Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2021 Aug 30;27(21):5599–5613. doi: 10.1111/gcb.15833

Nitrification is a minor source of nitrous oxide (N2O) in an agricultural landscape and declines with increasing management intensity

Di Liang 1,2,, G Philip Robertson 1,2
PMCID: PMC9291997  PMID: 34383336

Abstract

The long‐term contribution of nitrification to nitrous oxide (N2O) emissions from terrestrial ecosystems is poorly known and thus poorly constrained in biogeochemical models. Here, using Bayesian inference to couple 25 years of in situ N2O flux measurements with site‐specific Michaelis–Menten kinetics of nitrification‐derived N2O, we test the relative importance of nitrification‐derived N2O across six cropped and unmanaged ecosystems along a management intensity gradient in the U.S. Midwest. We found that the maximum potential contribution from nitrification to in situ N2O fluxes was 13%–17% in a conventionally fertilized annual cropping system, 27%–42% in a low‐input cover‐cropped annual cropping system, and 52%–63% in perennial systems including a late successional deciduous forest. Actual values are likely to be <10% of these values because of low N2O yields in cultured nitrifiers (typically 0.04%–8% of NH3 oxidized) and competing sinks for available NH4+ in situ. Most nitrification‐derived N2O was produced by ammonia‐oxidizing bacteria rather than archaea, who appeared responsible for no more than 30% of nitrification‐derived N2O production in all but one ecosystem. Although the proportion of nitrification‐derived N2O production was lowest in annual cropping systems, these ecosystems nevertheless produced more nitrification‐derived N2O (higher V max) than perennial and successional ecosystems. We conclude that nitrification is minor relative to other sources of N2O in all ecosystems examined.

Keywords: agriculture, ammonia‐oxidizing archaea (AOA), ammonia‐oxidizing bacteria (AOB), forest, greenhouse gas, nitrification, row crop, soil nitrogen


The long‐term contribution of soil nitrification to nitrous oxide (N2O) fluxes is poorly understood, especially under field conditions. In this study, we applied Bayesian inference to couple kinetics of nitrification‐derived N2O with 25 years of in situ N2O flux measurement in six ecosystems along a management intensity gradient. Our results revealed that nitrification (the frequency of the contribution of nitrification to N2O production) is a minor source of N2O in all ecosystems examined, and the relative importance of nitrification in contributing to N2O in situ decreases with increasing management intensity.

graphic file with name GCB-27-5599-g001.jpg

1. INTRODUCTION

Nitrous oxide (N2O) is a potent greenhouse gas with a 100‐year global warming potential ~300 times higher than CO2, and has the third largest radiative forcing among the biogenic greenhouse gases (Myhre et al., 2013). N2O also depletes stratospheric ozone (Revell et al., 2012). Globally, soils are the dominant sources of both anthropogenic and natural emissions of N2O, with 1.7–4.8 Tg N2O‐N year−1 emitted by agricultural soils and 3.3–9.0 Tg N2O‐N year−1 from soils under natural vegetation (Ciais et al., 2013).

Ammonia (NH3) oxidation, the rate‐limiting step of nitrification, is performed in soil mainly by aerobic ammonia‐oxidizing bacteria (AOB) and archaea (AOA), and releases N2O during conversion of NH3 to nitrite (NO2) and nitrate (NO3). Although the recently discovered complete ammonia oxidizers (comammox bacteria) can also produce N2O abiotically (Kits et al., 2019), only AOB and AOA are known for potentially significant contributions to global fluxes (Stein, 2020). Denitrification, performed in soil mainly by heterotrophic bacteria, releases N2O during the stepwise reduction of NO3 to N2O and thence dinitrogen (N2) when soils are anaerobic (Robertson & Groffman, 2021). Additionally, under hypoxic conditions, AOB that encode nitric oxide reductase (NorB) can reduce NO2 to N2O via NO through the nitrifier denitrification pathway (Stein, 2019). Nitrification and denitrification, including nitrifier denitrification, occur in most soils, and understanding the relative contributions of each is important for informing future N2O mitigation potentials and strategies, and as well for constraining uncertainties in biogeochemical models of N2O emissions.

Partitioning N2O emission pathways between nitrification and denitrification in situ have proved historically challenging. Both aerobic and anaerobic microsites occur within the same soil volume such that nitrification and denitrification often occur simultaneously (Kuenen & Robertson, 1994; Smith, 1980). In general, three types of approaches have been used to attribute N2O emission sources: specific inhibitors, stable isotope enrichment, and isotopomer analysis. Specific inhibitors have mainly been used in short‐term laboratory incubations, where acetylene (C2H2) can be used to selectively inhibit NH3 oxidation at 10 Pa and N2O reduction at 10 kPa (Robertson & Tiedje, 1987), and 1‐octyne can be used to selectively inhibit AOB ammonia monooxygenase (AMO; Taylor et al., 2013, 2015). Isotope enrichment approaches typically use either 15N‐NH4+ or 15N‐NO3 to differentiate nitrification and denitrification‐derived N2O in short‐term laboratory experiments (Stevens et al., 1997). Isotopomers of N2O reflect the differential intramolecular distribution (site preference, SP) of 15N at α and β positions of the N2O molecule (Nβ‐Nα‐O) and have been used to differentiate N2O sources in both the laboratory (Sutka et al., 2006) and field (Buchen et al., 2018; Opdyke et al., 2009; Ostrom et al., 2010).

Though helpful for identifying biochemical pathways, the use and interpretation of inhibitors and isotope enrichment approaches in situ suffer from the difficulty of achieving homogeneous distributions of added compounds in intact soils with their heterogeneously distributed microsites (Groffman et al., 2006). Artifacts of C2H2 use include further concerns of microbial C2H2 consumption (Terry & Duxbury, 1985; Topp & Germon, 1986), and as well heterotrophic nitrifiers are resistant to C2H2 (Hynes & Knowles, 1982; Schimel et al., 1984). 15N enrichment adds additional N to soils, potentially leading to overestimated rates of nitrification and denitrification especially in non‐agricultural soils (Baggs, 2008). The isotopomer approaches can be confounded by the overlap of SP values among different microbial processes. For example, N2O from fungal denitrification has an SP of 37‰, which is also within the range of nitrification (hydroxylamine oxidation; Sutka et al., 2008). An additional limitation of all three techniques is their short‐term nature in light of highly dynamic soil processes known to exhibit substantial temporal variation (Boone et al., 1999) with known effects on N2O emissions.

An alternative method for assessing the maximum potential importance of nitrification versus other N2O generating processes in soil is to combine soil‐specific kinetics of nitrification‐derived N2O with long‐term field N2O flux measurements. Nitrification kinetics measure a soil's existing potential to nitrify NH4+ to N2O and NO3 under conditions unconstrained by resource limitations (Norton & Stark, 2011; Stark & Firestone, 1996), thus allowing maximum potentials for nitrification‐derived N2O emissions to be estimated. Such potentials, if stable in time, might then be combined with field‐based measurements of N2O fluxes to allow calculation of the likely maximum percentage of nitrification‐derived N2O in relation to all other N2O sources.

Here we combine measured site‐specific nitrification kinetics for N2O production with over 25 years of field‐based N2O fluxes to estimate the maximum potential contribution of nitrification to N2O emissions along a long‐term management intensity gradient in the upper U.S. Midwest. Our replicated ecosystems range from intensively managed annual cropping systems to an unmanaged late successional deciduous forest. We first use short‐term laboratory incubations to build Michaelis–Menten kinetics models of N2O‐NH4+ relationships, and show them to be seasonally stable. Then we predict the potential maximum nitrification‐derived N2O of each ecosystem by assuming that all microbially available (soil solution phase) NH4+ can be oxidized into N2O. Finally, we use a Bayesian approach to calculate the maximum relative importance of nitrification for N2O emissions from each ecosystem based on long‐term field‐based N2O fluxes.

2. MATERIALS AND METHODS

2.1. Study site

This study was conducted in the Main Cropping System Experiment (MCSE) of the Kellogg Biological Station (KBS) LTER site located in southwest Michigan (42° 24'N, 85° 23'W). The MCSE was established in 1988 and includes, on the same soil series, ecosystems that form a management intensity gradient: annual cropping systems, perennial cropping systems, and unmanaged systems at different stages of ecological succession (Robertson & Hamilton, 2015). Most of the ecosystems are replicated in blocks as 1 ha (90 × 110 m) plots. KBS features a temperate climate with an average of 1005 mm annual precipitation distributed evenly throughout the year and a 10.1°C mean annual temperature (30‐year mean from 1981). Soils are well‐drained Alfisol loams (co‐mingled Kalamazoo and Oshtemo series Typic Hapludalfs), formed from glacial till and outwash with some intermixed loess (Crum & Collins, 1995; Luehmann et al., 2016). Average sand and clay contents in surface soils are 43% and 17%, respectively (Robertson & Hamilton, 2015).

We studied two annual cropping systems: conventionally managed (Conventional) and biologically managed (Biologically‐based) corn–soybean–winter wheat rotations; a hybrid poplar system (Poplar); and three successional systems of different ecological age: an early successional system (Early successional), a never‐tilled annually mown grassland system (Grassland), and a late successional deciduous forest (Deciduous forest). The Biologically‐based system is certified organic but receives no compost or manure. The two annual cropping systems and the Poplar and Early successional systems are replicated in each of six randomized blocks; four were selected for this study. The Grassland system is replicated four times and the Deciduous forest system is replicated three times.

