Abstract
NH4+ inhibition kinetics for CH4 oxidation were examined at near-atmospheric CH4 concentrations in three upland forest soils. Whether NH4+-independent salt effects could be neutralized by adding nonammoniacal salts to control samples in lieu of deionized water was also investigated. Because the levels of exchangeable endogenous NH4+ were very low in the three soils, desorption of endogenous NH4+ was not a significant factor in this study. The Km(app) values for water-treated controls were 9.8, 22, and 57 nM for temperate pine, temperate hardwood, and birch taiga soils, respectively. At CH4 concentrations of ≤15 μl liter−1, oxidation followed first-order kinetics in the fine-textured taiga soil, whereas the coarse-textured temperate soils exhibited Michaelis-Menten kinetics. Compared to water controls, the Km(app) values in the temperate soils increased in the presence of NH4+ salts, whereas the Vmax(app) values decreased substantially, indicating that there was a mixture of competitive and noncompetitive inhibition mechanisms for whole NH4+ salts. Compared to the corresponding K+ salt controls, the Km(app) values for NH4+ salts increased substantially, whereas the Vmax(app) values remained virtually unchanged, indicating that NH4+ acted by competitive inhibition. Nonammoniacal salts caused inhibition to increase with increasing CH4 concentrations in all three soils. In the birch taiga soil, this trend occurred with both NH4+ and K+ salts, and the slope of the increase was not affected by the addition of NH4+. Hence, the increase in inhibition resulted from an NH4+-independent mechanism. These results show that NH4+ inhibition of atmospheric CH4 oxidation resulted from enzymatic substrate competition and that additional inhibition that was not competitive resulted from a general salt effect that was independent of NH4+.
Atmospheric CH4 contributes substantially to the greenhouse effect, and the concentration of atmospheric CH4 has increased dramatically in the past century because of human activity associated with agriculture, land use changes, and industry (34, 35). Bacterial oxidation of atmospheric CH4 in well-drained soils is an important regulator of atmospheric CH4 concentration, yet the organisms responsible remain unidentified and the physiology of the process is poorly understood (9, 35, 36). Although soil CH4 consumption is inhibited by a wide variety of anthropogenic disturbances, such as agriculture, N deposition, and forestry (12, 17, 22, 23, 32, 43, 44), predictable inhibition patterns have failed to emerge, which has made it difficult to predict the effects of disturbance on soil CH4 flux in various ecosystems. The most commonly reported disturbance effect is that of NH4+ fertilizers, which can suppress soil CH4 consumption by up to 70% (1, 8, 10, 17, 22, 32, 33, 37, 38, 43). In the field, inhibition may occur immediately following fertilization, may be delayed for months to years, or may never occur despite years of chronic fertilization (9, 17). This variety of responses may stem at least in part from the distribution of physiologically diverse methane oxidizer populations across sites (17, 18, 20).
Of the various NH4+ inhibition patterns, immediate inhibition is the best documented. As in field studies, however, physiological laboratory studies have produced variable results, suggesting that there may be multiple inhibition mechanisms (15, 17, 26–28, 36, 39). Physicochemical similarities between CH4 and NH3 may permit these two compounds to compete for enzyme active sites so that fortuitous NH3 oxidation competitively inhibits CH4 oxidation (38). Although this mechanism has been demonstrated to occur in pure cultures of methanotrophic bacteria (6) and in a CH4-producing agricultural soil (15), it has not been demonstrated to occur in well-drained, nonagricultural mineral soils, which comprise the dominant terrestrial sink for atmospheric CH4 (14, 38, 45), nor has it been demonstrated to occur at near-atmospheric CH4 concentrations. In many cases, the kinetics of immediate NH4+ inhibition in soil cannot be reconciled easily with substrate competition (15, 16, 26–28, 39). An alternative mechanism has been proposed, whereby the toxicity of NO2− or NH2OH produced by fortuitous NH4+ oxidation suppresses methanotrophic activity (26, 27, 39). Hence, multiple inhibition mechanisms may be involved, and these mechanisms may vary with the physiology of different CH4 oxidizer populations (17).
Two physiologically distinct communities of CH4 oxidizers apparently exist in soil. One group, generally associated with atmospheric CH4 consumption, exhibits an extremely high affinity for CH4. Representatives of this group have yet to be cultivated or otherwise identified (9). The second group exhibits a much lower affinity for CH4 and is generally associated with common methanotrophs, such as those that have been studied in pure culture for many years (2, 9). In upland mineral soils, only high-affinity activity is usually detectable without artificial enrichment with high CH4 concentrations in the laboratory. However, the only prior study in which kinetic constants for NH4+ inhibition of soil CH4 oxidation were reported was conducted in a periodically moist, organic matter-rich agricultural soil with demonstrable methanogenesis (15, 16). Such a soil potentially harbors a rich community of CH4 oxidizers representing a continuum from low-affinity organisms to high-affinity organisms. Although this important investigation demonstrated that NH4+ inhibits CH4 oxidation via enzymatic substrate competition in an agricultural humisol, it is unclear to what extent its results apply to well-drained mineral soils lacking endogenous CH4 sources. Physiological studies of soil CH4 oxidation typically derive kinetic constants from oxidation rates at CH4 concentrations ranging from atmospheric levels (∼1.7 μl liter−1) to ≫Km for high-affinity CH4 oxidizers. Even in soil in which only high-affinity organisms are active, the CH4-oxidizing enzyme(s) could respond differently to NH4+ at high CH4 concentrations than at near-atmospheric concentrations (15, 39). Thus, to study NH4+ inhibition of high-affinity CH4 oxidizers per se, it would be preferable to examine inhibition kinetics at near-atmospheric CH4 concentrations in a soil with no apparent endogenous CH4 source.
A common shortcoming of NH4+ inhibition studies, regardless of the organisms involved, has been a lack of attention to nonammoniacal salt effects despite numerous reports of substantial inhibition by such salts (1, 10, 15, 17, 24). King and Schnell (28) examined the effects of several Cl− and SO42− salts and concluded that nonammoniacal salts indirectly inhibit CH4 oxidation by desorbing endogenous NH4+ from cation exchange sites in the soil, which then directly inhibit CH4 oxidation. Many N-limited soils, however, have extremely low concentrations of exchangeable NH4+, yet are substantially inhibited by nonammoniacal salts (17), suggesting that these salts have NH4+-independent effects on atmospheric CH4 oxidizers. Additional mechanisms may alter inhibition kinetics, thus hindering the diagnosis of NH4+-specific inhibition.
