program (information is available at http://ferret.pmel.noaa.gov/Ferret/). A grid point was defined as an open ocean grid point if its bottom depth was more than 200 m. An apparent N2O production rate for open ocean grid points was calculated in two ways (N2OPRdepth and N2OPRtemp) using the parametrizations developed by Freing et al. [10]. An apparent N2O production rate depending on an apparent oxygen utilization rate (AOUR) and depth was calculated as

(a). 2.1

where the coefficient values for the best fit are a1 = 0.0658, a2 = −0.0065, zsc = 20 000, z denotes the depth and AOUR = [AOU]/t, where t is the mean age of the water parcel calculated using the TTD method. An apparent N2O production rate depending on AOUR and temperature was calculated as

(a). 2.2

where the coefficient values for the best fit are a1 = 0.0665, a2 = −0.0032, Tsc = 20 000, T denotes the temperature and AOUR = [AOU]/t, where t is the mean age of the water parcel calculated using the TTD method.

To exclude any undue influence of short-term seasonal variations affecting the near-surface ocean in the estimates, all production rates were calculated only for grid fields below the mixed layer. The mixed layer depth was determined by using the mixed layer climatology of de Boyer Montégut et al. [19]. The mixed layer depth is determined by an absolute change in temperature of 0.2°C compared with the temperature at 10 m depth. This parametrization accounts only for N2O production by nitrification.

There are no CFC-12 data, and thus there is no mean age information available for the Arctic Ocean in the GLODAP dataset. A N2OPR for the Arctic Ocean (N2OPRarctic) was estimated using the 25 per cent-quantile of the overall N2OPR. This takes into account that, in general, the relevant biological productivity is comparatively low in the Arctic Ocean.

(b). Apparent nitrous oxide production rate on the continental shelves

A grid point was defined as a shelf grid point if its bottom depth was 200 m or less. An apparent shelf N2O production rate (N2OPRshelf) was calculated using 52 shelf data points from the N2O database MEMENTO (version March 2009; see https://memento.ifm-geomar.de/) [20]. Unlike for N2OPR, there is a good correlation between N2OPRshelf/AOUR and depth but not between N2OPRshelf/AOUR and temperature. Therefore, N2OPRshelf was estimated as

(b). 2.3

where the coefficient values for the best fit are a1 = 0.1795, a2 = 0.5374, zsc = 350 and z denotes the depth. This parametrization accounts only for N2O production by nitrification. The goodness of fit is illustrated in figure 1. Owing to lack of data, we did not attempt to calculate a production rate for the shelf of the Arctic Ocean.

Figure 1.

Figure 1.

Shelf nitrous oxide (N2O) production rates (nmol kg–1 yr–1) calculated from N2O data versus shelf N2O production rates (nmol kg–1 yr–1) calculated using equation (2.3).

(c). Denitrification

Denitrification can either produce or consume N2O, depending on the surrounding conditions: it produces N2O at the interface between suboxic and anoxic waters, while it consumes N2O under (close to) anoxic conditions when the complete process of denitrification is performed [6,21].

To estimate production owing to denitrification, we used the collection of N2O depth profiles archived in MEMENTO (version March 2009; figure 2) to determine the maximal and mean N2O (Inline graphic) concentration per depth and oxygen minimum zone. Following Codispoti et al. [22] and Stramma et al. [23], we defined suboxic zones as WOA fields with an oxygen content of less than 10 μmol kg−1. This results in three main oxygen minimum zones, in the Arabian Sea, the Eastern Tropical North Pacific and the Eastern Tropical South Pacific, which is in good agreement with Paulmier et al. [24]. The difference between Inline graphic as well as the difference between Inline graphic and the estimated [N2O] (= N2OPR × t + [N2O]eq]) resulting from nitrification were used as an estimate of the total amount of N2O produced by denitrification. When (Inline graphic) resulted in no N2O production via denitrification, Inline graphic and Inline graphic were used as a lower and upper boundary, respectively, for the total amount of N2O produced via denitrification. The N2O production rate owing to denitrification (N2OPRdenit) was then calculated as

(c). 2.4

Figure 2.

Figure 2.

Available nitrous oxide depth profiles extracted from the MEMENTO database, version March 2009 [20].

where tomz is the average time a water parcel has already spent in the respective oxygen minimum zone [2527] (table 1).

Table 1.

Average time of residence (years) per oxygen minimum zone (see text for details and references).

zone average oxygen age
Arabian Sea  5
ETNPa 60
ETSPa  5

aETNP and ETSP stand for eastern tropical North Pacific and eastern tropical South Pacific, respectively.

To estimate consumption owing to denitrification, we used the same collection of N2O depth profiles (figure 2) to determine the minimal and mean N2O (Inline graphic) concentration per depth and oxygen minimum zone. We used an oxygen content of 4 μmol kg−1 as a threshold for an environment supporting a denitrification sink of N2O, which marks out parts of the oxygen minimum zones described above. The difference between the estimated [N2O] resulting from nitrification and Inline graphic, and the difference between Inline graphic were used as an estimate of the total amount of N2O consumed by denitrification. When ([N2O]–Inline graphic) resulted in no N2O consumption via denitrification, Inline graphic were used as an upper and lower boundary, respectively, for the total amount of N2O consumed by denitrification. The N2O consumption rate owing to denitrification (N2OCRdenit) was then calculated as

(c). 2.5

Concentration differences in equation (2.5) indicating production were discarded for the purpose of this calculation.

(d). Integrated production estimates

To estimate annual production rates per square metre, the respective N2OPR was integrated over the water column. To estimate the magnitude of the global annual production of N2O, firstly the volume of each grid cell was determined by integrating over the three dimensions: latitude, longitude and depth. The calculated volume was then multiplied by the respective N2OPR to calculate the annual N2O yield per grid cell. The sum of the annual N2O yield of all ocean grid cells was used to estimate the global annual N2O production. The integrated volume of a grid cell Vgc in cubic metres used in our calculations can be described as

(d). 2.6

where φ1, φ2 and Θ1, Θ2, with Θ2 > Θ1, are the longitudinal and latitudinal boundaries, respectively, of the respective grid cell in degrees, D1, D2, with D2 > D1, are the depth boundaries of the respective grid cell in metres and Rearth is the radius of the Earth in metres.

