Introduction

Nitrite (NO2), as an intermediate of nearly all N transformations, is a crucial compound to understand the complexity of the N soil cycle with its many contributing pathways. Moreover, as a very reactive compound it usually occurs at very low concentrations, hence conveying information on currently active N transformations. The Ntrace model used for interpretation of 15N labelled soil studies has been recently expanded with the NO2 content and isotopic analyses, which vastly increased its interpretation perspectives1. Thanks to incorporation of NO2 dynamics in this model it appeared possible to distinguish and quantify three NO2 and N2O production pathways: denitrification, autotrophic nitrification and heterotrophic nitrification. Although 15N tracing studies can precisely identify various soil N transformations1,2, they require addition of 15N -labelled substances, which is associated with additional fertilization, soil disturbance, and potential problems with label distribution homogeneity3,4. Moreover, due to high costs and fast consumption of the 15N label, 15N tracing approach can be applied mostly for short-term and micro-plot studies5. Development of reliable methods for identifying N transformations based on natural abundance stable isotopes can overcome these problems and provide an approach allowing studies in undisturbed soil conditions ensuring original N transformation rates that can be traced in larger time and space scale.

Natural abundance NO2isotope studies are so far mostly applied in aquatic studies6,7,8,9 and appeared particularly informative for the oceanic oxygen deficient zones, where NO2 can be accumulated7,9. However, for soil studies the natural abundance NO2 analyses are so far lacking. Also in soils NO2 accumulation may happen and the monitoring of NO2 content in soils can provide important information to understand the N cycle10,11,12. In particular, NO2 plays a central role for N2O formation12,13. However, even in situations when NO2 accumulation is not observed, and the interpretation of, typically very low, soil NO2-contents is ambiguous, the N transformations can potentially be followed by the stable isotopic signature of NO2, which has neither been tested nor applied so far.

Nitrite can be formed during nitrate reduction (NAR) in the course of denitrification, ammonium oxidation (AOX) in the course of autotrophic nitrification and organic N oxidation (ORG) associated with heterotrophic nitrification, and consumed during nitrite reduction (NIR) to NO or N2O, and nitrite oxidation (NIOX) to NO31,7. Each of these sources and sinks are characterised by specific isotopic fractionation7,9,14, which makes it possible to trace them back to their origins and sinks of NO2, and consequently, for a better understanding of the N cycling7.

This study presents the first attempt to interpret the NO2 isotopic signatures (δ15NNO2− and δ18ONO2−) in agricultural soil to decipher soil transformation processes. Three laboratory incubations were performed: under oxic conditions with lower water content (L1), under oxic conditions with higher water content (L2) and under anoxic conditions (L3), to monitor the differences when various N transformation processes are enhanced. The incubations at natural abundance level (NA treatment) and under 15N enrichment (15NO3 treatment and 15NH4 treatment) were performed simultaneously. Based on the 15N treatments the Ntrace model1 was applied to determine NO2 sources and sinks. The results of NA treatment were used to construct the soil NO2 model, which is based on the model used for oceanic studies7, including the processes that have contributed to production and consumption of NO2 in soils. This study provides the first attempt to validate the results of NO2 isotope modelling with an independent 15N tracing approach.

Results

Soil NO2 characteristics

The oxic experiment was performed in two moisture treatments: L1 (dryer conditions) and L2 (wetter conditions), with water addition in the middle of experiment which increased the soil moisture from 61 to 68% water-filled pores space (WFPS) for L1a and L1b and from 72 to 81% WFPS for L2a and L2b, respectively. The detailed experimental conditions and information on general soil properties can be found in15 and in the supplement. NO2 content varied from 0.6 to 1.4 μmol N kg−1 soil for L1 and from 0.1 to 4.7 for L2, whereas the NO3 content was three orders higher and quite stable ranging from 1300 to 1700 μmol N kg−1 soil. The δ15NNO2− was similar for L1 and L2 with a mean of 3.2 ± 4.2‰ and 3.9 ± 4.2‰, respectively, whereas δ15NNO3− was very stable with a mean of 4.5 ± 0.4‰ and 4.7 ± 0.6‰, respectively. There was a negative correlation between δ15NNO2− and the NO2-content (Fig. 1A). Similar values for δ18ONO2− were found for both L1 and L2 with a mean of 11.8 ± 2.8‰ and 12.5 ± 5.0‰, respectively.

Figure 1
figure 1

Relationship between NO2 isotopic signature δ15NNO2− (A) or δ18ONO2− (B) and reciprocal NO2 content (Keeling plot analysis—see “Methods” section). For δ18ONO2− (B) the dashed line indicates the δ value of NO2 in full equilibrium with ambient water in 20 °C of 8.6‰23 and the arrows indicate the direction of change in δ values of NO2in course of equilibration with water (points ‘move’ towards full equilibrium). For L3 the first samples taken after 24 h are marked with circles and the second samples taken after 48 h are shown with crosses only. Note the logarithmic scale of the X-axis.