The Conventional agricultural system received standard rates of N fertilizer: 137 ± 20 kg N ha−1 year−1 for corn and 77 ± 17 kg N ha−1 year−1 for wheat (Gelfand et al., 2016). Soybeans received <5 kg N ha−1 year−1. Nitrogen fertilizer was mostly applied as urea‐ammonium nitrate (28‐0‐0). The Biologically‐based agricultural system received no N fertilizer; instead, winter cover crops included the legume red clover (Trifolium pratense L.) following wheat prior to corn, and annual rye grass (Lolium multiflorum L.) following corn prior to soybean. Red clover was frost‐seeded into wheat in March, lay dormant over winter, and was terminated just prior to planting corn the following spring. Over this period, it fixes ~35–53 kg N ha−1 (Snapp et al., 2017). Both red clover and ryegrass scavenge soil N otherwise leached or denitrified. Tillage for both systems included chisel plowing to a depth of 15–18 cm followed by secondary tillage. Herbicides were used to suppress weeds in the Conventional system and additional tillage provided weed control in the Biologically‐based system.

The Poplar system was planted in 1989 to Populus × canadensis Moench “Eugenei.” Fertilizer was applied as 123 kg N ha−1 ammonium nitrate in the establishment year and the first harvest was in 1999. After the second harvest in 2008 and one fallow year, Populus nigra × P. maximowiczii “NM6” was planted in 2009. Fertilizer was then applied once in 2011 at 157 kg N ha−1 as ammonium nitrate.

The Early successional system was abandoned from agriculture in 1989 and has been burned every spring since 1997 to exclude woody plants. Canada goldenrod (Solidago canadensis L.), Kentucky bluegrass (Poa pratensis L.), arrow leaved aster (Aster sagittifolius), and timothy grass (Phleum pratense L.) were dominants at the time of this study (https://lter.kbs.msu.edu/datatables/237). The Grassland system was established on a cleared woodlot ca. 1959 and has never been plowed, but likely received manure in the 1960s. Grass is mown annually to inhibit woody species. Current dominants include smooth brome grass (Bromus inermis Leyss.), Canada goldenrod (Solidago canadensis L.), tall oatgrass (Arrhenatherum elatius L.), blackberry (Rubus allegheniensis Porter), sassafras (Sassafras albidum), and Kentucky bluegrass (P. pratensis L.). The late successional Deciduous forest is unmanaged and has never been cleared or plowed. Overstory dominant species include red oak (Quercus rubra L.), pignut hickory (Carya glabra Mill.), white oak (Q. alba L.), and sugar maple (Acer saccharum Marsh.).

2.2. Soil sampling

Soils were sampled seasonally for testing nitrification‐derived N2O potentials, once for nitrification‐derived N2O kinetics, and once for solution‐phase NH4+ partitioning. For nitrification‐derived N2O potentials, soils from all systems but the Grassland were sampled in summer (late June 2016), winter (early December 2016), and spring (early May 2017). Grassland soils were sampled when determining the kinetics of nitrification‐derived N2O, for which samples were collected in 2017 from all systems from early fall (late September) to early winter (early December), after having first established no seasonal patterns for nitrification‐derived N2O potentials. For determining solution‐phase NH4+ partitioning, soil samples were collected in summer (late June) 2019 in all systems. For all experiments, five random samples were taken at either 0–15 cm (N2O potentials and N2O kinetics experiments) or 0–25 cm (solution‐phase NH4+ partitioning) depths and composited by field replicate. Soils were passed through a 4 mm mesh immediately and sieved soils were stored at 4°C before analysis within 4 days.

2.3. Nitrification potentials

To evaluate potentials for nitrification‐derived N2O, 5 g of freshly sieved soil was placed into a 155 ml Wheaton bottle amended with 50 ml deionized water containing 10 mM NH4Cl to maximize nitrification‐derived N2O emissions (Figure 1). We used 1‐octyne, a recently developed and tested chemical inhibitor of AOB AMO to distinguish relative contributions from AOA and AOB (Taylor et al., 2013, 2015). We used a gradient of octyne concentrations ranging from 0 to 10 µM aqueous concentration (C aq) to test for optimal inhibition and we found 4 µM C aq sufficient to inhibit AOB in all soils (Liang et al., 2020), which is in agreement with previous studies (Taylor et al., 2013). Capped bottles with or without 4 µM C aq octyne were immediately placed on a shaker table and shaken for 24 h at a constant speed of 200 rpm at room temperature (25°C). This method inhibits denitrification‐derived N2O as soil slurries are continuously aerated by high‐speed shaking.

FIGURE 1.

FIGURE 1

The kinetics of nitrification‐derived N2O in soils from different systems varying in management intensities. Michaelis–Menten models were fit to total nitrification‐derived N2O emissions (blue lines) and AOB‐derived N2O emissions (orange lines). Blue circles and orange triangles are the mean N2O emissions from total and AOB‐derived nitrification at each ammonium addition, respectively. Note y‐axis scale differs by system. Shaded bands represent 95% confidence intervals. Ammonium additions ranged from 0.05 and 15 mM for Poplar and annual cropping systems because N2O accumulation at 0.01 mM could not be reliably estimated. For all other systems, ammonium additions ranged from 0.01 to 15 mM. AOB, ammonia‐oxidizing bacteria; N2O, nitrous oxide

Samples for N2O were taken at 2 and 24 h and N2O emission rates were calculated based on N2O accumulations over 22 h. Slurry pH was buffered naturally as no apparent pH change was detected during the incubation. Emissions of N2O in the presence of octyne are attributed to AOA. Emissions of N2O from AOB are calculated as the difference between N2O without octyne (total nitrification‐derived N2O) minus N2O from AOA. Although comammox could also contribute to N2O emissions, recent evidence suggests that comammox plays only a very minor role in soil nitrification (Kits et al., 2019; Robertson & Groffman, 2021; Wang et al., 2020). N2O samples were stored over‐pressurized in 6 ml N2‐flushed glass vials (Exetainers, Labco Ltd). N2O was measured with a gas chromatograph (Agilent 7890A) coupled to an autosampler (Gerstel MPS2XL) and equipped with a 63Ni electron detector at 350°C and a Porapak Q column (1.8 m, 80/100 mesh) at 80°C (https://lter.kbs.msu.edu/protocols/159).

2.4. Nitrification kinetics

We placed 5 g of freshly sieved soil from each ecosystem into a 155 ml Wheaton bottle. We then added (NH4)2SO4 to make eight different NH4+ concentrations ranging from 0.01 to 15.0 mM (0.01, 0.05, 0.1, 0.5, 1, 5, 10, and 15 mM NH4+) with a final liquid volume of 50 ml. Bottles were capped and placed on a shaker table at a constant speed of 200 rpm at room temperature (25°C) and shaken for 24 h. Initial N2O samples were taken after 2 h, and we then added either 2.8 ml of octyne stock gas (see Taylor et al., 2013, for octyne stock gas preparation) to create 4 µM C aq concentrations or 2.8 ml of air without octyne. Another set of N2O samples were taken at 24 h. Nitrification kinetics were based on measured NH4+ concentrations, and included both added NH4+ as well as NH4+ produced from net N mineralization during the incubation. NH4+ concentrations were measured by a Lachat QuikChem 8500 flow injection analyzer (Hach).

Kinetics of nitrification‐derived N2O emissions were fit to Michaelis–Menten models using the equation:

V=VmaxSKm+S (1)

where V is the N2O emission rate from nitrification, V max is the maximum N2O emission rate from nitrification under conditions of unlimited substrate (NH4+), S is the NH4+ concentration, and K m is the half‐saturation constant that represents the NH4+ concentration when the N2O emission rate from nitrification is ½ V max. V max reflects the maximum capacity of a soil to oxidize NH4+ and produce nitrification‐derived N2O, and K m reflects the NH4+ affinity of soil AMO.

In addition, because nitrification can be inhibited at very high NH4+ concentrations (Suwa, 1994), we also fitted data with Haldane models when appropriate (Koper et al., 2010; Stark & Firestone, 1996):

V=VmaxSKm+S+S2/Ki (2)

The Haldane model introduces a third parameter K i that reflects the maximum NH4+ concentration at which nitrification‐derived N2O emissions rates are ½ V max. We performed an Akaike's information criterion (AIC)‐based model comparison, followed by an F‐test to determine model superiority between Michaelis–Menten and Haldane kinetics (Table 1).

TABLE 1.

Comparisons between Michaelis–Menten and Haldane kinetics models for total or AOB‐derived N2O emissions from nitrification

Ecosystem a Nitrification AIC b (Michaelis–Menten) AIC b (Haldane) F‐value c p‐value c
Poplar Total 111 113 0.188 0.668
AOB 105 106 0.488 0.491
Early successional Total 143 144 0.134 0.718
AOB 130 131 1.13 0.298
Grassland Total 27.9 28.1 1.70 0.202
AOB 30.2 30.6 1.50 0.233
Deciduous forest Total 109 111 0.001 0.980
AOB 106 108 0.049 0.827

Abbreviations: AIC, Akaike information criterion; AOB, ammonia‐oxidizing bacteria; N2O, nitrous oxide.

a

Data from Conventional and Biologically‐based systems were not fit to Haldane models because no signs of inhibition of nitrification‐derived N2O were found.

b

Models with lower AIC were considered superior.

c

Models were also compared based on F‐test. A p‐value > .05 supports the minimal model as the adequate model.

2.5. In situ N2O flux, soil NH4+, and soil bulk density

We used 25 years of in situ N2O flux data (from 1991 to 2016) to calculate the relative contribution of nitrification to N2O emissions within each system, except for the Grassland and Deciduous forest systems for which N2O fluxes were measured from 1992 to 2016 and 1993 to 2016, respectively. Most of these data have been previously published (Gelfand et al., 2016; Robertson et al., 2000). From 1991 to 2012, emissions were sampled every 2 weeks from March/April to November/December with the static chamber method (Holland et al., 1999). Additional winter samples were taken monthly starting from 2013. Square chambers (29 × 29 × 14 cm high) were placed on aluminum bases (28 × 28 × 10 cm high) semi‐permanently installed about 3 cm into soil. Gas samples were taken at approximately 20‐min intervals during a 1‐h sampling period. Volume‐based N2O fluxes were calculated by linearly regressing headspace N2O concentrations over time (µg N2O‐N L−1 min−1), which was then further converted to area‐based N2O fluxes by accounting for the volume of gas in the chamber and soil surface area covered by the chamber (g N2O‐N ha−1 day−1; Kahmark et al., 2020). The few headspace fluxes that exhibited nonlinearity were not used in the analysis.

