Next Article in Journal
Physicians’ and Nurses’ Knowledge in Palliative Care: Multidimensional Regression Models
Previous Article in Journal
Resilience Coping in Preschool Children: The Role of Emotional Ability, Age, and Gender
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Can N Fertilizer Addition Affect N2O Isotopocule Signatures for Soil N2O Source Partitioning?

1
School of Geography Science, Nanjing Normal University, Nanjing 210023, China
2
Jiangsu Center for Collaborative Innovation in Geographical Information Resource Development and Application, Nanjing Normal University, Nanjing 210023, China
3
Key Laboratory of Virtual Geographic Environment (Nanjing Normal University), Ministry of Education, Nanjing 210023, China
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2021, 18(9), 5024; https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph18095024
Submission received: 14 February 2021 / Revised: 22 April 2021 / Accepted: 27 April 2021 / Published: 10 May 2021

Abstract

:
Isotopocule signatures of N2O (δ15Nbulk, δ18O and site preference) are useful for discerning soil N2O source, but sometimes, N fertilizer is needed to ensure that there is enough N2O flux for accurate isotopocule measurements. However, whether fertilizer affects these measurements is unknown. This study evaluated a gradient of NH4NO3 addition on N2O productions and isotopocule values in two acidic subtropical soils. The results showed that N2O production rates obviously amplified with increasing NH4NO3 (p < 0.01), although a lower N2O production rate and an increasing extent appeared in forest soil. The δ15Nbulk of N2O produced in forest soil was progressively enriched when more NH4NO3 was added, while becoming more depleted of agricultural soil. Moreover, the N2O site preference (SP) values collectively elevated with increasing NH4NO3 in both soils, indicating that N2O contributions changed. The increased N2O production in agricultural soil was predominantly due to the added NH4NO3 via autotrophic nitrification and fungal denitrification (beyond 50%), which significantly increased with added NH4NO3, whereas soil organic nitrogen contributed most to N2O production in forest soil, probably via heterotrophic nitrification. Lacking the characteristic SP of heterotrophic nitrification, its N2O contribution change cannot be accurately identified yet. Overall, N fertilizer should be applied strictly according to the field application rate or N deposition amount when using isotopocule signatures to estimate soil N2O processes.

1. Introduction

The nitrous oxide (N2O) emitted from soil is governed by various pathways, which often occur simultaneously in different soil micro-sites [1,2]. To adopt appropriate mitigation strategies, attribution of the source of emitted N2O is very important. Recently, a position-specific nitrogen (N) isotope method measuring the intramolecular distribution of 15N in N2O (site preference, SP) has served as a useful tool to source partition N2O in various soils [3,4,5,6]. It has clear advantages, such as the minimal disturbance of soil over the 15N tracing method, independence of the isotopic signature of the substrate over the traditional natural abundance (15N, 18O) method and applicability in spatial–temporal scales with a low cost [7,8,9].
However, compared with bulk 15N (δ15Nbulk) and 18O (δ18O) measurements, obtaining highly accurate SP measurements is challenging, not only because it is indirectly determined by bulk 15N and α-15N (central position), but also because N2O isotopocules overlapping and 15N scrambling in an ion source both propagate analytical errors [5,10]. To ensure precise and robust isotopocule measurements, this method is mainly performed in soils with a high N2O concentration, such as agricultural and grassland soil, while it is seldom utilized in forest soil due to the relatively small N2O flux [3]. Only some tropical and frozen forest soils producing a significantly high N2O flux have been a constrained N2O source via this method [11,12,13]. To obtain a sufficient N2O quantity for isotopocule analysis, some researchers used an improved chamber technique with a molecular sieve to continuously trap N2O in field experiments [14], while others added N fertilizers (KNO3, urea, NH4NO3, etc.) to increase N2O flux in incubation experiments [15,16,17,18]. N fertilizer could enhance N2O production, but excess fertilizer might induce a priming effect, change the N2O efflux and bias the N2O source [19]. Agricultural and grassland soil is frequently applied with N fertilizer, commonly according to its actual application amount [3], but the appropriate N fertilizer application ratio in forest soil is hard to estimate. In the literature, the lowest application rate was 52.4 mg urea-N kg−1 (in line with the N deposition amount) in tropical lowland forest soil [15] and the highest rate was up to 1470 mg KNO3-N Kg−1 in temperate forest soil [18]. Such high application rates are comparable to that in agricultural soil, but whether they influence N2O isotopocule signatures and the subsequent N2O origin analysis has never been clearly defined. The purpose of this study was to investigate the effect of N fertilizer application on the isotopocule signatures of the N2O produced in forest soil and determine its potential influence on N2O source partition. Moreover, we selected an acidic forest soil in subtropical China that has been reported to have large N2O emissions [20,21], but its isotopocule values have not been reported.
Our aim was to test whether isotopocule values of soil-emitted N2O (δ15Nbulk, δ18O and SP) differed and whether the contributions of N2O pathways changed after applying a gradient of NH4NO3. Two types of acidic soil (agricultural and forest) in a subtropical area were selected.