The anoxic experiment L3 was performed to favour denitrification. NO2 content was much higher when compared to oxic conditions (L1 and L2) reaching 63.9 ± 12.5 μmol N kg−1 soil and 32.4 ± 8.9 μmol kg−1 soil after 24 h and 48 h of incubation, respectively. The average δ15NNO2− was − 24.8 ± 3.3‰ without significant differences between the two samplings, whereas δ18ONO2− showed significantly lower values of 4.4 ± 0.4‰ after 24 h compared to 6.6 ± 1.0‰ after 48 h (Fig. 1B).

In both 15N treatments (15NH4+ and 15NO3) in L1 and L2, we observed a sudden drop in 15N abundance in NO2 (a15NNO2−) from 12.7 to 5.1 at.% after water addition to the soil, whereas 15N abundance in NO3 (a15NNO3−) showed only slight decrease from 13.2 to 12.1 at.% (means for all treatments, individual values in Fig. S1 and Table S1). This indicates an incorporation of another source of unlabelled NO2 for the wet part of these experiments. In natural abundance isotopes this change was also reflected in a higher apparent isotope effect 15ηNO2−/ NO3−. For the wet part it was even positive with an average + 1.6‰, whereas for the dry part it was lower with − 4.4‰ (Table 1). Interestingly, this significant isotopic change in NO2 was not reflected in the N2O 15N abundance (a15NN2O) in the 15N treatments (Fig. S1, Table S1).

Table 1 Apparent isotopic fractionation factors for δ15N of NO3, NO2 and N2O.

Isotope effects between NO3 , NO2 and N2O

With the NA dataset for NO2 presented in this paper and the dataset for N2O presented in a previous paper for L1 and L215, and here for L3 (Table S1), we can investigate the relation between the isotopic characteristics of both N compounds and determine the apparent isotope effects between NO2 and N2O, and comparing them with isotope effects between NO3 and N2O. Therefore, we need the isotopic signatures of the produced N2O prior to isotopic fractionation due to N2O reduction. For L3 the incubations were partially conducted with N2O reduction inhibition (acetylated treatments) and we only report here the δN2O values of the inhibited treatment (Table S1) which represent the produced N2O isotopic signatures. For L1 and L2 a detailed study of N2O reduction was performed15 where the N2O reduced fraction (rN2O) was determined with 15N treatment, and the produced N2O (δN2O_p) can be calculated according to the equation:

$${\delta }_{N2O\_p}={\delta }_{N2{O}_{m}}-\mathrm{ln}{r}_{N2O}*{\varepsilon }_{red}$$

based on the measured N2O (δN2O_m) and the isotopic fractionation associated with N2O reduction (εred)15. The determined apparent N isotope effects (15η) was calculated as:

$${{}^{15}\eta }_{product-substrate}={\delta }^{15}{N}_{product}-{\delta }^{15}{N}_{substrate}$$

Determination of NO2 dominant source

Keeling plots were applied to identify the NO2 dominant source (see Methods Section for methodical explanation)15,16,17,18. For the oxic experiment, a significant linear fit between δ15NNO2− and reciprocal NO2 content was found, where an linear equation intercept of − 13.3‰ indicated the isotopic signature of the dominant NO2 source. It must be denitrification since the applied conditions of quite high soil moisture and nitrate amendment should have favoured denitrification. In a previous study, denitrification was identified as the dominant source for N2O15 and also the applied Ntrace model indicated the dominance of denitrification nitrate reduction (NAR) in the NO2-sources (fNAR of 0.53 and 0.55 for L1 and L2, respectively, Table 2). Hence, based on the value found from the Keeling plot (Fig. 1A) we can determine the nitrogen isotopic fractionation for denitrification (15εNAR) between δ15NNO3− (mean measured value) and δ15NNO2− (Keeling plot intercept) for this incubation experiment:

$$^{{{15}}} \varepsilon_{NAR} \left( {{\text{NO}}_{{2}}^{ - } /{\text{ NO}}_{{3}}^{ - } } \right) = - {13}.{3}\textperthousand{-}\left( { + {4}.{5}\textperthousand} \right) = - {17}.{8}\textperthousand $$

This value fits quite well in the literature range7,14 and is further used in the NO2 isotope model as 15εNAR.

Table 2 Nitrite stable isotope model to determine sources mixing proportions.

Under anoxic conditions, denitrification should be the only source of NO2, hence a typical Keeling correlation is not expected. We rather observed the opposite trend in L3 than under oxic conditions, i.e. lower δ15NNO2− values with lower NO2 contents (Fig. 1A). Most probably this reflects the variability of apparent isotope effects, which are typically larger for lower reaction rates19,20.

For δ18ONO2− values, beside sources mixing, we also deal with isotope exchange of O-atoms between NO2 and ambient water, hence the Keeling plot method cannot be applied. We observed that δ18ONO2− values were modified by the O exchange process, especially under anoxic conditions (L3), where lower NO2 content and incubation progress shifted δ18ONO2− values towards equilibrium with water (NO2 full eq, Fig. 1B). NO2 samples taken after 48 h showed more equilibrated δ18ONO2− values and lower NO2 content.