(e). Nitrous oxide concentration

[N2O] was predicted as

(e). 2.7

where f and g are switch functions, which determine whether denitrification contributes to [N2O]. They are defined as

(e). 2.8

and

(e). 2.9

N2OPR is calculated as N2OPRdepth or N2OPRtemp according to equation (2.1) or (2.2), respectively.

(f). Error estimates

The predicted AOUR varied between 0.01 and 49 μmol kg–1 yr–1, with an overall mean AOUR of 2.9 ± 0.5 μmol kg–1 yr–1 (1σ). These values are in good agreement with measurements [28,29]. This suggests that the use of a TTD-based mean age for rate calculations yields reliable rate estimates. Additionally, Tanhua et al. [30] found TTD-based estimates of anthropogenic carbon concentrations to be in good agreement with independent methods, suggesting the use of CFC-12 as tracers is a reliable basis for water mass age and hence estimation of background concentrations such as [N2O]eq.

In figure 3, the global TTD mean age in different water depths is shown. The TTD-derived mean ages presented here are in agreement with measurements of other tracers and model results [31,32].

Figure 3.

Figure 3.

Global transient time distribution mean age (in years) in (a) 75 m, (b) 250 m and (c) 700 m.

The CFC-12, AOU and temperature measurements originate neither from the same water mass nor necessarily even from roughly the same time. Additionally, they represent some kind of average over latitude, longitude and depth. This is likely to dominate the error in our estimates. To estimate this error, we used 5900 open ocean data points below the mixed layer and 46 shelf data points below the mixed layer from our database. Using the gridded CFC concentrations from GLODAP, we calculated the difference between the original [N2O] and [N2O]est with

(f). 2.10

where t is the mean age. The mean and median of the absolute percental differences can be found in table 2. It seems reasonable to use the median as a few very different data points can unduly influence the mean. It is likely that peculiar local conditions can lead to a considerable misfit but this does not adequately represent the overall goodness of fit. As the mean age is crucial for both terms in equation (2.10), we assume that both terms contribute to the overall percental error in the same way. Table 2 suggests that the parametrization based on temperature (see equation (2.2)) seems to do a slightly better job. The amount of N2O produced (consumed) by denitrification is estimated in a completely different fashion compared to nitrification rates (see §2c). Using the mean [N2O] from our database, the fraction of denitrification with regard to total annual production becomes insignificant (table 3). We therefore assume that the rate calculations based on Inline graphic, Inline graphic and Inline graphic (cf. §2c for details) represent reasonable lower and upper boundaries.

Table 2.

Average percental error in concentration estimates (see text for details).

parameter median (%) mean (%)
N2OPRdepth 24 34
N2OPRtemp 22 34
N2OPRshelf 42 54

Table 3.

Constituents of the annual production of nitrous oxide (N2O) below the mixed layer in Gmol N2O yr–1. The annual subsurface N2O yield in Tg N yr–1 is given in parentheses.

parameter depth temperature
N2OPR 101.6 ± 24.4 113.2 ± 24.9
N2OPRdenit 6.9 ± 6.9 6.8 ± 6.8
N2OCRdenit <0.1 ± <0.1 <0.1 ± <0.1
N2OPRarctic 1.2 ± 0.3 1.3 ± 0.3
N2OPRshelf 0.5 ± 0.2 0.5 ± 0.2
sum 110.1 ± 31.8 (3.1 ± 0.9) 121.7 ± 32.2 (3.4 ± 0.9)

3. Results and discussion

(a). Concentration and rate estimates

N2OPR varies between 0 and 3.3 nmol kg−1 yr−1 with an overall mean N2OPR of 0.2 ± 0.04 nmol kg−1 yr−1 (1σ). Between 100 and 500 m depth, N2OPR averages 0.4 ± 0.04 nmol kg−1 yr−1 (1σ), whereas below 500 m N2OPR averages 0.09 ± 0.01 nmol kg−1 yr−1 (1σ), indicating that the largest concentration of subsurface N2O is produced in the upper 500 m of the water column. The estimates of the two parametrizations differ only in the second significant figure.

The global distribution of predicted [N2O] varies between 4 and 31 nmol kg−1 and, where measurement data exist, is qualitatively very similar to the measured [N2O] of our database. Figure 4 shows a comparison of measured and predicted [N2O] along 110°W. In the oligotrophic part, the predicted concentrations agree quite well with the measured N2O; in the suboxic zone, the predicted [N2O] clearly underestimates [N2O]. This could be due to the use of annual mean dissolved [O2], which might not adequately reflect the location and extent of the oxygen minimum zone at the time of the data acquisition. Our parametrization does not account for the export of [N2O]xs from highly productive regions, which might also be a reason for the underestimation shown in figure 4. Qualitative patterns are represented very well by our prediction.

Figure 4.

Figure 4.

Section along 110° W (averaged between 100° W and 120° W). (a) Predicted [N2O] (nmol kg–1) according to N2OPRdepth. (b) Predicted [N2O] (nmol kg–1) according to N2OPRtemp. (c) Measured [N2O] (nmol kg–1) from Nevison et al. [33] and Farías et al. [34].

Figure 5 shows the estimated [N2O] distribution in 200 m depth. Concentrations exceeding 40 nmol kg−1, which are only found in the Arabian Sea, are masked out in figure 5 so as not to obscure the overall qualitative patterns. The eastern boundary upwelling systems are clearly displayed. Concentrations in the Antarctic are slightly elevated, owing to the cold temperatures increasing solubility and thus elevated [N2O]eq. Highest concentrations, 90 nmol kg−1, occur in the Arabian Sea. Concentrations are also comparatively high in the North Pacific Ocean and the Bay of Bengal. Considering the parametrization was solely derived from North Atlantic data, the systematic features of [N2O] in the world ocean are remarkably well displayed.

Figure 5.

Figure 5.

Global [N2O] (nmol kg–1) distribution in 200 m depth estimated using equations (2.7) with N2OPRdepth (a) and equation (2.7) using N2OPRtemp (b). (a,b) White areas in the Arabian Sea represent concentrations exceeding 40 nmol kg–1. Annual N2O production (μmol m–2 yr–1) via nitrification integrated over the water column estimated using (c) equation (2.1) and (d) equation (2.2).

[N2O] is almost always exclusively determined by production via nitrification and [N2O]eq, with the exception of the three oxygen minimum zones. In the oxygen minimum zone in the Arabian Sea, the fraction of N2O produced via denitrification varies between 2 per cent and 89 per cent between 200 and 900 m. In the oxygen minimum zone of the Eastern Tropical North Pacific, the average fraction of N2O produced via denitrification per depth varies between 2 per cent and 35 per cent between 400 and 500 m. In the oxygen minimum zone of the Eastern Tropical South Pacific, the average fraction of N2O produced via denitrification per depth varies between 1 per cent and 14 per cent between 150 and 400 m.