2. Materials and Methods

2.1. Soil Properties

The subtropical soils investigated were sampled from Jiangxi Province (27°59′N, 117°25′E), China, in May 2018, of which one was agricultural soil (with three replicates collected from three upland agricultural areas) and the other was forest soil (with three replicates). The total N2O emissions estimated from agronomy (maize, rice, wheat and vegetable crop) in Jiangxi Province were 0.03, 1.55, 0.01 and 1.13 Gg N yr1, respectively [22], while the annual cumulative N2O flux from forest was 0.95 ± 0.21 kg N ha1yr1 on average [23]. The dominant vegetation in forest soil was Pinus massoniana and Cunninghamia lanceolate. The agricultural soil was upland maize soil, which was established by clearing the native forest and has been applied at about 200–300 kg N ha−1 per year for 10 years. The two types of soil were different in parameters associated with N2O turnover, i.e., soil organic C and total nitrogen. The sampled area was a typical subtropical climate region, where the mean annual precipitation and temperature were 1785 mm and 18.4 °C, respectively. The surface soils (0–20 cm) were sampled at three randomly arranged places. Each type of soil sample was sieved through a 2-millimeter screen, mixed for homogeneity and then stored at 4 °C prior to incubation.
A 50-gram subsample of each soil type was air-dried for basic physicochemical properties’ investigation (Table 1). Soil pH was analyzed in a soil:water ratio of 1:2.5 (w/v). SOC, total N and C/N ratio were all determined by a CN element analyzer (Sercon Europa EA-GSL, UK). NH4-N and NO3-N were extracted using 2 M KCl in a soil:solution ratio of 1:5 (w/v) and analyzed by a continuous-flow analyzer (SA1000, Skalar, The Netherlands).

2.2. Experiments to Determine Effect of NH4NO3 Addition on Isotopocule Ratios of N2O

For each sample, 150 g fresh soil (oven-dry basis) was packed into a 500-milliliter conical flask. The soils were applied with 0, 20, 40, 80 and 160 mg N kg−1 soil NH4NO3 uniformly. This gradient of NH4NO3 was set according to the rates most frequently applied in soil experiments using N2O isotopocules. Then soils were adjusted to 60% maximum water-holding capacity using deionized water. Each treatment had 5 replicates. All flasks were sealed by rubber stops with plastic tubes as inlet and outlet flow lines. Before incubation and after each gas sampling, the flasks with soil samples were vacuumed and filled with synthetic air (N2O free) three times. The soils were incubated for 48 h at 25 °C and gas samples were collected at 24 h intervals. A 5-milliliter gas sample was collected from each flask for N2O concentration measurements. Another 50-milliliter gas sample was collected into a pre-evacuated bottle to determine the isotopocule ratios. The δ15Nbulk, δ18O and SP values from the gas samples were analyzed by a Delta V plus IRMS (Thermo Fisher Scientific, China), which was equipped with five cups to analyze both m/z 44, 45 and 46 of N2O molecules and m/z 30 and 31 of NO+ fragments. The scrambling factor in the ion source of this IRMS (0.085) was determined, as Röckmann [24] reported, using a series of β-labeled N2O. The laboratory’s N2O working standard was two-point, calibrated by two standards kindly offered by Dr. Reinhard Well and Dr. Anette Giesemann (Thünen Institute of Climate-Smart Agriculture, Germany). The nonlinear effect caused by various sample N2O amounts was corrected by a series of different standard gas mol fractions (0.3, 1, 5, 10 and 20 ppm), which were analyzed within each sample run. The δ15Nα, δ15Nbulk and SP were calculated according to Equations (1) and (2) [5]. The typical analytical precision for δ15Nbulk, δ18O and SP in this IRMS was 0.3, 0.6 and 0.9‰, respectively. The N2O concentrations were determined by a gas chromatograph fitted with an electron capture detector (Agilent 7890).
  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

2.3. Statistical Analysis

All statistical analyses were applied using SPSS 19. The data were checked for normality (Shapiro–Wilk test and Q–Q plot) and homogeneity of variances (Levine’s test) before statistical analysis. The N2O flux was log10 transformed to obtain a normal distribution. Then, one-way ANOVA analysis and the least significant difference (Tukey) at a level of p < 0.05 were used to compare the N2O production rates and isotopic signatures of N2O under different levels of fertilization.