NO2 isotope model

The model is constructed based on the NO2 isotope model proposed for oceanic studies7 and adapted for typical soil N pathways after the Ntrace model, designed for 15N labelled soil studies applying NO2 as a key intermediate in soil N transformations1, assuming steady state conditions. It takes into account three main NO2 sources (NAR, AOX and ORG) and two main NO2 sinks (NIR and NIOX), as well as δ18ONO2− equilibration with ambient water (Fig. 2), according to the following equations:

$${\delta }^{15}{N}_{NO2-}={\delta }^{15}{N}_{NAR}*{f}_{NAR}+{\delta }^{15}{N}_{AOX}*{f}_{AOX}+{\delta }^{15}{O}_{ORG}*{f}_{ORG}-{{}^{15}\varepsilon }_{NIR}*{f}_{NIR}-{{}^{15}\varepsilon }_{NIOX}*{f}_{NIOX}$$
(1)
$${\delta }^{18}{O}_{NO2-}={(\delta }^{18}{O}_{NAR}*{f}_{NAR}+{\delta }^{18}{O}_{AOX}*{f}_{AOX}+{\delta }^{18}{O}_{ORG}*{f}_{ORG}-{{}^{18}\varepsilon }_{NIR}*{f}_{NIR}-{{}^{18}\varepsilon }_{NIOX}*{f}_{NIOX})*\left(1-x\right)+{\delta }^{18}{O}_{eq}*x$$
(2)

where δNO2− is the measured residual NO2 isotopic signature, δNAR/AOX/ORG are the isotopic signatures of source NO2 calculated with the measured stable isotope values for NO2 substrates (NO3, NH4+, Norg, respectively for three sources, Table 2) and the characteristic isotopic fractionation associated with each NO2 formation pathway (εNAR/AOX/ORG, Table 2). εNIR/NIOX are the isotopic fractionation factors associated with NO2 sinks (εNIR/NIOX, Table 2). See also Methods Section for detailed description of isotope effects for particular processes; final values used in the model are shown in Table 2. The δ18Oeq stands for O isotopic signature of NO2 in complete equilibrium with water, which equals 8.6‰ for the incubation temperature of 20 °C7 and δ18OH2O of − 5‰. x is the extent of oxygen atom exchange between nitrate and ambient water determined with the 17O approach21 for N2O originating from denitrification processes under anoxic conditions (L3) and is equal 0.25 (see Methods Section). The exchange for NO2 cannot be higher than the value determined for N2O. Since most of the exchange observed for N2O is associated with the NO2-H2O isotope exchange21 this value was incorporated in the model calculations.

Figure 2
figure 2

General scheme of the NO2 stable isotope model. The isotopic signatures of NO2 sources shown are based on the measured mean isotopic signatures of substrates for L1 and L2 and the isotopic fractionation associated with NAR, AOX and ORG (Table 1). Dashed gray arrows illustrate the mixing of 3 NO2 sources with mean mixing proportions found in Ntrace study (fNAR = 0.55, fAOX = 0.15, fORG = 0.30, Table 1) resulting in the produced δNO2- (grey open point). This δ value can be modified by NIR (red arrow) and NIOX (blue arrow). The δ18ONO2- after reduction or oxidation (red and blue open point, respectively) is further modified by equilibration with ambient water with the extent of 0.25 of the equilibrated oxygen atoms (red and blue filled point, respectively). The ratio of NO2- reduction and oxidation processes (red-ox ratio, here 4:1, as mean from Ntrace study, Table 1) determines the final δNO2- (purple point).

For NAR we were able to determine the isotope effect (15εNAR =  − 17.8) based on the Keeling plot (Fig. 1A) for L1 and L2. This value of − 17.8‰ is on the lower range of previously determined values22,23,24,25,26,27,28,29 (see summary in the Methods Section) and was used in the model for L1 and L2. For anoxic experiment L3 the fractionation was larger (Fig. 1A) and we included the typical 15εNAR value of − 30‰ in the model. For 18εNAR the fractionation must be very low to obtain the observed range of δ18ONO2−: − 10‰ for L3 and no fractionation for L1 and L2. Similar ranges of 18εNAR values were modelled previously with indication of lower values for smaller reaction rates21. This is in accordance with our observations indicating much lower N transformation rates for the oxic experiments L1 and L2 than for the anoxic L3. Also the isotopic fractionation for NIR appears to be lower under anoxic conditions, where NO2 is accumulating. We obtained best fit between modelled and measured values for L3 when no fractionation associated with NIR was assumed (Table 2). This can be due to observed accumulation of NO2, indicating that the steady-state model assumption is not valid. In case of NO2 accumulation the isotopic fractionation associated with its reduction has a very low impact on the final isotopic signature of the residual NO2.