Soil cores for inorganic N determinations were taken approximately biweekly after the soils thawed in the spring, usually in March or April, and discontinued before soils froze, usually in November. Soils were sampled to 25 cm depth from 1989 to 2016 except from 1993 to 2016 for the Deciduous forest system. Soil was sieved through a 4 mm sieve and 10 g of fresh soil were extracted with 100 ml 1 M KCl to determine NH4+ concentrations. Soil bulk density (0–10 cm depth) was measured in 2013 when collecting deep core soil samples to a depth of 1 m with a hydraulic probe. Soil was sieved through a 4 mm sieve and then oven‐dried at 60°C for 48 h. When present, the weight of gravel (>4 mm) was recorded separately and then discarded. The gravel‐free bulk density was calculated as the dry mass of the soil (without gravel) divided by the volume of the core.

2.6. Microbially available (solution phase) soil NH4+

We partitioned long‐term KCl extracted soil NH4+ pools into sorbed‐phase (srNH4+) and solution‐phase (slNH4+) pools by performing an NH4+ sorption capacity assay modified from Venterea et al. (2015). We assume only slNH4+ is available to soil nitrifiers. Briefly, for each ecosystem, we added 10 g of sieved fresh soils into 100 ml of water containing an NH4+ gradient ranging from 0 to 50 mg NH4+‐N L−1 (0, 5, 10, 20, 30, 40, and 50 mg NH4+‐N L−1 generated by (NH4)2SO4 addition). Mixtures were shaken on an orbital shaker table at a constant speed of 100 rpm at room temperature (25°C) for 18 h. We centrifuged 10 ml aliquots at 10,000 g at room temperature (25°C) for 15 min. NH4+‐N was then analyzed by flow injection analysis as above after filtering aliquots through a 1 mm glass fiber filter. We calculated srNH4+ as the difference between added NH4+ (addNH4+) and the slNH4+ (measured as above) accounting for soil NH4+ contents (soilNH4+):

soilNH4+=NH4 KCl+NH4 0+ (3)
srNH4+=addNH4+slNH4++soilNH4+when addNH4+>0 (4)
srNH4+=soilNH4+when addNH4+=0 (5)

where NH4 KCl+ is the 1 M KCl extractable NH4+ concentrations and NH4 0+ is the water extractable NH4+ concentrations at 0 NH4+‐N L−1 addition. The relationship between srNH4+ (mg N kg−1) and slNH4+ (mM) is usually described by a Langmuir model:

srNH4+=μ×slNH4+K+slNH4+ (6)

where μ (mg N kg−1) is the maximum NH4+ content adsorbed by soil and K (mM) is the NH4+ concentration in solution phase at which srNH4+ is ½ μ. We modeled and plotted srNH4+ against slNH4+ (Figure S1), which allows one to convert total KCl‐based soil NH4+ values into slNH4+ for every NH4+ soil measurement taken between 1989 and 2016.

2.7. Statistical analysis

2.7.1. ANOVA for seasonal nitrification‐derived N2O

We converted gravimetric N2O emissions from the nitrification potential experiment into areal N2O emissions based on soil depth (15 cm) and bulk density:

N2Oarea=N2Omass×DP×BD10 (7)

where N2Oarea is expressed as g N2O‐N ha−1 day−1 and N2Omass is expressed as ng N2O‐N g−1 dry soil day−1, DP is the soil depth in cm, and BD (0–10 cm depth) is the bulk density expressed as g cm−3.

Potentials for nitrification‐derived N2O were analyzed with PROC GLIMMIX of SAS 9.4 (SAS Institute). The statistical model included 5 ecosystem types ×3 seasons × 2 sources of nitrification‐derived N2O, and the interaction among them was considered fixed factors. Field replicates nested within ecosystem types and the interaction between field replicates and seasons nested within ecosystem types were considered random factors. Analysis of variance (ANOVA) was performed by considering ecosystem types as a whole plot factor and season and sources of nitrification‐derived N2O as subplot and sub‐subplot factors. Homogeneity of variance assumptions was checked by Levene's test and normality of residuals was visually inspected. No violations of assumptions were detected. Pairwise comparisons among different ecosystems were conducted and we refer to p < .05 (two‐sided) as significantly different throughout the paper.

2.7.2. Model comparisons and kinetic parameters

Total or AOB‐derived N2O emissions from nitrification were fit to both Michaelis–Menten and Haldane kinetics models. We first used the “nls” function in R (version 3.5.0; R Core Team, 2020) to obtain AIC values for each kinetics model. Then we conducted an F‐test to further determine model superiority using the “anova” function. Models with lower AIC were considered superior, and a p‐value > .05 supports the minimal model (Michaelis–Menten) as the adequate model (Table 1). Once the appropriate kinetics model (Michaelis–Menten) was selected, V max and K m for total and AOB‐derived N2O emissions from nitrification for each ecosystem were estimated by the “nls” function (Table 3).

TABLE 3.

Michaelis–Menten kinetic parameters of total or AOB‐derived N2O emissions from nitrification. V max represents maximum nitrification‐derived N2O emissions (g N2O‐N ha−1 day−1) and K m represents half saturation constant (mM). Numbers within the parentheses represent standard errors

Ecosystem Nitrification V max K m
Conventional agriculture Total 12.7 (0.6) 0.20 (0.06)
AOB 11.4 (0.6) 0.24 (0.06)
Biologically‐based agriculture Total 15.1 (1.2) 0.079 (0.042)
AOB 13.8 (1.3) 0.088 (0.056)
Poplar Total 3.48 (0.40) 0.025 (0.019)
AOB 2.92 (0.36) 0.033 (0.026)
Early successional Total 4.54 (0.52) 0.009 (0.008)
AOB 3.31 (0.47) 0.012 (0.011)
Grassland Total 1.59 (0.08) 0.012 (0.004)
AOB 0.49 (0.09) 0.002 (0.002) a
Deciduous forest Total 4.12 (0.61) 0.031 (0.026)
AOB 3.01 (0.58) 0.042 (0.045)

Abbreviations: AOB, ammonia‐oxidizing bacteria; N2O, nitrous oxide.

a

K m value was estimated by constraining estimate >0.

2.7.3. Distribution for field N2O fluxes

In situ N2O fluxes typically show a highly skewed distribution with a long tail of high values, which makes constraining the range of the mean fluxes challenging (Cowan et al., 2017). N2O emissions can be assumed proportional to the product of the interactions of multiple biological and environmental variables such as population sizes and activities of soil nitrifiers and denitrifiers, soil moisture, soil temperature, soil inorganic N contents, and soil oxygen status. Thus, we consider multiplicative processes to influence N2O emissions, which follow log‐normal distributions (Limpert et al., 2001):

FN2Olognormx¯,s2 (8)

where x¯ and s are the mean and standard deviation of log‐transformed N2O emissions, respectively.

The mean of a log‐normal distribution (without log‐transformation) is usually described as follows:

μ=expx¯+s22 (9)

Here, we estimated log‐normal means of N2O fluxes using a Bayesian approach by evaluating the parameters in Equation (9). We chose vague prior probability distributions to reduce their impact on the inference.

Although fitting log‐normal distributions for N2O fluxes makes biological and theoretical sense, there are other distributions that describe continuous positive data with large variances well. Thus, we also fit N2O data with other candidate distributions including Gamma and Weibull distributions using the “fitdistrplus” package for R (Delignette‐Muller & Dutang, 2015; Table 2).

TABLE 2.

AIC of field‐based nitrous oxide fluxes from different ecosystems fitted with different distributions

Ecosystem Distribution
Log‐normal Gamma Weibull Normal
Conventional agriculture 4602 5038 4915 7922
Biologically‐based agriculture 5030 5489 5344 8629
Poplar 2303 2881 2659 6378
Early successional 2591 2804 2808 4392
Grassland 1733 1872 1865 3106
Deciduous forest 2452 2690 2648 4687

Abbreviation: AIC, Akaike information criterion.

2.7.4. Estimation of contributions from nitrification

Similar to N2O emissions from nitrification potentials, before fitting Michaelis–Menten models we converted gravimetric N2O emissions from each nitrification kinetics experiment into areal N2O emissions using Equation (7) based on soil depth (15 cm) and bulk density. We then used the “nls” function in R (version 3.5.0; R Core Team, 2020) to estimate V max and K m and their associated standard errors, which were then specified as prior information when we conducted a Markov Chain Monte Carlo simulation to sample posterior parameter distributions with the “jagsUI” package (Kellner, 2017) for R. We ran three chains of 15,000 iterations with 2000 burn‐in iterations with a thinning rate of three, which yielded 13,002 total samples for posterior distribution.

Based on the Michaelis–Menten model, we developed for each ecosystem, long‐term solution‐phase NH4+ data were applied to predict maximum potential N2O emissions from nitrification. The potential maximum contribution of nitrification to total N2O was estimated with the mean of the predicted nitrification‐derived N2O divided by the log‐normal mean of field N2O measurements for Conventional, Biologically‐based, Poplar, Grassland, and Deciduous forest systems. Because the contribution from nitrification cannot be >100%, we constrained our analysis with contributions ranging between 0 and 1. Overall, over 96% of the posterior distributions for contributions from total nitrification and over 99% of the posterior distributions for contributions from AOB‐derived nitrification were included.