With the limitations described above in mind, we used a simple steady-state kinetics approach to assess the mechanism of NH4+ inhibition of CH4 oxidation at near-atmospheric concentrations (1.8 to 15 μl liter−1) in three well-drained, N-limited forest soils that lack known endogenous CH4 sources. In addition, we examined the effects of nonammoniacal salts in parallel samples to judge the utility of these salts as experimental controls for neutralizing NH4+-independent salt effects.
MATERIALS AND METHODS
Field sites.
We studied soils from two temperate forests and one taiga forest, the major characteristics of which are listed in Table 1. The two temperate soils were from the Harvard Forest Long-Term Ecological Research site in western Massachusetts (29), where fertilizer inhibition of atmospheric CH4 consumption was first observed (43). The sites and their biogeochemical cycles have been described in detail previously (7, 8, 30). The taiga site is approximately 120 years postburn, and the understory is dominated by Rosa acicularis and Equisetum spp. The mineral soil consists of a uniform layer of silty glacial loess. The pedology, ecology, and biogeochemistry of this site are similar to the pedology, ecology, and biogeochemistry of nearby sites that have been described previously (17). All three sites are well drained and have never been observed to produce CH4 (7, 8, 19).
TABLE 1.
Sampling sites and characteristics of the soils examined in this study
Ecosystem type (dominant plant) | Location | Soil texture | Concn of extractable NH4+ (mg of N kg of dry soil−1)a | Water-holding capacity (g of H2O g of dry soil−1) |
---|---|---|---|---|
Birch taiga (Betula papyrifera) | University of Alaska, Fairbanks | Fine silt | 0.11 (0.014)b | 0.62 |
Temperate hardwood (Quercus velutina) | Harvard Forest, Petersham, Mass. | Sandy loam | 6.98 (2.91)c | 0.84 |
Temperate pine (Pinus resinosa) | Harvard Forest, Petersham, Mass. | Sandy loam | 7.83 (2.49)c | 0.73 |
concentration of NH4+ N in the upper 10 cm of the mineral horizon, which includes the zone of maximum CH4 oxidation. The values in parentheses are standard deviations.
Data from this study. The concentration of extractable NH4+ was determined as described previously (17).
Soil processing and bioassays.
All experiments were performed at the University of Alaska, Fairbanks. At each site, soil was collected in bulk from the upper 10 cm of the mineral soil, which included the zone of maximum CH4 oxidation, and stored in perforated plastic bags for transport to the laboratory. The soil was homogenized by sieving it through a 4-mm-mesh screen. The water-holding capacity of each soil type was determined as described previously (18), and the moisture was adjusted so that the final water content was 30 to 35% of the water-holding capacity (Table 1) after the final treatment with deionized water or salt solutions. This moisture level was determined previously to be optimal for atmospheric CH4 consumption in a wide variety of soils (18). Samples were treated with deionized water, K2SO4, (NH4)2SO4, Na2SO4 (taiga soil only), KCl, or NH4Cl. Each salt solution was added to a single bulk sample (0.1 ml g of dry soil−1), which was then mixed thoroughly and subdivided into individual samples. The salts were added at a rate of 5.6 μmol of cations per g of dry soil, so that all of the salts were equinormal with respect to cations. The resulting NH4+ additions were equivalent to 75 mg of N per kg of dry soil, which matched the treatments used in previous experiments (17). This amount of NH4+ was intended to overwhelm the endogenous soil N (Table 1) yet remain within the range of soil NH4+ concentrations reported for forest soils with various land use histories (13, 31, 42).
For each treatment, 12 subsamples (10 g of dry soil) were placed in 70-ml serum vials sealed with butyl rubber septa and allowed to equilibrate overnight. The following morning the vials were equilibrated with laboratory air (∼1.8 μl of CH4 liter−1) and sealed, and their headspace CH4 concentrations were adjusted by injecting appropriate volumes of 1% CH4 premixed with air (Scott Specialty Gases, Plumsteadville, Pa.); the headspace CH4 concentrations tested were approximately 1.8 (no CH4 added), 5, 10, and 15 μl liter−1. Three replicates for each treatment at each CH4 concentration were prepared. For the temperate soils, a single 2-h CH4 oxidation assay was carried out with the headspace CH4 concentration measured at the beginning and the end of the assay. The resulting consumption rate (d[CH4]/dt) was paired with the corresponding midpoint CH4 concentration in order to obtain a plot of oxidation rate versus CH4 concentration (Fig. 1b and c). For the birch taiga soil, a modified procedure was used because oxidation was 2 orders of magnitude slower in this soil than in the temperate soils (Table 2). On the first day of the experiment, a 3.3-h assay was carried out with the headspace CH4 concentration measured at the beginning and the end of the assay. The samples were kept sealed and were allowed to consume CH4 overnight, and the 3.3-h assay was repeated on the second day and again on the third day. Identical assays were then repeated every 48 h until either a threshold concentration was established or through the ninth day, whichever occurred first. As with the temperate soils, the rate (d[CH4]/dt) from each 3.3-h assay was plotted against the corresponding midpoint CH4 concentration in order to obtain a plot of oxidation rate versus CH4 concentration (Fig. 1a). CH4 was analyzed by gas chromatography as described previously (5, 17, 18).
FIG. 1.
CH4 oxidation kinetics in three upland forest soils, a birch taiga soil (a), a temperate hardwood soil (b), and a temperate pine soil (c). The data for the birch taiga soil are shown with linear regression lines, whereas the data for the temperate soils are shown with nonlinear regression curves fit to the Michaelis-Menten equation. Each point represents a single rate measurement.
TABLE 2.
