Next Article in Journal
Optimal Probability Distribution and Applicable Minimum Time-Scale for Daily Standardized Precipitation Index Time Series in South Korea
Previous Article in Journal
Analysis of Interannual Anomalies and Causes of Compound Extreme Wind and Precipitation Events in Spring over the Jiangsu–Anhui Region
Previous Article in Special Issue
Responses of Soil N2O Emission and CH4 Uptake to N Input in Chinese Forests across Climatic Zones: A Meta-Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

N2O Emissions from Aquatic Ecosystems: A Review

1
School of Atmospheric Sciences, Sun Yat-sen University, Zhuhai 519082, China
2
Guangdong Province Key Laboratory for Climate Change and Natural Disaster Studies, Sun Yat-sen University, Zhuhai 519082, China
3
Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai), Zhuhai 519082, China
*
Author to whom correspondence should be addressed.
Atmosphere 2023, 14(8), 1291; https://doi.org/10.3390/atmos14081291
Submission received: 14 July 2023 / Revised: 10 August 2023 / Accepted: 12 August 2023 / Published: 15 August 2023
(This article belongs to the Special Issue Nitrogen in a Changing Atmosphere)

Abstract

:
Emissions of nitrous oxide (N2O) from aquatic ecosystems are on the rise due to the dramatic increase in global reactive nitrogen input by anthropogenic activities (e.g., agricultural nitrogen fertilizer use). However, uncertainties exist in the estimation of aquatic N2O budgets due to limited knowledge of mechanisms involved in aquatic N2O emissions, as well as the N2O flux measurements and modelling. To give a full picture of aquatic N2O emissions, this review discusses the biotic and abiotic mechanisms involved in aquatic N2O emissions, common methods used in aquatic N2O flux measurements (including field measurement methods and formula simulation methods), and alternatives for aquatic N2O budget estimation. In addition, this review also suggests that stable isotope technology is promising in the application of aquatic N2O source partitioning.

1. Introduction

Nitrous oxide (N2O), whose single-molecular global warming potential is 273 times (100-year timescale) greater than carbon dioxide, is the third most important long-term greenhouse gas (LLGHG) and accounts for about 7% of the total radiative forcing of LLGHG in 2020 [1,2]. N2O has a long residence time of 110–180 years in the atmosphere [3]. N2O is stable in the troposphere, but when it escapes into the stratosphere, it is photolyzed into NO radical by ultraviolet light, which further destroys the stratospheric ozone layer [4,5]. After the prohibition of the usage of chlorofluorocarbon in The Montreal Protocol on Substances that Deplete the Ozone Layer, N2O has become the most important ozone-depleting substance currently being emitted [4,5].
N2O emissions increased rapidly during the Industrial Revolution [6], mainly due to the increased reactive nitrogen (Nr) input through the Haber–Bosch process [7,8]. According to the latest field network observation data, the global mean atmospheric N2O concentration reached a new high (333.2 ± 0.1 ppb) in 2020, which is 123% of the pre-industrial level, and is still increasing at an average annual growth rate of 0.25% [2].
Aquatic ecosystems, including inland waters, estuaries, and oceans, have long been considered as net N2O sources [9,10]. To meet the food demand of the dramatically growing world population, the application of chemical fertilizers increased from 11.46 Mt in 1961 to 113.29 Mt in 2020 to promote crop production (Figure 1). Only 15–70% of fertilizer N is taken up by plants, and some is leached into inland waters and further transported into estuaries and oceans [11,12,13]. Correspondingly, riverine N2O emissions went through a growth of 91.5% from 0.15 Tg N2O-N yr−1 in 1961 to 0.29 Tg N2O-N yr−1 in 2016 [14]. It is estimated that the total N2O emissions from global rivers and streams account for 10–15% of the total anthropogenic N2O emissions [15]. More than 80% of anthropogenic nitrogen emissions from aquatic ecosystems come from the northern mid-latitudes, consistent with the geographic distribution of nitrogen use and population [16]. The annual N2O emissions of the source streams and lakes are 0.36 and 0.52 Tg N2O yr−1, respectively, which are affected by the discharge of wastewater from agricultural activities, human life, and industrial production [10].
The global biogeochemical cycles of nitrogen in marginal oceans are being strongly interfered with, particularly in estuaries and coastal areas that serve as major transport channels between terrestrial and ocean ecosystems [17,18,19]. Although coastal areas account for only about 18% of the world’s ocean area, they contribute to about 41% of global marine N2O emissions [20,21,22]. The N2O emission of the global coastal continental shelf areas affected by human activities was 0.6 Tg N2O-N yr−1 [23]. The N2O emissions from estuaries and coastal areas are 0.23 and 0.28 Tg N yr−1, respectively, from 2007 to 2016 (Figure 2) [1,20], suggesting that coastal areas are an important source of global N2O emissions.
The ocean is the second natural source of N2O, but with the increase in atmospheric deposition of nitrogen compounds, the anthropogenic N input to the ocean is 1 Tg N yr−1 [24] and the Intergovernmental Panel on Climate Change (IPCC) assumed that the marine N2O emissions range from 2.5 to 4.3 Tg N yr−1 with an average of 3.4 Tg N yr−1 from 2007 to 2016 (Figure 2) [1,25,26,27]. The global ocean contribution to N2O emissions is estimated to be between 10% and 53%, with an average of 21%, but ongoing environmental changes are affecting ocean N2O cycling and emissions to the atmosphere [28,29]. Although it is clear that the ocean is a major natural contributor of N2O to the atmosphere, quantitative estimates remain uncertain [28,30,31].
This review aims to summarize the current state of knowledge of the role of aquatic ecosystems (inland waters–coastal regions–oceans) in global N2O emissions and discuss methods of flux measurement and budget quantification, interpret mechanisms involved in N2O production, and point out the power of stable isotope techniques in N2O source partitioning.
Figure 2. The N2O budgets of different aquatic ecosystems and their isotopic values. The N2O emission data of each aquatic ecosystem in units of Tg N2O-N yr−1 [1,10,20]. The δ15N isotope value in the figure represents the isotope ratio of 15N and 14N, and the site preference (SP) represents the enrichment difference between the proximal N and distal N of the O atom in the N2O molecule; data derived from [32,33,34,35,36].
Figure 2. The N2O budgets of different aquatic ecosystems and their isotopic values. The N2O emission data of each aquatic ecosystem in units of Tg N2O-N yr−1 [1,10,20]. The δ15N isotope value in the figure represents the isotope ratio of 15N and 14N, and the site preference (SP) represents the enrichment difference between the proximal N and distal N of the O atom in the N2O molecule; data derived from [32,33,34,35,36].
Atmosphere 14 01291 g002

2. Mechanisms Involved in N2O Production

2.1. Biological Pathways

N2O production refers to the conversion of any other N species into N2O [37]. In aquatic ecosystems, there are five different microbe-mediated pathways that are important for N2O production (Figure 3): (1) Nitrification is catalyzed by ammonia-oxidizing archaea (AOA) or ammonia-oxidizing bacteria (AOB). NH4+ is stepwise oxidized into hydroxylamine (NH2OH) and nitrite (NO2), which is further oxidized into NO3 by nitrite-oxidizing bacteria (NOB), and N2O is produced as a by-product of NH2OH oxidation [38]. (2) Denitrification includes bacterial denitrification and fungal denitrification. The former is usually an anaerobic dissimilatory reduction process of NO3 to NO2, NO, N2O, and N2 with reductase enzymes of nitrate reductase (Nar), nitrite reductase (Nir), nitric oxide reductase (Nor), and nitrous oxide reductase (Nos), while the latter lacks the gene encoding N2O reductase (NosZ) and fail to reduce N2O to N2. So far, the reduction of N2O to N2 by Nos during denitrification is the only known biological pathway of N2O consumption [4]. (3) Nitrate reduced to ammonium (DNRA) is a process that is primarily driven by nitrate-reducing bacteria and nitrite-reducing bacteria; N2O is produced as a byproduct of NO2 reduction [39]. (4) Nitrifier denitrification refers to the process in which denitrification is conducted by nitrifiers instead of denitrifiers [40]. (5) Coupled nitrification and denitrification is the oxidation of NH4+ to NO3 under an aerobic or micro-oxygen environment and the reduction of NO3 to N2O under anaerobic or hypoxic conditions [41,42,43,44,45,46].
Microbial nitrification and denitrification are biological pathways that contribute largely to aquatic N2O production [47,48,49]. In heavily polluted estuaries, the contribution of nitrification and denitrification to N2O emission ranges from 4.52% to 38.1% and from 61.9% to 80.3%, respectively [32,50]. The anoxic zone of the estuary accounts for 90% of the total estuarine N2O production, which is very likely from denitrification [35,51,52]. The contribution of denitrification to total N2O emissions is about 44%, 14%, 1%, and 11% in global continental shelves, marine oxygen minimum zones (OMZs), estuarine, and freshwater ecosystems (except groundwater), respectively [16]. When the estuary is affected by ocean acidification, N2O release from the sediment is stimulated mainly by bacterial denitrification, while in a neutral environment, N2O production is dominated by fungi [53].

2.2. Abiotic Pathways

Since the intermediates (NH2OH and NO2) of microbial nitrification and denitrification are highly chemically reactive, they are quickly chemically converted into N2O when leached into the environment [54]. Under acidic conditions, NO2 is protonated to HNO2, which is more reactive than NO2 and easily reduced to N2O by soil organic matter or reduced transitional metals in the environment [55]. The chemical pathway in which NO2 is chemically reduced to N2O is called chemodenitrification (Equations (1) and (2)). Ferrous iron (Fe (II)) and lignin-derived soil organic compounds are common reductants in chemodenitrification [44,56,57,58,59,60]. In general, structural iron plays a more important role in chemodenitrification compared with dissolved free Fe (II) ions [61]. In contrast to chemodenitrification, the chemical oxidation of NH2OH becomes quicker when the pH increases because NH2OH is more stable in acidic conditions [54]. When NH2OH is leached into the environment, it can be easily oxidized to N2O by oxidative transition metals, such as Fe (III) and Mn4+ (Equations (3) and (4)).
4 F e 2 + + 2 N O 2 + 5 H 2 O 2 F e 3 + + N 2 O + 6 H +
2 M n 2 + + 2 N O 2 + 6 H + 2 M n 4 + + N 2 O + 3 H 2 O
4 F e 3 + + 2 N H 2 O H 4 F e 2 + + N 2 O + H 2 O + 4 H +
2 M n 4 + + 2 N H 2 O H 2 M n 2 + + N 2 O + H 2 O + 4 H +
More and more studies demonstrate that chemodenitrification and chemical hydroxylamine oxidation play a significant role in global N2O production [45,62,63]. At least 15–25% of N2O formation in coastal marine sediments is caused by chemodenitrification [44]. In the reaction of Fe (Ⅱ) with NO2, the percentage of NO2 converted to N2O ranges from 11% to 52%, reflecting a considerable difference in the degree of chemodenitrification reactions [56]. In intertidal sediments, N2O production with NO2 as a precursor averaged from 70% to 80%, indicating that the N2O was largely catalyzed by fungal denitrification and abiotic reactions such as chemodenitrification [59]. Abiotic NH2OH oxidation is also an important source of N2O in coastal ecosystems; it has been reported that NH2OH produced by nitrifying bacteria is rapidly oxidized to N2O by active Mn (III/IV)-oxidizing minerals [64]. It has been proved that abiotic N2O production plays an important role in coastal N2O emissions, while their roles in streams and seawater are still unclear [37,44,59,64,65].