NO2 turnover

When δ18ONO2- values are not completely equilibrated with soil water, measured δ18ONO2− values can be used to estimate the rates of biological NO2 turnover relative to abiotic exchange7. This estimation is based on the abiotic equilibration rate as a function of temperature and pH7. Furthermore, we can determine the flux of NO2 oxygen atoms abiotic exchange as Feq = k*CNO2−. The NO2 flux of biological production (or consumption) can be determined from the δ18ONO2- isotope mass balance following the method proposed for oceanic studies7 adapted to soil NO2 transformations:

$${F}_{B}=\frac{{F}_{eq}({ \delta }^{18}{O}_{NO2-}-{\delta }^{18}{O}_{eq})}{{\delta }^{18}{O}_{NAR}*{f}_{NAR}+{\delta }^{18}{O}_{AOX}*{f}_{AOX}+{\delta }^{18}{O}_{ORG}*{f}_{ORG}-{{}^{18}\varepsilon }_{NIR}*{f}_{NIR}-{{}^{18}\varepsilon }_{NIOX}*{f}_{NIOX}-{ \delta }^{18}{O}_{NO2-}}$$
(3)

where δ18ONO2- is the measured NO2 , δ18ONAR/AOX/ORG are the calculated NO2 sources NAR, AOX and ORG, 18εNIR/NIOX are the isotopic fractionation associated with NO2 sinks NIR and NIOX, f are the respective contributions of NO2 sources determined by Ntrace model (Table 2), and δ18Oeq is the value for NO2 in complete equilibrium with ambient water (of 8.6‰ for this case study). In turnover rate calculations (Table 3) we have neglected NIOX because due to the inverse fractionation of this process for some cases the isotope mass balance did not work due to unrealistic discrepancies between calculated and measured δ18ONO2− values. Since 18εNIOX can be very low30 and the NIOX contribution is low for our case study (up to 30%, Table 2), this process has most probably little impact on the final δ18ONO2− values. In the NO2 isotope model, neglecting NIOX would result in higher final modelled δ18ONO2 values of 13.3‰ for both treatments, which would fit to the measured values equally well as when the NIOX fractionation is included (Table 2).

Table 3 Nitrite transformation fluxes (due to equilibration (Feq) and biological turnover (FB)) and residence time due to biological turnover (ΓB) or abiotic equilibration (Γeq) determined with δ18ONO2- values compared to nitrite turnover rate determined with Ntrace model (FNtrace).

Our results indicate that the NO2 flux in L2 is larger than in L1 (Table 3), which is reasonable since L2 was the wetter treatment showing more intensive nitrogen fluxes based N2O and N2 fluxes, which were twice as high in L2 compared to L115. Similar differences can be observed here for calculated FB values (Table 3), however this is not directly confirmed by the Ntrace results.

Discussion

We found very good congruity between Ntrace and the NA NO2 model. The modelled δ18ONO2− and δ15NNO2− values using measured source fractions provided by the Ntrace model differed up to 1.2‰ and 4.0‰, respectively, when compared to true measured values (Table 2). When we solely used the NA NO2 model to assess the fraction of NO2-sources contribution based on δ15NNO2−, i.e., fitting modelled δ15NNO2− values to the true measured values by adjusting the fraction of NO2-sources contribution, the fitted fractions are in good agreement with fractions provided by the Ntrace model (Table 2). Both results show similar dominance of NAR in NO2- production (fNAR of ca. 0.55) but the NA NO2 model indicates even higher contribution of heterotrophic vs autotrophic nitrification (fORG vs. fAOX).

The Ntrace approach, which is able to identify the contribution of ORG, was actually the first one that paid attention to this process in soils31. Here, with the NA NO2 model we get a confirmation of the potentially high ORG relevance in soil N transformations. Without this process the final δ15NNO2− and δ18ONO2− values could not be explained. Namely, if only considering two source processes: NAR and AOX, to meet the measured δ15NNO2− value we would need domination of NAR (fNAR > 0.75) and to meet the measured δ18ONO2− value we would need an unrealistically high contribution of AOX (fAOX > 0.70). Hence, the application of both isotope signatures (δ15NNO2− and δ18ONO2−) simultaneously allows for a proper identification of NO2 sources.

The presented NO2 isotope model may not be very typical, since for our case study we had exceptionally high δ15NNH4+ values (from 36 to 100‰), hence this worked partially as a naturally low level 15N tracing allowing for very clear separation of NAR and AOX with δ15NNO2− values. In case of similar δ values for substrate NO3 and NH4+, this separation would be very weak, but still, in combination with δ18ONO2− values, may be useful in assessing source contributions. The 15N enrichment of NH4+ was not purposely induced but was a consequence of the fast ammonium consumption. Ntrace analysis revealed that the dominant ammonium sink is immobilisation, responsible for more than 90% of ammonium consumption. This process is associated with pronounced enrichment of residual ammonium in 15N32,33,34, which we observed in this study. The very fast NH4+ immobilisation and its further release due to Norg oxidation to NO3 were unexpected in this study and cannot be fully explained. The Ntrace model assumes the existence of the labile Norg pool, which is associated with these extremely fast fluxes. In the NA model, the assumed substrate for ORG nitrite production is the measured δ15N of the bulk organic N pool. This is probably the largest uncertainty in the model, since the labile Norg pool may be isotopically different than the bulk Norg. Similar uncertainties may also be associated with the measured bulk δ15NNO3−, since this value may be significantly higher in the intensively denitrifying soil microsites.