CH4 oxidation kinetics in three upland forest soils
Soil | Treatment | Km(app) (μl liter−1)a | Vmax(app) (pmol g−1 h−1) | First-order rate constant (h−1)b
|
% Inhibition compared toc:
|
Implied inhibition mechanism | Curve fit (R2)d | ||
---|---|---|---|---|---|---|---|---|---|
True | Pseudo | H2O | K+ | ||||||
Birch taiga | H2O | 0.00711 | 0.994 | ||||||
K2SO4 | 0.00653 | 8.2 | ? | 0.996 | |||||
(NH4)2SO4 | 0.00526 | 26* | 19* | ? | 0.992 | ||||
KCl | 0.00078 | 89* | ? | 0.772 | |||||
NH4Cl | 0.00072 | 90* | 7.7 | ? | 0.771 | ||||
Temperate hardwood | H2O | 15.2 (1.3) | 5,807 (329) | 0.705 | 0.998 | ||||
K2SO4 | 9.8 (0.4) | 2,662 (584) | 0.510 | 28* | Uncompetitive | 0.999 | |||
(NH4)2SO4 | 22.4 (1.8) | 2,870 (157) | 0.235 | 67* | 54* | Mixed competitive | 0.998 | ||
KCl | 7.2 (0.6) | 786 (30) | 0.200 | 72* | Uncompetitive | 0.993 | |||
NH4Cl | 25.6 (4.9) | 1,174 (155) | 0.085 | 88* | 58* | Mixed competitive | 0.991 | ||
Temperate pine | H2O | 6.7 (0.9) | 2,594 (160) | 0.715 | 0.984 | ||||
K2SO4 | 6.1 (0.9) | 1,591 (98) | 0.485 | 32* | Noncompetitive | 0.978 | |||
(NH4)2SO4 | 9.8 (1.4) | 1,699 (123) | 0.320 | 55* | 34* | Mixed competitive | 0.986 | ||
KCl | 8.8 (4.0) | 573 (123) | 0.120 | 83* | Noncompetitive | 0.855 | |||
NH4Cl | 10.9 (7.2) | 515 (176) | 0.085 | 88 | 29 | Mixed competitive | 0.769 |
The values in parentheses are standard errors.
True first-order rate constants were calculated by linear regression of CH4 oxidation rates against midpoint CH4 concentrations. Pseudo-first-order rate constants were calculated by determining Vmax/Km after Km values were converted to picomoles of CH4 per bottle.
An asterisk indicates that the treatment value was statistically different from the corresponding control value (P ≤ 0.05).
The regression coefficients are linear for the birch taiga soil and nonlinear (Michaelis-Menten curve fit) for the two temperate soils.
Because Cl− inhibited CH4 oxidation much more than did SO42−, the effects of Cl− and SO42− salts on general microbial respiration in the birch taiga soil were examined. The amount of CO2 that accumulated was determined by measuring headspace CO2 concentrations, which never exceeded 2%, at the beginning and end of a 1-week incubation period. The amount of CO2 that accumulated in each salt treatment was compared to the amount of CO2 in water-treated controls. CO2 was analyzed by gas chromatography as described previously (5, 18).
Statistical analyses and calculations.
The effects of the salt treatments and CH4 concentration on oxidation rates in each soil were analyzed by analysis of covariance by using treatment as the independent factor and the initial CH4 concentration as a covariate; Bonferroni contrasts were used in multiple comparisons. Because the incubation times were the same for all treatments in a given soil, the treatments with higher oxidation rates consumed more substrate than the treatments with lower oxidation rates. For regression analyses, therefore, the oxidation rate from an individual assay was paired with the corresponding midpoint CH4 concentration (Fig. 1) rather than the initial concentration. This standard technique normalizes consumption rates for unequal substrate concentrations among treatments and also minimizes the deviation from standard Michaelis kinetics that can result from substrate depletion (41). First-order kinetics were modeled by linear regression, and the rate constants were estimated from the slope of the regression line, with both variables expressed in picomoles. Michaelis constants were obtained from a least-squares nonlinear regression fit of the data to the Michaelis-Menten equation. For treatments exhibiting Michaelis kinetics, pseudo-first-order rate constants were calculated as Vmax/Km, with both constants expressed in picomoles. Relative inhibition was calculated for each treatment as follows: relative inhibition = (1 − k2/k1) × 100, where k1 is the first-order (or pseudo-first-order) rate constant for the control sample and k2 is the first-order (or pseudo-first-order) rate constant for the treated sample.
Examining the relationship between relative inhibition and CH4 concentration requires calculating inhibition ratios for two treatments at specific CH4 concentrations. In doing this, care must be taken not to compare rates derived from substantially different midpoint CH4 concentrations, even when the initial concentration is the same for all treatments. This problem arises when a faster sample consumes substantially more substrate than a slower sample, resulting in a disparity between midpoint CH4 concentrations in the two assays. For this reason, the method used to calculate relative inhibition at specific CH4 concentrations varied according to the relative rates among treatments and the type of kinetics involved for each soil. In the birch taiga soil, which displayed first-order kinetics, inhibition ratios were calculated directly from the rates measured in the experiments on the first day of incubation. Because the oxidation rates were very low in this soil, the differences in midpoint CH4 concentrations among the treatments were trivial. The inhibition by each salt compared to the deionized water control was calculated for each of the initial CH4 concentrations (∼1.8, 5, 10, and 15 μl liter−1) and plotted against the midpoint CH4 concentration occurring in the control (Fig. 2a). In the temperate soils, which displayed Michaelis kinetics, the rates were high, and the midpoint CH4 concentrations varied among the treatments. Hence, estimated inhibition ratios were calculated by entering four different CH4 concentrations (1.8, 5, 10, and 15 μl liter−1) into the regression equation obtained for each treatment. The calculated oxidation rates at each concentration were then used to calculate inhibition ratios for each salt compared to the water control. Each ratio was then plotted against the CH4 concentration from which it was derived (Fig. 2b and c). The slope of the relationship between relative inhibition and CH4 concentration was estimated by linear regression.
FIG. 2.
Effect of CH4 concentration on inhibition of CH4 oxidation. General salt inhibition compared to water controls in the birch taiga soil (a) and in the temperate hardwood soil (b). (c) Specific NH4+ inhibition compared to K+ controls in the temperate hardwood soil.
RESULTS
Birch taiga soil.