It is worth noting that the qualitative features of [N2O] are controlled not only by the qualitative features of N2OPR but also by temperature (via solubility). [N2O]eq shows patterns very similar to those displayed by [N2O]. We find that [N2O]eq is strongly controlled by temperature. However, the qualitative features of N2OPR mainly reflect AOU/t and not temperature.

(b). Annual subsurface nitrous oxide yield

The integrated annual N2O production per square metre owing to nitrification is shown in figure 5 and listed in table 3. The qualitative features are similar for both parametrizations. N2OPRtemp estimates a slightly bigger N2O yield but both parametrizations agree within the error margins. The error analysis suggests that N2OPRtemp estimates the production slightly more accurately (cf. §2f; table 2).

The largest amount of N2O per square metre and per year is produced in the North Atlantic Ocean and off Argentina, which is more likely due to a potential data mismatch between the CFC12- and AOU-datasets in the deep rather than to truly elevated production. Profile data from the North Atlantic [35] suggest there is no significant N2O production this deep. However, assuming even that true North Atlantic subsurface production amounts only to half of what is predicted by our method would change the global production estimate only by approximately 5 per cent. Production is also very high in the upwelling regions off Mauritania and Chile, and is elevated in the North Pacific Ocean, the Southern Ocean and parts of the Indian Ocean. However, even if the N2O yield per square metre is relatively high in the Southern Ocean, then its overall contribution to the global annual production of N2O is small owing to the small area covered. Overall, the patterns of N2OPR in the deep ocean—and therefore, the patterns in the mean age in our case—govern the N2O yield of the water column. Larger water depth does, however, also lead to a larger N2O yield per square metre without N2OPR being significantly higher for the involved grid cells.

It should be noted that areas of extensive subsurface production must not necessarily coincide with areas of large sea-to-air fluxes because of the effects of advection. In addition, the yield per square metre does not necessarily illustrate the respective contribution of the different oceanic regions to the overall N2O source very well, as a high yield per square metre does not necessarily equal a large regional contribution, depending on the area of the respective region.

As denitrification is relevant only in suboxic environments, it has an impact only on the N2O yield of the suboxic zones in the Arabian Sea and in the Eastern Tropical Pacific. The production yield via denitrification is badly constrained (cf. table 3) and probably also highly variable, but—despite the locally limited influence of denitrification—it yields on average approximately 7 per cent of the total amount of N2O produced via nitrification (cf. table 3). However, these results are to be viewed with caution as the estimation of N2OPRdenit inherently depends on the predicted N2OPR. Nevertheless, our estimate is considerably lower than previous estimates of the global contribution of denitrification that range from 25 to 50 per cent [7,9].

Shelf areas account only for less than 0.5 per cent of the subsurface production of N2O (table 3). This is in line with the conclusions of Bange [36] and Barnes & Upstill-Goddard [37], who suggested that N2O production in coastal areas, which are not affected by upwelling, is found only in estuaries and river plumes but not on the open shelf. A large fraction of estuarine N2O comes from sedimentary sources (most probably denitrification). N2O production from estuaries is, therefore, an additional source of N2O; however, this is not taken into account in this study.

The overall global annual subsurface production of N2O amounts to 3.1 ± 0.9 Tg N yr−1 (110.1 ± 31.8 Gmol N2O yr−1) and 3.4 ± 0.9 Tg N yr−1 (121.7 ± 32.2 Gmol N2O yr−1), respectively, and is detailed in table 3. Our N2O production estimate is in good agreement with the estimate of 3.9 Tg N yr–1 by Suntharalingam & Sarmiento [38], but it is lower than the estimate of 5.8 ± 2 Tg N yr−1 by Nevison et al. [33].

Recent estimates based on gas-exchange parametrizations, compiled by Bange [4], estimate an oceanic source between 1.4 and 14 Tg N yr−1 with a mean oceanic source of 6.6 ± 3.6 Tg N yr−1 (1σ) based on gas-exchange parametrization, surface measurements and models. Rhee et al. [39] recently estimated global N2O emissions of 0.9–1.7 Tg N yr−1 based on extrapolation of measurements in the open Atlantic Ocean. The annual production of N2O in the ocean is obviously an upper boundary for the annual ocean atmosphere flux. Our production estimate is in agreement with the global emission estimates, especially in view of the uncertainties associated with the emission estimates.

4. Summary

  • — The predicted N2O production rates owing to nitrification according to both our suggested parametrizations, N2OPRtemp and N2OPRdepth, show the same qualitative features. However, overall N2OPRtemp estimates slightly higher production than N2OPRdepth. N2OPR in the deep ocean is comparatively uniform. The largest amount of subsurface N2O is produced in the upper 500 m of the water column.

  • — The predicted [N2O] is qualitatively very similar to measured [N2O]. Concentrations in the Antarctic Ocean are slightly elevated, whereas highest concentrations occur in the northeastern Pacific Ocean. Temperature is an important control of qualitative features of [N2O] due to its effect on solubility.

  • — Our estimates of N2O production via denitrification are badly constrained, but despite its locally limited influence, the production yield via denitrification on average amounts to approximately 7 per cent of the total amount of N2O produced via nitrification.

  • — Shelf areas may account for only less than 0.5 per cent of the global production of oceanic N2O; however, potential coastal sources of N2O from estuaries and upwelling areas are not taken into account in our study.

  • — The annual subsurface N2O yield amounts to between 3.1 ± 0.9 (N2OPRdepth) and 3.4 ± 0.9 Tg N yr−1 (N2OPRtemp), respectively. The annual yield of N2OPRtemp is generally slightly larger than that of N2OPRdepth. The error analysis suggests that N2OPRtemp estimates the production slightly more accurately.

  • — Our estimate of the total annual N2O subsurface yield is in agreement with of recent estimates of the global N2O emissions based on gas-exchange parametrization.

In summary, our findings emphasize that the marine N2O cycling, production and emissions are still not well understood and still associated with a high degree of uncertainty.