3. RESULTS

3.1. Seasonal N2O emissions from nitrification potential

Across all seasons examined, soils from the Conventional and Biologically‐based annual cropping systems had the highest nitrification‐derived N2O potentials (Figure 2), ranging from 17.6 to 24.8 and from 13.1 to 24.6 g N2O‐N ha−1 day−1, respectively. In comparison, Deciduous forest soils exhibited the lowest total and AOB‐derived N2O potentials: 2.39 ± 0.67 (standard error of the mean) and 2.98 ± 1.28 g N2O‐N ha−1 day−1, respectively, for spring, and 1.56 ± 0.60 and 2.93 ± 0.60 g N2O‐N ha−1 day−1 for winter. Although seasonal nitrification‐derived N2O potentials from the Conventional and Biologically‐based systems were significantly higher than from the Early successional or Deciduous forest (p < .05) systems, the differences between the two agricultural systems were not significant (p > .30) for two out of three seasons. Similarly, N2O potentials via nitrification were generally indistinguishable among Poplar, Early successional, and Deciduous forest systems (p > .15) in any given season.

FIGURE 2.

FIGURE 2

Seasonal potential N2O production from nitrification (total or AOB‐derived) across a management intensity gradient. Bars represent standard errors (for total, n = 4 except deciduous forest n = 3; for AOB, n = 3–4 except deciduous forest n = 2–3). No significant differences among seasons were detected (p = .30). AOB, ammonia‐oxidizing bacteria; N2O, nitrous oxide

No significant overall seasonal differences of nitrification‐derived N2O potentials were observed (p = .30, Figure 2). There were also no significant interaction effects between sources of N2O and seasons (p = .76) nor interactions among ecosystem types, sources of N2O, and seasons (p = .73).

3.2. Kinetics of nitrification‐derived N2O

Michaelis–Menten models fit nitrification‐derived N2O data well (Figure 1; Table 1). The Conventional and Biologically‐based cropping systems exhibited the highest values of V max (Table 3), ranging from 12.7 to 15.1 g N2O‐N ha−1 day−1 for total nitrification‐derived N2O, and 11.4 to 13.8 g N2O‐N ha−1 day−1 for AOB‐derived N2O. The Grassland system had the lowest V max, 1.59 ± 0.08 N2O‐N ha−1 day−1 and 0.49 ± 0.09 g N2O‐N ha−1 day−1 for total and AOB‐derived N2O, respectively, followed by Poplar but with a V max 2–6 times higher than the Grassland system. V max for Early successional and Deciduous forest systems were similar, ranging from 3.01 to 3.31 and 4.12 to 4.54 g N2O‐N ha−1 day−1 for AOB and total nitrification‐derived N2O, respectively.

K m values indicate how quickly NH4+ saturates nitrification‐derived N2O (Table 3). The Conventional agricultural system had the highest K m for both total and AOB‐derived N2O, reaching 0.20 ± 0.06 and 0.24 ± 0.06 mM NH4+, respectively, which was about 2.5 times higher than the Biologically‐based system, and 5–20 times higher than for all other systems.

3.3. The relative importance of AOA and AOB for nitrification‐derived N2O

Based on the posterior distributions of V max, we found that compared to AOA, AOB were the major contributors to nitrification‐derived N2O in most soils, accounting for more than 70% of total nitrification‐derived N2O (Figure 3) in all but the Grassland system, where the contribution from AOB averaged only 32 ± 4%. In addition, there was a decreasing trend of AOB’s contribution to N2O along the management gradient: about 90% of the nitrification‐derived N2O was from AOB in row crop systems, whereas in the Early successional and Deciduous forest systems, AOB’s contribution decreased to about 70% of total N2O. Concomitantly, the contribution of AOA to nitrification‐derived N2O generally increased from the intensively managed row crop to unmanaged Grassland and Deciduous forest.

FIGURE 3.

FIGURE 3

Relative contributions of AOA and AOB to nitrification‐derived N2O emissions in systems that differ in management intensities. Contributions from AOB (%, orange) were calculated with posterior distributions of V max derived from Michaelis–Menten models for AOB and total nitrification‐derived N2O kinetics. Contributions from AOA (%, blue) were calculated as 1 − AOB (%). The upper, mid, and lower lines of each boxplot indicate 25th, median, and 75th percentiles, respectively. The upper and lower whiskers indicate 1.5 × interquartile range. AOA, ammonia‐oxidizing archae; AOB, ammonia‐oxidizing bacteria; N2O, nitrous oxide

3.4. Contribution of nitrification to long‐term N2O emissions

Among all ecosystems, row crop systems appear to have the lowest maximum potential N2O contributed from nitrification. The percentage of 25th–75th posterior intervals from nitrification, assuming all soil NH4+ is available only to nitrifiers, ranged between 13.1% and 16.7% for the Conventional agricultural system and 27.4%–41.6% for the Biologically‐based system (Figure 4a). For the Poplar and Grassland systems, a maximum potential of 52.0% and 54.8% of field‐based N2O fluxes can be attributed to nitrification. The Deciduous forest system was associated with the highest maximum potential contribution from nitrification, with the percentage of 25th–75th posterior intervals ranging between 51.2% and 76.9% for total nitrification‐derived N2O and 27.2%–49.6% for AOB‐derived N2O (Figure 4a,b). For all ecosystems, the median maximum potential contributions of AOB to N2O were below 40%, ranging from 11.4% to 36.4% (Figure 4b).

FIGURE 4.

FIGURE 4

Contribution of nitrification to N2O production. Maximum relative contributions of (a) total nitrification and (b) AOB‐derived nitrification to long‐term field N2O emissions in systems that differ in management intensities assuming all solution‐phase in situ ammonium is oxidized and no nitrification‐derived N2O is reduced. Field‐based N2O fluxes were estimated assuming log‐normal distributions. Vertical lines indicate the median contribution for each system. Values in parentheses indicate the 25th–75th posterior intervals, respectively. Note that the Early successional system is not included as 95% of the posterior nitrification‐derived N2O was higher than the field fluxes. AOB, ammonia‐oxidizing bacteria; N2O, nitrous oxide

4. DISCUSSION

Soils from different ecosystems showed distinct patterns of Michaelis–Menten kinetics for nitrification‐derived N2O emissions, with highest and lowest V max and K m observed in the row crop and the Grassland ecosystems, respectively. Combining kinetic parameters with 25 years of in situ N2O flux and solution‐phase in situ soil NH4+ measurements suggests that nitrification is a minor source of N2O in these ecosystems. Results also show AOB rather than AOA are the dominant source of nitrification‐derived N2O in all ecosystems but the mown grassland.

4.1. Seasonal nitrification‐derived N2O emissions from AOA and AOB

Seasonal nitrification‐derived N2O potentials from AOB were 5–26 times higher than from AOA in Conventional and Biologically‐based systems (Figure S2), suggesting a greater capacity of AOB for emitting nitrification‐derived N2O from agricultural soils. Wang et al. (2016) have also reported the dominance of AOB over AOA for N2O produced in soils amended with inorganic ammonium fertilizer, although their study was conducted in static microcosms rather than in microcosms on shaker tables, so results could have been confounded by nitrifier denitrification since hypoxic conditions can develop in soil aggregates during static incubations (Lu et al., 2018; Stein, 2019).

Taken together, results suggest that low soil ammonium, in unfertilized systems derived primarily from soil organic matter mineralization, promotes a greater relative contribution of AOA to nitrification‐derived N2O as also found by Hink et al. (2018). Additionally, nitrifier community compositions in unfertilized systems could be very different from row crop systems, which, in turn, could affect relative N2O production. Upon fertilization, nitrifier community composition appears to favor AOB and in particular Nitrosospira spp., with no similar consistent changes in AOA yet identified (Bertagnolli et al., 2016; Kong et al., 2019; Phillips et al., 2000; Wu et al., 2011; Xue et al., 2016). Soil Nitrosospira spp. have been shown to positively respond to urea and as well are associated with increased N2O emissions (Cassman et al., 2019).

The absence of seasonal effects suggests that the composition and capacity for soil nitrifiers to produce N2O remain reasonably constant throughout any given year. These findings are consistent with a year‐round metagenomic study reporting remarkably stable nitrifier community composition and abundance in a US Midwest agricultural soil (Orellana et al., 2018). Similarly, both abundance and community structure of amoA genes of AOA and AOB have been shown to be stable across seasons in two acid forest soils (Qin et al., 2019). Thus, it seems reasonable to conclude that long‐term management practices in our ecosystems have selected soil nitrifier populations that are adapted to seasonal environmental fluctuations such as soil temperature (Séneca et al., 2020).

4.2. The responses of N2O kinetics to management intensities

The Conventional and Biologically‐based agricultural systems were associated with the highest values for V max and K m, suggesting a greater capacity of row crop soils to emit nitrification‐derived N2O than soils from our other systems. Notably, the Biologically‐based system had a similar V max but lower K m compared with the Conventional system. This difference may be because in the Biologically‐based system, the slower‐paced release of NH4+ from decomposing cover crop and other residues has selected nitrifier communities with high NH4+ affinities (Hink et al., 2017, 2018) and less tolerance for high NH4+ input as compared to nitrifiers from the Conventional system. The low V max and K m in Early successional, Grassland, and Deciduous forest systems may reflect their histories of no fertilizer inputs, resulting in a low capacity to produce nitrification‐derived N2O even under substrate‐unlimited conditions.

Existing studies of nitrification kinetics have mainly focused on the effects of NH4+ on NO2 + NO3 accumulation. Koper et al. (2010) reported that the V max of soils receiving ammonium sulfate at 200 kg N per hectare for 6 years was about twice higher than the V max of non‐fertilized soils, but no significant differences in K m were detected. It is possible that substrate affinity responds to fertilizer more slowly than maximum nitrification rate. In addition, although V max and K m of AOB and total nitrification could be boosted significantly within a month of fertilization, they can also decline rapidly within 3 months of fertilizer application (Ouyang et al., 2017). Together, these results suggest that long‐term management practices shaped differences in V max and K m responses among ecosystems varying in management intensity.