2.4. N2O Source Partition by the Two-End-Member Mixing Approach

The two-end-member mixing model [25] was used to estimate the contributions of soil-emitted N2O by denitrifier denitrification/nitrifer denitrification (lower SP values) and by autotrophic nitrification/fungal denitrification (higher SP values). Therefore, four cases were considered: Case 1 is autotrophic nitrification versus nitrifier denitrification, Case 2 is autotrophic nitrification versus denitrifier denitrification, Case 3 is fungal denitrification versus nitrifier denitrification and Case 4 is fungal denitrification versus denitrifier denitrification. As the SP values in this study were not low, we excluded the simultaneous occurrence of denitrifier denitrification and nitrifier denitrification, as per Zou et al. [25]. The contribution of N2O (x) from denitrifier denitrification in Case 1 can be calculated as Equation (3). If N2O reduction happens, its SP will increase along a slope of 1.2 ± 0.5 [26]. The contribution of each end member can be calculated by the intersection of the reduction line and the mixing line using Equation (3). The calculations of each end member’s contribution in Cases 2–4 are calculated as in Case 1. Admittedly, this method might overestimate the contributions from the two end members in each case.
S P s a m p l e = x S P d e n i t r i f i e r   d e n i t r i f i c a t i o n + ( 1 x ) S P a u t o t r o p h i c   n i t r i f i c a t i o n

3. Results

3.1. Soil Properties

The soil physical and chemical properties prior to the experiment are reported in Table 1. Both types of soil were acidic with a pH ≤ 4.8, and their dominant inorganic N was NO3. The soil organic carbon (SOC) and the C/N ratio in forest soil were twice as high as those in agricultural soil. However, higher NH4+ and NO3 concentrations were observed in agricultural soil, whose NO3 (60.7 mg kg−1) was even over four times higher than that in forest soil (13.8 mg kg−1).

3.2. N2O Production Rates

The N2O production rates in both agricultural and forest soils obviously increased with increasing NH4NO3 addition (p < 0.01), while the extent of its increase in agricultural soil was much greater than that in forest soil (Table 2 and Table 3). The N2O production rate in agricultural soil was 3.0 μg kg−1 d−1 without NH4NO3 addition; then, it significantly elevated by about 25% when 20 mg N kg−1 soil N was added (p < 0.01) and doubled (6.0 μg kg−1 d−1) when 160 mg N kg−1 soil N was added (p < 0.01). In forest soil, the N2O production rate was comparatively lower and slowly growing with increasing added NH4NO3. It was 1.8 μg kg−1 d−1 without NH4NO3 addition, gradually increased to 2.0 μg kg−1 d−1 when 40 mg N kg−1 soil N was added and stayed the same with more NH4NO3 added.

3.3. Isotopocule Values of N2O

The δ15Nbulk of N2O emitted in agricultural soil generally became progressively lighter with more NH4NO3 added, while it became gradually enriched in forest soil (Table 2 and Table 3). It was -26.4‰ in agricultural soil without adding NH4NO3, then decreased to −32.2‰ when 20 mg N kg−1 soil N was added and further depleted to −38.1‰ after applying 160 mg N kg−1 soil N. Negative correlations between the NH4NO3 level and δ15Nbulk values were observed in agricultural soil (r = −0.759, p < 0.001). The δ15Nbulk of N2O emitted in forest soil was higher than that in agricultural soil across 0–160 mg N kg−1 soil levels. It was −14.6‰ without adding NH4NO3 and slowly increased to −10.7‰ with addition of 160 mg N kg−1 soil N. Therefore, the difference in δ15Nbulk values of N2O emitted between agricultural and forest soil was about 12‰ without adding NH4NO3 and increased to 28‰ when 160 mg N kg−1 soil N was added. Compared with the obvious change in δ15Nbulk values, δ18O values of N2O in agricultural soil only slightly fluctuated with increasing NH4NO3, while they showed a slow growth trend in forest soil.
The SP values of N2O emitted in agricultural soil and forest soil slightly but significantly increased with increasing NH4NO3 addition (Table 2 and Table 3). In agricultural soil, the increment was not significant with 20, 40 and 80 mg N kg−1 soil addition (p > 0.05), except the 160 mg N kg−1 soil addition. Compared with agricultural soil, the effect of NH4NO3 addition was more obvious in forest soil. Its SP values significantly increased when 80 mg N kg−1 soil was applied (p < 0.05). Relative to agricultural soil, the SP values of N2O in forest soil were generally 10‰ lower across 0–160 mg N kg−1 soil levels.

3.4. N2O Source Contributions

The isotopocule values of N2O emitted in agricultural and forest soil are shown in the isotopocule map (Figure 1). All of the samples collected from agricultural soil were located between the mixing zones of the four processes, and their distributions were very close to the autotrophic nitrification and fungal denitrification processes. However, all of the samples obtained from forest soil were located outside the four processes’ mixing zone and they did not show a clear preference for each process. The two-end-member mixing results (Table 4), based on the δ15Nbulk-SP map, showed that the contributions of various N2O processes changed when more NH4NO3 was added. Autotrophic nitrification and fungal denitrification contributed just slightly more to N2O emission in agricultural soil without NH4NO3 but became dominant N2O pathways when amounts of 40 and 80 mg N kg−1 soil NH4NO3 were applied. With 160 mg N kg−1 soil NH4NO3, their contributions modestly decreased but were still higher than those without NH4NO3 addition in some cases. In forest soil, the contributions of N2O processes in Cases 2 and 4 clearly changed with the increasing NH4NO3 addition.