5. Outlook

(a). Nitrous oxide source in the mixed layer

Because the mixed layer is well oxygenated, a significant N2O source in the mixed layer originating from denitrification processes seems unlikely. Until recently, nitrification was thought to be inhibited by light [40]—making it a very unlikely occurrence in the mixed layer. The results of Bange [41], who concluded that the surface layer N2O concentration in the Arabian Sea is mainly controlled by gas exchange, entrainment of N2O from deeper layers and variability in the sea surface temperature, are in line with this. It should be noted, however, that assuming a uniform mixed layer depth of 50 m and a ventilation time of N2O of three weeks [42], a mixed layer source of 1 Tg N yr−1 would lead only to an accumulation of approximately 0.09 nmol kg−1 in the mixed layer. Assuming a uniform mixed layer depth of 20 m and a mixed layer source of 1 Tg N yr−1 would lead only to an accumulation of approximately 0.2 nmol kg−1 in the mixed layer. These changes could probably not be distinguished from natural variability in the surface, even if the method's precision allowed for their detection.

Model results by Yool et al. [43], based on nitrification measurements, suggest something very different. They suggest a significant nitrification activity in the euphotic zone. Clark et al. [44] measured NH4+ and NO2 oxidation rates on a north–south transect through the Atlantic Ocean. Their data suggest that in the oligotrophic Atlantic Ocean there is nitrification in the photic zone, which is of sufficient intensity to turn over the NO3-pool in one day. Wankel et al. [45] found that 17–25% of the nitrate-based productivity in the euphotic zone of Monterey Bay is supported by nitrification. Bianchi et al. [46] measured significant nitrification rates in the upper 100 m of the Indian sector of the Southern Ocean. The results of the studies introduced above indicate that there might be indeed significant nitrification activity in the euphotic zone on a global scale. However, there only a few studies that deal with the associated N2O yield by nitrification in the mixed layer: Dore and Karl [47] used in situ measurements of [N2O] at the ALOHA (A Long-Term Oligotrophic Habitat Assessment) station and the gas-exchange model of Wanninkhof [48] to calculate N2O ocean–atmosphere fluxes. They calculated the flux to the euphotic zone using concentration gradients and an eddy-diffusivity coefficient of 3.7 × 10−5 m2 s−1. They used these flux estimates to calculate a net N2O production rate in the euphotic zone of 1.68–7.94 μmol m−2 d−1, which they attribute to in situ nitrification. Assuming a N2O yield of 0.5 per cent during nitrification, nitrification estimates derived from the N2O production rate were on the same order of magnitude as their directly measured nitrification rates. Even taking into account the probably rather large uncertainties involved in all the calculations, these results clearly point towards a significant N2O production via nitrification in the euphotic zone. Slightly larger (super-)saturations of N2O in the upper 40 m of the water column might even have suggested near-surface production. Morell et al. [49], Charpentier et al. [50] and Law & Ling [51] calculated the air–sea N2O flux based on their data from the Atlantic Ocean/Caribbean Sea, the central and eastern South Pacific Ocean and the Australasian sector of the Southern Ocean, respectively. They also found a difference between cross-thermocline and air–sea N2O fluxes in the range of Dore & Karl [47]. They acknowledge that this difference might possibly be due to nitrification [49,51], as indicated by experimental evidence of Dore & Karl [47] and Dore et al. [52], or suggested a further hitherto unknown production process [50]. We found the cross-thermocline flux of N2O in the North Atlantic Ocean calculated from in situ N2O data to be an order of magnitude smaller than the N2O flux across the air–sea interface calculated using gas-exchange parametrizations [53].

While these results presented above seem to point towards a significant N2O production via nitrification in the euphotic zone, the probably rather large uncertainties involved in all these calculations need to be taken into account [51,54]. Air–sea fluxes are usually estimated using empirical air–sea gas-exchange parametrizations, which introduces large uncertainties into source estimates. A suite of rather different parametrizations exists, which all fit some experimental dataset, showing that gas exchange itself is highly variable and the governing mechanisms are not yet fully understood [55]: a recent study of both the N2O air–sea fluxes and the N2O diapycnal fluxes into the mixed layer revealed that, by using a common gas-exchange approach, the mean air–sea flux is about four times larger than the mean diapycnal flux into the mixed layer. Vertical advection or biological production was found not sufficient to compensate this discrepancy. Instead, flux calculations using an air–sea exchange parametrization that takes into account the effect of surfactants in the ocean surface microlayer are in good agreement with the diapycnal fluxes, indicating that surfactants, especially in areas with a high biological productivity, may have a large dampening effect on air–sea gas exchange of N2O [54].

There is, however, also an inherent problem with using any kind of gas-exchange model involving the air–sea gradient of some species to estimate long-term average fluxes. Commonly used gas-exchange parametrizations are a measure of an instant sea-to-air flux. After the transfer of molecules, the concentration gradient between the surface water and the atmosphere changes, changing the gas transfer in turn. While this is of no/little consequence to the instant/short-term flux across the air–sea interface, it becomes important when an instant gas-exchange flux is extrapolated over time. Lacking a term to account for the change in concentration over time, it is implicitly assumed in this kind of calculation that any amount of gas lost to the atmosphere from surface waters is instantly replaced from below. In addition, flux estimates based on gas-exchange calculations could be distinctly overestimated as they fail to take the annual cycle of the N2O flux across the air–sea interface into account [53]. Such an overestimation is even more likely as the underlying datasets are mostly seasonally biased. Both these effects can result in an overestimation of the sea-to-air flux, partly explaining the slight discrepancy we encountered here. Additionally, considering the nonlinear dependence on wind speed of almost all gas-exchange parametrizations, it is at least questionable if the use of averaged wind speeds results in a reasonable estimate of the average air–sea flux.

(b). Nitrous oxide and deoxygenation of the ocean

Stramma et al. [23] showed that the oxygen minimum zones of the intermediate layers (300–700 m water depth) in various regions of the ocean are expanding and have been losing oxygen during the past 50 years. This would result in an expansion of the zones supporting denitrification, which would probably have an impact on the production and decomposition of N2O. Whether it would have a net positive or net negative effect on N2O production remains unclear as the net behaviour of denitrification and its controlling mechanisms are not yet fully understood [21]. As N2O yields during both bacterial and archaeal nitrification have been found to be enhanced under low oxygen conditions in laboratory studies [56,57], this increase in volume will probably also lead to increased production of N2O owing to nitrification. Please note that Frame and Casciotti [58] recently showed that the bacterial N2O production via nitrification seems to be less sensitive to [O2] than previously thought [56]. This is in line with the results of two recent studies by Löscher et al. [57] and Santoro et al. [59] which indicate that oceanic N2O production is dominated by archaeal nitrification which, in turn, showed a considerably stronger [O2] sensitivity in culture experiments [57].