In the birch taiga soil, the CH4 oxidation kinetics at concentrations of ≤15 μl liter−1 were approximately first order (R2 > 0.99 except for Cl− salts) for all treatments, but the rate constants varied among treatments (Fig. 1a; Table 2). Analyses at higher CH4 concentrations (10 to 800 μl liter−1) (data not shown) yielded a Km(app) for oxidation in this soil of 39 μl liter−1 (57 nM in solution), which is typical for upland soils (2, 3, 36, 46, 47). Neither K2SO4 nor Na2SO4 significantly inhibited CH4 oxidation compared to deionized water (P = 0.78), and the curves for K2SO4 and Na2SO4 were indistinguishable (P = 0.91) (Na2SO4 data not shown). Specific NH4+ inhibition, calculated using K2SO4 as the control, was relatively weak (19%) but was statistically significant (P = 0.04). Both KCl and NH4Cl inhibited CH4 oxidation severely (∼90%; slopes were significantly different from zero at P < 0.01) (Fig. 1a). Unlike the comparison of K2SO4 and (NH4)2SO4, the effects of KCl and NH4Cl were indistinguishable (P = 0.84) (Fig. 1a). All four salts caused relative inhibition to increase as CH4 concentration increased (Fig. 2a). The slopes of the increases were similar for all salts regardless of which cation was added and regardless of the final soil NH4+ concentration. All salts inhibited total microbial respiration, but like CH4 oxidation, CO2 production was more sensitive to Cl− salts than to SO42− salts; the relative inhibition was ∼18% for both K2SO4 and (NH4)2SO4, whereas it was 22 to 25% for KCl and NH4Cl.
Temperate forest soils.
The relative inhibition patterns for the various salts in the temperate hardwood and pine forest soils were similar to the patterns in the birch taiga soil, except that CH4 oxidation conformed well to Michaelis-Menten kinetics (R2 > 0.98 in most cases) (Fig. 1b and c; Table 2). With minor differences in magnitude, the pine soil exhibited the same patterns as the hardwood soil. As in the birch taiga soil, the Cl− salts were the most inhibitory salts, followed by (NH4)2SO4 and then K2SO4. K2SO4 inhibition and specific NH4+ inhibition (relative to K+) were stronger in the temperate soils than in the taiga soil; the levels of specific NH4+ inhibition in the temperate hardwood and pine soils were 54 and 34%, respectively (Table 2). As in the birch taiga soil, inhibition of CH4 oxidation increased with the CH4 concentration when K+ salts were added. Unlike the taiga soil, however, inhibition in the temperate soils decreased as CH4 concentration increased when NH4+ salts were added (Fig. 2b) (pine forest results not shown). When specific NH4+ inhibition was calculated using K+ salts as controls, inhibition decreased sharply from ∼50% in the presence of 1.8 μl of CH4 liter−1 to ∼20 to 30% in the presence of 15 μl of CH4 liter−1 in the temperate hardwood soil (Fig. 2c).
The Michaelis parameters Km and Vmax exhibited similar patterns of responses to the various treatments in the two temperate soils (Table 2). In deionized water controls, the Km(app) values were 15 and 6.7 μl liter−1 (22 and 9.8 nM in solution) in the hardwood and pine soils, respectively. In the hardwood soil, the values of both Km and Vmax were about double the corresponding values in the pine soil, so the pseudo-first-order rate constants (Vmax/Km) were similar in the two soils (Table 2). K+ salts (irrespective of the anions) either decreased or had no effect on Km(app) values compared to water controls, whereas NH4+ salts always increased the Km(app). In contrast, Vmax(app) values decreased similarly in the presence of both K+ and NH4+ salts. Compared to K+ salts, however, NH4+ salts increased Km(app) but had no effect on Vmax(app).
DISCUSSION
Often, soil CH4 oxidation at near-atmospheric CH4 concentrations follows first-order reaction kinetics (3, 39, 46), as was the case in the birch taiga soil in this study (Fig. 1a). In fine-textured soils, first-order kinetics at lower CH4 concentrations may result from restricted gas diffusion from the atmosphere into the soil, a purely first-order process (14, 37, 45). The birch taiga soil studied is a fine silt soil and therefore strongly limits gas diffusion from the atmosphere to the CH4 oxidizers (14, 37), thus possibly increasing the Km(app) and creating a problem for studying low-concentration CH4 oxidation kinetics in this soil (Fig. 3). By contrast, the two temperate forest soils studied have a coarse sandy texture, which enhances CH4 diffusion, allowing uptake kinetics to reflect enzyme activity more closely and permitting standard kinetic analyses of NH4+ and salt inhibition of CH4 oxidation at near-atmospheric CH4 concentrations. The maximum CH4 concentration used in our experiments (15 μl liter−1) was similar to the Km(app) values in the temperate forest soils. The regression curves resulting from the kinetic analyses provided very good fits to the actual data (generally, R2 > 0.98) (Fig. 1; Table 2), indicating that the kinetic models which we used accurately described the process as measured in this study.
FIG. 3.
Potential effect of soil texture on CH4 oxidation kinetics in soil. Because gas transport by diffusion is purely first order, a fine-textured soil may exhibit first-order kinetics at lower CH4 concentrations, whereas a coarse-textured soil containing the same CH4-oxidizing enzyme or a similar enzyme may exhibit Michaelis-Menten kinetics at the same concentrations.
Although there have been numerous reports of steady-state kinetic constants for soil CH4 oxidation (2–4, 15, 36, 46, 47), only one previous study reported kinetic parameters for NH4+ inhibition (15). The soil studied previously was an agricultural humisol with an organic matter content of ∼70% and demonstrable methanogenic activity (15, 16). These conditions probably supported a much different CH4 oxidizer community than the community expected in upland mineral soils that lack an endogenous CH4 source. Indeed, the Km(app) values for our temperate forest soils (Table 2) were substantially lower than the Km(app) values reported in the previous study (15), suggesting that there were physiological differences in the CH4 oxidizer communities in the upland mineral soils and the agricultural humisol. In the present study we focused specifically on high-affinity CH4 oxidation in three non-CH4-producing upland soils from two North American biomes, subarctic taiga forest and northeastern temperate forest.
Salt effects.
Interpreting inhibition mechanisms based on kinetic parameters in a system that is as biologically and chemically complex as soil requires careful consideration of how ions added to the system may affect the process of interest, both directly and indirectly (15, 28). All of the ions used in this study potentially could affect CH4 oxidation in three basic ways. First, they could change the soil osmotic potential and impose water stress on the microbial community (18, 40); second, they could affect ion exchange, thereby altering NH4+ availability (28); and third, they could affect the CH4 oxidizers directly in a number of ways (11, 15, 21). Any of these factors could alter CH4 oxidation rates and kinetics and thus affect the interpretation of the specific NH4+ inhibition mechanism.