4. Discussion

In this study, we have shown that N fertilizer addition led to an obvious change in the N2O flux, N2O isotopocule signatures and N2O source contributions in agricultural and forest soil incubation experiments.
Earlier findings showed that N fertilizer addition in soil could increase the N2O production rate [27,28,29]. Our results reinforced that added NH4NO3 significantly enhanced the N2O production rate in both agricultural and forest soils with different magnitudes. However, with the increasing N2O flux in agricultural and forest soil, their isotopocule signatures obviously changed with NH4NO3 addition as well. The δ15Nbulk values in the two soils were very negative but exhibited opposite trends with increasing NH4NO3. In fact, the δ15Nbulk values in both soils were within the reported isotopocule signature ranges in the literature (−67.5 to 4.2‰ in forest soil and −66.7 to 6.0‰ in agricultural soil) [3]. Soil incubation studies commonly reported skew discrimination of 15N because substrate diffusion is not a limiting factor and microbial N2O production can get close to the maximum apparent isotope effect [3]. In the agricultural soil, the sufficient NH4+and NO3 substrates provided by NH4NO3 addition amplified such a 15N discrimination in N2O production, so its δ15Nbulk values were more depleted with the enhancing NH4NO3 level. In forest soil, our previous 15N tracing studies found that N2O was produced mainly from an organic N pool [21,30], which was not directly supplemented by added NH4NO3. Therefore, the 15N discrimination was reduced and the δ15Nbulk values were progressively enriched with increasing NH4NO3 when more organic N was consumed to produce N2O. Compared with the profound change in δ15Nbulk values, δ18O showed moderate changes in both soils. This is probably because the O isotopic composition of N2O not only depends on the substrate compounds but also on the O2 involved in ammonium/hydroxylamine oxidation and O exchange with H2O in denitrification [5,31].
It was somewhat surprising that the SP values of the produced N2O in the two types of soil collectively elevated with increasing NH4NO3. The enhancing SP probably indicated changing contributions from various N2O pathways, because SP is independent of the isotopic signatures of a substrate. The SP values of N2O production pathways can be divided into two groups: bacterial nitrification (average 31.4‰) and fungal denitrification (average 37‰) are specified with a higher SP, while nitrifier denitrification (average −3.8‰) and bacterial denitrifier denitrification (average −2‰) are characterized by a lower SP [8]. The high SP values (24.9~28‰) in the agricultural soil suggested that autotrophic nitrification or fungal denitrification contributed the most to N2O emission with or without NH4NO3 addition, but their contributions were obviously amplified when more NH4NO3 was applied (Table 4). It seemed that NH4NO3 addition would overestimate the influence of the two processes in agricultural soil. The SP values in forest soil (15~18.4‰) were relatively low, but all were located outside the mixing zone of the four processes (Figure 1). This phenomenon might occur when a large amount of N2O reduces to N2, or other N2O pathways whose SP values have not been illustrated as contributing the most. Since N2O and NO account for 80% of denitrification gas products even under very anaerobic conditions in this soil [32], it is probably heterotrophic nitrification that plays a dominant role in soil N2O production. Due to lacking the SP signature of heterotrophic nitrification, the two-end-member mixing model results were based on only four processes that might overestimate the impact of the denitrification processes. However, obvious contribution shifts in Cases 2 and 4 were observed after the NH4NO3 addition in forest soil.
Furthermore, we can use isotopocule and N2O flux data to investigate how exogenous NH4NO3 input alters N2O-producing processes. In the agricultural soil, the positive increase in the autotrophic nitrification and fungal denitrification was the result of NH4NO3 addition, which provided more NH4+ substrates for nitrifying bacteria and NO3 substrates for denitrifying fungal [33]. Therefore, its N2O production rates exhibited a positive elevation with added NH4NO3, but its δ15Nbulk of N2O became more depleted. In the forest soil, the N2O production rate increased when 20 and 40 mg N kg−1 NH4NO3 were added but did not further increase when more NH4NO3 was added. This indicated that a different priming mechanism was occurring in the forest soil. Some studies reported that increased labile N can trigger carbon limitations in microbes and then stimulate more extracellular enzyme production to break soil organic matter (SOM) to access SOM-C [34]. The simultaneously released SOM-N and SOM-C might provide available substrates for subsequent N2O emission [35]. However, it is hardly to determine which N2O pathway contributing most to the process without the SP signatures of heterotrophic nitrification. The only certainty is that N2O is mainly derived from SOM-N, because its δ15Nbulk showed an opposite increasing pattern with more NH4NO3 addition relative to agricultural soil.
Compared with the application rates of N fertilizer in the literature (ranging from 20 to 1600 mg N kg−1 soil), our added NH4NO3 rates were not high, but they still significantly altered the N2O isotopocule signatures and the N2O source contributions in both soils. N fertilizer can amplify N2O flux to reach the detection limits for accurate isotopocule measurements, but its impact on soil N2O production processes cannot be ignored. In our study, the contributions of autotrophic nitrification and fungal denitrification remarkably increased when only 20 mg N kg−1 soil NH4NO3 was applied in agricultural soil. In forest soil, added NH4NO3 acted as an external stimulus to produce more N2O from SOM, although its exact contribution shift was temporarily incalculable. In the few studies in the literature that used an isotopocule method to investigate forest soil N2O emission, much higher application rates (500 or 1470 mg N kg−1 soil) were applied in soil incubation experiments [15,18]. Such high inputs are apparently larger than most in agricultural or grassland soil. They would disturb natural forest soil ecosystems, but their impacts can vary depending on different soil ecosystems. Therefore, we propose that N fertilizer should be applied according to its real application rate in agricultural soil and should be avoided in forest soil. For those N deposition studies, N fertilizer should be applied strictly according to the real N deposition amount. To enlarge the N2O flux, increasing the incubation size of soil may be an appropriate alternative.