3. Assessment of Aquatic N2O Budget

3.1. N2O Flux Measurements

Floating chambers, including static and dynamic floating chambers, are the most commonly used methods to determine the spatiotemporal variation in aquatic N2O emissions [66]. The floating static chamber refers to using a top-sealed box placed on the surface of the water to collect the N2O emitted from the surface water through diffusion (Figure 4a). The N2O concentration in the chamber will be analyzed at regular time intervals in the laboratory [67,68], so that the N2O fluxes can be calculated according to the increasing rate of N2O concentration in the chamber over time, as follows [67,68,69,70]:
F N 2 O = n t n 0 A × t
where FN2O (mol m−2 d−1) is the N2O flux through the water–air interface; nt (mol) and n0 (mol) are the mole numbers of N2O in the chamber at time t and 0, respectively; A (m2) is the surface area of the chamber in contact with water; and t (day) is the time for which the concentration of the gas increases linearly.
In contrast, the dynamic floating chambers are based on continuous measurements of gases in the chamber (Figure 4b) [66,71]. At a given airflow rate, the dynamic chamber can reach a steady state in a short time after placement and it can reduce headspace concentration build-up in the gas chamber [66,72]. The gas flow rate extracted by the gas analyzer is about 0.52 L min−1, which is lower than the flow rate of N2-loaded gas controlled by the mass flow controller (7.72L min−1) [66]. In addition, an injection port is added to the intake carrier pipe (near the syringe in Figure 4b), which can be used to inject a known volume of greenhouse gases for calibration. In terms of ebullitive fluxes, the method can detect bubbles containing more than 1.6 × 10−3 mL of gas and a larger chamber surface area in contact with water will improve the ability to trap more bubbles in a given time [66]. It has been found that continuous monitoring data can accurately reflect the 24 h N2O emission concentration [71]. The flux can be determined by following equation [66]:
F = ( d ( θ d C D d t + C D ) d t + θ D d C D d t + C D C 0 θ D C ) × V D C A D C
where F indicates the instantaneous flux (g m−2 s−1); CD is the N2O concentration (g m−3) measured by the gas analyzer; C0 is the potential concentration of the influent gas (g m−3); θDC and θD are the residence time (s) in the dynamic chamber and the cavity of the gas analyzer, respectively; VDC is the volume (m3) of the headspace of the dynamic chamber; and ADC is the area (m2) of the dynamic chamber in contact with the aquatic ecosystem.
The uncertainty of static floating chamber methods mainly comes from the natural turbulence at the water–air interface when deploying floating chambers, so this method is recommended for low fraction and low wave conditions [10,20]. The impact of natural turbulence can be reduced by minimizing the chamber size, smoothing the surface, employing calm water conditions, and taking measurements during stable meteorological conditions. The gas detection accuracy of a dynamic floating chamber is affected by the carrier gas flow and the surface water area covered by the floating chamber. Therefore, it is necessary to set appropriate experimental parameters and select a detector with high sensitivity. For example, the internal pump of the ultra-portable greenhouse gas analyzer (UGGA, Model 915-0011, Los Gatos Research, Inc., Mountain View, CA, USA) can be used to extract the gas in the chamber and determine the concentration of greenhouse gases and water vapor at 1hz [66].
The water–air gas exchange model is another method to indirectly estimate the aquatic N2O flux (Figure 4c) [73]. It is based on the principle that the gas transfer at the water–air interface depends on the combination of the gas diffusion rate and concentration gradient [74]. The gas exchange model aims to calculate the N2O flux (FN2O) according to the N2O water–air gas transfer velocity (kw) using the following formulas [15,42,67,75,76,77]:
F N 2 O = k w ( C o b c C e q )
k w = k 600 ( S c 600 ) 0.5
S c = 2055.6 137.11 t + 4.3173 × t 2 0.054350 × t 3
where FN2O and kw are in the units of μmol·m−2·d−1 and m·d−1, respectively; Cobs is the measured concentration (μmol L−1) of dissolved N2O in the water, Ceq (μmol L−1) represents the air-equilibrated concentration of dissolved N2O, which is calculated according to the temperature, air pressure, and salinity of the sampling sites; Sc is the Schmidt coefficient in fresh water; and t refers to the in situ temperature. k600 is the normalized value to a Schmidt coefficient of 600 at 20 °C, which is generally calculated by empirical hydrology parameter models.
The uncertainties of the water–air gas exchange method are associated with how the wind or water turbulence flow affects gas exchange across the water–air interface [10]. Several different equations for k600 can be used to calculate a series of fluxes to reduce the deviation [15,20]. Equations that include the slope and velocity of a river are the best at predicting the speed of gas transfer, while equations that include depth terms have the strongest correlations (Table 1) [78], and other wind-speed-based or wind speed–hydrological models also used to calculate k600 (Table 1). The gas flux calculated using the multi-model average k600 is conservative but comparable to other models [15]. Field measurements of N2O obtained from the above bottom-up methods are often used to estimate N2O emissions in regional and global aquatic ecosystems but basically show high spatiotemporal heterogeneity [10,79].

3.2. Estimation Based on Emission Factor (EF)

IPCC Tier 1 assumes that the N2O emissions from aquatic systems can be estimated based on the emission factor (EF) of total N leached into aquatic systems [47,84,85,86,87]:
N 2 O N e m i s s i o n = E F × T o t a l N
where N2O-Nemission, Total N, and EF are in units of kg N2O-N yr−1, kg N yr−1, and kg N2O-N kg−1 N, respectively [1,85]. The default EF value was set at 0.0075 kg N2O-N/kg N by IPCC in 1997 and revised down to 0.0025 kg N2O-N/kg N for the convenient estimation of N2O emission from rivers [86,88,89]. But this recommendation was later revised to 0.0026 kg N2O-N/kg NO3-N for rivers, reservoirs, and down-streams, while the EFs for outer estuaries, coastal seawater, and open ocean are still missing [85,89].
In practical application, the data acquisition of N leaching and runoff required to determine the EF of all aquatic ecosystems based on the IPCC definition is incomplete and difficult [10,90,91,92]. Therefore, many studies choose to estimate the EF using the actual NO3 concentrations in focused aquatic systems [10,89,93,94]:
E F = C N 2 O N C N O 3 N
where CN2O−N (mg·L−1) and CNO3−N (mg·L−1) are concentrations measured in the water.
Uncertainty may be introduced when using EF-based estimation [42] because it ignores potential differences in spatial and temporal N delivery efficiency [90], and multiple sources of input may result in the supersaturation of N2O [7,78,95,96]. Moreover, N2O in aquatic ecosystems is mainly produced and consumed through nitrification and denitrification pathways, but the EF-based method does not take into account that these processes may differ significantly under different conditions in diverse waters [68,97]. It is suggested that N2ON/NH4+ should be used to calculate EF in river sections with the ammonium nitrogen pollution type, and N2ON/DIN (DIN ≈ NO3 + NH4+) should be used to calculate the EF when both NO3 and NH4+ are considerable N pollutants [98].
Data analysis based on global observations shows that the N2O fluxes in rivers are positively correlated with the concentrations of NH4+, NO3, and DIN [90]. N2O concentrations and EF values show spatial differences associated with different land use types along water bodies [89,99]. The values of EFs could range from 0.000028 to 0.022, corresponding to the variability in the regional environmental condition (Table 2). The concentrations of dissolved N2O and NO3 were usually higher in agricultural and urban rivers [89]. N2O flux may be determined by NH4+ concentration in some urban river networks, while NO3 plays a more important role in agricultural watersheds [42,67,100,101]. For example, in the Liaohe River basin in northeast China, sewage and aquaculture wastewater discharge leads to NH4+ (0.2–15.5mg/L) pollution in the river, while those polluted by NO3 (0.02–9.6mg/L) are mainly distributed in agricultural areas, and the EFs are 0.4456 and 0.0005, respectively [98]. Therefore, the EFs for inland waters and estuaries should fully consider the influence of specific factors such as river basin land use type and water pollution type, and adopting region-specific EF values is more appropriate for the estimation of regional and global aquatic N2O emissions [10,89,102].

3.3. Model Simulation

The modeling of N2O emissions from aquatic ecosystems can be mainly categorized into two types: statistical models and process-based models. Statistical models of N2O emissions, including empirical models and semi-empirical models, rely on multiple linear regression analysis which can reflect statistical relationships between emission data and their controlling factors [7,31,118,119,120]. For example, the global Nutrient Export from WaterShed (NEWS) model is the first global applicable semi-empirical model used to estimate aquatic N load rates and N2O emissions [47,119]. The model estimates N load rates as a function of human activities, such as nitrogen fertilizer use, human sewage, and atmospheric nitrogen deposition [31]. The estimated global N2O emission from rivers and estuaries based on this model is 1.32 Tg N yr−1 [119]. The global ocean N2O emissions are estimated to be 2.45 ± 0.80 Tg N yr−1 by a semi-empirical model based on the relationship between ocean primary productivity and the N cycle [30]. Machine learning algorithms, such as Random Forest or Monte Carlo simulation, can be used to identify complex relationships between multiple variables [9,10,26].
Process-based models for N2O emissions are based on the biogeochemical and physical processes that drive N2O emissions. For example, the Dynamic Land Ecosystem Model (DLEM) riverine module simulates N2O emissions according to nitrification and denitrification rates, N inputs, N retention and release rates, oxygen consumption, and organic matter decomposition and estimated riverine N2O emission to be 0.3 ± 0.06Tg N yr−1 [14]. The Bern3D Earth system model calculates N2O emissions based on denitrification and organic matter consumption and estimates an oceanic N2O budget of 4.5 ± 1.0 Tg N yr−1 [25]. Process-based models show good advantages to estimate aquatic N2O emissions in the absence of measured data and to predict the response of N2O fluxes to multifactor climate and environmental changes [25,79,121,122]. In addition, the process-based N2O emission model can be integrated with hydrological models to predict changes in N2O emissions in response to land use changes beside the water bodies [37,118,123].
Even though multiple models have been developed to predict aquatic N2O emissions, their results vary from model to model, regarding both natural and anthropogenic sources. The total N2O emission in rivers and oceans ranges largely from 0.03 to 1.05 Tg N yr−1 and from 2.45 to 4.50 Tg N yr−1, respectively (Table 3). The uncertainties of N2O-predicting models mainly come from the fact that (1) measurement data are highly limited due to the difficulties of field sampling and (2) the mechanisms involved in aquatic N2O emissions are not yet fully interpreted [47].