With the salt additions used in this study, the water potential in the birch taiga soil was approximately −0.2 MPa, which is the optimum water potential for atmospheric CH4 oxidation in a wide variety of upland soils (18, 40). Because the birch taiga soil had the lowest water-holding capacity of the soils used in this study (i.e., the lowest water/salt ratio [Table 1]), its osmotic potential should have been the most sensitive to the salt additions. Thus, salt-related inhibition of atmospheric CH4 oxidation in this study did not appear to be related to water stress.
It is unlikely that K+ salts indirectly produced the inhibition observed in this study by desorbing NH4+ from cation exchange sites, as proposed elsewhere (28). This mechanism requires that an untreated soil contain sufficient exchangeable NH4+ to account for the inhibition observed with nonammoniacal salts, yet the soils we studied had very low concentrations of exchangeable NH4+ relative to our NH4+ additions (Table 1). Cl− salts consistently inhibit soil CH4 consumption to a far greater extent than do SO42− salts (28; this study). King and Schnell (28) attributed this phenomenon to greater NH4+ adsorption to cation exchange sites in the presence of SO42− than in the presence of Cl−. In the present study, however, it was impossible for the KCl treatments to produce free NH4+ concentrations approaching those of the (NH4)2SO4 treatments, because we added 1 to 3 orders of magnitude more NH4+ than was potentially available in the untreated soils (Table 1). Even so, KCl inhibition was far greater than (NH4)2SO4 inhibition in all three soils (Fig. 1; Table 2). KCl and NH4Cl produced similar levels of inhibition in each of the soils, despite the fact that the NH4Cl treatments necessarily resulted in much higher free NH4+ concentrations. Similarly, the results of King and Schnell show that NaCl caused inhibition equal to or greater than the inhibition that equinormal NH4Cl caused in another temperate forest soil (28). Again, this result could not have been dependent on NH4+ concentrations. Hence, desorption of endogenous NH4+ cannot account for the extremely inhibitory effects of Cl− salts in a variety of soils, and it is clear that Cl− salts should be avoided in NH4+ inhibition studies, unless it can be demonstrated that Cl− is not toxic to CH4 oxidizers in a particular soil.
Unlike KCl, K2SO4 inhibited CH4 oxidation less than (NH4)2SO4 inhibited CH4 oxidation, raising the possibility that there is indirect inhibition by cation exchange when SO42− salts are used. Compared to water, K2SO4 inhibition was 32 to 58% of (NH4)2SO4 inhibition in the three soils (Table 2). Assuming that the desorption of endogenous soil NH4+ was 100%, which is unlikely, the K2SO4 treatments would have produced maximum NH4+ concentrations that were between ∼0.1 and 10% of the amount added in the (NH4)2SO4 treatment (Table 1). Hence, compared to the (NH4)2SO4 treatments, the ratios of relative inhibition to potential NH4+ concentration obtained with the K2SO4 treatments seem unlikely. More importantly, as discussed below, steady-state kinetic parameters indicate that K2SO4 and (NH4)2SO4 inhibited CH4 oxidation via different physiological mechanisms, which would not be the case if K+ acted indirectly via NH4+ desorption.
The ubiquity of NH4+-independent salt effects and the variety of salts that induce similar responses suggest that a fundamental physiological process is involved. Roslev et al. (36) found that high-affinity CH4 oxidizers in a temperate forest soil efficiently incorporated 14CH4-C into biomass. Adding NH4Cl to the soil not only decreased CH4 oxidation rates but also reduced the C assimilation efficiency and dramatically increased the proportion of 14C oxidized to CO2. It is impossible to know whether this response was to NH4+ or Cl− or both, as the experiments did not include parallel salt controls. However, because we found that Cl− overwhelmingly dominated NH4Cl inhibition in all of our soils, the response that Roslev et al. observed may also have been predominantly due to Cl−. Killham (25) reported that NaCl additions had the same effect on microbial assimilation and respiration of [14C]glucose in soil and found that an increase in the ratio of respired C to assimilated C was a sensitive index of physiological stress within a soil heterotroph community. Shifts in the ratio were attributed to an increase in the maintenance energy required for the cells to cope with the imposed stress. If active transport of ions out of the cell or some other energy-intensive coping strategy were required by energy-limited CH4 oxidizers exposed to a salt, then cellular reductant might be diverted to this process, making less reductant available for growth and potentially to the CH4-oxidizing enzyme, thus decreasing the CH4 oxidation rates. This scenario is plausible for an extremely energy-limited population and is reconcilable with the inhibition kinetics reported here, as diverting reductant away from the CH4-oxidizing enzymes should reduce the catalytic efficiency of the extant enzyme pool, thereby potentially altering Km(app) and Vmax(app) as described below. Gulledge et al. (17) observed an apparent growth response of the atmospheric CH4 oxidizer community in samples obtained from depths of 20 to 40 cm in another forest soil. The in situ CH4 concentrations at depths below 20 cm were chronically <0.5 μl liter−1. After 14 days of exposure to ambient atmospheric CH4 in the laboratory, the CH4 consumption rates in water-treated samples increased severalfold compared to the rates measured after only 5 days of exposure. In K2SO4-treated samples a less pronounced increase occurred, and in (NH4)2SO4-treated samples no increase occurred, suggesting that the effects of NH4+ and salt were synergistic. These results also are consistent with a cellular stress response by an energy-limited population and may illustrate why atmospheric CH4 oxidizers have a limited capacity to recover from soil fertilization (32, 33, 39).
If salts generally inhibit soil CH4 oxidation by an NH4+-independent mechanism, then it seems appropriate to quantify specific NH4+ inhibition based on a parallel salt control rather than a deionized water control. This approach has been challenged by the view that other cations may have unique inhibition mechanisms that make them ineffective as experimental controls (28). Although this hypothesis is plausible, no differential toxicity of potential control cations, such as Na+ and K+, has been reported for soil CH4 oxidation. King and Schnell (28) found that KCl inhibited CH4 uptake by pure cultures of Methylosinus trichosporium more than did NaCl, but they observed no difference in soil CH4 consumption in the presence of these two salts. Similarly, we observed no difference in the effects of K2SO4 and Na2SO4 in the birch taiga soil in the present study (the temperate soils were not tested with Na+). Moreover, it is equally plausible that similar cations, such as NH4+, K+, and Na+, exert equivalent nonspecific effects that, in conjunction with counteranion effects, account for the nonammoniacal inhibition observed with salts in general. Since K+ and Na+ salts inhibit soil CH4 oxidation to the same extent (28; this study), this hypothesis appears to be sound. Our view, therefore, is that parallel salt controls must be employed when NH4+ inhibition is examined, because there is no other way to account for the nonammoniacal effects that salts clearly have on soil CH4 consumption. In some cases, salt effects can be substantial compared to specific NH4+ inhibition and therefore probably interfere with kinetic analysis of the NH4+ inhibition mechanism. In the present study we used both deionized water and nonammoniacal salt controls in order to examine the relative efficacies of the two approaches for elucidating the mechanism of NH4+ inhibition.