5. Conclusions

In conclusion, while only two types of soil from an acidic subtropical area were involved in this study, our data suggest that adding NH4NO3 significantly increased N2O production, changed N2O isotopocule signatures and altered N2O source contributions. The increased N2O production in the agricultural soil was predominantly derived from added NH4NO3, while it mainly came from SOM-N in the forest soil. Overall, the results presented here provide a basis for conducting soil incubation experiments for N2O source partition using an isotopocule method. N fertilizer should be applied according to its field application rate in agricultural soil, while it should be avoided or applied based on the N deposition amount in forest soil. As an alternative, amplifying the soil incubation size would help to achieve enough N2O flux for isotopocule measurements. Since heterotrophic nitrification is a major N2O source in acidic subtropical forest soil, further experiments are needed to elucidate its isotopocule signatures.

Author Contributions

Investigation, writing—original draft preparation, P.Z. and Y.H.; conceptualization, writing—reviewing and editing, T.W.; editing and reviewing the manuscript, J.Z. and Z.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China (grant nos. 41977084 and 41501254) and a project funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions.

Institutional Review Board Statement

This study did not require ethical approval.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

There is no supplemental data in this manuscript.

Acknowledgments

We sincerely thank Reinhard Well and Anette Giesemann for invaluable assistance in N2O isotopocule method establishment.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Denk, T.R.; 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]
  2. Wu, X.; Zang, S.; Ma, D.; Ren, J.; Chen, Q.; Dong, X. Emissions of CO2, CH4, and N2O Fluxes from Forest Soil in Permafrost Region of Daxing’an Mountains, Northeast China. Int. J. Environ. Res. Public. Health 2019, 16, 2999. [Google Scholar] [CrossRef] [Green Version]
  3. Toyoda, S.; Yoshida, N.; Koba, K. Isotopocule Analysis of Biologically Produced Nitrous Oxide in Various Environments. Mass Spectrom. Rev. 2015, 36, 135–160. [Google Scholar] [CrossRef]
  4. Well, R.; Flessa, H.; Xing, L.; Ju, X.T.; Romheld, V. Isotopologue Ratios of N2O Emitted from Microcosms with NH4+ Fertilized Arable Soils under Conditions Favoring Nitrification. Soil Biol. Biochem. 2008, 40, 2416–2426. [Google Scholar] [CrossRef]
  5. Yu, L.; Harris, E.; Lewicka-Szczebak, D.; Barthel, M.; Blomberg, M.R.A.; Harris, S.J.; Johnson, M.S.; Lehmann, M.F.; Liisberg, J.; Müller, C.; et al. What Can We Learn from N2O Isotope Data?—Analytics, Processes and Modelling. Rapid Commun. Mass Spectrom. 2020, 34, e8858. [Google Scholar] [CrossRef] [PubMed]
  6. Bracken, C.J.; Lanigan, G.J.; Richards, K.G.; Müller, C.; Tracy, S.R.; Well, R.; Carolan, R.; Murphy, P.N.C. Development and Verification of a Novel Isotopic N2O Measurement Technique for Discrete Static Chamber Samples Using Cavity Ring-down Spectroscopy. Rapid Commun. Mass Spectrom. 2021, 35, e9049. [Google Scholar] [CrossRef]
  7. 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] [PubMed]
  8. Decock, C.; Six, J. How Reliable Is the Intramolecular Distribution of 15N in N2O to Source Partition N2O Emitted from Soil? Soil Biol. Biochem. 2013, 65, 114–127. [Google Scholar] [CrossRef]
  9. Yoshida, N.; Toyoda, S. Constraining the Atmospheric N2O Budget from Intramolecular Site Preference in N2O Isotopomers. Nature 2000, 405, 330–334. [Google Scholar] [CrossRef]
  10. Mohn, J.; Wolf, B.; Toyoda, S.; Lin, C.T.; Liang, M.C.; Bruggemann, N.; Wissel, H.; Steiker, A.E.; Dyckmans, J.; Szwec, L.; et al. Interlaboratory Assessment of Nitrous Oxide Isotopomer Analysis by Isotope Ratio Mass Spectrometry and Laser Spectroscopy: Current Status and Perspectives. Rapid Commun. Mass Spectrom. 2014, 28, 1995–2007. [Google Scholar] [CrossRef]