3.4. Uncertainties

Most N2O emission inventories rely on simple EF methods or model studies [47]. Field measurements provide the most realistic estimates of N2O fluxes and can provide spatiotemporal datasets for model calibration and verification [10,37,96]. While more field measurements can yield more accurate EFs, they also bring uncertainty due to aquatic conditions and the heterogeneity in land use types along water bodies, leading to a wide range of N2O estimations [14,90,120,125,126,127]. Moreover, the large-scale deployment of EF measurements is difficult and labor-intensive, and the assessment of aquatic ecosystem EFs using the IPCC method needs detailed N-relevant data, which are difficult to obtain and often missing in many studies [37,70,127]. Despite the limitations, field measurements play an important role in providing accurate estimates of N2O fluxes and improving model frameworks. In the future, it is necessary to take into account river hydrological characteristics and measure more data for the revision of IPCC EFs [67,125].
There is considerable uncertainty in the assessment of N2O emission models at both regional and global scales [7,9,118,120]. Uncertainties come mainly from the complexity of the aquatic ecosystems and inadequate measurement data [123,128]. More data networks need to be established to improve model prediction and validation performance [25,28,79]. The N2O model comparison project is proposed to better identify and ultimately reduce these uncertainties between different models [121], and additional long-term studies measuring results in aquatic ecosystems are needed [28,47].

4. N2O Source Partitioning with Stable Isotope Technique

4.1. Stable Isotopes of N2O

Isotopes are atoms of the same element that have different numbers of neutrons, resulting in different atomic masses, such as 14N and 15N and 16O and 18O. The stable isotope technique is widely used in the qualitative measurement of N-transforming processes, and it is also an effective method to study the pathway of N2O production [36]. N2O is an asymmetric linear molecule with the proximal and distal N of the O atom called Nα and Nβ, respectively. The difference in 15N enrichment of Nα and Nβ is defined as site preference (SP) of N2O [4]:
δ 15 N b u l k = ( δ 15 N α + δ 15 N β ) 2
S P = δ 15 N α δ 15 N β
where δ15Nα and δ15Nβ denote the relative 15N abundance of Nα and Nβ, respectively.
Isotopes substituted by 15N or 18O are usually studied to provide information about the formation and decomposition processes of N2O [129,130].
Obvious differences in δ15N and SP values were detected among each aquatic ecosystem (Figure 2). The δ15N-N2O values of freshwater and seawater are −4.64 ± 9.84‰ and 6.63 ± 3.50‰, respectively [34]. The isotopic characteristics of N2O are controlled by biological, chemical, and physical fractionations, as well as isotopic characteristics of N substrates [101]. Due to the isotope fractionation effect of each N2O production pathway, the stable isotopic values can be used to analyze the relative contributions of the N2O production pathways [43,57].

4.2. Isotopic Characteristics of Biological Processes

During nitrification, the isotope effect of N2O production (δ15N-N2O) by the oxidation of NH4+ is −56.6 ± 7.3‰, but when NH2OH is the only substrate, the δ15N-N2O increases to −5.1 ± 12.0‰ [131]. In aquatic environments, the isotope effect of NO3 reduced to NO2 is −14.3 ± 9.7‰, while the δ15N-N2O for denitrification and nitrifier denitrification from NO2 to N2O can be represented by −14.9 ± 6.7‰ and −34.5 ± 0.7‰, respectively [131,132]. The ranges of δ18O-N2O produced by fungal denitrification and NH4+ by bacterial nitrification are between 27‰ and 42‰ and 20‰ and 26‰, respectively. The δ18O-N2O difference is partly due to the different isotopic compositions of O-containing sources. In the process by which bacteria produce N2O by oxidizing NH4+, the O comes from oxygen and water vapor in the atmosphere. On the contrary, oxygen atoms are derived from NO3 or NO2 substrates during denitrification [133].
Tracing the path of N2O production by using these SP values has become a research hotspot [6,41,132,134]. The SP values of N2O produced by different paths have certain differences, ranging from −11‰ to 36.9‰ (Table 4 and Table 5). Unlike δ18O and δ15N, SP is thought to reflect the N2O production mechanism while remaining independent of the substrate’s isotopic signature [135]. For example, when nitrifying bacteria produce N2O, there may be a positive correlation between SP and δ18O-N2O, resulting from nitrification by mixing high SP and 18O-enriched N2O from hydroxylamine decomposition with low SP and 18O-depleted N2O from nitrifying denitrification [43]. Compared with aerobic denitrification and bacterial denitrification, the N2O process produced by fungal denitrification and nitrification has a higher SP value [36,136]. The SP value of the coupled nitrification and denitrification process is similar to that of denitrification, with a value of 0.1 ± 1.7‰, indicating that the isotopic SP value can distinguish nitrification from coupled nitrification and denitrification [132]. In addition, differences in the experimental environment or the presence of an unknown N2O production pathway in the sample will influence the measurement of the SP value. Therefore, it is necessary to reduce uncertainties and limitations through more experiments in the future.

4.3. Isotopic Characteristics of Abiotic Processes

Dual isotope ecosystems of N and O have also made further progress in their application to distinguish abiotic processes. The SP reflects the difference in the regulatory reaction mechanism, and the end value is related to the final concentration and yield of N2O [56]. For the abiotic reduction of NO2 by Fe (Ⅱ), the δ15N and δ18O of N2O ranges from −19.8‰ to −3.0‰ and from 29.3‰ to 46.4‰, respectively [56]. The SP values of N2O produced by NH2OH oxidation and NO2 reduction processes are 30.1‰ and 29.5‰, respectively [135], while in the process of NO2 reduction catalyzed by Fe (Ⅱ), the SP value ranges from 0.4‰ to 35‰ [56,57,137]. In addition, the reported SP of chemodenitrification ranges from −4‰ to 26.5‰, which overlaps with the SP value of biological processes [45,56,57,61]. At present, although the SP value is an important indicator to identify the pathways of N2O production, most of the studies are pure culture experiments conducted in the laboratory [56,57,61], and the differentiation characteristics of N and O isotope effects on N2O produced by the coupling of NH2OH and NO2 with Fe/Mn in various natural aquatic environments remain unclear, which limits the understanding of the mechanism of N2O production.
Table 4. Summary of N2O isotope characteristics produced by each pathway.
Table 4. Summary of N2O isotope characteristics produced by each pathway.
N2O SourcesReaction Typeδ15N-N2O (‰)δ18O-N2O (‰)SP (‰)Ref.
Ocean-8.860.735.7[138]
-−6~21--[139]
Biotic3.5~5.535.5~41.50–8[33]
Natural or agricultural soilBiotic−38~6--[139]
Antarctic soilBiotic + Abiotic−82 ± 123 ± 122 ± 1[62]
11 rivers in southern Ontario, Canada-−22.6~10.747.4~51.5-[101]
Estuarine salt marshBiotic−1.41~2.9939.66~50.384.09~13.69[129]
Troposphere-7.0 ± 0.643.7 ± 0.918.7 ± 2.2[140]
Stratosphere-11.448.321.3
Weighted average of marine and land sourcesBiotic + Abiotic−25.5~1.09.8~37.4−0.5 ~15.1
“-” indicates no measurement.
Table 5. Summary of N2O isotope characterization of biotic and abiotic processes.
Table 5. Summary of N2O isotope characterization of biotic and abiotic processes.
Reaction ProcessN2O Sourcesδ15N-N2O (‰)δ18O-N2O (‰)SP (‰)Ref.
BioticOxidation of NH2OH and NH4+--32~35[132]
Denitrification of NO2 and NO3--0
NO2 reduction−90~213~35−11~0[6]
NH2OH oxidation-−10-[41]
−68~1922~24.513~37[6]
--32.5~35.6[132]
Nitrification−5.1 ± 12.0--[131]
−56.6 ± 7.3--[131]
Bacterial denitrification--23.3 ± 4.2, −5.1 ± 1.8[134]
Fungal denitrification-37.3 ± 1.3, 31.5 ± 0.537.1 ± 2.5, 36.9 ± 2.8[133]
--15.8~36.7[52]
Nitrifier denitrification--0.1 ± 1.7[132]
AbioticNO2 reduction--30.1 ± 1.7[134]
NH2OH oxidation--29.5 ± 1.1[134]
Chemodenitrification2~114~1026.5 ± 0.8[61]
−19.8~−3.029.3~46.40.4~26.0[56]
--10~22[57]
“-” indicates no measurement.

5. Research Prospect

Aquatic ecosystems are important global N2O sources, but uncertainties still exist in the quantification of the aquatic N2O budget. To understand more precisely the aquatic N2O emissions, further studies should be conducted to (1) establish more N2O flux measurements in aquatic ecosystems, especially data-inadequate areas such as oceans and arctic regions; (2) interpret more clearly the mechanisms involved in aquatic N2O production and its controlling environmental factors; (3) decrease the uncertainties of N2O-predicting models and make the results of different models more comparable; and (4) advance the application of stable isotopic techniques in aquatic N2O emissions regarding both mechanism studies and model establishment.

Author Contributions

Methodology, H.P. and J.W.; Validation, H.P., Z.Z. and S.Z.; Investigation, H.P., Z.Z., S.Z. and F.W.; Data curation, H.P.; Writing—original draft, H.P. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Guangdong Major Project of Basic and Applied Basic Research (2020B0301030004), the National Natural Science Foundation of China (Nos. U21A6001; 42203004), Innovation Group Project of Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai) (No. 311021009), and the Guangdong Provincial Department of Science and Technology (2019ZT08G090).