(c). Nitrous oxide from anammox

The traditional view that denitrification is the dominant process in suboxic oxygen minimum zone has been challenged by the finding that in the oxygen minimum zone of the Eastern Tropical South Pacific and off Namibia the N2 loss might be almost exclusively performed by bacterial anammox (i.e. anaerobic ammonia oxidation: NO2 + NH4+ → N2) [60,61]. However, so far, N2O has only been found to be produced during anammox in relatively small quantities during nitric oxide detoxification (NO2 → NO → N2O), which seems to be performed by the anammox bacterium Kuenenia stuttgartiensis as a side reaction [62]. Thus, the role of marine anammox bacteria in N2O cycling in suboxic oxygen minimum zones remains to be verified.

(d). Nitrous oxide from coastal hypoxia

Estuaries and nitrogen rich coastal zones probably contribute to a significant degree to the total oceanic emissions of N2O [63,64]. Estuaries and their adjacent regions are fertilized to an increasing degree by river run-off carrying a high load of organic nitrogen. This fertilization may lead to enhanced primary production and thus enhanced nitrification and N2O formation in an increasing number of [O2] depleted (i.e. hypoxic) coastal areas [65]. Additionally, the changing flux of organic material reaching the sediments might change O2 concentrations in the sediment, changing N2O production in turn. And, unlike in the open ocean, the shallow depths of coastal regions allow for N2O produced in the sediments to reach the atmosphere.

(e). Warming of the ocean

There is evidence that the oceans are warming [66,67]. As marine autotrophic and heterotrophic processes display sensitivities to temperature (to varying degrees), ocean warming might result in changes of the bacterial community structure and hence in changes of N2O production. Changes in ocean temperature also affect the solubility of N2O. Rising ocean temperature implies that the N2O long-term storage capacity of the deep ocean will be reduced. Additionally, this effect will temporarily strengthen the N2O source, as apparent supersaturations created by warming water masses after their last contact with the atmosphere will increase the ocean–atmosphere gradient when these water masses finally get into contact with the atmosphere again. This strengthening effect will disappear again once the temperature change levels off and the system reaches equilibrium again. In addition, non-uniform warming could induce enhanced stratification of the water column. This could potentially diminish the oceanic N2O source by keeping N2O-rich water from intermediate depths from reaching the air–sea interface.

(f). Ocean acidification

The ongoing increase of CO2 in the atmosphere also causes a decrease of the oceanic pH (i.e. ocean acidification). This, in turn, results in a shift of the NH3–NH4+ equilibrium towards NH4+. In a recent study by Beman et al. [68], it was shown that nitrification rates decreased significantly when the pH was lowered to values expected to occur in the future ocean. (One explanation for the pH sensitivity of nitrification rates is that the ammonia monooxygenase enzyme uses NH3 rather than NH4+ as substrate in the first step of the nitrification sequence [69].) On the basis of these results, it was suggested by Beman et al. [68] that future oceanic N2O production during nitrification should be decreased as well. However, this scenario might be not that straightforward: nitrification is part of organic matter remineralization (i.e. oxidation of organic matter with O2 to CO2). Therefore, both pH and O2 are decreasing during organic matter remineralization. However, it is well known that decreasing O2 concentrations lead to increasing N2O production during nitrification [57] and obviously there seems to be only a minor effect of decreasing pH on N2O production during nitrification as part of the organic matter remineralization process. Laboratory experiments to verify the effect of ocean acidification on N2O production via nitrification are missing.

Acknowledgements

H.W.B. is grateful for the invitation to the Theo Murphy Meeting ‘Nitrous oxide: the forgotten greenhouse gas’ at the Kavli Royal Society International Centre, 23–24 May 2011. We thank two anonymous reviewers for their constructive comments. The authors are indebted to each of the individual contributors who generously submitted their data to the marine methane and nitrous oxide database (MEMENTO). We thank Toste Tanhua for helpful discussions about the TTD method. This study was supported by the German Science Foundation (DFG) by research grants no. DFG BA1990/7 and BMBF grant no. 03F0462A (SOPRAN).