  11. Goldberg, S.D.; Borken, W.; Gebauer, G. N2O Emission in a Norway Spruce Forest Due to Soil Frost: Concentration and Isotope Profiles Shed a New Light on an Old Story. Biogeochemistry 2010, 97, 21–30. [Google Scholar] [CrossRef]
  12. Pérez, T.; Garcia-Montiel, D.; Trumbore, S.; Tyler, S.; de Camargo, P.; Moreira, M.; Piccolo, M.; Cerri, C. Nitrous Oxide Nitrification and Denitrification 15N Enrichment Factors from Amazon Forest Soils. Ecol. Appl. 2006, 16, 2153–2167. [Google Scholar] [CrossRef] [Green Version]
  13. Pérez, T.; Trumbore, S.E.; Tyler, S.C.; Davidson, E.A.; Keller, M.; de Camargo, P.B. Isotopic Variability of N2O Emissions from Tropical Forest Soils. Glob. Biogeochem. Cycles 2000, 14, 525–535. [Google Scholar] [CrossRef] [Green Version]
  14. Smemo, K.A.; Ostrom, N.E.; Opdyke, M.R.; Ostrom, P.H.; Bohm, S.; Robertson, G.P. Improving Process-Based Estimates of N2O Emissions from Soil Using Temporally Extensive Chamber Techniques and Stable Isotopes. Nutr. Cycl. Agroecosyst. 2011, 91, 145–154. [Google Scholar] [CrossRef]
  15. Koehler, B.; Corre, M.D.; Steger, K.; Well, R.; Zehe, E.; Sueta, J.P.; Veldkamp, E. An In-Depth Look into a Tropical Lowland Forest Soil: Nitrogen-Addition Effects on the Contents of N2O, CO2 and CH4 and N2O Isotopic Signatures down to 2-m Depth. Biogeochemistry 2012, 111, 695–713. [Google Scholar] [CrossRef]
  16. Menyailo, O.V.; Hungate, B.A.; Lehmann, J.; Gebauer, G.; Zech, W. Tree Species of the Central Amazon and Soil Moisture Alter Stable Isotope Composition of Nitrogen and Oxygen in Nitrous Oxide Evolved from Soil. Isotopes Environ. Health Stud. 2003, 39, 41–52. [Google Scholar] [CrossRef]
  17. Menyailo, O.V.; Hungate, B.A. Stable Isotope Discrimination during Soil Denitrification: Production and Consumption of Nitrous Oxide. Glob. Biogeochem. Cycles 2006, 20. [Google Scholar] [CrossRef]
  18. Snider, D.M.; Schiff, S.L.; Spoelstra, J. 15N/14N and 18O/16O Stable Isotope Ratios of Nitrous Oxide Produced during Denitrification in Temperate Forest Soils. Geochim. Cosmochim. Acta 2009, 73, 877–888. [Google Scholar] [CrossRef]
  19. Kuzyakov, Y. Priming Effects: Interactions between Living and Dead Organic Matter. Soil Biol. Biochem. 2010, 9, 1363–1371. [Google Scholar] [CrossRef]
  20. Zhang, Y.; Cai, Z.; Zhang, J.; Müller, C. C:N Ratio Is Not a Reliable Predictor of N2O Production in Acidic Soils after a 30-Day Artificial Manipulation. Sci. Total Environ. 2020, 725, 138427. [Google Scholar] [CrossRef] [PubMed]
  21. Zhang, Y.; Zhao, W.; Zhang, J.; Cai, Z. N2O Production Pathways Relate to Land Use Type in Acidic Soils in Subtropical China. J. Soils Sediments 2016, 17, 306–314. [Google Scholar] [CrossRef]
  22. Aliyu, G.; Luo, J.; Di, H.J.; Lindsey, S.; Liu, D.; Yuan, J.; Chen, Z.; Lin, Y.; He, T.; Zaman, M.; et al. Nitrous Oxide Emissions from China’s Croplands Based on Regional and Crop-Specific Emission Factors Deviate from IPCC 2006 Estimates. Sci. Total Environ. 2019, 669, 547–558. [Google Scholar] [CrossRef] [PubMed]
  23. Li, X.; Cheng, S.; Fang, H.; Yu, G.; Dang, X.; Xu, M.; Wang, L.; Si, G.; Geng, J.; He, S. The Contrasting Effects of Deposited NH4+ and NO3 on Soil CO2, CH4 and N2O Fluxes in a Subtropical Plantation, Southern China. Ecol. Eng. 2015, 85, 317–327. [Google Scholar] [CrossRef]
  24. Röckmann, T.; Kaiser, J.; Brenninkmeijer, C.A.; Brand, W.A. Gas Chromatography/Isotope-ratio Mass Spectrometry Method for High-precision Position-dependent 15N and 18O Measurements of Atmospheric Nitrous Oxide. Rapid Commun. Mass Spectrom. 2003, 17, 1897–1908. [Google Scholar] [CrossRef] [PubMed]
  25. Zou, Y.; Hirono, Y.; Yanai, Y.; Hattori, S.; Toyoda, S.; Yoshida, N. Isotopomer Analysis of Nitrous Oxide Accumulated in Soil Cultivated with Tea (Camellia Sinensis) in Shizuoka, Central Japan. Soil Biol. Biochem. 2014, 77, 276–291. [Google Scholar] [CrossRef]