Specific NH4+ inhibition.
Determining the physiological mechanism of specific, immediate NH4+ inhibition has proven to be difficult (6, 15, 17, 26–28). Dunfield and Knowles (15) demonstrated that NH4+ inhibited CH4 oxidation by enzymatic substrate competition in an agricultural humisol assayed at high CH4 concentrations. The kinetics varied between samples, however, indicating that an additional mechanism may have been involved. King and Schnell (27, 39) examined NH4Cl inhibition at low CH4 concentrations and found that relative inhibition increased with CH4 concentration. They concluded that this phenomenon resulted from the fortuitous oxidation of NH4+ to toxic NO2− or NH2OH, which in turn reduced the activity of the methanotroph population (39). They did not examine Michaelis constants or comparable kinetic parameters. We observed similar increases in inhibition with increasing CH4 concentrations in all three of the soils we examined. In the taiga soil, this phenomenon occurred with nonammoniacal salts as well as NH4+ salts, and the slope of the increase was not affected by the NH4+ concentration (Fig. 2a). In the temperate soils, K+ salts caused inhibition to increase, whereas NH4+ salts caused inhibition to decrease as the CH4 concentration increased (Fig. 2b and c). The same pattern occurred whether Cl− or SO42− salts were added, indicating that it was not specific to a particular counterion (Fig. 2a and b). These results indicate that the increase in inhibition did not result from NH4+ or its by-products. Thus, although NO2− undoubtedly inhibits atmospheric CH4 oxidation when it is added directly to soil (19, 26), an increase in NH4+ salt inhibition when the CH4 concentration increases more likely results from a general salt effect than from by-products of fortuitous NH4+ oxidation.
A net increase in inhibition with an increase in the CH4 concentration in response to NH4+ salts may actually indicate that specific NH4+ inhibition is weak or absent. For instance, in the birch taiga soil, in which NH4+ inhibition was relatively weak (Table 2), both NH4+ and K+ salts caused similar increases in inhibition as the CH4 concentration increased (Fig. 2a). However, in the temperate hardwood soil, in which NH4+ inhibition was relatively strong (Table 2), only K+ salts caused inhibition to increase, whereas NH4+ salts caused inhibition to decrease as the concentration of CH4 increased (Fig. 2b). Specific NH4+ inhibition, isolated by using K+ salts as controls, declined precipitously as the CH4 concentration increased (Fig. 2c). Hence, salts generally caused increases in inhibition, whereas NH4+ caused decreases in inhibition as the CH4 concentration increased, indicating that there are separate inhibition mechanisms for NH4+ specifically and salts generally. In our soils, the relative strengths of these two mechanisms were apparent from the slopes of the plots of (NH4)2SO4 inhibition (relative to deionized water) versus CH4 concentration; a positive slope indicated a stronger salt effect, whereas a negative slope indicated a stronger NH4+ effect.
The soils which we examined were relatively acidic (pH ∼3.5 to 4.5). Because NH3, rather than NH4+, is probably the competitive inhibitor of CH4 oxidation, salt effects may be more prevalent in acidic soils, whereas competitive inhibition may be relatively more important in neutral to alkaline soils, such as the agricultural humisol investigated by Dunfield and Knowles (15). Despite the intuitive appeal of this hypothesis, there is no obvious relationship between pH and the degree of NH4+ inhibition in soils with different pH values, suggesting that other cross-site variables are generally more important (17). Moreover, it is clear from the results obtained with the temperate hardwood soil, in which NH4+ accounted for 58% of the total inhibition, that NH4+ inhibition can be dominant in acidic soils. Perhaps the intracellular pH, which should be near neutral regardless of the soil pH, is the relevant control on NH3/NH4+ ratios at the enzyme level.
Compared to the Km in the water controls, the Km(app) either decreased (hardwood soil) or remained unchanged (pine soil) when K+ salts were added, but it always increased when NH4+ salts were added (Table 2). Again, this pattern supports the hypothesis that there are different inhibition mechanisms for NH4+ and salts in general and also eliminates the possibility that K+ salts acted indirectly by desorbing soil-bound NH4+ into solution, as concluded previously for another temperate forest soil (28). If K+ ions acted indirectly via NH4+, then K+ and NH4+ salts should have produced similar inhibition kinetics, yet they had different effects on Km(app) compared to deionized water. In contrast to Km(app), Vmax(app) decreased in response to both K+ and NH4+ salts (Table 2). Thus, the kinetic constants suggest that there is a partial mixed-type inhibition for NH4+ salts, with both a competitive component (increasing Km) and a noncompetitive or uncompetitive component (decreasing Vmax) that is independent of NH4+ (41) (Table 2). If general salt effects account for the additional inhibition, then using K2SO4 as a control rather than deionized water should isolate the specific NH4+ effect. Indeed, compared to K+ salts, NH4+ salts caused Km(app) to increase substantially, whereas they had no effect on Vmax(app) (Table 2). The Cl−-salt pair produced the same kinetic pattern as the SO42− pair despite the greater inhibition by Cl−. These consistent results strongly indicate that NH4+ inhibited atmospheric CH4 oxidation in the two temperate forest soils via simple enzyme substrate competition.
Summary and conclusions.