References

  • 1.IPCC 2007. Climate change 2007: the physical science basis. Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change (eds Solomon S., Qin D., Manning M., Chen Z., Marquis M., Averyt K. B., Tignor M., Miller H. L.), p. 996 Cambridge, UK: Cambridge University Press [Google Scholar]
  • 2.WMO 2011. Scientific assessment of ozone depletion: 2010, global ozone research and monitoring project. Report no. 52 Geneva, Switzerland: WMO [Google Scholar]
  • 3.Forster P., et al. 2007. Changes in atmospheric constituents and in radiative forcing. In Climate change 2007: the physical science basis. Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change (eds Solomon S., Qin D., Manning M., Chen Z., Marquis M., Averyt K. B., Tignor M., Miller H. L.), pp. 129–234 Cambridge, UK: Cambridge University Press [Google Scholar]
  • 4.Bange H. W. 2006. New directions: the importance of the oceanic nitrous oxide emissions. Atmos. Environ. 40, 198–199 10.1016/j.atmosenv.2005.09.030 (doi:10.1016/j.atmosenv.2005.09.030) [DOI] [Google Scholar]
  • 5.Denman K. L., et al. 2007. Couplings between changes in the climate system and biogeochemistry. In Climate change 2007: the physical science basis. Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change (eds Solomon S., Qin D., Manning M., Chen Z., Marquis M., Averyt K. B., Tignor M., Miller H. L.), pp. 499–587 Cambridge, UK: Cambridge University Press [Google Scholar]
  • 6.Bange H. W., Freing A., Kock A., Löscher C. R. 2010. Marine pathways to nitrous oxide. In Nitrous oxide and climate change (ed. Smith K.), pp. 36–62 London, UK: Earthscan [Google Scholar]
  • 7.Bange H. W., Andreae M. O. 1999. Nitrous oxide in the deep waters of the world's oceans. Glob. Biogeochem. Cycles 13, 1127–1135 10.1029/1999GB900082 (doi:10.1029/1999GB900082) [DOI] [Google Scholar]
  • 8.Codispoti L. A. 2010. Interesting times for marine N2O. Science 327, 1339–1340 10.1126/science.1184945 (doi:10.1126/science.1184945) [DOI] [PubMed] [Google Scholar]
  • 9.Suntharalingam P., Sarmiento J. L., Toggweiler J. R. 2000. Global significance of nitrous oxide production and transport from oceanic low-oxygen zones: a modeling study. Glob. Biogeochem. Cycles 14, 1353–1370 10.1029/1999GB900100 (doi:10.1029/1999GB900100) [DOI] [Google Scholar]
  • 10.Freing A., Wallace D. W. R., Tanhua T., Walter S., Bange H. W. 2009. North Atlantic production of nitrous oxide in the context of changing atmospheric levels. Glob. Biogeochem. Cycles 23, GB4015. 10.1029/2009GB003472 (doi:10.1029/2009GB003472) [DOI] [Google Scholar]
  • 11.Locarnini R. A., Mishonov A. V., Antonov J. I., Boyer T. P., Garcia H. E. 2006. World Ocean Atlas 2005, Volume 1: Temperature (ed. Levitus S.). Washington, DC: NOAA [Google Scholar]
  • 12.Antonov J. I., Locarnini R. A., Boyer T. P., Mishonov A. V., Garcia H. E. 2006. World Ocean Atlas 2005, Volume 2: Salinity (ed. Levitus S.). Washington, DC: NOAA [Google Scholar]
  • 13.Garcia H. E., Locarnini R. A., Boyer P., Antonov J. I. 2006. World Ocean Atlas 2005, Volume 3: Dissolved oxygen, apparent oxygen utilization and oxygen saturation (ed. Levitus S.). Washington, DC: NOAA [Google Scholar]
  • 14.Key R. M., et al. 2004. A global ocean carbon climatology: results from Global Data Analysis Project (GLODAP). Glob. Biogeochem. Cycles 18, GB4031. 10.1029/2004GB002247 (doi:10.1029/2004GB002247) [DOI] [Google Scholar]
  • 15.Matsumoto K. 2007. Radiocarbon-based circulation of the world oceans. J. Geophys. Res. 112, C09004. 10.1029/2007JC004095 (doi:10.1029/2007JC004095) [DOI] [Google Scholar]
  • 16.Machida T., Nakazawa T., Fujii Y., Aoki S., Watanabe O. 1995. Increase in the atmospheric nitrous oxide concentration during the last 250 years. Geophys. Res. Lett. 22, 2921–2924 10.1029/95GL02822 (doi:10.1029/95GL02822) [DOI] [Google Scholar]
  • 17.Battle M., et al. 1996. Atmospheric gas concentrations over the past century measured in air from firn at the South Pole. Nature 383, 231–235 10.1038/383231a0 (doi:10.1038/383231a0) [DOI] [Google Scholar]
  • 18.Holland E. A., Lee-Taylor J., Nevison C., Sulzman J. 2005. Fluxes and N2O mixing ratios originating from human activity. See http://www.daac.ornl.gov
  • 19.de Boyer Montégut C., Madec G., Fischer A. S., Lazar A., Iudicone D. 2004. Mixed layer depth over the global ocean: an examination of profile data and a profile-based climatology. J. Geophys. Res. 109, C12003. 10.1029/2004JC002378 (doi:10.1029/2004JC002378) [DOI] [Google Scholar]
  • 20.Bange H. W., Bell T. G., Cornejo M., Freing A., Uher G., Upstill-Goddard R. C., Zhang G. 2009. MEMENTO: a proposal to develop a database of marine nitrous oxide and methane measurements. Environ. Chem. 6, 195–197 10.1071/EN09033 (doi:10.1071/EN09033) [DOI] [Google Scholar]
  • 21.Devol A. H. 2008. Denitrification including anammox. In Nitrogen in the marine environment, 2nd edn. (eds Capone D. G., Bronk D. A., Mulholland M. R., Carpenter E. J.), pp. 263–301 Amsterdam, The Netherlands: Elsevier [Google Scholar]
  • 22.Codispoti L. A., Elkins J. W., Yoshinari T., Friederich G. E., Sakamoto C. M., Packard T. T. 1992. On the nitrous oxide flux from productive regions that contain low oxygen waters. In Oceanography of the Indian Ocean (ed. Desai B. N.), pp. 271–284 Rotterdam, The Netherlands: A.A. Balkema [Google Scholar]
  • 23.Stramma L., Johnson G. C., Sprintall J., Mohrholz V. 2008. Expanding oxygen minimum zones in the tropical oceans. Science 320, 655–658 10.1126/science.1153847 (doi:10.1126/science.1153847) [DOI] [PubMed] [Google Scholar]
  • 24.Paulmier A., Ruiz-Pinto D. 2009. Oxygen minimum zones in the modern ocean. Prog. Oceanogr. 80, 113–128 10.1016/j.pocean.2008.08.001 (doi:10.1016/j.pocean.2008.08.001) [DOI] [Google Scholar]