  26. Koba, K.; Osaka, K.; Tobari, Y.; Toyoda, S.; Ohte, N.; Katsuyama, M.; Suzuki, N.; Itoh, M.; Yamagishi, H.; Kawasaki, M.; et al. Biogeochemistry of Nitrous Oxide in Groundwater in a Forested Ecosystem Elucidated by Nitrous Oxide Isotopomer Measurements. Geochim. Cosmochim. Acta 2009, 73, 3115–3133. [Google Scholar] [CrossRef]
  27. Chai, L.L.; Hernandez-Ramirez, G.; Dyck, M.; Pauly, D.; Kryzanowski, L.; Middleton, A.; Powers, L.-A.; Lohstraeter, G.; Werk, D. Can Fertigation Reduce Nitrous Oxide Emissions from Wheat and Canola Fields? Sci. Total Environ. 2020, 745, 141014. [Google Scholar] [CrossRef]
  28. Stehfest, E.; Bouwman, L. N2O and NO Emission from Agricultural Fields and Soils under Natural Vegetation: Summarizing Available Measurement Data and Modeling of Global Annual Emissions. Nutr. Cycl. Agroecosyst. 2006, 74, 207–228. [Google Scholar] [CrossRef]
  29. Fang, K.; Yi, X.; Dai, W.; Gao, H.; Cao, L. Effects of Integrated Rice-Frog Farming on Paddy Field Greenhouse Gas Emissions. Int. J. Environ. Res. Public. Health 2019, 16, 1930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Xu, Y.B.; Xu, Z.H.; Cai, Z.C.; Reverchon, F. Review of Denitrification in Tropical and Subtropical Soils of Terrestrial Ecosystems. J. Soils Sediments 2013, 13, 699–710. [Google Scholar] [CrossRef]
  31. Buchen, C.; Lewicka-Szczebak, D.; Flessa, H.; Well, R. Estimating N2O Processes during Grassland Renewal and Grassland Conversion to Maize Cropping Using N2O Isotopocules. Rapid Commun. Mass Spectrom. 2018, 32, 1053–1067. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, J.; Sun, W.; Zhong, W.; Cai, Z. The Substrate Is an Important Factor in Controlling the Significance of Heterotrophic Nitrification in Acidic Forest Soils. Soil Biol. Biochem. 2014, 76, 143–148. [Google Scholar] [CrossRef]
  33. Bouwman, A.F.; Boumans, L.J.M.; Batjes, N.H. Emissions of N 2 O and NO from Fertilized Fields: Summary of Available Measurement Data: Summary of NO and N2O Measurement Data. Glob. Biogeochem. Cycles 2002, 16. [Google Scholar] [CrossRef]
  34. Daly, E.J.; Hernandez-Ramirez, G. Sources and Priming of Soil N2O and CO2 Production: Nitrogen and Simulated Exudate Additions. Soil Biol. Biochem. 2020, 149, 107942. [Google Scholar] [CrossRef]
  35. Liu, X.-J.A. Labile Carbon Input Determines the Direction and Magnitude of the Priming Effect. Appl. Soil Ecol. 2017, 109, 7–13. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Relations between SP and δ15Nbulk of N2O produced in agricultural soil (a) and forest soil (b) without NH4NO3 addition. The boxes indicate the expected ranges of N2O produced by autotrophic nitrification, fungal denitrification, nitrifier denitrification and denitrifier denitrification. The dash lines denote the mixing zone of Case 1 (autotrophic nitrification and nitrifier denitrification are end members); solid lines denote the mixing zone of Case 2 (nitrifier denitrification and denitrifier denitrification are end members); dash dot lines denote the mixing zone of Case 3 (fungal denitrification and nitrifier denitrification are end members); short dot lines denote the mixing zone of Case 4 (fungal denitrification and denitrifier denitrification are end members). The black circles denote the N2O samples collected in agricultural soil and the black squares denote the N2O samples collected in forest soil.
Figure 1. Relations between SP and δ15Nbulk of N2O produced in agricultural soil (a) and forest soil (b) without NH4NO3 addition. The boxes indicate the expected ranges of N2O produced by autotrophic nitrification, fungal denitrification, nitrifier denitrification and denitrifier denitrification. The dash lines denote the mixing zone of Case 1 (autotrophic nitrification and nitrifier denitrification are end members); solid lines denote the mixing zone of Case 2 (nitrifier denitrification and denitrifier denitrification are end members); dash dot lines denote the mixing zone of Case 3 (fungal denitrification and nitrifier denitrification are end members); short dot lines denote the mixing zone of Case 4 (fungal denitrification and denitrifier denitrification are end members). The black circles denote the N2O samples collected in agricultural soil and the black squares denote the N2O samples collected in forest soil.