Data Availability Statement

Data is available under request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IPCC. Global Carbon and other Biogeochemical Cycles and Feedbacks. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2021; pp. 673–816. [Google Scholar] [CrossRef]
  2. Basso, L.; Crotwell, A.; Dolman, H.; Gatti, L.; Gerbig, C.; Griffith, D.; Hall, B.; Jordan, A.; Krummel, P.; Loh, Z.; et al. WMO Greenhouse Gas Bulletin. The State of Greenhouse Gases in the Atmosphere Based on Global Observations through 2020; World Meteorological Organization: Geneva, Switzerland, 2021. [Google Scholar]
  3. Bange, H.W.; Rapsomanikis, S.; Andreae, M.O. Nitrous oxide in coastal waters. Glob. Biogeochem. Cycles 1996, 10, 197–207. [Google Scholar] [CrossRef] [Green Version]
  4. Pauleta, S.R.; Carepo, M.S.P.; Moura, I. Source and reduction of nitrous oxide. Coord. Chem. Rev. 2019, 387, 436–449. [Google Scholar] [CrossRef]
  5. Ravishankara, A.R.; Daniel, J.S.; Portmann, R.W. Nitrous oxide (N2O): The dominant ozone-depleting substance emitted in the 21st century. Science 2009, 326, 123–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Humbert, G.; Sébilo, M.; Fiat, J.; Lang, L.; Filali, A.; Vaury, V.; Spérandio, M.; Laverman, A.M. Isotopic evidence for alteration of nitrous oxide emissions and producing pathways contribution under nitrifying conditions. Biogeosciences 2019, 17, 979–993. [Google Scholar] [CrossRef] [Green Version]
  7. Beaulieu, J.J.; Tank, J.L.; Hamilton, S.K.; Wollheim, W.M.; Hall, R.O., Jr.; Mulholland, P.J.; Peterson, B.J.; Ashkenas, L.R.; Cooper, L.W.; Dahm, C.N.; et al. Nitrous oxide emission from denitrification in stream and river networks. Proc. Natl. Acad. Sci. USA 2011, 108, 214–219. [Google Scholar] [CrossRef] [PubMed]
  8. Raymond, P.A.; Zappa, C.J.; Mulholland, P.; Laursen, A.E.; Butman, D.; Bott, T.L.; Potter, J.; McDowell, W.H.; Newbold, D. Scaling the gas transfer velocity and hydraulic geometry in streams and small rivers. Limnol. Oceanogr. Fluids Environ. 2012, 2, 41–53. [Google Scholar] [CrossRef]
  9. Marzadri, A.; Amatulli, G.; Tonina, D.; Bellin, A.; Shen, L.Q.; Allen, G.H.; Raymond, P.A. Global riverine nitrous oxide emissions: The role of small streams and large rivers. Sci. Total Environ. 2021, 776, 145148. [Google Scholar] [CrossRef] [PubMed]
  10. Zheng, Y.; Wu, S.; Xiao, S.; Yu, K.; Fang, X.; Xia, L.; Wang, J.; Liu, S.; Freeman, C.; Zou, J. Global methane and nitrous oxide emissions from inland waters and estuaries. Glob. Chang. Biol. 2022, 28, 4713–4725. [Google Scholar] [CrossRef]
  11. Gruber, N.; Galloway, J.N. An Earth-system perspective of the global nitrogen cycle. Nature 2008, 451, 293–296. [Google Scholar] [CrossRef] [Green Version]
  12. Menegat, S.; Ledo, A.; Tirado, R. Greenhouse gas emissions from global production and use of nitrogen synthetic fertilisers in agriculture. Sci. Rep. 2022, 12, 14490. [Google Scholar] [CrossRef]
  13. Harris, E.; Yu, L.; Wang, Y.P.; Mohn, J.; Henne, S.; Bai, E.; Barthel, M.; Bauters, M.; Boeckx, P.; Dorich, C.; et al. Warming and redistribution of nitrogen inputs drive an increase in terrestrial nitrous oxide emission factor. Nat. Commun. 2022, 13, 4310. [Google Scholar] [CrossRef] [PubMed]
  14. Yao, Y.; Tian, H.; Shi, H.; Pan, S.; Xu, R.; Pan, N.; Canadell, J.G. Increased global nitrous oxide emissions from streams and rivers in the Anthropocene. Nat. Clim. Chang. 2020, 10, 138–142. [Google Scholar] [CrossRef]
  15. Wang, X.; Yu, L.; Liu, T.; He, Y.; Wu, S.; Chen, H.; Yuan, X.; Wang, J.; Li, X.; Li, H.; et al. Methane and nitrous oxide concentrations and fluxes from heavily polluted urban streams: Comprehensive influence of pollution and restoration. Environ. Pollut. 2022, 313, 120098. [Google Scholar] [CrossRef] [PubMed]
  16. Seitzinger, S.; Harrison, J.A.; Böhlke, J.K.; Bouwman, A.F.; Lowrance, R.; Peterson, B.; Tobias, C.; Drecht, G.V. Denitrification across Landscapes and Waterscapes: A Synthesis. Ecol. Appl. 2006, 16, 2064–2090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Chen, C.T.A.; Wang, S.L.; Lu, X.X.; Zhang, S.R.; Lui, H.K.; Tseng, H.C.; Wang, B.J.; Huang, H.I. Hydrogeochemistry and greenhouse gases of the Pearl River, its estuary and beyond. Quat. Int. 2008, 186, 79–90. [Google Scholar] [CrossRef]
  18. Lequy, E.; Asmala, E.; Ibrom, A.; Loubet, B.; Massad, R.S.; Markager, S.; Garnier, J. Contribution from a eutrophic temperate estuary to the landscape flux of nitrous oxide. Water Res. 2022, 222, 118874. [Google Scholar] [CrossRef] [PubMed]
  19. Zhang, W.X.; Yue, F.J.; Wang, Y.; Li, Y.; Lang, Y.C.; Li, S.L. Dynamic N transport and N2O emission during rainfall events in the coastal river. Sci. Total Environ. 2023, 23, 166206. [Google Scholar] [CrossRef] [PubMed]
  20. Murray, R.H.; Erler, D.V.; Eyre, B.D. Nitrous oxide fluxes in estuarine environments: Response to global change. Glob. Chang. Biol. 2015, 21, 3219–3245. [Google Scholar] [CrossRef]
  21. Wells, N.S.; Maher, D.T.; Erler, D.V.; Hipsey, M.; Rosentreter, J.A.; Eyre, B.D. Estuaries as Sources and Sinks of N2O Across a Land Use Gradient in Subtropical Australia. Glob. Biogeochem. Cycles 2018, 32, 877–894. [Google Scholar] [CrossRef]
  22. Yan, W.; Yang, L.; Wang, F.; Wang, J.; Ma, P. Riverine N2O concentrations, exports to estuary and emissions to atmosphere from the Changjiang River in response to increasing nitrogen loads. Glob. Biogeochem. Cycles 2012, 26, GB4006. [Google Scholar] [CrossRef]
  23. Seitzinger, S.P.; Kroeze, C.; Styles, R.V. Global distribution of N2O emissions from aquatic systems: Natural emissions and anthropogenic effects. Glob. Chang. Sci. 2000, 2, 267–279. [Google Scholar] [CrossRef]
  24. Syakila, A.; Kroeze, C. The global nitrous oxide budget revisited. Greenh. Gas Meas. Manag. 2011, 1, 17–26. [Google Scholar] [CrossRef]
  25. Battaglia, G.; Joos, F. Marine N2O Emissions From Nitrification and Denitrification Constrained by Modern Observations and Projected in Multimillennial Global Warming Simulations. Glob. Biogeochem. Cycles 2018, 32, 92–121. [Google Scholar] [CrossRef] [Green Version]
  26. Yang, S.; Chang, B.X.; Warner, M.J.; Weber, T.S.; Bourbonnais, A.M.; Santoro, A.E.; Kock, A.; Sonnerup, R.E.; Bullister, J.L.; Wilson, S.T.; et al. Global reconstruction reduces the uncertainty of oceanic nitrous oxide emissions and reveals a vigorous seasonal cycle. Proc. Natl. Acad. Sci. USA 2020, 117, 11954–11960. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, X.; Ma, X.; Gu, X.; Liu, S.; Song, G.; Jin, H.; Zhang, G. Seasonal and spatial variations of N2O distribution and emission in the East China Sea and South Yellow Sea. Sci. Total Environ. 2021, 775, 145715. [Google Scholar] [CrossRef]
  28. Bange, H.W.; Arévalo-Martínez, D.L.; de la Paz, M.; Farías, L.; Kaiser, J.; Kock, A.; Law, C.S.; Rees, A.P.; Rehder, G.; Tortell, P.D.; et al. A Harmonized Nitrous Oxide (N2O) Ocean Observation Network for the 21st Century. Front. Mar. Sci. 2019, 6, 157. [Google Scholar] [CrossRef]
  29. Leng, P.; Li, Z.; Zhang, Q.; Koschorreck, M.; Li, F.; Qiao, Y.; Xia, J. Deciphering large-scale spatial pattern and modulators of dissolved greenhouse gases (CO2, CH4, and N2O) along the Yangtze River, China. J. Hydrol. 2023, 623, 129710. [Google Scholar] [CrossRef]
  30. Buitenhuis, E.T.; Suntharalingam, P.; Le Quéré, C. Constraints on global oceanic emissions of N2O from observations and models. Biogeosciences 2018, 15, 2161–2175. [Google Scholar] [CrossRef] [Green Version]
  31. Que, Z.; Wang, X.; Liu, T.; Wu, S.; He, Y.; Zhou, T.; Yu, L.; Qing, Z.; Chen, H.; Yuan, X. Watershed land use change indirectly dominated the spatial variations of CH4 and N2O emissions from two small suburban rivers. J. Hydrol. 2023, 619, 129357. [Google Scholar] [CrossRef]
  32. Gao, Y.; Jia, Y.; Yu, G.; He, N.; Zhang, L.; Zhu, B.; Wang, Y. Anthropogenic reactive nitrogen deposition and associated nutrient limitation effect on gross primary productivity in inland water of China. J. Clean. Prod. 2019, 208, 530–540. [Google Scholar] [CrossRef]
  33. Popp, B.N.; Westley, M.B.; Toyoda, S.; Miwa, T.; Dore, J.E.; Yoshida, N.; Rust, T.M.; Sansone, F.J.; Russ, M.E.; Ostrom, N.E.; et al. Nitrogen and oxygen isotopomeric constraints on the origins and sea-to-air flux of N2O in the oligotrophic subtropical North Pacific gyre. Glob. Biogeochem. Cycles 2002, 16, 12-1–12-10. [Google Scholar] [CrossRef]