4.3. Contribution of AOA and AOB to V max along the management intensity gradient

We used a Bayesian approach to calculate the relative contributions of AOA versus AOB to nitrification‐derived N2O based on posterior distributions of V max for each ecosystem, which is different from the traditional method of separating AOA from AOB based on 1 mM NH4+ addition (Lu et al., 2015; Ouyang et al., 2016; Taylor et al., 2010). As noted earlier, 1 mM NH4+ additions did not always yield the highest N2O emissions in our systems (Figure 1), especially for agricultural soils. Thus, partitioning sources of nitrification‐derived N2O with V max derived from substrate kinetics aligns with the concept of nitrification potential assays, which reflect the maximum nitrification‐derived N2O from nitrifier communities (Norton & Stark, 2011).

The declining importance of AOB to N2O production along the management intensity gradient likely reflects different strategies of soil nitrifiers’ responding to different agronomic practices. First, the Conventional system constantly receives high N inputs, which favor AOB activity or population size in agricultural soils (Habteselassie et al., 2013; Jia & Conrad, 2009; Shen et al., 2008; Taylor et al., 2010, 2013). In contrast, AOA’s contribution is more important in systems where the major NH4+ source is via decomposition of soil organic matter. Thus, the speed of NH4+ supply to soil seems important for shaping the dynamics of AOA versus AOB N2O‐generating activities. Indeed, Hink et al. (2018) observed that AOA dominated nitrification‐derived N2O in incubated soils receiving slow‐release fertilizer instead of free urea.

A second major difference between row crop and unfertilized systems is the history of tillage. Both the Conventional and Biologically‐based systems have been either moldboard or chisel‐plowed since well before 1988. In contrast, the Early successional and Poplar systems have been untilled since 1989 and the Deciduous forest and Grassland systems have never been tilled. Tillage accelerates soil organic matter turnover, which results in more pulse‐like releases of NH4+ in soil compared with non‐tilled systems. As a result, AOB likely also outcompetes AOA following tillage‐induced pulses of NH4+.

The dominance of AOA for nitrification‐derived N2O in the Grassland system seems anomalous and might be attributed to differential inhibition of AOB versus AOA induced by root‐released nitrification inhibitors known to occur in at least one grass species. While we have no direct evidence of inhibitors produced by grasses in our study sites, in a 3‐year field study, Subbarao et al. (2009) showed that brachialactone, a root exudate isolated from the forage grass Brachiaria sp., inhibited 90% of in situ NH4+ oxidation and over 90% of cumulative N2O emissions in a tropical pasture. Moreover, the inhibition seemed to be specific to AOB rather than AOA. Historically, among all of our ecosystems, the Grassland system has always had the highest monthly soil NH4+ concentrations and exhibited the lowest relative nitrification potentials (Millar & Robertson, 2015). Since root exudates of Bromus spp., a dominant species in the Grassland system, have been reported to significantly inhibit nitrification in vitro in both AOB culture and whole soils (O'Sullivan et al., 2017), we suspect AOB inhibition in the Grassland system.

4.4. Long‐term contribution of nitrification to in situ N2O fluxes

Seasonally stable nitrification‐derived N2O fluxes allow us to apply kinetics models to predict potential maximum N2O emissions from nitrification and, subsequently, the theoretical maximum relative contribution of nitrification to field‐based N2O emissions assuming nitrifiers has exclusive access to solution‐phase NH4+. Since the kinetics results are based on aerobic incubations of shaken soil slurries that eliminate both N2O reduction and N2O from nitrifier denitrification (Wrage et al., 2001; Wrage‐Mönnig et al., 2018), N2O rates can be considered nitrifier nitrification rather than nitrifier denitrification, and when applied to historical solution‐phase in situ NH4+ pools, reveal maximum potential nitrification‐derived N2O in situ.

An important consideration in whole‐soil kinetic assays is that they ignore the likelihood that some taxa will be nitrifying at rates lower than their maximum possible as nitrifiers exhibit significant phylogenetic and physiological diversity (Hazard et al., 2021). That said, whole‐community incubations under laboratory conditions that favor nitrification in general, allow us to identify the maximum likely rates of whole‐soil nitrification, were such conditions possible in the field. So though our controlled laboratory conditions might be suboptimal for some taxa, the assay overall seems a reasonable, conservative proxy for obtaining maximum whole‐community nitrification rates under different substrate conditions.

The finding that total nitrification contributed a theoretical maximum of 13%–17% of field‐based N2O fluxes in the Conventional agricultural system suggests that nitrification is unlikely to be a significant source of N2O in long‐fertilized systems. That a theoretical maximum of only 27%–42% of field‐based fluxes were nitrification‐derived in the Biologically‐based system suggests that nitrification is likewise unlikely to be a dominant N2O source in even unfertilized annual cropping systems. Using N2O SP analysis, Opdyke et al. (2009) and Zou et al. (2014) reported a small role for nitrification in N2O produced by agricultural soils (including ours), although these studies were short‐term snapshots. Similarly, AOB‐derived nitrification is unlikely to be the major process leading to N2O production in any of our ecosystems regardless of management. These results are also consistent with Buchen et al. (2018), who also used SP in situ to suggest that >80% of N2O can be attributed to denitrification (whether heterotrophic or nitrifier‐derived) in managed grasslands.

Since our Michaelis–Menten models were necessarily developed under laboratory conditions that favored nitrification, the calculated contributions of nitrification to N2O reflect maximum in situ potentials that assume all solution‐phase NH4+ is available exclusively to nitrifiers and no nitrification‐derived N2O is further denitrified to N2. Neither of these assumptions are realistic in situ. Soils are rarely completely aerobic, and even if in situ nitrification emitted N2O equivalent to the amount from shaken soil slurries, some of the N2O will be captured by denitrifiers and reduced to N2 before being emitted to the atmosphere (Decock & Six, 2013; Lewicka‐Szczebak et al., 2017; Shcherbak & Robertson, 2019).

Malhi and McGill (1982) estimated that the daily maximum NH4+‐N oxidation rate is <10% of available NH4+‐N (100 µg N g−1) based on laboratory incubations. Prosser et al. (2020) reported pure culture N2O yields for AOB and AOA to be only 0.1%–8% and 0.04%–0.3%, respectively, although a greater diversity of nitrifiers in situ (Amann et al., 1995) will reflect a wider range. Hence, our assumption of 100% of daily NH4+ is oxidized and consequently eligible for transformation to N2O is undoubtedly an overestimate by a factor of 10 to 100 or more. That said, our conclusion of nitrification being a minor source of N2O in these ecosystems is conservative by nature. Actual contributions of nitrification to measured N2O fluxes in situ are likely to be only 0.1%–10% of the potential maximum rates we identify.

By way of example, the least‐constrained nitrifier contribution to N2O fluxes was measured in Early successional and Deciduous forest soils where 51%–77% of total N2O fluxes might potentially derive from nitrification in the Deciduous forest system (Figure 4a), and over 95% of the predicted nitrification‐derived N2O was higher than the field fluxes in the Early successional system. But here, perhaps especially, the extrapolation assumptions seem severe. The Early successional and Deciduous forest soils have high concentrations of macroaggregates (2000–8000 µm; Grandy & Robertson, 2007) and thus a larger volume fraction of anoxic centers (Schlüter et al., 2018), which contribute to high measured denitrification rates (Robertson & Tiedje, 1984). So even in our systems with the greatest percentage of N2O contributed by nitrifiers based on Michaelis–Menten kinetics, actual results will be but a fraction.

Overall, we conclude that nitrification is a minor source of N2O emissions in all of the systems examined. This finding has significant implications for biogeochemical N2O flux models that assume a significant fraction of emissions are nitrifier derived (e.g. Parton et al., 2001). Our findings further suggest that taxa‐specific N2O mitigation might better target processes other than nitrification, except insofar as nitrification makes nitrate available to denitrifiers.

CONFLICT OF INTEREST

The authors declare no conflict of financial interests.

Supporting information

Supplementary Material

ACKNOWLEDGEMENTS

Support for this study was provided by the U.S. Department of Energy, Office of Science, Office of Biological and Environmental Research (Awards DE‐SC0018409 and DE‐FC02‐07ER64494), by the National Science Foundation Long‐Term Ecological Research Program (DEB 1637653) at the Kellogg Biological Station, by the USDA Long‐Term Agroecosystem Research Program, by an ESPP Urban Environment Summer Research Fellowship, and by Michigan State University AgBioResearch. We are indebted to S. Kravchenko, S. Evans, and N. Ostrom for valuable feedback during the course of our study. We are also indebted to R. Venterea for advice on partitioning solution‐phase from sorbed‐phase NH4+, and to S. VanderWulp and C. McMinn for laboratory and field assistance.

Liang, D. , & Robertson, G. P. (2021). Nitrification is a minor source of nitrous oxide (N2O) in an agricultural landscape and declines with increasing management intensity. Global Change Biology, 27, 5599–5613. 10.1111/gcb.15833

DATA AVAILABILITY STATEMENT

KBS MCSE long‐term field N2O flux measurements can be accessed at https://lter.kbs.msu.edu/datatables/28; long‐term in situ soil NH4+‐N concentrations are available at https://lter.kbs.msu.edu/datatables/55. Soil bulk densities of deep core soil samples are available from https://lter.kbs.msu.edu/datatables/308. Species composition of Early successional and Deciduous forest systems is available from https://lter.kbs.msu.edu/datatables/237 and https://lter.kbs.msu.edu/datatables/238. All other data used in this study are available at datadryad.org (https://doi.org/10.5061/dryad.37pvmcvkz).

Code availability: The code for estimating log‐normal mean of in situ N2O fluxes and the maximum contribution of nitrification to total N2O are shown in Supporting Information.