The NO2 isotopic signature time series in the 15N treatments (Fig. S1), strongly indicates the appearance of new unlabelled NO2 for L1 and L2 after water addition in the course of the incubation. Ntrace clearly indicated an increase in ORG contribution after water addition (from 0.11 to 0.49 and from 0.07 to 0.33 for L1 and L2, respectively, Table S3) and the NA NO2 model confirms this finding (from 0.11 to 0.52 and from 0.21 to 0.31 for L1 and L2, respectively, Table S3). This suggests that NA NO2 analyses can be used to trace the dynamic changes in soil N transformations.

The 15N treatment indicated that N2O 15N enrichment (a15NN2O) follows rather a15NNO3− than a15NNO2− (Fig. S1). This indicates that mostly NAR NO2 is further reduced and emitted as N2O and suggests that the ORG NO2 forms an isolated NO2 pool, as also suggested earlier1, which is probably not further reduced to N2O in significant magnitude. The calculated aP_N2O value representing the 15N enrichment of the 15N -pool derived N2O is higher than a15NN2O due to the contribution of non-labelled N2O to the total N2O flux, so that a15NN2O = fP_N2O * aP_N2O + (1- fP_N2O ) * aNA , where aNA is the 15N abundance of the natural abundance samples (0.367 at.%). However, aP_N2O values always have a higher 15N abundance than found for any soil N-pool (Table S1). This indicates that in denitrification soil microsites, where the 15N-pool derived N2O is produced, we deal with higher a15NNO2− and a15NNO3− values than the mean analysed values. This confirms the N2O emission originating from various isolated soil N-pools. The fraction of the 15N -pool N2O is dominating – from 0.7 to 1.0—with higher values for higher soil moisture (Table S1). This is in contrast to NO2 which gets 15N depleted after water addition (Fig. S1) and fNAR is only around 0.5.

Regarding the NA isotope effects, we can see that for pure denitrification processes under anoxic conditions in L3, we deal with very low 15ηNO2−NO3 values (Table 1) indicating a pronounced isotope effect between NO3 and NO2, whereas for L1 and L2, where NO2 is formed not only due to NAR but also AOX and ORG (Table S2), this effect is much smaller (as indicated by 15ηNO2−NO3 closer to 0, Table 1). Interestingly, in the wetter parts of the experiments (L1b, L2b), when increased contribution of ORG NO2 occurs, even an inverse effect is observed, i.e. NO2 is 15N enriched compared to NO3. As a result, for L1 and L2 very similar isotope effects for N2O production with both substrates are observed (15ηN2O-NO3 and 15ηN2O-NO2 are not significantly different). This is in contrast with L3, where 15ηN2O-NO2 is much lower compared to 15ηN2O-NO3. For oxic conditions, clearly the highest isotope effect for N2O production with both substrates is noted for L2b, for which the 15ηN2O-NO3 is nearest to the values typical for denitrification, as found in L3. Indeed, for L2b the denitrification pool derived fraction (fP_N2O, Table S1) is the highest. Also the previous study15 indicated that the N2O fraction produced due to denitrification, including both bacterial and fungal denitrification, is near 1 for this part of the experiment (L2b)15. However, the 15ηN2O-NO2 values are much lower for L2b than for L3 which is probably caused by admixture of other nitrite sources present for L2b but absent for L3.

Despite the fact that NO2−− turnover rates determined with δ18ONO2− (FB) differed somewhat from the FNtrace results (Table 3), the congruence can be considered to be adequate because very plausible ranges for NO2 turnover rates were observed. Both methods provide a similar range of values, however, the Ntrace model does not reflect the significant difference in turnover rates between L1 and L2 (Table 3). This may be due to lower sensitivity of the Ntrace approach in precise NO2 fluxes determination, since these are determined as a result of complex modelling of all N pools and the final result is an average of the best fit fluxes for both treatments (15NH4+ and 15NO3). This turnover rate estimation provides a unique opportunity to predicting process rates based on natural abundance isotopic measurements.

Summing up, the natural abundance signatures of NO2 can be applied for identification of NO2 sources by applying the NA isotope model, which allows to estimate the contribution of the main pathways: NAR, AOX and ORG. This study showed that these are in a very good agreement with the results provided by the Ntrace model. Moreover, analysis of δ18ONO2− values allows for estimation of NO2 turnover rates. The natural abundance signatures of NO2 may potentially be used in linking the soil N transformations with gaseous emissions in the form of N2O. However, this connection still cannot be fully understood and needs further studies.