  • 25.Naqvi S. W. A., Shailaja M. S. 1993. Activity of the respiratory electron transport system and respiration rates within the oxygen minimum layer of the Arabian Sea. Deep-Sea Res. Part II 40, 687–695 10.1016/0967-0645(93)90052-O (doi:10.1016/0967-0645(93)90052-O) [DOI] [Google Scholar]
  • 26.Olson D. B., Hitchcock G. L., Fine R. A., Warren B. A. 1993. Maintenance of the low-oxygen layer in the central Arabian Sea. Deep-Sea Res. Part II 40, 673–685 10.1016/0967-0645(93)90051-N (doi:10.1016/0967-0645(93)90051-N) [DOI] [Google Scholar]
  • 27.Karstensen J., Stramma L., Visbeck M. 2008. The oxygen minimum zones in the eastern tropical Atlantic and Pacific Oceans. Prog. Oceanogr. 77, 331–350 10.1016/j.pocean.2007.05.009 (doi:10.1016/j.pocean.2007.05.009) [DOI] [Google Scholar]
  • 28.Feely R. A., Sabine C. L., Schlitzer R., Bullister J. L., Mecking S., Greeley D. 2004. Oxygen utilization and organic carbon remineralization in the upper water column of the Pacific Ocean. J. Oceanogr. 60, 45–52 10.1023/B:JOCE.0000038317.01279.aa (doi:10.1023/B:JOCE.0000038317.01279.aa) [DOI] [Google Scholar]
  • 29.Jenkins W. J., Wallace D. W. R. 1992. Tracer based inferences of new primary production in the sea. In Primary productivity and biogeochemical cycles (eds Falkowski P. G., Woodhead A. D.), pp. 299–316 New York, NY: Plenum Press [Google Scholar]
  • 30.Tanhua T., Körtzinger A., Friis K., Waugh D. W., Wallace D. W. R. 2007. An estimate of anthropogenic CO2 inventory from decadal changes in oceanic carbon content. Proc. Natl Acad. Sci. USA 104, 3037–3042 10.1073/pnas.0606574104 (doi:10.1073/pnas.0606574104) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.England M. H. 1995. The age of water and ventilation timescales in a global ocean model. J. Phys. Oceanogr. 25, 2756–2777 (doi:10.1175/1520-0485(1995)025<2756:TAOWAV>2.0.CO;2) [DOI] [Google Scholar]
  • 32.Peacock S., Maltrud M. 2006. Transit-time distributions in a global model. J. Phys. Oceanogr. 36, 474–495 10.1175/JPO2860.1 (doi:10.1175/JPO2860.1) [DOI] [Google Scholar]
  • 33.Nevison C., Butler J. H., Elkins J. W. 2003. Global distribution of N2O and ΔN2O–AOU yield in the subsurface ocean. Glob. Biogeochem. Cycles 17, 1119. 10.029/2003GB002068 (doi:10.029/2003GB002068) [DOI] [Google Scholar]
  • 34.Farías L., Paulmier A., Gallegos M. 2007. Nitrous oxide and N-nutrient cycling in the oxygen minimum zone off northern Chile. Deep-Sea Res. Part I 54, 164–180 10.1016/j.dsr.2006.11.003 (doi:10.1016/j.dsr.2006.11.003) [DOI] [Google Scholar]
  • 35.Walter S., Bange H. W., Breitenbach U., Wallace D. W. R. 2006. Nitrous oxide in the North Atlantic Ocean. Biogeosciences 3, 607–619 10.5194/bg-3-607-2006 (doi:10.5194/bg-3-607-2006) [DOI] [Google Scholar]
  • 36.Bange H. W. 2006. Nitrous oxide and methane in European coastal waters. Estuar. Coastal Shelf Sci. 70, 361–374 10.1016/j.ecss.2006.05.042 (doi:10.1016/j.ecss.2006.05.042) [DOI] [Google Scholar]
  • 37.Barnes J., Upstill-Goddard R. C. 2011. N2O seasonal distribution and air–sea exchange in UK estuaries: implications for tropospheric N2O source from European coastal waters. J. Geophys. Res. 116, G01006. 10.1029/2009JG001156 (doi:10.1029/2009JG001156) [DOI] [Google Scholar]
  • 38.Suntharalingam P., Sarmiento J. L. 2000. Factors governing the oceanic nitrous oxide distribution: simulations with an ocean general circulation model. Glob. Biogeochem. Cycles 14, 429–454 10.1029/1999GB900032 (doi:10.1029/1999GB900032) [DOI] [Google Scholar]
  • 39.Rhee T. S., Kettle A. J., Andreae M. O. 2009. Methane and nitrous oxide emissions from the ocean: a reassessment using basin-wide observations in the Atlantic. J. Geophys. Res. 114, D12304. 10.1029/2008JD011662 (doi:10.1029/2008JD011662) [DOI] [Google Scholar]
  • 40.Horrigan S. G., Carlucci A. F., Williams P. M. 1981. Light inhibition of nitrification in sea-surface films. J. Mar. Res. 39, 557–565 [Google Scholar]
  • 41.Bange H. W. 2004. Air–sea exchange of nitrous oxide and methane in the Arabian Sea: a simple model of the seasonal variability. Indian J. Mar. Sci. 33, 77–83 [Google Scholar]
  • 42.Najjar R. G. 1992. Marine biogeochemistry. In Climate system modeling (ed. Trenberth K. E.), pp. 241–280 Cambridge, UK: Cambridge University Press [Google Scholar]
  • 43.Yool A., Martin A. P., Fernández C., Clark D. R. 2007. The significance of nitrification for oceanic new production. Nature 447, 999–1002 10.1038/nature05885 (doi:10.1038/nature05885) [DOI] [PubMed] [Google Scholar]
  • 44.Clark D. R., Rees A. P., Joint I. 2008. Ammonium regeneration and nitrification rates in the oligotrophic Atlantic Ocean: implications for new production estimates. Limnol. Oceanogr. 53, 52–62 10.4319/lo.2008.53.1.0052 (doi:10.4319/lo.2008.53.1.0052) [DOI] [Google Scholar]
  • 45.Wankel S. D., Kendall C., Pennington J. T., Chavez F. P., Paytan A. 2007. Nitrification in the euphotic zone as evidenced by nitrate isotopic composition: observation from Monterey Bay, California. Glob. Biogeochem. Cycles 21, GB2009. 10.1029/2006GB002723 (doi:10.1029/2006GB002723) [DOI] [Google Scholar]
  • 46.Bianchi M., Feliatra F., Treguer P., Vincendeau M.-A., Morvan J. 1997. Nitrification rates, ammonium and nitrate distribution in upper layers of the water column and in sediments of the Indian sector of the Southern Ocean. Deep-Sea Res. Part II 44, 1017–1032 10.1016/S0967-0645(96)00109-9 (doi:10.1016/S0967-0645(96)00109-9) [DOI] [Google Scholar]
  • 47.Dore J. E., Karl D. M. 1996. Nitrification in the euphotic zone as a source for nitrite, nitrate, and nitrous oxide at station ALOHA. Limnol. Oceanogr. 41, 1619–1628 10.4319/lo.1996.41.8.1619 (doi:10.4319/lo.1996.41.8.1619) [DOI] [Google Scholar]