  34. Snider, D.M.; Venkiteswaran, J.J.; Schiff, S.L.; Spoelstra, J. From the ground up: Global nitrous oxide sources are constrained by stable isotope values. PLoS ONE 2015, 10, e0118954. [Google Scholar] [CrossRef] [Green Version]
  35. Su, X.; Cui, L.; Tang, Y.; Wen, T.; Yang, K.; Wang, Y.; Zhang, J.; Zhu, G.; Yang, X.; Hou, L.; et al. Denitrification and N2O Emission in Estuarine Sediments in Response to Ocean Acidification: From Process to Mechanism. Environ. Sci. Technol. 2022, 56, 14828–14839. [Google Scholar] [CrossRef] [PubMed]
  36. Toyoda, S.; Yoshida, N.; Koba, K. Isotopocule analysis of biologically produced nitrous oxide in various environments. Mass Spectrom. Rev. 2017, 36, 135–160. [Google Scholar] [CrossRef] [PubMed]
  37. Quick, A.M.; Reeder, W.J.; Farrell, T.B.; Tonina, D.; Feris, K.P.; Benner, S.G. Nitrous oxide from streams and rivers: A review of primary biogeochemical pathways and environmental variables. Earth Sci. Rev. 2019, 191. [Google Scholar] [CrossRef]
  38. Thakur, I.S.; Medhi, K. Nitrification and denitrification processes for mitigation of nitrous oxide from waste water treatment plants for biovalorization: Challenges and opportunities. Bioresour. Technol. 2019, 282, 502–513. [Google Scholar] [CrossRef] [PubMed]
  39. Pandey, C.B.; Kumar, U.; Kaviraj, M.; Minick, K.J.; Mishra, A.K.; Singh, J.S. DNRA: A short-circuit in biological N-cycling to conserve nitrogen in terrestrial ecosystems. Sci. Total Environ. 2020, 738, 139710. [Google Scholar] [CrossRef] [PubMed]
  40. Wrage-Mönnig, N.; Horn, M.A.; Well, R.; Müller, C.; Velthof, G.; Oenema, O. The role of nitrifier denitrification in the production of nitrous oxide revisited. Soil Biol. Biochem. 2018, 123, A3–A16. [Google Scholar] [CrossRef]
  41. Baggs, E.M. A review of stable isotope techniques for N2O source partitioning in soils: Recent progress, remaining challenges and future considerations. Rapid Commun. Mass Spectrom. 2008, 22, 1664–1672. [Google Scholar] [CrossRef]
  42. Beaulieu, J.J.; Arango, C.P.; Hamilton, S.K.; Tank, J.L. The production and emission of nitrous oxide from headwater streams in the Midwestern United States. Glob. Chang. Biol. 2008, 14, 878–894. [Google Scholar] [CrossRef]
  43. Frame, C.H.; Casciotti, K.L. Biogeochemical controls and isotopic signatures of nitrous oxide production by a marine ammonia-oxidizing bacterium. Biogeosciences 2010, 7, 2695–2709. [Google Scholar] [CrossRef] [Green Version]
  44. Otte, J.M.; Blackwell, N.; Ruser, R.; Kappler, A.; Kleindienst, S.; Schmidt, C. N2O formation by nitrite-induced (chemo)denitrification in coastal marine sediment. Sci. Rep. 2019, 9, 10691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Xu, C.; Qi, M.; Lin, W.; Li, X. Nitrous Oxide from Abiotic Processes of Hydroxylamine and Nitrite in Estuarine and Coastal Ecosystems: A Review. J. Mar. Sci. Eng. 2022, 10, 623. [Google Scholar] [CrossRef]
  46. Deng, W.; Wu, J.; Hong, Y.; Liu, X.; Hu, Y. The diversity distribution and N2O production driven by fungal denitrification in different natural ecosystems. Acta Microbiol. Sin. 2021, 61, 1551–1566. (In Chinese) [Google Scholar] [CrossRef]
  47. Ivens, W.P.M.F.; Tysmans, D.J.J.; Kroeze, C.; Löhr, A.J.; van Wijnen, J. Modeling global N2O emissions from aquatic systems. Curr. Opin. Environ. Sustain. 2011, 3, 350–358. [Google Scholar] [CrossRef]
  48. Müller, D.; Bange, H.W.; Warneke, T.; Rixen, T.; Müller, M.; Mujahid, A.; Notholt, J. Nitrous oxide and methane in two tropical estuaries in a peat-dominated region of northwestern Borneo. Biogeosciences 2016, 13, 2415–2428. [Google Scholar] [CrossRef] [Green Version]
  49. Tian, Y.; Yang, P.; Yang, H.; Wang, H.; Zhang, L.; Tong, C.; Lai, D.Y.F.; Lin, Y.; Tan, L.; Hong, Y.; et al. Diffusive nitrous oxide (N2O) fluxes across the sediment-water-atmosphere interfaces in aquaculture shrimp ponds in a subtropical estuary: Implications for climate warming. Agric. Ecosyst. Environ. 2023, 341, 108218. [Google Scholar] [CrossRef]
  50. Li, X.; Qi, M.; Gao, D.; Liu, M.; Sardans, J.; Penuelas, J.; Hou, L. Nitrous oxide emissions from subtropical estuaries: Insights for environmental controls and implications. Water Res. 2022, 212, 118110. [Google Scholar] [CrossRef]
  51. Yang, J.T.; Hsu, T.C.; Tan, E.; Lee, K.; Krom, M.D.; Kang, S.; Dai, M.; Hsiao, S.S.; Yan, X.; Zou, W.; et al. Sedimentary processes dominate nitrous oxide production and emission in the hypoxic zone off the Changjiang River estuary. Sci. Total Environ. 2022, 827, 154042. [Google Scholar] [CrossRef]
  52. Maeda, K.; Spor, A.; Edel-Hermann, V.; Heraud, C.; Breuil, M.C.; Bizouard, F.; Toyoda, S.; Yoshida, N.; Steinberg, C.; Philippot, L. N2O production, a widespread trait in fungi. Sci. Rep. 2015, 5, 9697. [Google Scholar] [CrossRef] [Green Version]
  53. Su, X.; Wen, T.; Wang, Y.; Xu, J.; Cui, L.; Zhang, J.; Xue, X.; Ding, K.; Tang, Y.; Zhu, Y.-g. Stimulation of N2O emission via bacterial denitrification driven by acidification in estuarine sediments. Glob. Chang. Biol. 2021, 27, 5564–5579. [Google Scholar] [CrossRef] [PubMed]
  54. Wei, J.; Zhang, X.; Xia, L.; Yuan, W.; Zhou, Z.; Bruggmann, N. Role of chemical reactions in the nitrogenous trace gas emissions and nitrogen retention: A meta-analysis. Sci. Total Environ. 2022, 808, 152141. [Google Scholar] [CrossRef] [PubMed]
  55. Wei, J.; Amelung, W.; Lehndorff, E.; Schloter, M.; Vereecken, H.; Brüggemann, N. N2O and NOx emissions by reactions of nitrite with soil organic matter of a Norway spruce forest. Biogeochemistry 2017, 132, 325–342. [Google Scholar] [CrossRef]
  56. Buchwald, C.; Grabb, K.; Hansel, C.M.; Wankel, S.D. Constraining the role of iron in environmental nitrogen transformations: Dual stable isotope systematics of abiotic NO2 reduction by Fe(II) and its production of N2O. Geochim. Cosmochim. Acta 2016, 186, 1–12. [Google Scholar] [CrossRef] [Green Version]
  57. Jones, L.C.; Peters, B.; Lezama Pacheco, J.S.; Casciotti, K.L.; Fendorf, S. Stable isotopes and iron oxide mineral products as markers of chemodenitrification. Environ. Sci. Technol. 2015, 49, 3444–3452. [Google Scholar] [CrossRef]
  58. Picardal, F. Abiotic and Microbial Interactions during Anaerobic Transformations of Fe(II) and NOx. Front. Microbiol. 2012, 3, 112. [Google Scholar] [CrossRef] [Green Version]
  59. Wankel, S.D.; Ziebis, W.; Buchwald, C.; Charoenpong, C.; de Beer, D.; Dentinger, J.; Xu, Z.; Zengler, K. Evidence for fungal and chemodenitrification based N2O flux from nitrogen impacted coastal sediments. Nat. Commun. 2017, 8, 15595. [Google Scholar] [CrossRef] [Green Version]
  60. Wei, J.; Knicker, H.; Zhou, Z.; Eckhardt, K.-U.; Leinweber, P.; Wissel, H.; Yuan, W.; Brüggemann, N. Nitrogen immobilization caused by chemical formation of black- and amide-N in soil. Geoderma 2023, 429, 116274. [Google Scholar] [CrossRef]
  61. Grabb, K.C.; Buchwald, C.; Hansel, C.M.; Wankel, S.D. A dual nitrite isotopic investigation of chemodenitrification by mineral-associated Fe(II) and its production of nitrous oxide. Geochim. Cosmochim. Acta 2017, 196, 388–402. [Google Scholar] [CrossRef] [Green Version]
  62. Peters, B.; Casciotti, K.L.; Samarkin, V.A.; Madigan, M.T.; Schutte, C.A.; Joye, S.B. Stable isotope analyses of NO2, NO3, and N2O in the hypersaline ponds and soils of the McMurdo Dry Valleys, Antarctica. Geochim. Cosmochim. Acta 2014, 135, 87–101. [Google Scholar] [CrossRef]
  63. Zhu-Barker, X.; Cavazos, A.R.; Ostrom, N.E.; Horwath, W.R.; Glass, J.B. The importance of abiotic reactions for nitrous oxide production. Biogeochemistry 2015, 126, 251–267. [Google Scholar] [CrossRef]
  64. Cavazos, A.R.; Taillefert, M.; Tang, Y.; Glass, J.B. Kinetics of nitrous oxide production from hydroxylamine oxidation by birnessite in seawater. Mar. Chem. 2018, 202, 49–57. [Google Scholar] [CrossRef] [Green Version]
  65. Su, X.; Yang, L.; Yang, K.; Tang, Y.; Wen, T.; Wang, Y.; Rillig, M.C.; Rohe, L.; Pan, J.; Li, H.; et al. Estuarine plastisphere as an overlooked source of N2O production. Nat. Commun. 2022, 13, 3884. [Google Scholar] [CrossRef] [PubMed]
  66. Gerardo-Nieto, O.; Vega-Penaranda, A.; Gonzalez-Valencia, R.; Alfano-Ojeda, Y.; Thalasso, F. Continuous Measurement of Diffusive and Ebullitive Fluxes of Methane in Aquatic Ecosystems by an Open Dynamic Chamber Method. Environ. Sci. Technol. 2019, 53, 5159–5167. [Google Scholar] [CrossRef] [PubMed]
  67. Hinshaw, S.E.; Dahlgren, R.A. Dissolved nitrous oxide concentrations and fluxes from the eutrophic San Joaquin River, California. Environ. Sci. Technol. 2013, 47, 1313–1322. [Google Scholar] [CrossRef]