Methods

Laboratory incubations

Oxic incubations: L1 and L2 experiment

Silt loam soil Albic Luvisol from arable cropland of Merklingsen experimental station located near Soest (North Rhine-Westphalia, Germany, 51° 34′ 15.5″ N, 8° 00′ 06.8″ E) was used in the incubations (0.87 silt, 0.11 clay, 0.02 sand). The soil density of intact cores was 1.3 g cm−3, pH value 6.8, total C content 0.0130, total N content 0.0016, organic matter content 0.0214, initial NO3 content 864 μmol N kg−1 dry soil and initial NH4+ content 50 μmol N kg−1 dry soil . The soil, upper 30 cm soil layer, was collected on the 18.01.2018 and the incubation was conducted from 19.02.2018 to 05.03.2018. The soil was air dried and sieved at 4 mm mesh size. Afterwards, the soil was rewetted to achieve a water content equivalent to 60% water-filled pore space (WFPS) and fertilised with 20 mg N per kg soil, added as NaNO3 (10 mg N) and NH4Cl (10 mg N). Three treatments were prepared: natural abundance (NA), labelled with 15N nitrate (15NO3) and labelled with 15N ammonium (15NH4). For the 15NO3 treatment, NaNO3 solution with 72 atom % 15N was added and for the 15NH4 treatment, NH4Cl solution with 63 atom % 15N was added. Then soils were thoroughly mixed to obtain homogenous distribution of water and fertilizer and an equivalent of 1.69 kg dry soil was repacked into each incubation column with bulk density of 1.3 g cm−3.

For each treatment 14 soil columns were prepared, and half of them received additional water injected on the top of the column (100 mL water added) to prepare two moisture treatments: L1 (61% WFPS) and L2 (72% WFPS). The incubation lasted 12 days. In the meantime, on the 6th day of incubation, water addition on the top of each column was repeated (80 mL water added) to increase the soil moisture in both treatments to ca. 68% WFPS in L1 and ca. 81% WFPS in L2. The strategy of adding water on the top of the column to achieve target water content was necessary to allow mixing and compaction at a suitable (low) water content of the soil and thus to optimise homogeneity of water and fertilizer distribution3. The incubation temperature was 20 °C. The columns were continuously flushed with a gas mixture with reduced N2 content to increase the measurements sensitivity (2% N2 and 21% O2 in He35) with a flow of 9 mL min−1. Gas samples were collected daily into two 12 mL septum-capped Exetainer vials (Labco Limited, Ceredigion, UK) connected to the vents of the incubation columns. Soil samples were collected 5 times during the incubation by sacrificing one incubation column per sampling event, which was then divided into three subsamples (replicate samples of mixed soil).

Anoxic incubations: L3 experiment

The same soil was used for the static incubations performed under an anoxic atmosphere (N2) in closed, gas-tight vessels, where denitrification products accumulated in the headspace. The incubation was conducted from 13.07.2020 to 15.07.2020. The soil was air dried and sieved at 4 mm mesh size. Afterwards, the soil was rewetted to achieve a water content equivalent to 70% water-filled pore space (WFPS) and fertilised with 100 mg N per kg soil, added as NaNO3 using natural Chile saltpetre (NaNO3, Chili Borium Plus, Prills-Natural origin, supplied by Yara, Dülmen, Germany, δ18O = 56‰, Δ17O = 21.8‰) to prepare 12 incubation soil samples of NA treatment and Na15NO3 to prepare 6 incubation soil samples of 15NO3 treatment. The soil was thoroughly mixed to obtain a homogenous distribution of water and fertilizer and an equivalent of 85 g of dry soil was repacked into each incubation jar at bulk densities of 1.3 g cm−3. The 0.5 dm3 Mason jars were used with airtight rubber seals and with two three-way valves installed in their cover to enable sampling and flushing. The jars were flushed with N2 at approximately 500 cm3 min−1 (STP: 273.15 K, 100 kPa) for 10 min to create anoxic conditions. In 6 NA vessels and three 15NO3 vessels 50 dm3 of headspace N2 was replaced with 50 dm3 of acetylene to inhibit N2O reduction to N2. Half of the incubation vessels of each treatment was incubated for 45 h and the other half was finished after 21 h for destructive sampling for soil mineral N analyses. The incubation temperature was 20 °C. Four gas samples were collected in 10 to 12 h-intervals by transferring 30 cm3 of headspace gases into two pre-evacuated 12 cm3 Exetainer vials (Labco Limited, Ceredigion, UK). The excess 3 cm3 of headspace gas in each vial ensured that no ambient air entered the vials. The removed sample volume was immediately replaced by pure N2 gas.

Soil analyses

All soil samples were homogenized. Soil water content was determined by weight loss after 24 h drying at 110 °C. Soil pH was determined in 0.01 mol CaCl2 solution (ratio 1:5). Nitrate and ammonium concentrations were determined by extraction in 2 M KCl in 1:4 ratio by 1 h shaking. Nitrite concentration was determined in alkaline extraction solution of 2 M KCl with addition of 2 M KOH (25 mL per L) in 1:1 ratio for 1 min of intensive shaking36. The amount of added KOH was adjusted to keep the alkaline conditions in extracts (pH over 8). After shaking, the samples were centrifuged for 5 min and filtered. The extracts for NO2 measurements were stored at − 4 °C and analyzed within 5 days. NO3, NH4+ and NO2 concentrations were determined colorimetrically with an automated analyser (Skalar Analytical B.V., Breda, the Netherlands).