Ijerph 18 05024 g001
Table 1. Soil properties of agricultural and forest soil (means ± SD, n = 3).
Table 1. Soil properties of agricultural and forest soil (means ± SD, n = 3).
Soil
Type
pHTN
(g N kg−1)
SOC
(g C kg−1)
C/N
Ratio
NH4-N
(mg kg−1)
NO3-N
(mg kg−1)
Agricultural4.8 ± 0.10.8 ± 0.09.5 ± 0.211.5 ± 0.617.1 ± 1.560.7 ± 2.9
Forest4.6 ± 0.21.0 ± 0.121.2 ± 4.421.5 ± 3.16.3 ± 1.413.8 ± 5.0
Table 2. Isotopocule ratios of N2O flux with different NH4NO3 application rates in agricultural soils.
Table 2. Isotopocule ratios of N2O flux with different NH4NO3 application rates in agricultural soils.
NH4NO3 Application
(mg N kg−1 Soil)
N2O Flux
(μg kg−1 d−1)
δ15Nbulk
(‰)
δ18O
(‰)
SP
(‰)
03.0 ± 0.2a−26.4 ± 1.8d39.0 ± 1.1b24.9 ± 1.3a
203.7 ± 0.2b−32.2 ± 0.9c37.8 ± 0.2a26.1 ± 2.8ab
404.7 ± 0.3c−35.2 ± 1.8b38.2 ± 0.5ab26.6 ± 0.4ab
805.5 ± 0.8d−39.2 ± 1.0a38.1 ± 0.4ab26.4 ± 1.3ab
1606.0 ± 0.5d−38.1 ± 1.4a39.0 ± 0.4bc28.0 ± 1.1b
Identical letters indicate no significant differences in average values. ± represents standard deviation.
Table 3. Isotopocule ratios of N2O flux with different NH4NO3 application rates in forest soils.
Table 3. Isotopocule ratios of N2O flux with different NH4NO3 application rates in forest soils.
NH4NO3 Application
(mg N kg−1 Soil)
N2O Flux
(μg kg−1 d−1)
δ15Nbulk
(‰)
δ18O
(‰)
SP
(‰)
01.8 ± 0.1a−14.6 ± 0.6a 34.5 ± 0.8a15.0 ± 1.4a
201.9 ± 0.2ac−14.0 ± 1.2ab35.5 ± 1.1ab15.0 ± 2.4a
402.0 ± 0.2bc−12.3 ± 0.5bc 35.8 ± 0.5b14.3 ± 2.4a
802.0 ± 0.2c−12.0 ± 1.4c36.5 ± 0.6cd18.0 ± 1.8b
1602.0 ± 0.1c−10.7 ± 0.8c37.3 ± 0.3d18.4 ± 1.1b
Identical letters indicate no significant differences in average values. ± represents standard deviation.
Table 4. Contributions of different pathways of N2O production assuming that N2O from different N2O pathways mixed before reduction.
Table 4. Contributions of different pathways of N2O production assuming that N2O from different N2O pathways mixed before reduction.
Soil
Type
NH4NO3 Application (mg N kg−1 soil)Case 1Case 2Case 3Case 4
Contribution to N2O Production (%)Contribution to N2O Production (%)Contribution to N2O Production (%)Contribution to N2O Production (%)
Bacterial NitrificationNitrifier
Denitrification
Bacterial NitrificationDenitrifier DenitrificationFungal DenitrificationNitrifier DenitrificationFungal DenitrificationDenitrifier Denitrification
Agricultural soil 059 (5)41 (6)70 (6)30 (5)54 (7)46 (5)61 (8)39 (6)
2061 (6)39 (3)76 (6)24 (5)73 (6)27 (5)85 (8)15 (4)
4073 (7)27 (5)83 (8)17 (4)77 (8)23 (6)93 (7)7(4)
8078 (7)22 (6)88 (8)13 (6)100 0 100 (6)0 (6)
16067 (4)33 (4)70 (5)30 (8)55 (7)45 (7)65 (8)35 (7)
Forest soil00 100 29 (5)71 (7)0 100 42 (7)58 (7)
200 100 0 100 0 100 0 100
400 100 0 100 0 100 1 (5)99 (6)
800 100 10 (5)90 (8)0 100 100 0
1600 100 12 (6)88 (8)0 100 18 (4)82 (6)
The uncertainties of contributions are shown in brackets.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, P.; Wen, T.; Hu, Y.; Zhang, J.; Cai, Z. Can N Fertilizer Addition Affect N2O Isotopocule Signatures for Soil N2O Source Partitioning? Int. J. Environ. Res. Public Health 2021, 18, 5024. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph18095024

AMA Style

Zhang P, Wen T, Hu Y, Zhang J, Cai Z. Can N Fertilizer Addition Affect N2O Isotopocule Signatures for Soil N2O Source Partitioning? International Journal of Environmental Research and Public Health. 2021; 18(9):5024. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph18095024

Chicago/Turabian Style

Zhang, Peiyi, Teng Wen, Yangmei Hu, Jinbo Zhang, and Zucong Cai. 2021. "Can N Fertilizer Addition Affect N2O Isotopocule Signatures for Soil N2O Source Partitioning?" International Journal of Environmental Research and Public Health 18, no. 9: 5024. https://0-doi-org.brum.beds.ac.uk/10.3390/ijerph18095024

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