  68. Zhang, L.; Zhang, S.; Xia, X.; Battin, T.J.; Liu, S.; Wang, Q.; Liu, R.; Yang, Z.; Ni, J.; Stanley, E.H. Unexpectedly minor nitrous oxide emissions from fluvial networks draining permafrost catchments of the East Qinghai-Tibet Plateau. Nat. Commun. 2022, 13, 950. [Google Scholar] [CrossRef] [PubMed]
  69. Beaulieu, J.J.; Shuster, W.D.; Rebholz, J.A. Nitrous Oxide Emissions from a Large, Impounded River: The Ohio River. Environ. Sci. Technol. 2010, 44, 7527–7533. [Google Scholar] [CrossRef]
  70. Webb, J.R.; Clough, T.J.; Quayle, W.C. A review of indirect N2O emission factors from artificial agricultural waters. Environ. Res. Lett. 2021, 16, 043005. [Google Scholar] [CrossRef]
  71. Yang, W.B.; Yuan, C.S.; Huang, B.Q.; Tong, C.; Yang, L. Emission Characteristics of Greenhouse Gases and Their Correlation with Water Quality at an Estuarine Mangrove Ecosystem—The Application of an In-situ On-site NDIR Monitoring Technique. Wetlands 2018, 38, 723–738. [Google Scholar] [CrossRef]
  72. Denmead, O.T. Approaches to measuring fluxes of methane and nitrous oxide between landscapes and the atmosphere. Plant Soil 2008, 309, 5–24. [Google Scholar] [CrossRef]
  73. Wu, S.; Chen, J.; Li, C.; Kong, D.; Yu, K.; Liu, S.; Zou, J. Diel and seasonal nitrous oxide fluxes determined by floating chamber and gas transfer equation methods in agricultural irrigation watersheds in southeast China. Environ. Monit. Assess. 2018, 190, 122. [Google Scholar] [CrossRef] [PubMed]
  74. Gao, J.; Zheng, X.; Wang, R.; Liao, T.; Zou, J. Preliminary Comparison of the Static Floating Chamber and the Diffusion Model Methods for Measuring Water–Atmosphere Exchanges of Methane and Nitrous Oxide from Inland Water Bodies. Clim. Environ. Res. 2014, 19, 290–302. (In Chinese) [Google Scholar] [CrossRef]
  75. Clough, T.J.; Bertram, J.E.; Sherlock, R.R.; Leonard, R.L.; Nowicki, B.L. Comparison of measured and EF5-r-derived N2O fluxes from a spring-fed river. Glob. Chang. Biol. 2006, 12, 477–488. [Google Scholar] [CrossRef]
  76. Liu, S.; Gao, Q.; Wu, J.; Xie, Y.; Yang, Q.; Wang, R.; Zhang, J.; Liu, Q. Spatial distribution and influencing mechanism of CO2, N2O and CH4 in the Pearl River Estuary in summer. Sci. Total Environ. 2022, 846, 157381. [Google Scholar] [CrossRef] [PubMed]
  77. Xia, X.; Zhang, L.; Wang, G.; Wang, J.; Zhang, L.; Zhang, S.; Li, Z. Nitrogen loss from a turbid river network based on N2 and N2O fluxes: Importance of suspended sediment. Sci. Total Environ. 2021, 757, 143918. [Google Scholar] [CrossRef] [PubMed]
  78. Rosamond, M.S.; Thuss, S.J.; Schiff, S.L. Dependence of riverine nitrous oxide emissions on dissolved oxygen levels. Nat. Geosci. 2012, 5, 715–718. [Google Scholar] [CrossRef]
  79. Tian, H.; Xu, R.; Canadell, J.G.; Thompson, R.L.; Winiwarter, W.; Suntharalingam, P.; Davidson, E.A.; Ciais, P.; Jackson, R.B.; Janssens-Maenhout, G.; et al. A comprehensive quantification of global nitrous oxide sources and sinks. Nature 2020, 586, 248–256. [Google Scholar] [CrossRef] [PubMed]
  80. Yu, Z.; Deng, H.; Wang, D.; Ye, M.; Tan, Y.; Li, Y.; Chen, Z.; Xu, S. Nitrous oxide emissions in the Shanghai river network: Implications for the effects of urban sewage and IPCC methodology. Glob. Chang. Biol. 2013, 19, 2999–3010. [Google Scholar] [CrossRef]
  81. Jiang, L.Q.; Cai, W.J.; Wang, Y. A comparative study of carbon dioxide degassing in river- and marine-dominated estuaries. Limnol. Oceanogr. 2008, 53, 2603–2615. [Google Scholar] [CrossRef] [Green Version]
  82. Borges, A.V.; Vanderborght, J.-P.; Schiettecatte, L.-S.; Gazeau, F.; Ferrón-Smith, S.; Delille, B.; Frankignoulle, M. Variability of the gas transfer velocity of CO2 in a macrotidal estuary (the Scheldt). Estuaries 2004, 27, 593–603. [Google Scholar] [CrossRef] [Green Version]
  83. Yu, Z.; Wang, D.; Li, Y.; Deng, H.; Hu, B.; Ye, M.; Zhou, X.; Da, L.; Chen, Z.; Xu, S. Carbon dioxide and methane dynamics in a human-dominated lowland coastal river network (Shanghai, China). J. Geophys. Res. Biogeosciences 2017, 122, 1738–1758. [Google Scholar] [CrossRef] [Green Version]
  84. Aho, K.S.; Fair, J.H.; Hosen, J.D.; Kyzivat, E.D.; Logozzo, L.A.; Weber, L.C.; Yoon, B.; Zarnetske, J.P.; Raymond, P.A. An intense precipitation event causes a temperate forested drainage network to shift from N2O source to sink. Limnol. Oceanogr. 2022, 67, S242–S257. [Google Scholar] [CrossRef]
  85. IPCC. 2019 Refinement to the 2006 IPCC Guidelines for National Greenhouse Gas Inventories. Volume 4 Agriculture, Forestry and Other Land Use. Chapter 11: N2O Emissions from Managed Soils, and CO2 Emissions from Lime and Urea Application; IPCC: Geneva, Switzerland, 2019. [Google Scholar]
  86. IPCC. 2006 IPCC Guidelines for National Greenhouse Gas Inventories. Volume 4: Agriculture, Forestry and Other Land Use; IPCC: Geneva, Switzerland, 2006. [Google Scholar]
  87. Zhou, Y.; Huang, M.; Tian, H.; Xu, R.; Ge, J.; Yang, X.; Liu, R.; Sun, Y.; Pan, S.; Gao, Q.; et al. Four decades of nitrous oxide emission from Chinese aquaculture underscores the urgency and opportunity for climate change mitigation. Environ. Res. Lett. 2021, 16, 114038. [Google Scholar] [CrossRef]
  88. Wang, J.; Chen, N.; Yan, W.; Wang, B.; Yang, L. Effect of dissolved oxygen and nitrogen on emission of N2O from rivers in China. Atmos. Environ. 2015, 103, 347–356. [Google Scholar] [CrossRef]
  89. Wang, J.; Wang, G.; Zhang, S.; Xin, Y.; Jiang, C.; Liu, S.; He, X.; McDowell, W.H.; Xia, X. Indirect nitrous oxide emission factors of fluvial networks can be predicted by dissolved organic carbon and nitrate from local to global scales. Glob. Chang. Biol. 2022, 28, 7270–7285. [Google Scholar] [CrossRef] [PubMed]
  90. Hu, M.; Chen, D.; Dahlgren, R.A. Modeling nitrous oxide emission from rivers: A global assessment. Glob. Chang. Biol. 2016, 22, 3566–3582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Outram, F.N.; Hiscock, K.M. Indirect nitrous oxide emissions from surface water bodies in a lowland arable catchment: A significant contribution to agricultural greenhouse gas budgets? Environ. Sci. Technol. 2012, 46, 8156–8163. [Google Scholar] [CrossRef] [PubMed]
  92. Liu, F.; Zhu, Q.; Wang, Y.; Lai, X.; Liao, K.; Guo, C. Storages and leaching losses of soil water dissolved CO2 and N2O on typical land use hillslopes in southeastern hilly area of China. Sci. Total Environ. 2023, 886, 163780. [Google Scholar] [CrossRef]
  93. Hama-Aziz, Z.Q.; Hiscock, K.M.; Cooper, R.J. Indirect Nitrous Oxide Emission Factors for Agricultural Field Drains and Headwater Streams. Environ. Sci. Technol. 2017, 51, 301–307. [Google Scholar] [CrossRef] [Green Version]
  94. Zhou, Y.; Toyoda, R.; Suenaga, T.; Aoyagi, T.; Hori, T.; Terada, A. Low nitrous oxide concentration and spatial microbial community transition across an urban river affected by treated sewage. Water Res. 2022, 216, 118276. [Google Scholar] [CrossRef]
  95. Garnier, J.; Billen, G.; Vilain, G.; Martinez, A.; Silvestre, M.; Mounier, E.; Toche, F. Nitrous oxide (N2O) in the Seine river and basin: Observations and budgets. Agric. Ecosyst. Environ. 2009, 133, 223–233. [Google Scholar] [CrossRef]
  96. Hu, M.; Li, B.; Wu, K.; Zhang, Y.; Wu, H.; Zhou, J.; Chen, D. Modeling Riverine N2O Sources, Fates, and Emission Factors in a Typical River Network of Eastern China. Environ. Sci. Technol. 2021, 55, 13356–13365. [Google Scholar] [CrossRef] [PubMed]
  97. Tian, L.; Cai, Y.; Akiyama, H. A review of indirect N2O emission factors from agricultural nitrogen leaching and runoff to update of the default IPCC values. Environ. Pollut. 2019, 245, 300–306. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, M.; Li, Y.; Lei, K.; Yang, L. Concentration, Flux, and Emission Factor of N2O in River with Different Nitrogen Pollution Features. Environ. Sci. 2018, 39, 5400–5409. (In Chinese) [Google Scholar] [CrossRef]
  99. He, Y.; Wang, X.; Chen, H.; Yuan, X.; Wu, N.; Zhang, Y.; Yue, J.; Zhang, Q.; Diao, Y.; Zhou, L. Effect of watershed urbanization on N2O emissions from the Chongqing metropolitan river network, China. Atmos. Environ. 2017, 171, 70–81. [Google Scholar] [CrossRef]
  100. Audet, J.; Wallin, M.B.; Kyllmar, K.; Andersson, S.; Bishop, K. Nitrous oxide emissions from streams in a Swedish agricultural catchment. Agric. Ecosyst. Environ. 2017, 236, 295–303. [Google Scholar] [CrossRef]
  101. Baulch, H.M.; Schiff, S.L.; Thuss, S.J.; Dillon, P.J. Isotopic character of nitrous oxide emitted from streams. Environ. Sci. Technol. 2011, 45, 4682–4688. [Google Scholar] [CrossRef] [PubMed]