REFERENCES

  1. Amann, R. I. , Ludwig, W. , & Schleifer, K. H. (1995). Phylogenetic identification and in situ detection of individual microbial cells without cultivation. Microbiological Reviews, 59(1), 143–169. 10.1128/mr.59.1.143-169.1995 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Baggs, E. M. (2008). A review of stable isotope techniques for N2O source partitioning in soils: Recent progress, remaining challenges and future considerations. Rapid Communications in Mass Spectrometry, 22(11), 1664–1672. 10.1002/rcm.3456 [DOI] [PubMed] [Google Scholar]
  3. Bertagnolli, A. D. , McCalmont, D. , Meinhardt, K. A. , Fransen, S. C. , Strand, S. , Brown, S. , & Stahl, D. A. (2016). Agricultural land usage transforms nitrifier population ecology. Environmental Microbiology, 18(6), 1918–1929. 10.1111/1462-2920.13114 [DOI] [PubMed] [Google Scholar]
  4. Boone, R. D. , Grigal, D. F. , Sollins, P. , Ahrens, R. J. , & Armstrong, D. E. (1999). Soil sampling, preparation, archiving, and quality control. In Robertson G. P., Coleman D. C., Bledsoe C. S., & Sollins P. (Eds.), Standard soil methods for long‐term ecological research (pp. 3–28). Oxford University Press. [Google Scholar]
  5. Buchen, C. , Lewicka‐Szczebak, D. , Flessa, H. , & Well, R. (2018). Estimating N2O processes during grassland renewal and grassland conversion to maize cropping using N2O isotopocules. Rapid Communications in Mass Spectrometry, 32(13), 1053–1067. 10.1002/rcm.8132 [DOI] [PubMed] [Google Scholar]
  6. Cassman, N. A. , Soares, J. R. , Pijl, A. , Lourenço, K. S. , van Veen, J. A. , Cantarella, H. , & Kuramae, E. E. (2019). Nitrification inhibitors effectively target N2O‐producing Nitrosospira spp. in tropical soil. Environmental Microbiology, 21(4), 1241–1254. 10.1111/1462-2920.14557 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Ciais, P. , Sabine, C. , Bala, G. , Bopp, L. , Brovkin, V. , Canadell, J. , Chhabra, A. , DeFries, R. , Galloway, J. , Heimann, M. , Jones, C. , Le Quéré, C. , Myneni, R. B. , Piao, S. , & Thornton, P. (2013). Carbon and other biogeochemical cycles. In Stocker T. F., Qin D., Plattner G.‐K., Tignor M., Allen S. K., Boschung J., Nauels A., Xia Y., Bex V., & Midgley P. M. (Eds.), Climate change 2013—The physical science basis: Contribution of working group I to the fifth assessment report of the intergovernmental panel on climate change (pp. 465–570). Cambridge University Press. [Google Scholar]
  8. Cowan, N. J. , Levy, P. E. , Famulari, D. , Anderson, M. , Reay, D. S. , & Skiba, U. M. (2017). Nitrous oxide emission sources from a mixed livestock farm. Agriculture, Ecosystems & Environment, 243, 92–102. 10.1016/j.agee.2017.04.014 [DOI] [Google Scholar]
  9. Crum, J. R. , & Collins, H. P. (1995). KBS soils. 10.5281/zenodo.2581504 [DOI] [Google Scholar]
  10. Decock, C. , & Six, J. (2013). How reliable is the intramolecular distribution of 15N in N2O to source partition N2O emitted from soil? Soil Biology and Biochemistry, 65, 114–127. 10.1016/j.soilbio.2013.05.012 [DOI] [Google Scholar]
  11. Delignette‐Muller, M. L. , & Dutang, C. (2015). fitdistrplus: An R package for fitting distributions. Journal of Statistical Software, 64(4). 10.18637/jss.v064.i04 [DOI] [Google Scholar]
  12. Gelfand, I. , Shcherbak, I. , Millar, N. , Kravchenko, A. N. , & Robertson, G. P. (2016). Long‐term nitrous oxide fluxes in annual and perennial agricultural and unmanaged ecosystems in the upper Midwest USA. Global Change Biology, 22(11), 3594–3607. 10.1111/gcb.13426 [DOI] [PubMed] [Google Scholar]
  13. Grandy, A. S. , & Robertson, G. P. (2007). Land‐use intensity effects on soil organic carbon accumulation rates and mechanisms. Ecosystems, 10(1), 59–74. 10.1007/s10021-006-9010-y [DOI] [Google Scholar]
  14. Groffman, P. M. , Altabet, M. A. , Böhlke, J. K. , Butterbach‐Bahl, K. , David, M. B. , Firestone, M. K. , Giblin, A. E. , Kana, T. M. , Nielsen, L. P. & Voytek, M. A. (2006). Methods for measuring denitrification: Diverse approaches to a difficult problem. Ecological Applications, 16(6), 2091–2122. 10.1890/1051-0761(2006)016[2091:Mfmdda]2.0.Co;2 [DOI] [PubMed] [Google Scholar]
  15. Habteselassie, M. , Xu, L. , & Norton, J. (2013). Ammonia‐oxidizer communities in an agricultural soil treated with contrasting nitrogen sources. Frontiers in Microbiology, 4(326). 10.3389/fmicb.2013.00326 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Hazard, C. , Prosser, J. I. , & Nicol, G. W. (2021). Use and abuse of potential rates in soil microbiology. Soil Biology and Biochemistry, 157, 108242. 10.1016/j.soilbio.2021.108242 [DOI] [Google Scholar]
  17. Hink, L. , Gubry‐Rangin, C. , Nicol, G. W. , & Prosser, J. I. (2018). The consequences of niche and physiological differentiation of archaeal and bacterial ammonia oxidisers for nitrous oxide emissions. The ISME Journal, 12(4), 1084–1093. 10.1038/s41396-017-0025-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Hink, L. , Nicol, G. W. , & Prosser, J. I. (2017). Archaea produce lower yields of N2O than bacteria during aerobic ammonia oxidation in soil. Environmental Microbiology, 19(12), 4829–4837. 10.1111/1462-2920.13282 [DOI] [PubMed] [Google Scholar]
  19. Holland, E. A. , Robertson, G. P. , Greenberg, J. , Groffman, P. M. , Boone, R. D. , & Gosz, J. R. (1999). Soil CO2, N2O, and CH4 exchange. In Robertson G. P., Coleman D. C., Bledsoe C. S., & Sollins P. (Eds.), Standard soil methods for long‐term ecological research (pp. 185–201). Oxford University Press. [Google Scholar]
  20. Hynes, R. K. , & Knowles, R. (1982). Effect of acetylene on autotrophic and heterotrophic nitrification. Canadian Journal of Microbiology, 28(3), 334–340. 10.1139/m82-049 [DOI] [Google Scholar]
  21. Jia, Z. , & Conrad, R. (2009). Bacteria rather than Archaea dominate microbial ammonia oxidation in an agricultural soil. Environmental Microbiology, 11(7), 1658–1671. 10.1111/j.1462-2920.2009.01891.x [DOI] [PubMed] [Google Scholar]
  22. Kahmark, K. , Millar, N. , & Robertson, G. P. (2020). Static chamber method for measuring soil greenhouse gas lluxes. Zenodo. 10.5281/zenodo.4630396 [DOI] [Google Scholar]
  23. Kellner, K. (2017). jagsUI: A wrapper around ‘rjags’ to streamline ‘JAGS’ analyses (Version 1.4.9). https://cran.r‐project.org/web/packages/jagsUI/index.html [Google Scholar]
  24. Kits, K. D. , Jung, M.‐Y. , Vierheilig, J. , Pjevac, P. , Sedlacek, C. J. , Liu, S. , Herbold, C. , Stein, L. Y. , Richter, A. , Wissel, H. , Brüggemann, N. , Wagner, M. , & Daims, H. (2019). Low yield and abiotic origin of N2O formed by the complete nitrifier Nitrospira inopinata . Nature Communications, 10(1), 1836. 10.1038/s41467-019-09790-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Kong, Y. , Ling, N. , Xue, C. , Chen, H. , Ruan, Y. , Guo, J. , Zhu, C. , Wang, M. , Shen, Q. , & Guo, S. (2019). Long‐term fertilization regimes change soil nitrification potential by impacting active autotrophic ammonia oxidizers and nitrite oxidizers as assessed by DNA stable isotope probing. Environmental Microbiology, 21(4), 1224–1240. 10.1111/1462-2920.14553 [DOI] [PubMed] [Google Scholar]
  26. Koper, T. E. , Stark, J. M. , Habteselassie, M. Y. , & Norton, J. M. (2010). Nitrification exhibits Haldane kinetics in an agricultural soil treated with ammonium sulfate or dairy‐waste compost. FEMS Microbiology Ecology, 74(2), 316–322. 10.1111/j.1574-6941.2010.00960.x [DOI] [PubMed] [Google Scholar]
  27. Kuenen, J. G. , & Robertson, L. A. (1994). Combined nitrification‐denitrification processes. FEMS Microbiology Reviews, 15(2), 109–117. 10.1111/j.1574-6976.1994.tb00129.x [DOI] [Google Scholar]
  28. Lewicka‐Szczebak, D. , Augustin, J. , Giesemann, A. , & Well, R. (2017). Quantifying N2O reduction to N2 based on N2O isotopocules—Validation with independent methods (helium incubation and 15N gas flux method). Biogeosciences, 14(3), 711–732. 10.5194/bg-14-711-2017 [DOI] [Google Scholar]
  29. Liang, D. , Ouyang, Y. , Tiemann, L. , & Robertson, G. P. (2020). Niche differentiation of bacterial versus archaeal soil nitrifiers induced by ammonium inhibition along a management gradient. Frontiers in Microbiology, 11(2753). 10.3389/fmicb.2020.568588 [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Limpert, E. , Stahel, W. A. , & Abbt, M. (2001). Log‐normal distributions across the sciences: Keys and clues. BioScience, 51(5), 341–352. 10.1641/0006-3568(2001)051[0341:Lndats]2.0.Co;2 [DOI] [Google Scholar]