To determine isotopic signatures of mineral nitrogen in NA treatments, microbial analytical methods were applied. For nitrate, the bacterial denitrification method with Pseudomonas aureofaciens was applied37,38. For nitrite, the bacterial denitrification method for selective nitrite reduction with Stenotrophomonas nitritireducens was applied6, also for 15N -enriched samples from 15N treatments. For ammonium, a chemical conversion to nitrite with hypobromite oxidation39 followed by bacterial conversion of nitrite after pH adjustment was applied40. δ15N of the organic N was analysed in the flushed and dried soil sample after mineral N extractions by EA combustion coupled to Delta Plus mass spectrometer (Thermo Finnigan, Bremen, Germany).

In 15N treatments, 15N abundances of NO3 (aNO3−) and NH4+ (aNH4+) were measured as described in Eschenbach, et al.41. NO3 was reduced to NO by Vanadium-III chloride (VCl3) and NH4+ was oxidized to N2 by hypobromite (NaOBr). NO and N2 were used as measurement gas. Measurements were performed on isotope ratio mass spectrometer (Delta Plus, Thermo Finnigan, Bremen, Germany).

Soil water was extracted with the method described by Königer, et al.42 and the δ18O of water samples (with respect to VSMOW) was measured using cavity ringdown spectrometer Picarro L1115-i (Picarro Inc., Santa Clara, USA). The measurement repeatability (1σ) of the internal standards (three calibrated waters with known δ18O: − 19.67‰, − 8.60‰, + 1.37‰) was below 0.1‰. The overall error associated with the soil water extraction method determined as standard deviation (1σ) of the 5 samples replicates was below 0.5‰.

All isotopic values are expressed as ‰ deviation from the 15N/14N and 18O/16O ratios of the reference materials (i.e. atmospheric N2 and Vienna Standard Mean Ocean Water (VSMOW), respectively).

Gas analyses

The samples for gas concentration analyses were collected in Exetainer vials (Labco Limited, Ceredigion, UK) and were analysed using an Agilent 7890A gas chromatograph (GC) (Agilent Technologies, Santa Clara, CA, USA) equipped with an electron capture detector (ECD). Measurement repeatability as given by the relative standard deviation (1σ) of four standard gas mixtures was typically 1.5%.

The gas samples collected from 15N treatments were analyzed for a15NN2O (15N abundance in the emitted N2O), aP_N2O (15N abundance in the 15N-pool derived N2O) and fP_N2O (15N-pool derived fraction of N2O)15 with a modified GasBench II preparation system coupled to MAT 253 isotope ratio mass spectrometer (Thermo Scientific, Bremen, Germany) according to Lewicka-Szczebak et al.43. In this set-up, N2O is converted to N2 during in-line reduction, and stable isotope ratios 29R (29N2/28N2) and 30R (30N2/29N2), of N2 are determined.

The gas samples of the NA treatment were analysed for N2O isotopocules (δ15NN2O, δ18ON2O, δ15NSPN2O) using a Delta V isotope ratio mass spectrometer (Thermo Scientific, Bremen, Germany), coupled to an automatic preparation system with Precon + Trace GC Isolink (Thermo Scientific, Bremen, Germany), where N2O was pre-concentrated, separated and purified, and m/z 44, 45, and 46 of the intact N2O+ ions as well as m/z 30 and 31 of NO+ fragment ions were determined. The results were evaluated accordingly44,45,46 which allows the determination of average δ15N, δ15Nα (δ15N of the central N position of the N2O molecule), and δ18O. δ15Nβ (δ15N of the peripheral N position of the N2O molecule) was calculated as δ15N = ( δ15Nα + δ15Nβ)/2 and 15N site preference (δ15NSP) as δ15NSP = δ15Nαδ15Nβ.

Determination of Δ17O excess in N2O and NO3 and estimation of O-atoms exchange (x)

N2O samples collected in the L3 NA treatment and N2O produced from soil NO3 by the bacterial denitrifier method were analysed for Δ17O after microwave equilibration in a sapphire tube and separation of N2 and O2 on a mole sieve column47. The 17O excess, Δ17O, is defined as48:

$$ \Delta^{17} {\text{O}} = \frac{{{1} + \delta^{{{17}}} {\text{O}}}}{{{(1} + \delta^{{{18}}} {\text{O)}}^{{{0}{\text{.5279}}}} }} - 1 $$
(4)

The measurement repeatability (1σ) of the international standards (USGS34, USGS35) was typically 0.5‰ for Δ17O.

The extent of isotope exchange (x) was determined based on the comparison of Δ17O in soil nitrate and produced N2O. It requires the application of nitrate characterised by high Δ17O. Therefore, for this determination, soils in L3 were amended with natural NaNO3 Chile saltpetre showing high Δ17O (of 21.8‰) and the Δ17O of the N2O product was measured. Δ17O of soil water was assumed to be 0‰.

The magnitude of oxygen isotope exchange (x) was calculated as:

$$ x = 1 - \frac{{\Delta^{17} {\text{O}}({\text{N}}_{{2}} {\text{O}})}}{{\Delta^{17} {\text{O}}({\text{NO}}_{3}^{-} )}} $$
(5)

The accuracy of x determination was better than 1%.