Our approach using K+ salts as controls, and the resulting interpretation, provided a plausible NH4+ inhibition mechanism that is consistent with the data presented here and can also account for the contrasting results of previous studies (15, 27, 39). Whereas Dunfield and Knowles (15) observed competitive inhibition kinetics in their agricultural humisol, Schnell and King observed increasing inhibition with increasing CH4 concentrations in a temperate forest soil, a result that, by itself, is inconsistent with competitive inhibition. Both phenomena occurred simultaneously in our temperate forest soils and could be explained by a mixed- type inhibition resulting from at least two independent mechanisms, enzymatic substrate competition by NH4+ and one or more noncompetitive or uncompetitive mechanisms common to salts in general. Although the inhibition mechanism in the birch taiga soil could not be determined directly because it displayed first-order kinetics (Fig. 1a), the relative inhibition pattern for the various treatments was consistent with the patterns obtained with the two temperate soils, so that all three soils may have shared the same mechanisms. Moreover, it is notable that despite very different soil characteristics, we found essentially the same inhibition mechanism that Dunfield and Knowles (15) found in an agricultural humisol. This convergence of physiological responses in ecologically diverse environments suggests that enzymatic substrate competition is an important NH4+ inhibition mechanism in a wide variety of soils.
Although the results readily explain immediate inhibition of soil CH4 oxidation, delayed inhibition, which has been observed in both field and laboratory studies (17), remains enigmatic. Delayed inhibition probably results from shrinkage of the CH4 oxidizer population over time rather than from decreases in the specific activities of individual CH4 oxidizers (17). NH4+ and salt effects may act synergistically to impose whole-cell stress that increases maintenance energy requirements, thereby diverting reductant from growth, even if sufficient reductant for the CH4-oxidizing enzyme remains available. This scenario might diminish a population’s ability to replace dying biomass, yet might not slow the oxidation of CH4 until the population begins to shrink, resulting in a delayed inhibition response (17). Hence, multiple physiological mechanisms may contribute synergistically to both immediate and delayed NH4+ fertilizer inhibition of atmospheric CH4 consumption in soil. Moreover, nonammoniacal salts in the environment, especially KCl and NaCl (both of which are used heavily in agriculture and industry), may be as problematic as NH4+ fertilizers for soil CH4 consumption.
ACKNOWLEDGMENTS
This work was supported by funds from the National Science Foundation through the Bonanza Creek Taiga Long-Term Ecological Research project. J.G. is currently a DOE-Energy Biosciences Postdoctoral Fellow of the Life Sciences Research Foundation.
We thank P. A. Steudler for access to his research plots in the Harvard Forest and K. Newkirk for assistance with soil sampling. In addition, we thank an anonymous reviewer for comments that improved the manuscript.
REFERENCES
- 1.Adamsen A P S, King G M. Methane consumption in temperate and subarctic forest soils: rates, vertical zonation, and responses to water and nitrogen. Appl Environ Microbiol. 1993;59:485–490. doi: 10.1128/aem.59.2.485-490.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Bender M, Conrad R. Kinetics of CH4 oxidation in oxic soils exposed to ambient air or high CH4 mixing ratios. FEMS Microbiol Ecol. 1992;101:261–270. [Google Scholar]
- 3.Bender M, Conrad R. Kinetics of methane oxidation in oxic soils. Chemosphere. 1993;26:687–696. [Google Scholar]
- 4.Benstead J, King G M. Response of methanotrophic activity in forest soil to methane availability. FEMS Microbiol Ecol. 1997;23:333–340. [Google Scholar]
- 5.Breitenbeck G A. Sampling the atmospheres of small vessels. Soil Sci Soc Am J. 1990;54:1794–1797. [Google Scholar]
- 6.Carlsen H N, Joergensen L, Degn H. Inhibition by ammonia of methane utilization in Methylococcus capsulatus (Bath) Appl Microbiol Biotechnol. 1991;35:124–127. [Google Scholar]
- 7.Castro M S, Mellilo J M, Steudler P A, Chapman J W. Soil moisture as a predictor of methane uptake by temperate forest soils. Can J For Res. 1994;24:1805–1810. [Google Scholar]
- 8.Castro M S, Steudler P A, Mellilo J M. Factors controlling atmospheric methane consumption by temperate forest soils. Global Biogeochem Cyc. 1995;9:1–10. [Google Scholar]
- 9.Conrad R. Soil microorganisms as controllers of atmospheric trace gases (H2, CO, CH4, OCS, N2O, and NO) Microbiol Rev. 1996;60:609–640. doi: 10.1128/mr.60.4.609-640.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Crill P M, Martikainen P J, Nykänen H, Silvola J. Temperature and N fertilization effects on methane oxidation in a drained peatland soil. Soil Biol Biochem. 1994;26:1331–1339. [Google Scholar]
- 11.Dinesh R, Ramanathan G, Singh H. Influence of chloride and sulphate ions on soil enzymes. J Agron Crop Sci. 1995;175:129–133. [Google Scholar]
- 12.Dobbie K E, Smith K A. Comparison of CH4 oxidation rates in woodland, arable and set aside soils. Soil Biol Biochem. 1996;28:1357–1365. [Google Scholar]
- 13.Donaldson J M, Henderson G S. Nitrification potential of secondary-succession upland oak forest. I. Mineralization and nitrification during laboratory incubations. Soil Sci Soc Am J. 1990;54:892–897. [Google Scholar]
- 14.Dörr H, Kattruff L, Levin I. Soil texture parameterization of the methane uptake in aerated soils. Chemosphere. 1993;26:697–713. [Google Scholar]
- 15.Dunfield P, Knowles R. Kinetics of inhibition of methane oxidation by nitrate, nitrite, and ammonium in a humisol. Appl Environ Microbiol. 1995;61:3129–3135. doi: 10.1128/aem.61.8.3129-3135.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Dunfield P F, Topp E, Archambault C, Knowles R. Effect of nitrogen fertilizers and moisture content on CH4 and N2O fluxes in a humisol: measurements in the field and intact soil cores. Biogeochemistry. 1995;29:199–222. [Google Scholar]
- 17.Gulledge J, Doyle A P, Schimel J P. Different NH4+-inhibition patterns of soil CH4 consumption: a result of distinct CH4 oxidizer populations across sites? Soil Biol Biochem. 1997;29:13–21. [Google Scholar]
- 18.Gulledge J, Schimel J P. Moisture control over atmospheric CH4 consumption and CO2 production in diverse Alaskan soils. Soil Biol Biochem. 1998;30:1127–1132. [Google Scholar]
- 19.Gulledge, J., and J. P. Schimel. Unpublished data.