  31. Lu, X. , Bottomley, P. J. , & Myrold, D. D. (2015). Contributions of ammonia‐oxidizing archaea and bacteria to nitrification in Oregon forest soils. Soil Biology and Biochemistry, 85, 54–62. 10.1016/j.soilbio.2015.02.034 [DOI] [Google Scholar]
  32. Lu, X. , Nicol, G. W. , & Neufeld, J. D. (2018). Differential responses of soil ammonia‐oxidizing archaea and bacteria to temperature and depth under two different land uses. Soil Biology and Biochemistry, 120, 272–282. 10.1016/j.soilbio.2018.02.017 [DOI] [Google Scholar]
  33. Luehmann, M. D. , Peter, B. G. , Connallon, C. B. , Schaetzl, R. J. , Smidt, S. J. , Liu, W. , Kincare, K. A. , Walkowiak, T. A. , Thorlund, E. , & Holler, M. S. (2016). Loamy, two‐storied soils on the outwash plains of southwestern lower Michigan: Pedoturbation of loess with the underlying sand. Annals of the American Association of Geographers, 106(3), 551–572. 10.1080/00045608.2015.1115388 [DOI] [Google Scholar]
  34. Malhi, S. S. , & McGill, W. B. (1982). Nitrification in three Alberta soils: Effect of temperature, moisture and substrate concentration. Soil Biology and Biochemistry, 14(4), 393–399. 10.1016/0038-0717(82)90011-6 [DOI] [Google Scholar]
  35. Millar, N. , & Robertson, G. P. (2015). Nitrogen transfers and transformations in row‐crop ecosystems. In Hamilton S. K., Doll J. E., & Robertson G. P. (Eds.), The ecology of agricultural landscapes: Long‐term research on the path to sustainability (pp. 213–251). Oxford University Press. [Google Scholar]
  36. Myhre, G. , Shindell, D. , Bréon, F.‐M. , Collins, W. , Fuglestvedt, J. , Huang, J. , Koch, D. , Lamarque, J. F. , Lee, D. , Mendoza, B. , Nakajima, T. , Robock, A. , Stephens, G. , Takemura, T. & Zhang, H. (2013). Anthropogenic and natural radiative forcing. In Stocker T. F., Qin D., Plattner G.‐K., Tignor M., Allen S. K., Boschung J., Nauels A., Xia Y., Bex V., & Midgley P. M. (Eds.), Climate change 2013—The physical science basis: Contribution of working group I to the fifth assessment report of the intergovernmental panel on climate change (pp. 659–740). Cambridge University Press. [Google Scholar]
  37. Norton, J. M. , & Stark, J. M. (2011). Regulation and measurement of nitrification in terrestrial systems. In Klotz M. G. (Ed.), Methods in enzymology (Vol. 486, pp. 343–368). Academic Press. [DOI] [PubMed] [Google Scholar]
  38. Opdyke, M. R. , Ostrom, N. E. , & Ostrom, P. H. (2009). Evidence for the predominance of denitrification as a source of N2O in temperate agricultural soils based on isotopologue measurements. Global Biogeochemical Cycles, 23(4). 10.1029/2009gb003523 [DOI] [Google Scholar]
  39. Orellana, L. H. , Chee‐Sanford, J. C. , Sanford, R. A. , Löffler, F. E. , & Konstantinidis, K. T. (2018). Year‐round shotgun metagenomes reveal stable microbial communities in agricultural soils and novel ammonia oxidizers responding to fertilization. Applied and Environmental Microbiology, 84(2), e01646–e11617. 10.1128/AEM.01646-17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Ostrom, N. E. , Sutka, R. , Ostrom, P. H. , Grandy, A. S. , Huizinga, K. M. , Gandhi, H. , von Fischer, J. C. , & Robertson, G. P. (2010). Isotopologue data reveal bacterial denitrification as the primary source of N2O during a high flux event following cultivation of a native temperate grassland. Soil Biology and Biochemistry, 42(3), 499–506. 10.1016/j.soilbio.2009.12.003 [DOI] [Google Scholar]
  41. O'Sullivan, C. A. , Whisson, K. , Treble, K. , Roper, M. M. , Micin, S. F. , & Ward, P. R. (2017). Biological nitrification inhibition by weeds: Wild radish, brome grass, wild oats and annual ryegrass decrease nitrification rates in their rhizospheres. Crop and Pasture Science, 68(8). 10.1071/cp17243 [DOI] [Google Scholar]
  42. Ouyang, Y. , Norton, J. M. , & Stark, J. M. (2017). Ammonium availability and temperature control contributions of ammonia oxidizing bacteria and archaea to nitrification in an agricultural soil. Soil Biology and Biochemistry, 113, 161–172. 10.1016/j.soilbio.2017.06.010 [DOI] [Google Scholar]
  43. Ouyang, Y. , Norton, J. M. , Stark, J. M. , Reeve, J. R. , & Habteselassie, M. Y. (2016). Ammonia‐oxidizing bacteria are more responsive than archaea to nitrogen source in an agricultural soil. Soil Biology and Biochemistry, 96, 4–15. 10.1016/j.soilbio.2016.01.012 [DOI] [Google Scholar]
  44. Parton, W. J. , Holland, E. A. , Del Grosso, S. J. , Hartman, M. D. , Martin, R. E. , Mosier, A. R. , Ojima, D. S. , & Schimel, D. S. (2001). Generalized model for NOx and N2O emissions from soils. Journal of Geophysical Research: Atmospheres, 106(D15), 17403–17419. 10.1029/2001jd900101 [DOI] [Google Scholar]
  45. Phillips, C. J. , Harris, D. , Dollhopf, S. L. , Gross, K. L. , Prosser, J. I. , & Paul, E. A. (2000). Effects of agronomic treatments on structure and function of ammonia‐oxidizing communities. Applied and Environmental Microbiology, 66(12), 5410–5418. 10.1128/aem.66.12.5410-5418.2000 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Prosser, J. I. , Hink, L. , Gubry‐Rangin, C. , & Nicol, G. W. (2020). Nitrous oxide production by ammonia oxidizers: Physiological diversity, niche differentiation and potential mitigation strategies. Global Change Biology, 26(1), 103–118. 10.1111/gcb.14877 [DOI] [PubMed] [Google Scholar]
  47. Qin, H. , Xing, X. , Tang, Y. , Hou, H. , Yang, J. , Shen, R. , Zhang, W. , Liu, Y. I. , & Wei, W. (2019). Linking soil N2O emissions with soil microbial community abundance and structure related to nitrogen cycle in two acid forest soils. Plant and Soil, 435(1–2), 95–109. 10.1007/s11104-018-3863-7 [DOI] [Google Scholar]
  48. R Core Team . (2020). R: A language and environment for statistical computing. R Foundation for Statistical Computing. https://www.R‐project.org/ [Google Scholar]
  49. Revell, L. E. , Bodeker, G. E. , Smale, D. , Lehmann, R. , Huck, P. E. , Williamson, B. E. , Rozanov, E. , & Struthers, H. (2012). The effectiveness of N2O in depleting stratospheric ozone. Geophysical Research Letters, 39(15). 10.1029/2012GL052143 [DOI] [Google Scholar]
  50. Robertson, G. P. , & Groffman, P. M. (2021). Nitrogen transformations: Fixation, mineralization‐immobilization, nitrification, denitrification, and movement. In Paul E. A. & Frey S. D. (Eds.), Soil microbiology, ecology, and biochemistry (5th ed.). Elsevier. [Google Scholar]
  51. Robertson, G. P. , & Hamilton, S. K. (2015). Long‐term ecological research at the kellogg biological station LTER site: Conceptual and experimental framework. In Hamilton S. K., Doll J. E., & Robertson G. P. (Eds.), The ecology of agricultural landscapes: Long‐term research on the path to sustainability (pp. 1–32). Oxford University Press. [Google Scholar]
  52. Robertson, G. P. , Paul, E. A. , & Harwood, R. R. (2000). Greenhouse gases in intensive agriculture: Contributions of individual gases to the radiative forcing of the atmosphere. Science, 289(5486), 1922–1925. 10.1126/science.289.5486.1922 [DOI] [PubMed] [Google Scholar]
  53. Robertson, G. P. , & Tiedje, J. M. (1984). Denitrification and nitrous oxide production in successional and old‐growth Michigan forests. Soil Science Society of America Journal, 48(2), 383–389. 10.2136/sssaj1984.03615995004800020032x [DOI] [Google Scholar]
  54. Robertson, G. P. , & Tiedje, J. M. (1987). Nitrous oxide sources in aerobic soils: Nitrification, denitrification and other biological processes. Soil Biology and Biochemistry, 19(2), 187–193. 10.1016/0038-0717(87)90080-0 [DOI] [Google Scholar]
  55. Schimel, J. P. , Firestone, M. K. , & Killham, K. S. (1984). Identification of heterotrophic nitrification in a Sierran forest soil. Applied and Environmental Microbiology, 48(4), 802–806. 10.1128/aem.48.4.802-806.1984 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Schlüter, S. , Henjes, S. , Zawallich, J. , Bergaust, L. , Horn, M. , Ippisch, O. , Vogel, H.‐J. , & Dörsch, P. (2018). Denitrification in soil aggregate analogues‐effect of aggregate size and oxygen diffusion. Frontiers in Environmental Science, 6(17). 10.3389/fenvs.2018.00017 [DOI] [Google Scholar]