  • 48.Wanninkhof R. 1992. Relationship between wind speed and gas exchange over the ocean. J. Geophys. Res. 97, 7373–7382 10.1029/92JC00188 (doi:10.1029/92JC00188) [DOI] [Google Scholar]
  • 49.Morell J. M., Capella J., Mercado A., Bauzá J., Corredor J. E. 2001. Nitrous oxide fluxes in Caribbean and tropical Atlantic waters: evidence for near surface production. Mar. Chem. 74, 131–143 10.1016/S0304-4203(01)00011-1 (doi:10.1016/S0304-4203(01)00011-1) [DOI] [Google Scholar]
  • 50.Charpentier J., Farias L., Pizarro O. 2010. Nitrous oxide fluxes in the central and eastern South Pacific. Glob. Biogeochem. Cycles 24, GB3011. 10.1029/2008GB003388 (doi:10.1029/2008GB003388) [DOI] [Google Scholar]
  • 51.Law C. S., Ling R. D. 2001. Nitrous oxide flux and response to increased iron availability in the Antarctic Circumpolar Current. Deep-Sea Res. Part II 48, 2509–2527 10.1016/S0967-0645(01)00006-6 (doi:10.1016/S0967-0645(01)00006-6) [DOI] [Google Scholar]
  • 52.Dore J. E., Popp B. N., Karl D. M., Sansone F. J. 1998. A large source of atmospheric nitrous oxide from subtropical North Pacific surface waters. Nature 396, 63–66 10.1038/23921 (doi:10.1038/23921) [DOI] [Google Scholar]
  • 53.Freing A. 2009. Production and emissions of oceanic nitrous oxide. PhD thesis, University of Kiel, Kiel, Germany [Google Scholar]
  • 54.Kock A., Schafstall J., Brandt P., Dengler M., Bange H. W. 2011. Air–sea and diapycnal nitrous oxide fluxes in the eastern tropical North Atlantic Ocean. Biogeosci. Dis. 8, 10 229–10 246 [Google Scholar]
  • 55.Wanninkhof R., Asher W. E., Ho D. T., Sweeney C., McGillis W. R. 2009. Advances in quantifying air–sea gas exchange and environmental forcing. Annu. Rev. Mar. Sci. 1, 213–244 10.1146/annurev.marine.010908.163742 (doi:10.1146/annurev.marine.010908.163742) [DOI] [PubMed] [Google Scholar]
  • 56.Goreau T. J., Kaplan W. A., Wofsy S. C., McElroy M. B., Valois F. W., Watson S. W. 1980. Production of NO2 and N2O by nitrifying bacteria at reduced concentrations of oxygen. Appl. Environ. Microbiol. 40, 526–532 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Löscher C., Kock A., Könneke M., LaRoche J., Bange H. W., Schmitz R. Submitted Ammonia-oxidizing archaea dominate oceanic nitrous oxide production in the OMZ off Mauritania. Biogeosciences. [Google Scholar]
  • 58.Frame C. H., Casciotti K. L. 2010. Biogeochemical controls and isotope signatures of nitrous oxide production by a marine ammonia-oxidizing bacterium. Biogeosciences 7, 2695–2709 10.5194/bg-7-2695-2010 (doi:10.5194/bg-7-2695-2010) [DOI] [Google Scholar]
  • 59.Santoro A. E., Buchwald C., McIlvin M. R., Casciotti K. L. 2011. Isotopic signature of N2O produced by marine ammonia-oxidizing archaea. Science 333, 1282–1285 10.1126/science.1208239 (doi:10.1126/science.1208239) [DOI] [PubMed] [Google Scholar]
  • 60.Lam P., et al. 2009. Revising the nitrogen cycle in the Peruvian oxygen minimum zone. Proc. Natl Acad. Sci. USA 106, 4752–4757 10.1073/pnas.0812444106 (doi:10.1073/pnas.0812444106) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Kuypers M. M. M., Lavik G., Woebken D., Schmid M., Fuchs B. M., Amann R., Jørgensen B. B., Jetten M. S. M. 2005. Massive nitrogen loss from Benguela upwelling system through anaerobic ammonium oxidation. Proc. Natl Acad. Sci. USA 102, 6478–6483 10.1073/pnas.0502088102 (doi:10.1073/pnas.0502088102) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Kartal B., Kuypers M. M. M., Lavik G., Schalk J., Op den Camp H. J. M., Jetten M. S. M., Strous M. 2007. Anammox bacteria disguised as denitrifiers: nitrate reduction to dinitrogen gas via nitrite and ammonium. Environ. Microbiol. 9, 635–642 10.1111/j.1462-2920.2006.01183.x (doi:10.1111/j.1462-2920.2006.01183.x) [DOI] [PubMed] [Google Scholar]
  • 63.Kroeze C., Dumont E., Seitzinger S. P. 2005. New estimates of global emissions of N2O from rivers and estuaries. Environ. Sci. 2, 159–165 10.1080/15693430500384671 (doi:10.1080/15693430500384671) [DOI] [Google Scholar]
  • 64.Zhang G-L., Zhang J., Liu S-M., Ren J-L., Zhao Y.-C. 2010. Nitrous oxide in the Changjiang (Yangtze River) estuary and its adjacent marine area: riverine input, sediment release and atmospheric fluxes. Biogeosciences 7, 3505–3516 10.5194/bg-7-3505-2010 (doi:10.5194/bg-7-3505-2010) [DOI] [Google Scholar]
  • 65.Naqvi S. W. A., Bange H. W., Farías L., Monteiro P. M. S., Scranton M. I., Zhang J. 2010. Marine hypoxia/anoxia as a source of CH4 and N2O. Biogeosciences 7, 2159–2190 10.5194/bg-7-2159-2010 (doi:10.5194/bg-7-2159-2010) [DOI] [Google Scholar]
  • 66.Barnett T. P., Pierce D. W., Achuta Rao K. M., Gleckler P. J., Santer B. D., Gregory J. M., Washington W. M. 2005. Penetration of human induced warming into the world's oceans. Science 309, 284–287 10.1126/science.1112418 (doi:10.1126/science.1112418) [DOI] [PubMed] [Google Scholar]
  • 67.Levitus S., Antonov J. I., Boyer P. 2005. Warming of the world ocean, 1955–2003. Geophys. Res. Lett. 32, L02604. 10.1029/2004GL021592 (doi:10.1029/2004GL021592) [DOI] [Google Scholar]
  • 68.Beman J. M., Chow C-E., King A. L., Feng Y., Fuhrman J. A., Andersson A., Bates N. R., Popp B. N., Hutchins D. A. 2011. Global declines in oceanic nitrification rates as consequence of ocean acidification. Proc. Natl Acad. Sci. USA 108, 208–213 10.1073/pnas.1011053108 (doi:10.1073/pnas.1011053108) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Ward B. B. 2008. Nitrification in marine systems. In Nitrogen in the marine environment, 2nd edn. (eds Capone D. G., Bronk D. A., Mulholland M. R., Carpenter E. J.), pp. 199–261 Amsterdam, The Netherlands: Elsevier [Google Scholar]

Articles from Philosophical Transactions of the Royal Society B: Biological Sciences are provided here courtesy of The Royal Society

RESOURCES