Application of the Keeling plot

The original idea for Keeling plot application applies for mixing of the background low level (atmospheric CO2) and one dominant source responsible for the significant increase of the CO2 concentration16. In such a case, plotting the δ values against the reciprocal CO2 concentration reveals the isotopic signature of the dominant as intercept of the linear fit16. Afterwards, the application of Keeling approach to isotopic studies has expanded to the other environments and substances, including nitrates source identification17,18,49. In these studies the requirement of only two sources is not necessarily fulfilled, but the occurrence of a clear linear relation between isotopic signature and reciprocal concentration of the studied substance indicates that there is a dominant source which can be isotopically characterised49. This is clearly the case for our nitrite samples, where we find a very significant linear relation (Fig. 1A). Nitrite contents in soils are typically very low and only rarely accumulate, mostly as a result of intensified nitrification or denitrification processes11,12,49,50,51. Hence, with Keeling plot we can isotopically identify the dominant NO2 source and identify the pathway responsible for this accumulation.

Isotope fractionation factors for the nitrite model

The isotope fractionation factors are always expressed as:

$$ \varepsilon_{{{\text{product}}/{\text{substrate}}}} = \, \delta_{{{\text{product}}}} - \, \delta_{{{\text{substrate}}}} $$

Hence, negative ε values inform about normal isotope effect resulting in product depletion in heavy isotopes.

Nitrite sources

NAR is associated with quite high isotopic fractionation of N and O, resulting in significant depletion in 15N and 18O in the product NO2. The nitrate reductase enzymatic experiments showed a mean 15εNAR of − 26.6 ± 0.2‰, similar to 18εDEN with a mean of − 24.9 ± 0.3‰25. In pure culture bacterial studies much larger variations of 15εDEN were observed, i.e. ranging from − 30.5 to − 5.4‰22,24,26 and it has been suggested that the range from − 15 to − 10‰ is most representative for typical cellular nitrate reduction rates for bacterial strains27. The strongest fractionation was found for pure culture fungal studies with a mean 15εNAR of − 37.8 ± 6.6‰20. Similar values were found for 18εDEN in pure culture studies: ranging between − 30 and − 25‰ for bacterial denitrification23 and between − 30 and − 10‰ for fungal denitrification20 . In the sediment denitrification experiments 15εDEN ranged from − 24.4 to − 18.9‰ and 18εDEN from − 21.9 to − 15.8‰8,29. A slightly lower 15εDEN of − 29.4 ± 2.4‰ was determined for soil studies28.

Nitrite produced from AOX is depleted in 15N compared to its ammonium substrate. Bacterial ammonia oxidation show a mean 15εAOX of − 25.8 ± 9.8‰52, similar to archaeal ammonia oxidation with a mean 15εAOX of − 22 ± 5‰53. δ18ONO2- from AOX depends on δ18OO2 (+ 23.5‰), δ18OH2O (− 5‰) and δ18ON2Oeq (8.6‰) according to the equation7,54:

$${\delta }^{18}{O}_{AOX}=0.5*\left({\delta }^{18}{O}_{O2}+{\delta }^{18}{O}_{H2O}+20\right)*0.92+\left({\delta }^{18}{O}_{H2O}+{\delta }^{18}{O}_{NO2-eq}\right)*0.08$$
(6)

Nitrite produced from ORG show much lower 15N enrichment with a mean 15εORG of about − 2‰ as measured for marine sediments fractionation55. δ18ONO2− from ORG was assumed to be the same as for AOX according to Eq. (3) (+ 18.4‰).

Nitrite sinks

Two major nitrite sinks—reduction and oxidation—show opposite isotopic fractionation. Nitrite reduction is associated with normal isotope effect resulting in enrichment in 15N and 18O of the nitrite pool, whereas nitrite oxidation is characterised by inverse isotope effect, where heavy isotopes are preferentially transferred to the oxidised product leaving nitrite pool depleted in 15N and 18O7. For NIR different fractionation may be associated with various nitrite reductases involved, showing a 15εNIR of − 22 ± 2‰ and an 18εNIR of − 2 ± 2‰ for Cu-NIR and − 8 ± 2‰ and − 6 ± 2‰ respectively for Fe-NIR56. In batch experiments with environmental bacterial communities a 15εDNIR ranging from − 15 to − 10‰ was observed when nitrite was investigated as an intermediate product but much lower when nitrite was a substrate29. Here we probably also observe this for L3—where nitrite is accumulating we get the best fit with the measured values when no fractionation associated with NIR is assumed (Table 2).

For nitrite oxidation the inverse isotope effects with a 15εNOX of + 12.857 and an 18εNOX of + 5‰30 were found.

Nitrite equilibration with water

The oxygen isotope signature of NO2 is additionally modified by the abiotic equilibrium exchange with ambient water23. The magnitude of this exchange is governed by the equilibrium isotope effect between NO2 and water (εeq) which is a function of temperature7,23 and the extend of O atoms exchange. εeq for the incubation temperature of 20 °C equals 13.63, δ18OH2O is − 5‰, consequently, the δ18O of nitrite in complete equilibrium with water is 8.6‰. The extend of O atoms exchange was determined with the 17O approach21 for N2O originating for denitrification processes in anoxic experiment L3 and equalled 0.25.