  57. Séneca, J. , Pjevac, P. , Canarini, A. , Herbold, C. W. , Zioutis, C. , Dietrich, M. , Simon, E. , Prommer, J. , Bahn, M. , Pötsch, E. M. , Wagner, M. , Wanek, W. , & Richter, A. (2020). Composition and activity of nitrifier communities in soil are unresponsive to elevated temperature and CO2, but strongly affected by drought. The ISME Journal, 14(12), 3038–3053. 10.1038/s41396-020-00735-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Shcherbak, I. , & Robertson, G. P. (2019). Nitrous oxide (N2O) from subsurface soils of agricultural ecosystems. Ecosystems, 22(7), 1650–1663. 10.1007/s10021-019-00363-z [DOI] [Google Scholar]
  59. Shen, J.‐P. , Zhang, L.‐M. , Zhu, Y.‐G. , Zhang, J.‐B. , & He, J.‐Z. (2008). Abundance and composition of ammonia‐oxidizing bacteria and ammonia‐oxidizing archaea communities of an alkaline sandy loam. Environmental Microbiology, 10(6), 1601–1611. 10.1111/j.1462-2920.2008.01578.x [DOI] [PubMed] [Google Scholar]
  60. Smith, K. A. (1980). A model of the extent of anaerobic zones in aggregated soils, and its potential application to estimates of denitrification. Journal of Soil Science, 31(2), 263–277. 10.1111/j.1365-2389.1980.tb02080.x [DOI] [Google Scholar]
  61. Snapp, S. , Wilke, B. , Gentry, L. E. , & Zoellner, D. (2017). Compost legacy down‐regulates biological nitrogen fixation in a long‐term field experiment. Agronomy Journal, 109(6), 2662–2669. 10.2134/agronj2017.03.0152 [DOI] [Google Scholar]
  62. Stark, J. M. , & Firestone, M. K. (1996). Kinetic characteristics of ammonium‐oxidizer communities in a California oak woodland‐annual grassland. Soil Biology and Biochemistry, 28(10), 1307–1317. 10.1016/S0038-0717(96)00133-2 [DOI] [Google Scholar]
  63. Stein, L. Y. (2019). Insights into the physiology of ammonia‐oxidizing microorganisms. Current Opinion in Chemical Biology, 49, 9–15. 10.1016/j.cbpa.2018.09.003 [DOI] [PubMed] [Google Scholar]
  64. Stein, L. Y. (2020). The long‐term relationship between microbial metabolism and greenhouse gases. Trends in Microbiology, 28(6), 500–511. 10.1016/j.tim.2020.01.006 [DOI] [PubMed] [Google Scholar]
  65. Stevens, R. J. , Laughlin, R. J. , Burns, L. C. , Arah, J. R. M. , & Hood, R. C. (1997). Measuring the contributions of nitrification and denitrification to the flux of nitrous oxide from soil. Soil Biology and Biochemistry, 29(2), 139–151. 10.1016/S0038-0717(96)00303-3 [DOI] [Google Scholar]
  66. Subbarao, G. V. , Nakahara, K. , Hurtado, M. P. , Ono, H. , Moreta, D. E. , Salcedo, A. F. , Yoshihashi, A. T. , Ishikawa, T. , Ishitani, M. , Ohnishi‐Kameyama, M. , Yoshida, M. , Rondon, M. , Rao, I. M. , Lascano, C. E. , Berry, W. L. , & Ito, O. (2009). Evidence for biological nitrification inhibition in Brachiaria pastures. Proceedings of the National Academy of Sciences of the United States of America, 106(41), 17302–17307. 10.1073/pnas.0903694106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Sutka, R. L. , Adams, G. C. , Ostrom, N. E. , & Ostrom, P. H. (2008). Isotopologue fractionation during N2O production by fungal denitrification. Rapid Communications in Mass Spectrometry, 22(24), 3989–3996. 10.1002/rcm.3820 [DOI] [PubMed] [Google Scholar]
  68. Sutka, R. L. , Ostrom, N. E. , Ostrom, P. H. , Breznak, J. A. , Gandhi, H. , Pitt, A. J. , & Li, F. (2006). Distinguishing nitrous oxide production from nitrification and denitrification on the basis of isotopomer abundances. Applied and Environmental Microbiology, 72(1), 638–644. 10.1128/aem.72.1.638-644.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Suwa, Y. (1994). Ammonia‐oxidizing bacteria with different sensitivities to (NH4)2SO4 in activated sludges. Water Research, 28(7), 1523–1532. 10.1016/0043-1354(94)90218-6 [DOI] [Google Scholar]
  70. Taylor, A. E. , Taylor, K. , Tennigkeit, B. , Palatinszky, M. , Stieglmeier, M. , Myrold, D. D. , Schleper, C. , Wagner, M. , & Bottomley, P. J. (2015). Inhibitory effects of C2 to C10 1‐alkynes on ammonia oxidation in two Nitrososphaera species. Applied and Environmental Microbiology, 81(6), 1942–1948. 10.1128/aem.03688-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Taylor, A. E. , Vajrala, N. , Giguere, A. T. , Gitelman, A. I. , Arp, D. J. , Myrold, D. D. , Sayavedra‐Soto, L. , & Bottomley, P. J. (2013). Use of aliphaticn‐alkynes to discriminate soil nitrification activities of ammonia‐oxidizing thaumarchaea and bacteria. Applied and Environmental Microbiology, 79(21), 6544–6551. 10.1128/aem.01928-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Taylor, A. E. , Zeglin, L. H. , Dooley, S. , Myrold, D. D. , & Bottomley, P. J. (2010). Evidence for different contributions of archaea and bacteria to the ammonia‐oxidizing potential of diverse Oregon soils. Applied and Environmental Microbiology, 76(23), 7691–7698. 10.1128/AEM.01324-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Terry, R. E. , & Duxbury, J. M. (1985). Acetylene decomposition in soils. Soil Science Society of America Journal, 49(1), 90–94. 10.2136/sssaj1985.03615995004900010018x [DOI] [Google Scholar]
  74. Topp, E. , & Germon, J. C. (1986). Acetylene metabolism and stimulation of denitrification in an agricultural soil. Applied and Environmental Microbiology, 52(4), 802–806. 10.1128/aem.52.4.802-806.1986 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Venterea, R. T. , Clough, T. J. , Coulter, J. A. , Breuillin‐Sessoms, F. , Wang, P. , & Sadowsky, M. J. (2015). Ammonium sorption and ammonia inhibition of nitrite‐oxidizing bacteria explain contrasting soil N2O production. Scientific Reports, 5, 12153. 10.1038/srep12153 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Wang, Q. , Zhang, L.‐M. , Shen, J.‐P. , Du, S. , Han, L.‐L. , & He, J.‐Z. (2016). Nitrogen fertiliser‐induced changes in N2O emissions are attributed more to ammonia‐oxidising bacteria rather than archaea as revealed using 1‐octyne and acetylene inhibitors in two arable soils. Biology and Fertility of Soils, 52(8), 1163–1171. 10.1007/s00374-016-1151-3 [DOI] [Google Scholar]
  77. Wang, X. , Wang, S. , Jiang, Y. , Zhou, J. , Han, C. , & Zhu, G. (2020). Comammox bacterial abundance, activity, and contribution in agricultural rhizosphere soils. Science of the Total Environment, 727, 138563. 10.1016/j.scitotenv.2020.138563 [DOI] [PubMed] [Google Scholar]
  78. Wrage, N. , Velthof, G. L. , van Beusichem, M. L. , & Oenema, O. (2001). Role of nitrifier denitrification in the production of nitrous oxide. Soil Biology and Biochemistry, 33(12), 1723–1732. 10.1016/S0038-0717(01)00096-7 [DOI] [Google Scholar]
  79. Wrage‐Mönnig, N. , Horn, M. A. , Well, R. , Müller, C. , Velthof, G. , & Oenema, O. (2018). The role of nitrifier denitrification in the production of nitrous oxide revisited. Soil Biology and Biochemistry, 123, A3–A16. 10.1016/j.soilbio.2018.03.020 [DOI] [Google Scholar]
  80. Wu, Y. , Lu, L. U. , Wang, B. , Lin, X. , Zhu, J. , Cai, Z. , Yan, X. , & Jia, Z. (2011). Long‐term field fertilization significantly alters community structure of ammonia‐oxidizing bacteria rather than archaea in a paddy soil. Soil Science Society of America Journal, 75(4), 1431–1439. 10.2136/sssaj2010.0434 [DOI] [Google Scholar]
  81. Xue, C. , Zhang, X. U. , Zhu, C. , Zhao, J. , Zhu, P. , Peng, C. , Ling, N. , & Shen, Q. (2016). Quantitative and compositional responses of ammonia‐oxidizing archaea and bacteria to long‐term field fertilization. Scientific Reports, 6, 28981. 10.1038/srep28981 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Zou, Y. , Hirono, Y. , Yanai, Y. , Hattori, S. , Toyoda, S. , & Yoshida, N. (2014). Isotopomer analysis of nitrous oxide accumulated in soil cultivated with tea (Camellia sinensis) in Shizuoka, central Japan. Soil Biology and Biochemistry, 77, 276–291. 10.1016/j.soilbio.2014.06.016 [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Material

Data Availability Statement

KBS MCSE long‐term field N2O flux measurements can be accessed at https://lter.kbs.msu.edu/datatables/28; long‐term in situ soil NH4+‐N concentrations are available at https://lter.kbs.msu.edu/datatables/55. Soil bulk densities of deep core soil samples are available from https://lter.kbs.msu.edu/datatables/308. Species composition of Early successional and Deciduous forest systems is available from https://lter.kbs.msu.edu/datatables/237 and https://lter.kbs.msu.edu/datatables/238. All other data used in this study are available at datadryad.org (https://doi.org/10.5061/dryad.37pvmcvkz).

Code availability: The code for estimating log‐normal mean of in situ N2O fluxes and the maximum contribution of nitrification to total N2O are shown in Supporting Information.


Articles from Global Change Biology are provided here courtesy of Wiley

RESOURCES