  102. Cooper, R.J.; Wexler, S.K.; Adams, C.A.; Hiscock, K.M. Hydrogeological controls on regional-scale indirect nitrous oxide emission factors for rivers. Environ. Sci. Technol. 2017, 51, 10440–10448. [Google Scholar] [CrossRef] [Green Version]
  103. Li, X.; Sardans, J.; Qi, M.; Ni, X.; Zhang, M.; Peñuelas, J.; Yue, K.; Wu, F. Nitrous oxide concentration and flux in Min River Basin of southeast China: Effects of land use, stream order and water variables. J. Hydrol. 2022, 614, 128507. [Google Scholar] [CrossRef]
  104. Qi, M.; Luo, L.; Hong, X.; Qian, W.; Tong, C.; Li, X. Spatial and temporal variations and influencing factors of nitrous oxide emissions from surface water in Min River Estuary. Acta Sci. Circumstantiae 2022, 42, 489–500. (In Chinese) [Google Scholar] [CrossRef]
  105. Guan, D.; Zhao, H.; Yao, Z. The distribution and flux of nitrous oxide in the Liaohe Estuary area. Acta Oceanol. Sin. 2009, 31, 85–90. (In Chinese) [Google Scholar]
  106. Li, X.; Yue, F.; Zhou, B.; Wang, X.; Hu, J.; Chen, S.; Li, S. Study on N2O release from typical gate-controlled rivers in Bohai Bay. China Environ. Sci. 2022, 42, 356–366. (In Chinese) [Google Scholar] [CrossRef]
  107. Hu, B.; Wang, D.; Zhou, J.; Meng, W.; Li, C.; Sun, Z.; Guo, X.; Wang, Z. Greenhouse gases emission from the sewage draining rivers. Sci. Total Environ. 2018, 612, 1454–1462. [Google Scholar] [CrossRef] [PubMed]
  108. Xia, Y.; Li, Y.; Li, X.; Guo, M.; She, D.; Yan, X. Diurnal pattern in nitrous oxide emissions from a sewage-enriched river. Chemosphere 2013, 92, 421–428. [Google Scholar] [CrossRef]
  109. Wu, W.; Wang, J.; Zhou, X.; Yuan, B.; Guo, M.; Ren, L. Spatiotemporal distribution of nitrous oxide (N2O) emissions from cascade reservoirs in Lancang-Mekong River Yunnan section, Southwestern China. River Res. Appl. 2020, 37, 1055–1069. [Google Scholar] [CrossRef]
  110. Li, X.; Yu, Y.; Fan, H.; Tang, C. Intense denitrification and sewage effluent result in enriched 15N in N2O from urban polluted rivers. J. Hydrol. 2022, 608, 127631. [Google Scholar] [CrossRef]
  111. Lin, H.; Dai, M.; Kao, S.-J.; Wang, L.; Roberts, E.; Yang, J.-Y.T.; Huang, T.; He, B. Spatiotemporal variability of nitrous oxide in a large eutrophic estuarine system: The Pearl River Estuary, China. Mar. Chem. 2016, 182, 14–24. [Google Scholar] [CrossRef]
  112. Grossel, A.; Bourennane, H.; Ayzac, A.; Pasquier, C.; Henault, C. Indirect emissions of nitrous oxide in a cropland watershed with contrasting hydrology in central France. Sci. Total Environ. 2021, 766, 142664. [Google Scholar] [CrossRef]
  113. Borges, A.V.; Darchambeau, F.; Lambert, T.; Morana, C.; Allen, G.H.; Tambwe, E.; Sembaito, A.T.; Mambo, T.; Wabakhangazi, J.N.; Descy, J.-P.; et al. Variations in dissolved greenhouse gases (CO2, CH4, N2O) in the Congo River network overwhelmingly driven by fluvial-wetland connectivity. Biogeosciences 2019, 16, 3801–3834. [Google Scholar] [CrossRef] [Green Version]
  114. Mwanake, R.M.; Gettel, G.M.; Aho, K.S.; Namwaya, D.W.; Masese, F.O.; Butterbach-Bahl, K.; Raymond, P.A. Land Use, Not Stream Order, Controls N2O Concentration and Flux in the Upper Mara River Basin, Kenya. J. Geophys. Res. Biogeosci. 2019, 124, 3491–3506. [Google Scholar] [CrossRef]
  115. Hershey, R.; Nandan, S.B.; Jayachandran, P.R.; Vijay, A.; Sudheesh, V. Nitrous oxide flux from a Tropical estuarine system (Cochin estuary, India). Reg. Stud. Mar. Sci. 2019, 30, 100725. [Google Scholar] [CrossRef]
  116. Nirmal Rajkumar, A.; Barnes, J.; Ramesh, R.; Purvaja, R.; Upstill-Goddard, R.C. Methane and nitrous oxide fluxes in the polluted Adyar River and estuary, SE India. Mar. Pollut. Bull. 2008, 56, 2043–2051. [Google Scholar] [CrossRef] [PubMed]
  117. Clough, T.J.; Buckthought, L.E.; Casciotti, K.L.; Kelliher, F.M.; Jones, P.K. Nitrous oxide dynamics in a braided river system, New Zealand. J. Environ. Qual. 2011, 40, 1532–1541. [Google Scholar] [CrossRef]
  118. Gao, X.; Ouyang, W.; Lin, C.; Wang, K.; Hao, F.; Hao, X.; Lian, Z. Considering atmospheric N2O dynamic in SWAT model avoids the overestimation of N2O emissions in river networks. Water Res. 2020, 174, 115624. [Google Scholar] [CrossRef] [PubMed]
  119. Kroeze, C.; Dumont, E.; Seitzinger, S. Future trends in emissions of N2O from rivers and estuaries. J. Integr. Environ. Sci. 2010, 7, 71–78. [Google Scholar] [CrossRef]
  120. Maavara, T.; Lauerwald, R.; Laruelle, G.G.; Akbarzadeh, Z.; Bouskill, N.J.; Van Cappellen, P.; Regnier, P. Nitrous oxide emissions from inland waters: Are IPCC estimates too high? Glob. Chang. Biol. 2018, 25, 473–488. [Google Scholar] [CrossRef]
  121. Tian, H.; Yang, J.; Lu, C.; Xu, R.; Canadell, J.G.; Jackson, R.B.; Arneth, A.; Chang, J.; Chen, G.; Ciais, P.; et al. The Global N2O Model Intercomparison Project. Bull. Am. Meteorol. Soc. 2018, 99, 1231–1251. [Google Scholar] [CrossRef] [Green Version]
  122. Wang, C.; Schürz, C.; Zoboli, O.; Zessner, M.; Schulz, K.; Watzinger, A.; Bodner, G.; Mehdi-Schulz, B. N2O Emissions from Two Austrian Agricultural Catchments Simulated with an N2O Submodule Developed for the SWAT Model. Atmosphere 2021, 13, 50. [Google Scholar] [CrossRef]
  123. Marzadri, A.; Dee, M.M.; Tonina, D.; Bellin, A.; Tank, J.L. Role of surface and subsurface processes in scaling N2O emissions along riverine networks. Proc. Natl. Acad. Sci. USA 2017, 114, 4330–4335. [Google Scholar] [CrossRef]
  124. Kroeze, C.; Seitzinger, S.P. Nitrogen inputs to rivers, estuaries and continental shelves and related nitrous oxide emissions in 1990 and 2050: A global model. Nutr. Cycl. Agroecosystems 1998, 52, 195–212. [Google Scholar] [CrossRef]
  125. Reay, D.S.; Davidson, E.A.; Smith, K.A.; Smith, P.; Melillo, J.M.; Dentener, F.; Crutzen, P.J. Global agriculture and nitrous oxide emissions. Nat. Clim. Chang. 2012, 2, 410–416. [Google Scholar] [CrossRef]
  126. Kroeze, C.; Dumont, E.; Seitzinger, S.P. New estimates of global emissions of N2O from rivers and estuaries. Environ. Sci. 2005, 2, 159–165. [Google Scholar] [CrossRef]
  127. Qin, X.; Li, Y.; Goldberg, S.; Wan, Y.; Fan, M.; Liao, Y.; Wang, B.; Gao, Q.; Li, Y. Assessment of Indirect N2O Emission Factors from Agricultural River Networks Based on Long-Term Study at High Temporal Resolution. Environ. Sci. Technol. 2019, 53, 10781–10791. [Google Scholar] [CrossRef] [PubMed]
  128. Borges, A.V.; Darchambeau, F.; Teodoru, C.R.; Marwick, T.R.; Tamooh, F.; Geeraert, N.; Omengo, F.O.; Guérin, F.; Lambert, T.; Morana, C.; et al. Globally significant greenhouse-gas emissions from African inland waters. Nat. Geosci. 2015, 8, 637–642. [Google Scholar] [CrossRef] [Green Version]
  129. Gao, D.; Hou, L.; Li, X.; Liu, M.; Zheng, Y.; Yin, G.; Yang, Y.; Liu, C.; Han, P. Exotic Spartina alterniflora invasion alters soil nitrous oxide emission dynamics in a coastal wetland of China. Plant Soil 2019, 442, 233–246. [Google Scholar] [CrossRef]
  130. Sutka, R.L.; Ostrom, N.E.; Ostrom, P.H.; Gandhi, H.; Breznak, J.A. Nitrogen isotopomer site preference of N2O produced by Nitrosomonas europaea and Methylococcus capsulatus Bath. Rapid Commun. Mass Spectrom. 2003, 17, 738–745. [Google Scholar] [CrossRef] [PubMed]
  131. Denk, T.R.A.; Mohn, J.; Decock, C.; Lewicka-Szczebak, D.; Harris, E.; Butterbach-Bahl, K.; Kiese, R.; Wolf, B. The nitrogen cycle: A review of isotope effects and isotope modeling approaches. Soil Biol. Biochem. 2017, 105, 121–137. [Google Scholar] [CrossRef] [Green Version]
  132. Sutka, R.L.; Ostrom, N.E.; Ostrom, P.H.; Breznak, J.A.; Gandhi, H.; Pitt, A.J.; Li, F. Distinguishing nitrous oxide production from nitrification and denitrification on the basis of isotopomer abundances. Appl. Environ. Microbiol. 2006, 72, 638–644. [Google Scholar] [CrossRef] [Green Version]
  133. Sutka, R.L.; Adams, G.C.; Ostrom, N.E.; Ostrom, P.H. Isotopologue fractionation during N2O production by fungal denitrification. Rapid Commun. Mass Spectrom. 2008, 22, 3989–3996. [Google Scholar] [CrossRef]
  134. Toyoda, S.; Mutobe, H.; Yamagishi, H.; Yoshida, N.; Tanji, Y. Fractionation of N2O isotopomers during production by denitrifier. Soil Biol. Biochem. 2005, 37, 1535–1545. [Google Scholar] [CrossRef]