- 20.Gulledge J, Steudler P A, Schimel J P. Effect of CH4-starvation on atmospheric CH4 oxidizers in taiga and temperate forest soils. Soil Biol Biochem. 1998;30:1463–1467. [Google Scholar]
- 21.Kamekura M, Kushner D J. Effect of chloride and glutamate ions on in vitro protein synthesis by the moderate halophile Vibrio costicola. J Bacteriol. 1984;160:385–390. doi: 10.1128/jb.160.1.385-390.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Keller M, Mitre M E, Stallard R F. Consumption of atmospheric methane in tropical soils of central Panama: effects of agricultural development. Global Biogeochem Cyc. 1990;4:21–28. [Google Scholar]
- 23.Keller M, Veldkamp E, Weltz A M, Reiners W A. Effect of pasture age on soil trace-gas emissions from a deforested area of Costa Rica. Nature. 1993;365:244–246. [Google Scholar]
- 24.Kightley D, Nedwell D B, Cooper M. Capacity for methane oxidation in landfill cover soils measured in laboratory-scale soil microcosms. Appl Environ Microbiol. 1995;61:592–601. doi: 10.1128/aem.61.2.592-601.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Killham K. A physiological determination of the impact of environmental stress on the activity of microbial biomass. Environ Pollut Ser A Ecol Biol. 1985;38:283–294. [Google Scholar]
- 26.King G M, Schnell S. Ammonium and nitrite inhibition of methane oxidation by Methylobacter albus BG8 and Methylosinus trichosporium OB3b at low methane concentrations. Appl Environ Microbiol. 1994;60:3508–3513. doi: 10.1128/aem.60.10.3508-3513.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.King G M, Schnell S. Effect of increasing atmospheric methane concentration on ammonium inhibition of soil methane consumption. Nature. 1994;370:282–284. [Google Scholar]
- 28.King G M, Schnell S. Effects of ammonium and non-ammonium salt additions on methane oxidation by Methylosinus trichosporium OB3b and Maine forest soils. Appl Environ Microbiol. 1998;64:253–257. doi: 10.1128/aem.64.1.253-257.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Long Term Ecological Research Program. 1998. http://lternet.edu.
- 29a.Magill, A. Personal communication.
- 30.Magill A H, Aber J D, Hendricks J J, Bowden R D, Melillo J M, Steudler P A. Biogeochemical response of forest ecosystems to simulated chronic nitrogen deposition. Ecol Appl. 1997;7:402–415. [Google Scholar]
- 31.Montangnini F, Buschbacher R. Nitrification rates in two undisturbed tropical rain forests and three slash-and-burn sites of the Venezuelan Amazon. Biotropica. 1989;21:9–14. [Google Scholar]
- 32.Mosier A, Schimel D, Valentine D, Bronson K, Parton W. Methane and nitrous oxide fluxes in native, fertilized and cultivated grasslands. Nature. 1991;350:330–332. [Google Scholar]
- 33.Nesbit S P, Breitenbeck G A. A laboratory study of factors influencing methane uptake by soils. Agric Ecosyst Environ. 1992;41:39–54. [Google Scholar]
- 34.Prather M, Derwent R, Ehhalt D, Fraser P, Sanhueza E, Zhou X. Other trace gases and atmospheric chemistry. In: Houghton J T, Meire Filho L G, Bruce J, Lee J, Callander B A, Haites E, Harris N, Maskell K, editors. Climate change 1994. Cambridge, United Kingdom: Cambridge University Press; 1995. pp. 77–126. [Google Scholar]
- 35.Reeburgh W S. “Soft spots” in the global methane budget. In: Lidstrom M E, Tabita F R, editors. Microbial growth on C1 compounds. Dordrecht, The Netherlands: Kluwer; 1996. pp. 334–342. [Google Scholar]
- 36.Roslev P, Iversen N, Henriksen K. Oxidation and assimilation of atmospheric methane by soil methane oxidizers. Appl Environ Microbiol. 1997;63:874–880. doi: 10.1128/aem.63.3.874-880.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Saari A, Martikainen P J, Ferm A, Ruusdanen J, De Boer W, Troelstra S R, Laanbroek H J. Methane oxidation in soil profiles of Dutch and Finnish coniferous forests with different soil texture and atmospheric nitrogen deposition. Soil Biol Biochem. 1997;29:1625–1632. [Google Scholar]
- 38.Schimel J P, Holland E A, Valentine D. Controls on methane flux from terrestrial ecosystems. In: Rolston D E, Harper L A, Mosier A R, Duxbury J M, editors. Agricultural ecosystem effects on trace gases and global climate change. Madison, Wis: American Society of Agronomy; 1993. pp. 167–182. [Google Scholar]
- 39.Schnell S, King G M. Mechanistic analysis of ammonium inhibition of atmospheric methane consumption in forest soils. Appl Environ Microbiol. 1994;60:3514–3521. doi: 10.1128/aem.60.10.3514-3521.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Schnell S, King G M. Responses of methanotrophic activity in soils and cultures to water stress. Appl Environ Microbiol. 1996;62:3203–3209. doi: 10.1128/aem.62.9.3203-3209.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Segel I H. Enzyme kinetics behavior and analysis of rapid equilibrium and steady-state enzyme systems. New York, N.Y: John Wiley & Sons; 1975. [Google Scholar]
- 42.Stams A J, Marnette E C L. Investigation of nitrification in forest soils with soil percolation columns. Plant Soil. 1990;125:135–141. [Google Scholar]
- 43.Steudler P A, Bowden R D, Melillo J M, Aber J D. Influence of nitrogen fertilization on methane uptake in temperate forest soils. Nature. 1989;341:314–316. [Google Scholar]
- 44.Steudler P A, Melillo J M, Bowden R D, Castro M S. The effects of natural and human disturbances on soil nitrogen dynamics and trace gas fluxes in a Puerto Rican wet forest. Biotropica. 1991;23:356–363. [Google Scholar]
- 45.Striegl R G. Diffusional limits to the consumption of atmospheric methane by soils. Chemosphere. 1993;26:715–720. [Google Scholar]
- 46.Whalen S C, Reeburgh W S. Moisture and temperature sensitivity of CH4 oxidation in boreal soils. Soil Biol Biochem. 1996;28:1271–1281. [Google Scholar]
- 47.Whalen S C, Reeburgh W S, Sandbeck K A. Rapid methane oxidation in a landfill cover soil. Appl Environ Microbiol. 1990;56:3405–3411. doi: 10.1128/aem.56.11.3405-3411.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]