  135. Schmidt, I.; van Spanning, R.J.; Jetten, M.S. Denitrification and ammonia oxidation by Nitrosomonas europaea wild-type, and NirK- and NorB-deficient mutants. Microbiology 2004, 150, 4107–4114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Ma, P.; Li, X.; Chen, F.; Liu, S.; Hou, C. The isotopomer ratios of N2O in the Shaying River, the upper Huai River network, Eastern China: The significances of mechanisms and productions of N2O in the heavy ammonia polluted rivers. Sci. Total Environ. 2019, 687, 1315–1326. [Google Scholar] [CrossRef]
  137. Heil, J.; Wolf, B.; Brüggemann, N.; Emmenegger, L.; Tuzson, B.; Vereecken, H.; Mohn, J. Site-specific 15N isotopic signatures of abiotically produced N2O. Geochim. Cosmochim. Acta 2014, 139, 72–82. [Google Scholar] [CrossRef]
  138. Li, Y.H.; Peng, T.H.; Broecker, W.S.; Ostlund, H.C. The average vertical mixing coefficient for the oceanic thermocline. Tellus B Chem. Phys. Meteorol. 1984, 36, 212. [Google Scholar] [CrossRef] [Green Version]
  139. Toyoda, S.; Kuroki, N.; Yoshida, N.; Ishijima, K.; Tohjima, Y.; Machida, T. Decadal time series of tropospheric abundance of N2O isotopomers and isotopologues in the Northern Hemisphere obtained by the long-term observation at Hateruma Island, Japan. J. Geophys. Res. Atmos. 2013, 118, 3369–3381. [Google Scholar] [CrossRef]
  140. Yoshida, N.; Toyoda, S. Constraining the atmospheric N2O budget from intramolecular site preference in N2O isotopomers. Nature 2000, 405, 330–334. [Google Scholar] [CrossRef]
Figure 1. Trends of global population, fertilizer application, and global riverine N2O emission. Riverine N2O emission data from Yao et al., 2020 [14]; global population historical data from world-population; global fertilizer application data from FAOSTAT.
Figure 1. Trends of global population, fertilizer application, and global riverine N2O emission. Riverine N2O emission data from Yao et al., 2020 [14]; global population historical data from world-population; global fertilizer application data from FAOSTAT.
Atmosphere 14 01291 g001
Figure 3. Schematic diagram of biological and abiotic pathways of N2O production.
Figure 3. Schematic diagram of biological and abiotic pathways of N2O production.
Atmosphere 14 01291 g003
Figure 4. Diagram of N2O flux measurement with static floating chamber (a), dynamic floating chamber (b), and water–air gas exchange model (c). The red arrow in Figure (b) represents the direction of the gas flow in the carrier pipe, and the “1” represents the electric fan in the chamber, which allows the gas to flow.
Figure 4. Diagram of N2O flux measurement with static floating chamber (a), dynamic floating chamber (b), and water–air gas exchange model (c). The red arrow in Figure (b) represents the direction of the gas flow in the carrier pipe, and the “1” represents the electric fan in the chamber, which allows the gas to flow.
Atmosphere 14 01291 g004
Table 1. Equations for predicting the k600 (m d−1) based on stream velocity (V, in m s−1), slope (S, unitless), depth (D, in meters), discharge (Q, in m3 s−1), the Froude number (Fr = V/(gD)0.5), and wind speed at a 10 m height (W, in m s−1). All p-values for the regressions are 0.001. “-” indicates no measurement.
Table 1. Equations for predicting the k600 (m d−1) based on stream velocity (V, in m s−1), slope (S, unitless), depth (D, in meters), discharge (Q, in m3 s−1), the Froude number (Fr = V/(gD)0.5), and wind speed at a 10 m height (W, in m s−1). All p-values for the regressions are 0.001. “-” indicates no measurement.
Model EquationR2Slopey-InterceptRef.
k600 = (VS)0.89 × D0.54 × 50370.720.92 ± 0.0240.98 ± 0.17[8]
k600 = 5937 × (1–2.54 × Fr2) × (VS)0.89 × D0.580.760.94 ± 0.0220.76 ± 0.16
k600 = 1162 × V0.85S0.770.540.91 ± 0.0360.91 ± 0.24
k600 = (VS)0.76 × 951.50.530.82 ± 0.0370.92 ± 0.24
k600 = VS × 2841 + 2.020.551.0 ± 0.038−4.8 × 10−3 ± 0.26
k600 = 929 × (VS)0.75 × Q0.0110.530.92 ± 0.0360.81 ± 0.24
k600 = 4725 × (VS)0.86 × D0.66 × Q−0.140.760.95 ± 0.0230.57 ± 0.17
k600 = 1.91 × e0.35W---[80]
k600 = 0.314 × W2 − 0.436 × W + 3.99---[81]
k600 = 1.0 + 1.719 × (V/D)0.5 + 2.58 × W---[82]
k600 = 17.19 × V0.5 × D−0.5 + 2.58 × W + 1.0---[83]
Table 2. N2O fluxes and EFs of different rivers.
Table 2. N2O fluxes and EFs of different rivers.
RegionsAquatic EcosystemsN2O Flux
(μmol·m−2·d−1)
NH4+ (mg/L)NO3
(mg/L)
EFs
(%)
Ref.
Southeast ChinaMin River Basin−0.84–3.120.01–2.31.9–11.90.043–0.93 a[103]
Estuary of Min River−2.9–4.30.01–0.20.4–2.70.029–0.25 a[104]
Northeast ChinaLiao River Basin9.36–8539.20. 2–15.50.02–9.60.01–2.2 b[98]
Daliao River and Estuary4.6–145.1---[105]
North ChinaLower Haihe River Basin−7.2–160.80.1–0.20.3–0.50.20–0.39 a[106]
Beitang Drainage River and Dagu Drainage River44.6 ± 39.4 3.3–10.6-0.007–0.1 b[107]
Duliujian River, Yongdingxin River, and Nanyuhe River26.9 ± 31.2 3.3–10.6-0.003–0.09 b[107]
Eastern ChinaMiddle and lower reaches of the Yangtze River0.1–35.00.2–0.32.6–6.90.033–0.053 a[22]
Chaohu River Basin 0.4–2102.60.3–12.50.5–0.70.027–0.80 a[88]
Xin’an Tang River in Taihu region35.3–86.70.5–1.23.0–7.20.040–0.044 a[108]
Southwest ChinaChongqing metropolitan river network4.5–1566.8-3.7 ± 2.50.47 a[99]
Qingshui Stream, Taohua Stream, Panxi Stream, Xiaojia River, Fenghuang Stream, and Jiuqu River in Chongqing10–49400.1–1.61.1–9.60.14–0.29 a[15]
Western ChinaLancang River7.70–26.000.18–0.450.26–0.530.41–0.61 a[109]
The upper reaches of Yangtze River, Yellow River, Lancang River, and Nu River9.4 ± 6.2--0.1 a[68]
South ChinaShenzhen River, Dashahe River, Xixiang River-0–9.60.3–15.80.0028–0.44 a[110]
Estuary of Pearl River31.9 ± 7.5---[76]
37 ± 150–8.460–14.260.13 a[111]
USASan Joaquin River22.6–177.4-0.2–15.50.003–0.21 a[67]
Kalamazoo River−7.7–228.70–0.40.003–27.40.05–0.014 a[42]
Connecticut River28.90.03–0.070.5–2.30.09–0.44 a[84]
CanadaGrand River−35–4200---[68]
Ontario Streams0.48–199.20.002–0.40.5–12.8-[101]
SwedenUppsala watershed92.5–150<0.14.4–22.10.11 a[100]
JapanTama River-0.4–0.78.0–30.90.019 a[94]
FranceHaut-Loir watershed-0.03–0.0527.9–50.50.014–0.095 a[112]
Seine River2.3–193.2-22.1–310.01–0.028 a[95]
Sub-Saharan
Africa
Congo River−52–3190.040.40.0024 a[113]
KenyaMara River13.70.02–0.190.65–1.770.031–0.04 a[114]
MalaysiaLupar and Saribas Rivers21.30.012–0.120.3–2.30.1–1.19 a[48]
South AsiaCochin Estuary0.5–29.70.02–0.50.4–2.60.21–0.37 a[115]
Adyar River and Estuary0.24–122.47.6–770.3–2.80.03–1.76 a[116]
UKUpper Thurne River129.80.47.10.11 a[91]
Wensum, Eden, and Avon Rivers50.0-19.8–37.70.006–0.24 a[102]
New
Zealand
LII River20.6–273.4-11.1–23.50.023–0.048 a[75]
Ashburton River8.6–16.3-0.09–6.020.17–1.69 a[117]
a The N2O concentration in water divided by the NO3 concentration. b The N2O concentration in water divided by the concentration of NO3 and NH4+. “-” indicates no measurement.
Table 3. Estimates of global aquatic emissions of N2O by different models or methods in units of Tg N yr−1. “-” indicates no measurement.
Table 3. Estimates of global aquatic emissions of N2O by different models or methods in units of Tg N yr−1. “-” indicates no measurement.
N2O EmissionNaturalAnthropogenicTotalModels/MethodsRef.
Global9.7 (8.0–12.0)7.3 (4.2–11.4)17Process-based model (DLEM)[79]
Inland and coastal waters0.3 (0.3–0.4)0.5 (0.2–0.7)0.8Process-based model (DLEM)[79]
Rivers and estuaries 0.071.251.32Semi-empirical model (NEWS)[23]
Continental shelves0.50.10.6Semi-empirical model (NEWS)[23]
Estuaries --0.148–0.277Process-based N balance model[120]
Rivers, reservoirs, lakes, ponds, streams--0.94Empirical Monte Carlo simulation[10]
Estuaries--0.26Empirical Monte Carlo simulation[10]
Riverine--0.3 ± 0.06Process-based model (DLEM) [14]
Riverine--0.073Semi-empirical model[9]
Riverine--0.03–0.035Empirical multiple regression model[90]
Riverine0.1050.9451.05Semi-empirical model (NEWS)[124]
Oceans3.4 (2.5–4.3)0.1 (0.1–0.2)3.5Process-based model (DLEM)[79]
Oceans4.2 ± 1.0-4.2 ± 1.0Empirical Random Forest model[26]
Oceans--4.5 ± 1.0 Bern3D Earth System model[25]
Oceans--2.45± 0.8Semi-empirical N cycle model [30]
Oceans3.503.5Semi-empirical model (NEWS)[23]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pan, H.; Zhou, Z.; Zhang, S.; Wang, F.; Wei, J. N2O Emissions from Aquatic Ecosystems: A Review. Atmosphere 2023, 14, 1291. https://doi.org/10.3390/atmos14081291

AMA Style

Pan H, Zhou Z, Zhang S, Wang F, Wei J. N2O Emissions from Aquatic Ecosystems: A Review. Atmosphere. 2023; 14(8):1291. https://doi.org/10.3390/atmos14081291

Chicago/Turabian Style

Pan, Huixiao, Zheyan Zhou, Shiyu Zhang, Fan Wang, and Jing Wei. 2023. "N2O Emissions from Aquatic Ecosystems: A Review" Atmosphere 14, no. 8: 1291. https://doi.org/10.3390/atmos14081291

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop