gg £3 kg «if . a. u . . ‘ V . , s, , a»; «1% $1”: 3 I. u .u‘ .r 3 v. ‘ . 3.3: . . E 5“ . mm. $4 fi. w.“ m3 WWW. :4 . . .. a". "may” efiwhfimmwnz . :50 a 3.. WNW .# .Mw 33$ :1 .23.! .. .. an 2.2539 in ~ a» 9.! t; .1: ~ I . Ev a: .. 43%“. .s _ .wu .u. . “NI-3‘“ :5 . in,“ 534-! 1 .135- .75 2. Wsnfin .5 ‘ T , u‘ .. mm“? 1...: ,flhymu; . $ 4" u ‘ . , A » ,. Jar... . , .A d». ,. av ‘. ‘ é. . ”Emu ‘ . 53v .K d... . V ‘2 . . I u .Pé....£.:fl. . n .5. I ‘r i a TB . : pi” i I 535%? . 43...”: s 1 41:35 2cm This is to certify that the dissertation entitled THE RELATIONSHIP BETWEEN DENITRIFICATION AND NITROUS OXIDE FLUX FROM SOIL presented by Timothy Todd Bergsma has been accepted towards fulfillment of the requirements for Ph. D. degree in Ecology (DMM M ajorHesTsor Date 9— mm: MS U is an Affirmative Action/Equal Opportunity Institution 0- 12771 LIIRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE ”a 9.; 553 SEP 0 4 2007 11; ”9132832 11/00 c/CIRC/Dateotnpsfm,“ THE RELATIONSHIP BETWEEN DENITRIFICATION AND NITROUS OXIDE FLUX FROM SOIL By Timothy Todd Bergsma A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Crop and Soil Sciences 2000 ABSTRACT THE RELATIONSHIP BETWEEN DENITRIFICATION AND NITROUS OXIDE FLUX FROM SOIL By Timothy Todd Bergsma Denitrification in soil is the major source of atmospheric nitrous oxide (N20), a potent contributor to global warming and a regulator of stratospheric ozone. Flux of N20 to the atmosphere is poorly understood. There is a serious imbalance in the global N20 budget (missing sources), and N20 flux at the field scale is difficult to predict, even when the rate of denitriflcation has been adequately characterized. In this dissertation, I review the literature pertaining to the relationship between denitrification and N20 flux, emphasizing a major source of uncertainty: relative production of N20 and N2 (dinitrogen) during denitrification. I explore the ecological factors that influence the nitrous oxide mole fraction ( N201[ N20 + N2] ), an aspect of denitrification that expresses relative production of N20 and N2. Second, I develop theory that facilitates the evaluation of N20 and N2 flux using 15N-labeled compounds and mass spectrometry. My heuristic model of labeled N-gas flux from soil simplifies the process of drawing inferences from isotope data. Third, I report new procedures for measuring fluxes of labeled N20 and N2 for the same incubation using mass spectrometry. My procedures could lead to improved estimates of the nitrous oxide mole fraction. l illustrate these procedures with field data. Finally, I describe a laboratory experiment in which I combine traditional and isotope methods to test for effects of moisture history (antecedent soil moisture) and ecosystem management history on nitrous oxide mole fraction, while controlling soil type and moisture. I find that response of mole fraction to differences in short-term (48 h) soil moisture history is different for soils from ecosystems with different management histories. A cropped soil had a high (~0.9) mole fraction after rapid transition from air-dry to 85% water-filled pore space but a low (~0.3) mole fraction when 80% of added moisture was applied 48 h in advance of the incubation. However, soil from a successional system generated a nitrous oxide mole fraction of about 0.3 regardless of short term moisture history. Progress in understanding the relationship between denitrification and nitrous oxide flux from soil seems still to be methods-limited. For Qiaobing. W0 ai ni! ACKNOWLEDGMENTS I thank God for creating a world that is both interesting and (often!) intelligible. I thank Jesus for rescuing me from sin and for inspiring me, by his Spirit, to pursue a life of service and gratitude. I thank my wife Qiaobing (Crystal) who never complained when I came home from the lab later than I said I would (which was always). I thank her also for giving me a marriage and a family, and for picking up more than her share of household responsibility so that I could finish the degree program represented by this dissertation. I thank David for helping me with field work (we have the pictures!) and I thank Maarten for a thousand smiles. Special thanks are due to Yiyan Liang, my mother-in- law, who cared for Maarten while I was finishing my lab work and writing. I thank my parents for their encouragement and support. Deepest thanks are extended to my graduate committee. My advisor, Phil Robertson, artfully guided my degree program. He never pushed me for results, which was just the right strategy to get me to push myself. I am deeply grateful for all he has taught me about planning research and especially about communicating it in writing. Nathaniel Ostrom spent uncounted hours teaching me mass spectrometry and non-lethal use of a vacuum line. More importantly, Nathaniel taught me the value of goal-directed problem solving by modeling a deft economy of lab effort. Even without his training, this dissertation would have been impossible apart from Nathaniel’s spectrometry skills. Thanks also are due to honorary committee member Peggy Ostrom, whose distance from the project and research savvy resulted in many critical suggestions that cut through my technical or conceptual waywardness. Steve Hamilton was an inspirational role model as well as effective critic of my experimental designs. I thank him also for access to critical lab equipment and services. Eldor Paul offered the seasoned perspective of an entire research community. Some of the fundamental concepts of my research project grew out of his coursework assignments. Appreciation is due also to the constellation of KBS faculty for creating a nurturing environment for research. I thank Per Ambus for introducing me to field sampling and to data management. Kevin Kosola conspired with me to contrive all manner of beastly field equipment; I admire his inventiveness and thank him for encouragement. Heather Reynolds listened patiently to my ideas about spatial variability of soil resources, and provided excellent guidance. Claire McSwiney laughed and cried with me over the torments of understanding anything about nitrogenous trace gases - a kindred spirit. Jane Boles was a sister in more than a strictly academic sense; I thank her for her friendship and admire her ability to think outside her field. Michel Cavigelli was something of a hero, as well as mentor and confidant. Martha Tomacek was a pillar of emotional support. Stuart Grandy and Pongthep Suwanwaree shared ideas and lab space. I thank all the K88 graduate students - and not a few from Zoology, Botany, and Crop & Soils - for friendship and collegiality. In particular, I thank Michelle Pruyn for introducing me to Qiaobing. Time and space fail me if I try to mention all the support staff that contributed substantially. I thank the staffs - past and present - of the Robertson, Hamilton, and Klug labs. Special thanks to Olivia Daemon, Kay Baergen, Sandy Halstead, John Furguson, Christine Easley, Sandy Marsh, and Madhav Machavaram. My field assistant Jennifer Siekierski patiently executed many bizarre - and sometimes futile - field treatments for me, with the courtesy of not voicing the obvious. I thank Barb Fox for handling innumerable administrative details. Special thanks to Nina Consolatti, who took a genuine interest in my equipment needs, and shared with me membership in the secret circle of those who really know how the weather station operates. John Gorentz cheerfully solved myriad computing problems, and shared my interest in the process as well as the solution. Randy De Jong, Nate Dom, and Greg Houseman were faithful friends; they listened patiently to my sad stories when experiments were floundering and gave much needed encouragement. I am inspired by their dedication and enthusiasm. My REU student Mike Case had as much influence on me as I had on him, I’m sure. I thank him for his energy, creativity, and persistence, and for teaching me more about nitrification than I ever thought I wanted to know. I thank Gabriel Vogeli for extensive help and advice on career development. Support from the following sources is gratefully acknowledged: the USDA-NRI Program, the NSF RTG and LTER Programs, the Michigan vii Agricultural Experiment Station, and the C. S. Mott Fellowship of Sustainable Agriculture. viii TABLE OF CONTENTS LIST OF TABLES ........................................................................................................ xiii LIST OF FIGURES .................................................................................................... xvii INTRODUCTION .................................. 1 Chapter 1 FACTORS INFLUENCING THE RELATIONSHIP BETWEEN DENITRIFICATION AND NITROUS OXIDE FLUX FROM SOIL. ........................... 3 Background ............................................................................................................... 3 Challenges for research ........................................................................................ 4 Factors influencing N20 mole fraction during denitrification ........................... 6 Theory .................................................................................................................... 6 Temperature ......................................................................................................... 9 Available carbon ................................................................................................... 9 pH ......................................................................................................................... 1 0 Nitrate and nitrite ................................................................................................ 11 Soil Water and Oxygen ...................................................................................... 12 Enzyme status .................................................................................................... 13 Time ..................................................................................................................... 13 Chapter 2 A HEURISTIC MODEL FOR THE CALCULATION OF DINITROGEN AND NITROUS OXIDE FLUX FROM NlTROGEN-15-LABELED SOIL ....................... 20 Abstract .................................................................................................................... 21 Measuring flux ........................................................................................................ 21 Graphical representation ..................................................................................... 21 Equafions ................................................................................................................ 22 Expectation of error due to multiple pools ........................................................ 23 Discussion .............................................................................................................. 25 Notes ........................................................................................................................ 27 Chapter 3 A NOVEL METHOD FOR DIRECT DETERMINATION OF NITROUS OXIDE AND DINITROGEN FLUX FROM 15N-LABELED SOIL ......................................... 29 Summary ................................................................................................................. 29 Introduction ............................................................................................................ 30 Materials and Methods ......................................................................................... 32 Overview .............................................................................................................. 32 Vessels and Sampling ..................................................................................... 33 Analysis ............................................................................................................... 34 Data Processing ................................................................................................ 36 Verification ........................................................................................................... 37 Laboratory Denitrification ................................................................................. 39 Field Demonstration ......................................................................................... 41 Results .................................................................................................................... 42 Verification ........................................................................................................... 42 Laboratory Denitrification ................................................................................. 43 Field Demonstration ......................................................................................... 44 Precision and Detection Limits ....................................................................... 45 Discussion .............................................................................................................. 47 Notes ........................................................................................................................ 52 Chapter 4 NITROUS OXIDE MOLE FRACTION DURING DENITRIFICATION IN SOIL: RESPONSE TO RECENT MOISTURE HISTORY VARIES AMONG ECOSYSTEMS ............................................................................................................ 72 Summary ................................................................................................................. 72 Introduction ............. . ............................................................................................... 73 Materials and Methods ......................................................................................... 74 Soil collection and processing ........................................................................ 74 Experiment and treatments ............................................................................. 75 Sampling and analysis ..................................................................................... 76 Results .................................................................................................................... 78 Experimental design ......................................................................................... 78 N20 mole fraction ............................................................................................... 79 Isotopic data ....................................................................................................... 80 Discussion .............................................................................................................. 82 Experimental Design ........................................................................................ 82 N20 mole fraction ............................................................................................... 84 Isotope data ........................................................................................................ 90 Conclusions ........................................................................................................... 94 OVERALL CONCLUSION ....................................................................................... 109 APPENDIX ................................................................................................................. 1 1 1 Literature Cited ......................................................................................................... 116 xii LIST OF TABLES Chapter 1 Table 1.1. Summary of factors potentially influencing N20 mole fraction. The generally-accepted direction of the correlation is indicated. WFPS = water-filled pore space. .................................................................... 17 Chapter 3 Table 3.1. Test for stability of measured ratios with varying sample size. Variations were simulated by adjusting the strength of the major beam for analysis of N20 laboratory standard gas. 2/1 and SH refer to mlz ratios 45/44 and 46/44, respectively. SE is standard error, n is number of samples. ....................................................................... 55 Table 3.2. Test for a memory effect during analysis. Highly enriched N20 was prepared by mixing various quantities of labeled and unlabeled N20. Analysis of laboratory standard N20 immediately followed each analysis of enriched gas. Since the laboratory standard is the same as the reference gas, a value of O %o is expected if there is no memory effect. ........................................................... 56 Table 3.3. Summary of analyses of high-enrichment laboratory mixtures of N20. 44:45:46 represents the mixing ratio of natural abundance N20, “N20, and 46N20. Predictions 15a and 45N20 adjust for gas purity. ‘53 is the atom fraction of 15N in the sample, consistent with the notation of Arah et al. (1992). 45N20 is the molecular fraction of xiii mass 45 in the sample. Measured 15a and “N20 are means for all samples where number of samples n is more than 1. ........................ 57 Table 3.4. Test for effects of the LiOH water trap, water in LiOH, and time on measured enrichment of a 1:12 mixture of natural abundance N20, “N20, and 46N20. 'Hydrous' refers to hydrous LiOH used as a chemical trap, 'dried' refers to hydrous LiOH dried on a vacuum line overnight. “N20 and 4‘5N2O represent the mean calculated molecular fraction for N20 of masses 45 and 46. SE is standard error. ..................................................................................................................... 58 Table 3.5. Test for an effect of collecting sample gas quickly using pre- evacuated vessels. Air samples were collected by diffusion into open vessels or by rapid filling of pre-evacuated vessels when stopcocks were opened. Samples were analyzed for N20 using the unmodified head amplifier. 2/1 and 3/1 are mlz ratios of 45l44 and 46/44, respectively. Means :I: standard errors are reported; n is 3. P > m is the significance level for Student's t test of differences in means. Least significant number (LSN) is the smallest number of samples needed to demonstrate significant differences in means at a confidence level of a = 0.05 .............................. 59 Table 3.6. 15N enrichment of soil N03' pools for laboratory denitrification experiments, calculated and measured . Calculations are based on the enrichments and mixing ratios of natural abundance KNOa' and highly enriched KNOa' (99.93 atom %). Measurements use the equations of Bergsma et al. (1999). 'n' is number of samples. Means :I: standard errors are reported for separate experiments measuring N2 and N20. P>|t| is the significance level for Student's ttest. ..................................................................................................................... 60 Table 3.7. Comparison of spectrometer precision by study. Precision is reported for analysis N2 or N20. 2/1 refers to mlz 29/28 or 45l44. 3/1 refers to 30/28 or 46144, except as noted. STD is standard deviation and n is number of samples .......................................................... 61 Chapter 4 The isotope data show that the mineral N pool undergoing denitrification was isotopically uniform in most cases. There is no conclusive explanation for the strong differences between estimates of N2 production by acetylene inhibition and by 15N isotope dilution. Perhaps, under the experimental conditions described, there existed an alternative substrate for production of N2, but not for N20; the 15N method may then have reported only production of labeled N2, while acetylene inhibition would have reflected gross production of N2. Factors likely contributing to the difference between methods are the complexity of the soil environment and the dynamic nature of N transformations during rapid re-wetting of soil. ....................................................................................................................... 95 Table 4.1. Nitrous oxide mole fraction ( N20 l [ N20 + N2] ) analyzed by ecosystem and recent soil moisture history. Estimates are mean :I: standard error, in pg N - g dry soil '1. Effects are P values (Prob. > F). “Ecosystem by history” is the interaction term. .................................. 96 Table 4.2. Comparison, by ecosystem and replicate, of predicted soil NO; enrichment (atom fraction 15N) with apparent enrichment of the soil pool undergoing denitrification. “Predicted” is calculated by mass balance from the extractable NO; levels in stock soil and the known addition of KNO3. Apparent enrichments are from N2 isotope data (“by Ng”), from N20 isotope data (“by N20”), or “average” of N2 and N20 estimates. ............................................................... 98 Table A1. Tests of significance (of differences among means) for Figure 4.2: probabilities of finding the observed differences in means by chance alone. Each row represents a single model of “Response”, incorporating the effects for which there are column entries. In most cases, models returning P > 0.05 for an effect were re-run without that effect ("-"). Dropping the effect (“-") is a notational convenience only; statistically it is no different from grouping the responses across levels (“grouped"). Levels of block are [4, 5, 6]. Levels of ecosystem are [cropped successional]. Levels of history are [pre-wet, control]. “NA" means the effect (column) is not meaningful for testing the response (row) ................................................................................................. 1 12 LIST OF FIGURES Chapter 1 Figure 1.1. Conceptual diagram: factors influencing flux of N20 from soils. WFPS = water-filled pore space. C = organic carbon availability. P = phosphorus availability. ............................. 18 Chapter 2 Fig. 1. Illustration of the equivalence of ternary and Cartesian plots. (a): One thousand simulated N2 samples, with each molecular fraction independently randomized, plotted as a ternary graph. (b): These same one thousand samples plotted as 29x vs. 15a. .............................................................................. 23 Fig. 2. Visualization of the heuristic model (Eq. [1], [2] and [3]). A represents the atmosphere and Mrepresents the sample: positions of both are determined analytically. The ray drawn from A through M intersects the equilibrium curve, implicating a pool P of gas derived from the soil, having an equilibrium distribution of masses. ................................................... 24 Fig. 3. Visualization of underestimation due to multiple soil N pools of differing enrichment. If P is in fact a mixture of equilibrium gases (resulting from multiple pools) then it Pcalculated necessarily falls below the equilibrium curve. Thus is necessarily an overestimate of Padua', when 15a,, > 15a... Since the segment AP‘i‘k‘miW’“l is longer than AP‘W“, flux is necessarily underestimated (since flux cc AM 1 AP) ......................... 25 Fig. 4. Visualization of a method to circumvent the problem of pool non-uniformity. After an initial incubation in which A, and M are sampled, the chamber is flushed, closed, and spiked with ”N2. A2 is displaced relative to A1. Measurement of N12 gives the trajectory of a second ray that converges on the first at Pam’a'. The principal assumption is constant isotopic character of the evolving gas. (Scale is illustrative only.) ............... 26 Chapter 3 Figure 3.1. Gas purification system connected to mass spectrometer. Valves are configured such that during purification, He bypasses the cold trap and travels onto the GC column, while sample passes into the cold trap (N2) or through the cold trap and then to vent via the mass flow controller (N20). During analysis, helium passes through the cold trap and onto the GC column, while the sample vessel remains open to vacuum via the mass flow controller. During N2 analysis the sample trap is not chilled. Different columns for N2 and N20 are used. Interface helium is used only for N2 analysis. The Penning valve passes about 10% of the gas stream to the spectrometer. ......................................................................................... 62 xviii Figure 3.2. Measured vs. calculated molecular fraction of 45N20 for moderate enrichments of N20, prepared volumetrically in the laboratory from enriched and unenriched standards; analyzed with the unmodified head amplifier ................................... 64 Figure 3.3. Isotopic character of nine N20 samples from laboratory denitrification of prepared ‘5N03' label, plotted against the equilibrium curve. As expected, these samples are in equilibrium. Vertical bars intersect the curve at predicted enrichments, open circles represent measured values ....................................................................................................... 66 Figure 3.4. N; and N20 fluxes with cumulative precipitation for nine field incubations over the same highly labeled plot ............... 68 Figure 3.5. Estimated soil enrichment, comparison of MS and GC, and N20 mole fraction for nine incubations represented in Figure 3.4 ..................................................................... 70 Chapter 4 Figure 4.1. Change in headspace concentrations (volume/volume) of nitrous oxide during incubation, with selected (for clarity) standard error bars. Dashed lines = successional soils, solid lines cropped soils. Open symbols = control, filled symbols = pre-wet. Gray squares = + acetylene, black diamonds = - acetylene. Control and pre-wet curves diverge strongly for cropped soils (solid lines with open or filled diamonds) but not for successional soils (dashed lines with open or filled diamonds). ........................ 99 Figure 4.2. Summary of production of N2 and N20. “Total denitrification" is N20 + N2 by acetylene inhibition, “N20” is production of N20 in the absence of acetylene, and “N2" is the difference. Even though total denitrification did not differ significantly between moisture histories for cropped soils (bars with solid borders), the control incubations produced significantly more N20 and correspondingly less N2. Bars with dashed borders represent soils from successional plots ........................................................................................................ 101 Figure 4.3. Isotopic character of N2 for labeled and unlabeled jars at the end of the incubation, paired (by line segment) to represent incubation units. Dashed lines = successional soils, solid lines = cropped soils. Open symbols = controls, filled symbols = pre-wet. Circles = Replicate 4, squares = Replicate 5, triangles = Replicate 6. Due to the overwhelming abundance of unlabeled N2 from the atmosphere relative to labeled soil—derived N2, displacement of isotopic character during the incubation (i.e., length of the line segments) is much smaller than for N20 (Figure 4.4). Only a small portion of the equilibrium curve (downward-opening parabola in Figure 4.4) is visible at this scale. For cropped soils, control segments (solid lines with open symbols) are much shorter than pre-wet segments (solid lines with closed symbols) indicating less N2 production for controls. For successional soils (dashed lines) the opposite effect or no effect is seen. ............................... 103 Figure 4.4. Isotopic character of N20 for labeled and unlabeled jars at the end of the incubation, paired (by line segment) to represent incubation units. Dashed lines = successional soils, solid lines = cropped soils. Open symbols = controls, filled symbols = pre-wet. Circles = Replicate 4, squares = Replicate 5, triangles = Replicate 6. Due to the overwhelming abundance of N20 from soil relative to unlabeled atmospheric N20, enriched samples represent essentially the isotopic character of N20 derived from soil. Samples falling on or near the equilibrium curve (downward-opening parabola) indicate a source in isotopic equilibrium, implying a single, uniformly labeled substrate pool (eg. a homogenous mixture of native NO; and ”NO; label). ..................................................................................................... 105 Figure 4.5. N2 production calculated independently by the acetylene inhibition technique (AIT) and by mass spectrometry (MS), for both cropped and successional soils. In every case but one, MS gave a much smaller estimate of N2 production than did AIT. ........................................... 107 INTRODUCTION Denitrification in soil is the major source of atmospheric nitrous oxide (N20), a potent contributor to global warming and to the destruction of stratospheric ozone. Flux of N20 to the atmosphere is poorly understood. There is a substantial imbalance in the global N20 budget (missing sources), and N20 flux at the field scale is difficult to predict, even when the rate of denitrification has been adequately characterized. Much of the uncertainty regarding N20 fluxes from soil may arise from the highly variable relative proportions of N20 and N2 produced during denitrification, commonly expressed as the N20 mole fraction. The primary objective of my dissertation research was to explore the influence of various ecological factors on N20 mole fraction. A supporting objective was to develop robust methods for characterizing N20 mole fraction in the laboratory and field. The structure of this dissertation reflects the objectives listed above. Chapter 1 reviews current knowledge regarding factors that influence N20 mole fraction directly and indirectly. Chapter 2 is reprinted from a publication in which we present theory and equations that facilitate the interpretation of 15N data for N-gas fluxes from soil. Chapter 3 explains our development of methods for measuring N-gas flux and the N20 mole ratio. We combine chamber methods and mass spectrometry in a way that allows very sensitive measurement of N2 and N20 flux from 15N labeled soil. The goals were to improve on the statistical uncertainty of traditional methods that require separate chambers for N2 and N20 measurements and to mitigate the inherent bias of 15N methods by using identical methods for N2 and N20 when estimating ratios. Chapter 4 deploys both traditional and isotope methods to examine an empirical question: what are the effects of short term moisture history and ecosystem management history on N20 mole fraction? Though many questions remain, I am convinced that we have made significant progress. Chapter 1 FACTORS INFLUENCING THE RELATIONSHIP BETWEEN DENITRIFICATION AND NITROUS OXIDE FLUX FROM SOIL. Background Human domination of ecosystems and biogeochemical cycles has become widely recognized (Vitcusek et al. 1997). Anthropogenic changes in ecosystems can generate changes in the Earth’s atmosphere, which in turn can have widespread consequences. Nitrous oxide is one of several atmospheric trace gases that have attracted the attention of the scientific community. It is produced in soils by microbial denitrification (Robertson 1999), nitrification (Firestone and Davidson 1989) and other biological processes (Robertson and Tiedje 1987). The concentration of N20 in the Earth’s atmosphere is currently about 312 ppb, (= 312 - 10'9 L - L") and is increasing at a rate of 0.5 ppb per year (IPCC 1996). Pre-industrial levels were < 290 ppb, suggesting that current levels reflect a significant anthropogenic influence. Nitrous oxide has a long (~120 y) half-life in the stratosphere. lt indirectly catalyzes the destruction of stratospheric ozone, increasing the hazard of UV exposure at the Earth’s surface (Hahn and Crutzen 1982). Nitrous oxide is also a radiatively active gas; about 5% of the observed radiative forcing of climate change over the last 100 years can be attributed to N20 (Bouwman 1990) International concern over rising levels of atmospheric nitrous oxide has stimulated research regarding its global sources and sinks. An estimated 14 T9 NZO-N are consumed or stored in the atmosphere annually, while only about 7.9 Tg of sources have been identified at the Earth’s surface (Davidson 1991). A large part of the imbalance in the nitrous oxide budget may be due to uncertainty in flux estimates, especially for tropical regions of the world. In particular, contributions from agriculture may be underestimated (Mosier et al. 1998, Kroeze et al. 1999). Most of the increase in nitrous oxide flux to the atmosphere may be related to land use change, especially conversion to agriculture. Deforestation produces pulses of nitrous oxide, whether by clear-cutting (Robertson and Tiedje 1988) or burning (Davidson 1991). Nearly 100 T9 of anthropogenic fertilizer N is applied to crops globally each year (Eichner 1994). Most of the anthropogenic N applied to watersheds in the North Atlantic drainage basin is eventually denitrified (Howarth et al. 1996). If even a small fraction of fertilizer N results in N20 flux, contribution to the global budget could be significant. Eichner (1994) estimates that as much as 2.1 T9 N20-N - y'1 may derive from fertilizer, similarly Matthews (1994) estimates the fertilizer source at 2.0 T9 N20- N - y". Mosier et al. (1998) place emission of NZO-N from agricultural soils for 1989 at 2.1 T9. Challenges for research Although denitrification is well understood, the relationship between denitrification and N20 flux from soil is not simple. First, other processes, especially nitrification, produce N20 in soil (Firestone and Davidson 1989). While nitrification dominates in well-aerated soils and denitrification dominates under anaerobic conditions, both can contribute simultaneously to N20 flux in soils of intermediate aeration (e.g. Stevens et al. 1997, Panek et al. 2000), or even low aeration (Wolf and Russow 2000). The relationship between soil moisture and relative importance of nitrification and denitrification for N20 production is uncertain (Davidson 1991). Over large temporal and spatial scales, flux of N20 from nitrification is smaller and more predictable, while denitrification accounts for the bulk of N20 flux but is highly variable temporally and spatially. Second, denitrification has another major product: dinitrogen. Dinitrogen, unlike nitrous oxide, is formed only by denitrification, and represents closure of the nitrogen cycle. The relative proportions of N20 and N2 produced during denitrification vary widely. Field studies have shown the inadequacy of using a simple conversion factor to estimate N20 flux from denitrification (Vinther 1984) or denitrification from N20 flux (Weier et al. 1993). The N20 mole fraction during denitrification (i.e., N20l[N20 + N2] ) can vary from zero to one (e.g. Rolston et al. 1982), is influenced by a suite of physical and chemical factors (Firestone and Davidson 1989), and is unstable over time (Letey et al. 1980a). Third, N20 flux during denitrification in soil is partly uncoupled from N20 production per se by factors which regulate the transfer of N gases from soil solution to the atmosphere. N20 may be released from soil hours or days after it is formed, while N2 produced at the same time may diffuse away from the source more quickly due to lower solubility in soil water (Letey et al. 1980b). Other factors, such as freeze/thaw cycles and sudden changes in barometric pressure, may also modulate N20 flux independently of N20 production. Fourth, measurement of denitrification is difficult. Chamber methods and gas chromatography suffice for estimates of N20 flux. However, changes in headspace concentration of N2 are nearly impossible to detect against the high atmospheric background. Two methods of N2 measurement have been recommended: acetylene inhibition and 1SN dilution (Mosier and Klemedtsson 1994). Both have disadvantages, which vary in importance depending on the specifics of implementation. This dissertation focuses on the second and fourth of the challenges listed above: factors influencing N20 mole fraction and methods for measurement of denitrification. N20 mole fraction is briefly reviewed below. The two chapters following explore theory and methods, respectively, for measuring denitrification products. A final chapter advances understanding of the relationship between denitrification and N20 flux by testing for ecosystem differences and effects of soil moisture history on N20 mole fraction. Factors influencing N20 mole fraction during denitrification Theory Denitrification is an anaerobic microbial process that converts nitrate to gases (Paul and Clark 1996) by means of the cellular enzymes nitrate reductase (NR), nitrite reductase (NiR), nitric oxide reductase (NOR), and nitrous oxide reductase (NOS). It can be summarized as follows: 2[ N03'15-R—> 2[ N02'1-’&> 2N0 fl» N20 N°S > N2 Nitrate serves as an electron acceptor, most commonly for the anaerobic respiration of organic matter. Almost all denitrifying organisms use 02 preferentially, with the result that denitrification is restricted to anoxic sites (Paul and Clark 1996). Genera with denitrifiers are found among the organotrophs, chemolithotrophs, photolithotrophs, diazotrophs, thermophiles, and archaea, although organotrophs are the principal denitrifiers in soil. In the denitrification sequence (above), NO and N20 are free intermediates that may or may not be further reduced, depending on conditions. Proximally, N20 production depends on the relative status of NOR and N08. lf N08 is fully active, denitrification may result in NZ only; however, if NOS is inhibited relative to NOR activity, N20 is produced. It is convenient to express the relationship between N20 and N2 production as N20 mole fraction: N20 / [N20 + N2]. Alternative formulations of mole fraction giving essentially the same information appear in the literature (9.9. N; / N20 etc.). A suite of proximal controls on N20 mole fraction have been consistently identified, with few variations (Colbourn and Dowdell 1984, Sahrawat and Keeney 1986, Firestone and Davidson 1989, Arah and Smith 1990, Bouwman 1990, Aulakh et al. 1992, Hutchinson and Davidson 1993). Table 1.1 lists factors hypothesized or demonstrated to influence N20 mole fraction. Not surprisingly, most of the factors that influence N20 mole fraction also influence the overall rate of denitrification (e.g., Paul and Clark 1996). By what mechanisms do soil physical and chemical factors influence N20 mole fraction? Change in N20 mole fraction at the cellular level results from a change in the relative rates of production and consumption of N20. Environmental factors can change N20 mole fraction by altering the effective availability of substrates, by altering the relative production and maintenance of enzymes, or by inhibiting the function of enzymes. Betlach and Tiedje (1981) demonstrated the importance of substrate availability. They manipulated soil characteristics and followed each pool in the reductive sequence using 13N label. They found that their results could be predicted qualitatively using a simple Michaelis-Menten model for each reductive step (cf. Dendooven et al. 1994). Experimentation with the model system led to the generalization that any change that slows the overall rate of denitrification leads to an accumulation of N20. Additionally, they concluded that this effect alone was sufficient to account for their experimental results; differential inhibition of enzymes need not be invoked. The Betlach and Tiedje paradigm is consistent with most other experimental findings. An important implication of the Betlach and Tiedje paradigm is that, for many controlling factors, influence on denitrification rate and influence on N20 mole fraction are twin aspects of a single phenomenon: general modification of reaction rates. However, conceptual independence is warranted because there are important exceptions (eg. the effect of nitrate concentration on N20 mole fraction). The hole—in-the-pipe model (Firestone and Davidson 1989, Davidson 1991) maintains a functional distinction between denitrification and N20 mole fraction as controls on N20 flux. According to the model, N20 flux is subject to controls at 3 levels: factors influencing overall rates of nitrification and denitrification (flow through the pipes), factors influencing N20 mole fraction (the size of the holes in the pipes) and factors influencing transport of gases out of the soil. Figure 1.1 shows explicit relationships among these three levels of influence and the proximal factors listed in Table 1.1. Notice that if a factor has opposite effects on rate and N20 mole fraction, changes in N20 flux tend to be buffered. Sahrawat and Keeney (1986) provide a rather comprehensive review of these factors. Below, their work is summarized, with supplemental material from other sources. Temperature Temperature generally has a positive effect on total denitrification and a negative effect on N20 mole fraction. Some have questioned the importance of temperature for predicting N20 mole fraction (Lensi and Chalamet 1982). Temperature has a general influence on a wide range of biochemical processes; its influence is likely to be complicated (Sahrawat and Keeney 1986). Available carbon Organic carbon added to soils can stimulate denitrification and decrease N20 mole fraction (Sahrawat and Keeney 1986). Greater supply of electron donors may create a demand for electron acceptors, causing N20 to be reduced. This is consistent with early work by Nommik (1956). Weier et al (1993) reported a decrease in N20 mole fraction with increasing carbon availability. However, Dendooven et al. (1996a) discovered an increase in N20 mole fraction with added glucose. pH Soil pH is an important regulator of enzyme activity. Early work showed an increase in N20 mole fraction at pH < 6 (leer and Delwiche 1954, Nommik 1956); this finding was supported by later research (see Sahrawat and Keeney 1986, and references therein). Focht (1974) modeled N20 flux from soils using Nommik’s data: he incorporated an explicit effect of pH on N20 mole fraction. Blackmer and Bremner (1978) disputed Focht’s model; they found that pH had no effect in the absence of nitrate. Whereas leer and Delwiche (1954) had suggested that low pH inhibits the reduction of N20 to N2, Blackmer and Bremner (1978) claimed rather that nitrate inhibits the activity of nitrous oxide reductase although it stimulates its production (cf. Blackmer and Bremner 1979) and that low pH interacts with nitrate to increase its inhibitory effect. Koskinen and Keeney (1982) attribute the pH effect on N20 mole fraction to either the greater sensitivity of N20 reductase to pH or to changes in species diversity at low pH. They note (sensu Blackmer and Bremner 1978) that there is no pH effect at low nitrate concentrations. They suggest that low pH may inhibit NOS indirectly by favoring high nitrite concentrations. Weier and Gilliam (1986) noted large fluxes of N20 from low pH soils; fluxes were correlated with 10 accumulation of N021 In general, pH control of N20 mole fraction has not been demonstrated in the field (Davidson 1991). Nitrate and nitrite It is universally reported that high nitrate concentrations increase N20 mole fraction (Nommik 1956, Cooper and Smith 1963, Blackmer and Bremner 1978, Vinther 1984, Weier et al. 1993). Firestone et al. (1979) found, however, that nitrite was a more potent inhibitor of nitrous oxide reduction than nitrate (cs. Gaskell et al. 1981, cited in Coulboum and Dowdell 1984), and suggested that small amounts of nitrite produced from large pools of nitrate may in fact be responsible for the inhibition usually observed. Since nitrate, when limiting (Sahrahwat and Keeney 1986) is positively correlated with denitrification and yet increases N20 mole fraction at high concentrations, it does not conform in a strict sense to the Betlach and Tiedje model. This suggests that its influence is not via simple substrate kinetics. Probably nitrate concentration is not high and limiting at the same time; nitrate may have a small but unreported kinetic effect on N20 mole fraction at low concentrations but must exert influence by a different mechanism at high concentrations. Two proposed mechanisms are that nitrate competes with N20 as an electron acceptor or that it directly or indirectly inhibits NOS. Firestone et al. (1979) favored the first mechanism for nitrate but the second for nitrite. This question has not been resolved, although the fact that pH interacts with nitrate concentration suggests the priority of enzyme inhibition. 11 Soil Water and Oxygen Three mechanisms are responsible for the influence of soil water on N20 mole fraction. First, at very low soil moistures, metabolic stress may limit denitrification (Sahrawat and Keeney 1986) and N20 mole fraction may be influenced concomitantly. Second, at high soil moistures, Oz availability in soil microsites is limited by diffusional transport of oxygen; denitrification and N20 mole fraction are affected. Third, high soil water content limits the diffusion of N gases away from the sites of formation, increasing the probability of further reduction of NZO-N. The first two mechanisms are consistent with the kinetic model of denitrification (Betlach and Tiedje 1981) and all three proceed in the same direction with respect to the effect of soil moisture on N20 mole fraction. Water-filled pore space, when expressed as a percentage of the total pore space, is the most informative way of describing soil moisture because it integrates absolute water content (eg. gravimetric) with bulk density (Linn and Doran 1984). It is probably the single most useful predictor of N20 mole fraction, largely because of its inverse relationship with soil aeration. Denitrification is anaerobic, and all denitrification enzymes are sensitive to (inhibited by) oxygen. Lower soil water contents favor N20 production over N2 (Nommik 1956, Weier et al. 1993). However, the effect of water status is inseparable from its effect on soil aeration. Oxygen inhibits denitrification but increases N20 mole fraction (Firestone et al. 1979). Focht (1974) included aeration as a 12 determinant of N20 mole fraction in his model of denitrification. Implications for enzyme status are discussed below. Enzyme status The principal effect of oxygen is to inhibit denitrification enzymes. The most sensitive of these is the terminal enzyme NOS. As oxygen increases, reduction of N20 to N2 slows sooner than reduction of NO to N20, thereby increasing N20 mole fraction (McKenney et al. 1994). Notwithstanding, the Betlach and Tiedje model predicts increasing N20 mole fraction with increasing Oz, regardless of enzyme sensitivity. If a well-aerated soil is sufficiently wetted that it becomes anaerobic, denitrification enzymes are induced sequentially (see Letey et al. 1980a). Thus some N20 accumulates before NOS is completely induced, and N20 mole fraction should decrease with time. The ever-present potential for synthesis and inactivation of enzymes makes N20 mole fraction inherently dynamic. Enzymes may be induced, repressed, de-repressed, re-repressed, or destroyed. Any difference in the time-dependent behavior of the last two reduction enzymes leads to a time dependent difference in accumulation of the last two reductive products - N20 and N2. Just as water status is usually inseparable from Oz status, enzyme status is inseparable from its temporal dependency. Time Changes in N20 mole fraction may lag behind changes in factors that control it, creating complex temporal dependency. Rolston et al. (1978) found that N20 mole fraction was highest at the initiation of denitrification, and 13 decreased thereafter. Jacinthe et al. (2000) manipulated water tables in soil columns (various drainage classes) to stimulate denitrification; N20 mole fraction was 0.95 four days after raising a water table to 10 cm below the soil surface, but dropped to 0.35 one week later. Firestone and Tiedje (1979) reported that after the onset of anaerobiosis, N20 mole fraction was initially low, then rose significantly, and then declined. They attributed this effect to the staggered synthesis of enzymes, and speculated that it reflected the sequence of events following 02 depletion due to rainfall or irrigation. Letey et al. (1980a) reported high initial values for N20 mole fraction during denitrification, with a subsequent decrease to zero. They attributed this effect to differential rates of induction of nitrate reductase and nitrous oxide reductase. Rolston et al. (1982) found that “nitrous oxide mole fractions tended to be smallest immediately after irrigation and increased as the soil water redistributed and became less anoxic”. They also found a decrease in N20 mole fraction with successive irrigation cycles. Hallmark and Terry (1985) reported that N20 mole fraction following irrigation was initially high, then dropped over a period of six weeks. Weier et al. (1993) reported N2/N20 ratios rising (i.e., N20 mole fraction dropping) over a 5-day period. Dendooven and Anderson (1994) found that de novo synthesis of nitrate reductase and nitrous oxide reductase began one and 16 hours after anaerobiosis was imposed, respectively. Dendooven and Anderson (1995) found that upon return of soil to aerobic conditions, N20 mole fraction increased with time over a 70 day period (from 51% to 100%). Conversely, soil cores submerged for 96 h had smaller subsequent N20 mole 14 fraction than soils submerged for 6 h (Dendooven et al. 1996b). In general, it is difficult to predict N20 mole fraction without knowing both the antecedent water regime and the time since change in water regime. Effects of moisture cycles on denitrification and N20 mole fraction merit consideration. Arnold (1954) suggested that moisture fluctuation should stimulate gaseous N loss from soils. High denitrification rates are sometimes attributed to extreme drying-wetting cycles (Peterjohn and Schlesinger 1991). Smith and Patrick (1983) found that 7 and 14 day aerobic/anaerobic cycles produced far more N20 than either condition alone. A simple explanation is that the cycling systems represent intermediate levels of aeration, such that denitrification is supported, but not complete reduction to N2. However, moisture cycles may support more complex dynamics. Groffman and Tiedje (1988) showed that hysteresis (dependence of the response on the direction of change) is important for the overall rate of denitrification. Hutch et al. (1999) found N20 emission under a fluctuating moisture treatment 4 to 9 times higher than under constant low or high moisture. The final chapter of this dissertation describes an investigation of the temporal dependence of N20 mole fraction on soil moisture, i.e., the effect of moisture history. Prospectus The task of understanding the relationship between N20 flux and denitrification will have been accomplished when N20 from denitrification can be predicted reasonably well from a discrete set of measurable factors. Soil water status will likely continue to be a focus of investigation because it is 15 easily measurable and correlates well with oxygen status. Oxygen status, though more difficult to measure in spatially heterogeneous soil environments, is the most important proximal regulator of denitrification enzyme status. Dynamic, differential variations in enzyme status are the principal cause of uncertainty regarding the relationship between N20 flux and denitrification because of the potential for production of the alternative product N2. From an empirical perspective, useful approaches will be (1) to investigate the response of N20 and N2 flux to differences in soil moisture (2) to assess the temporal dependence of the relationship between flux and moisture, and (3) to test for the robustness of the perceived patterns across varying environments. In the work that follows, I seek to implement these approaches by developing techniques for observing N20 and N2 flux, and by testing explicitly the temporal dependence of flux on change in moisture, for soils from contrasting ecosystems. 16 Table 1.1. Summary of factors potentially influencing N20 mole fraction. The generally-accepted direction of the correlation is indicated. WFPS = water-filled pore space. factor correlation temperature (0 - 30 °C) - water (60-100% WFPS) - Time since water addition - nitrate, nitrite + available organic carbon - pH (4 - 7) - 902 “I“ enzyme status variable rhizospherel plants ? depth of activity - soil structure + 17 Figure 1.1. Conceptual diagram: factors influencing flux of N20 from soils. WFPS = water-filled pore space. C = organic carbon availability. P = phosphorus availability. 18 . F. F 2:9“. \_. SeaEzcocoma 208 ONZ _ l + new . - snags. his E E E .I Cm «MZONZ l a/TLI- 3 z..+ + + -+« +\®\ + SEEEZ 30 can Beaumont“ a use onz +/$/ \_ gafiécouofi 22a ouz l + mam - w + - 3.52... a . E B -E E s 302 a- - + BESEEQ a cam era. #3 E5 19 Chapter 2 A HEURISTIC MODEL FOR THE CALCULATION OF DINITROGEN AND NITROUS OXIDE FLUX FROM NITROGEN-15-LABELED SOIL Published in Soil Science Society of America Journal 63: 1709-1716 (1999). 20 Wham theSoilSeiaccSna‘eoro/mm Volume 63. no. 6. Nam-Dec. 1999 6775m5eguM.MadimV/153711USA A Heuristic Model for the Calculation of Dinitogen and Nitrous Oxide Flux from Nitrogen-IS-Laheled Soil Timothy T. Bergsma.‘ Qiaobing C. Bergsma. Nathaniel E. Ostrom. and G. Philip Robertson ABSTRACT VeryseasitivematsongaadNfiflufronsoilan Mute-gandvedfromuN-lfleledsoilisaaalyzedbyisotope ratiomspectronetry.1'hisapproachisusefulforstudyiagthe Hedaiuugenlertilizerandforstudyingsoilmicrobialproceases coatriiuting to the atmospheric increase ofnitrotnoxideJradiatively activetracegasthatcanconuibutetoyobalwamingandoaoae depletion. Most system of equations that relate isotopic analysis to psfluxaresulficiently complexthatcerta’n limitatiousaadpoteatials olthe ”N approach may be overlooked. We describe a graphical representationollabeledN-gaslluxthatMeatheequatioas-d Woitialth‘nkingregardingtheinplemeatationdrelated eaperineatsflismodelisusedtointerpretanderestiatatioathat msifhxderivesfromnalt’quepoolsofdifleringenrichmenA statistialderivationispreaented lorapreviously published simulation of-derast'nafioadueto-ultiplepoohnesaneequatioasare qphed to field data to explore whether temporal variation in soil fiateearichneatislihelytocausesigflficant nderestinafiomTwo “WmWM-ayehhatethempfioa otpoolflordttherebyeiaa’aatiagapoteatialmoluderesti- nation. D lNlTROGEN AND NITROUS oxror: are alternative end products of microbial denitrification. Quantifying their flux from soil can help explain fertilizer losses from agricultural systems (Mosier et al.. 1986: Eichner. 1990'. Weier et al.. 1993) as well as the atmospheric buildup of N20 — an important greenhouse gas in the tropo sphere and ozone-destructive catalyst in the strato- sphere (Bouwman. 1990: lPCC. 1996). However. field studies of denitrification have been hampered by the insensitivity of standard instrumentation to N2 increases. e.g., under soil covers (Mosier and Klemedtsson. 1994). Mass spectrometric analysis of gas from 'SN-labeled soil is a sensitive method for quantifying the flux of N2. as well as NO. because of the low natural abundance of '5N. Isotopic data for headspace gases can also esti- mate the enrichment of the source N pool. can help identify the source of N for N20. and can be used to check assumptions of the flux method (e.g., Stevens et al.. 1997 : Arah. 1997). Recent advances in spectrometer sensitivity and affordability have generated new interest in using "N for soil biogeochemical investigations. Given the complexity of the system of equations nor- mally used for interpreting the isotope data. it is useful T.T. Bergsma and G.P. Robertson. W.l(. Kellogg Biological Station and Dep. of Crop and Soil Sciences. Michigan State Univ.. Hickory Corners. Ml 49060. QC. Bergsma. MedFocus Clinical Research Con~ sulting Services. Des Plaines. ll. “18: NE. Ostrom. Dep. of Geologi- cal Sciences. Michigan State Univ.. East Lansing. MI 48823. Received 16 Oct 1998. ‘Corresponding author (tbergsma@kbs.msu.edu). Published in Soil Sci. Soc. Am. J. 63:1709-1716 (1999). 1709 21 to have conceptual tools which make the principles of the system more intuitive. We present here a heuristic model that illustrates the estimation of source pool enrichment and estimation of soil-derived headspace gas for N; or NO from 1’N- labeled soil. A convenient graphical representation of N isotope data is identified. from which a complete set of equations is derived by geometric inference. These equations are similar in form and identical in function to those of others (Siegel et al.. 1982; Mulvaney. 1984; Arah. 1992). To illustrate utility, the model is applied to the problem of underestimation that occurs when flux derives from multiple pools of differing enrichment. Measuring flux Flux of N2 produced by denitrification in a "N- enriched soil can be measured by monitoring the in- crease in enrichment of headspace gas in a chamber placed over the soil. Proper analysis requires measure- ment of the abundance of all three molecular masses of N3 (28, 29. 30). When atoms of lsN are distributed randomly among a sample of N2 molecules. measure- ment of any two masses suffices. because the abundance of the third mass can be predicted statistically. However. a mixture of N; from two differently labeled sources (e.g.. enriched N; from the soil and unenriched atmo- spheric N2 in an enclosure) is not in isotopic equilibrium (Hauck et al.. 1958); that is. the isotopes of N are not randomly distributed among the three molecular frac- tions. This means that all three masses must be mea- sured. which has the additional advantage of providing an indirect estimate of the average enrichment of the soil N pool (Hauck and Bouldin, 1961). The estimation of enrichment is convenient because it is non-destructive and is a time-weighted mean. Equations for the determinations of flux and source enrichment, by isotope ratio mass spectrometry. are well established (Siegel et al., 1982; Mulvaney and Boast, 1986; Mulvaney, 1984; Arah. 1992). These equations are designed to measure total flux of N2 and assume that the gas is derived from a single, unifomly labeled pool of soil N. (In the absence of pool uniformity. flux derived from added label—cg, fertilizer—can be estimated. but not by these equations.) The same principles apply for N30; for simplicity, most of this discussion is limited to N2. Graplical Representation We adopt, wherever possible, the notations and defi- nitions of Arah (1992). Briefly. the "N atom fraction l’a of any sample of N is the total number of l5N atoms 1710 son. scr. soc. AM. 1. VOL 63. NOVEMBER-DECEMBER 1999 divided by the total number of N atoms (informally. the enrichment). The molecular fractions 21‘x. 2"x. and ”x are the fractions of the total number of N; molecules in a sample with masses 28. 29. and 30. respectively. A sam- ple of N3 is in isoropic equilibrium if the molecular frac- tions follow a binomial distribution: 28x = (Na)! [la] 2"x = 2(“a) (‘50) [1b] -‘°x = ("of [1c] where " = l - 15a [1d] All N: (or NO) derived from uniform pools of mole- cules with single N atoms (e.g., nitrate) by microbial and chemical processes (e.g., denitrification) are expected to be in isotopic equilibrium. even if isotopic fractionation is significant (i.e., even if one isotope is inherently fa- vored by the process). All mixtures of two or more equilibrium samples with different "N atom fractions are necessarily nonequilibrium mixtures. Atmospheric N2 is assumed to be in isotopic equilibrium. Additionally, we define the isotopic character of an N; sample as the relative proportions of the three molecular fractions. We represent the isotopic character of N; by plotting ”1 versus ”a. The fact that such a plot is equiva- lent to a ternary plot of ”x, ”x. and ”r: (Fig. 1; Note 1) is both a mathematical convenience and an informal proof that ”x and "a completely characterize N2 isotopi- cally. The isotopic character of a single sample may be expressed as the Coordinate pair (”a. ”x ). Figure 2 represents graphically the calculation of source pool enrichment and flux from a plot of isotopic character. The downward-opening parabola is called the equilibrium curve. and has the form of Eq. [1b]; it repre- sents the isotopic characters of all possible N2 samples that are in isotopic equilibrium. Whenever N atoms are paired randomly from a source of given enrichment. ”N3 will be a quadratic function of IN. with no ”N, produced when all or none of the source atoms have mass 15. The symbol A represents the isotopic character of atmospheric N3 already present in a chamber headspace. P represents the isotopic character of N2 derived from the soil pool: since the soil pool consists of uniformly enriched mineral N. Component P is initially in isotopic equilibrium. M represents the isotopic character of a mixture of atmospheric and soil-derived N2 in a chamber headspace (this mixture is not in isotopic equilibrium). Plotted thus. the isotOpic character of any mixture is a linear interpolation between its two constituents. and its distance from either is inversely proportional to the relative contribution from that constituent. Thus M falls on a line between A and P. and its position along that line indicates the mixing ratio of A and P. In practice. a soil cover is deployed. and gas samples are taken at the beginning and end of an incubation. A and M are the isotopic compositions of these samples. respectively. P is unknown initially. but must fall some- where on the ray drawn from A through M. and must 22 also fall on the equilibrium curve. Thus. the intersection of the curve and the ray identifies the character of P. The atom fraction of the soil N pool, identical (barring fractionation) to that of P. is displayed on the horizontal axis. The relative contribution of P to the mix is given by the “travel” of M along the ray (the length of segment AM). divided by the total length of the segment AP. Since the ray has constant slope. the relative contribu- tion can be determined simply from the relative enrich- ments ("N) of A. M. and P (i.e., the proportion collapses to a single axis). Equations From the concept illustrated in Fig. 2. equations for gas flux and enrichment of source pool can easily be derived by geometric inference. Formally, the isotopic character of a gas mixture and the assumed or measured character of one of its pre-mixing components (e.g., atmospheric N2 already present in a chamber head- space) can be used to calculate both “a,. the atom frac- tion of the second component (e.g., soil-derived N2), and d, the fractional contribution of the second compo- nent to the mixture. This is achieved by assuming that the second component is initially in isotopic equilibrium (as stated earlier). The atom fraction of the second com- ponent is found by simultaneous solution of an equation for the mixing line ”x = C + 5% [2a] and an equation for the equilibrium curve [1b], which results in a quadratic expression: "a, = {-3 z (132 - 4AC)"’] I 2A [2b] where A = 2 [2c] 8 = s - 2 [2d] C = ”x. - 5050.) = ”x. - 3("0.) [2c] = (29x- ._ 29x.)/(IS . _ 15“,) [2f] and the Subscripts a. p, and m refer respectively to the initial component (atmospheric N), the second compo- nent (soil pool N). and the mixture. C is the intercept for the mixing line, and s is the slope of the mixing line. The relative contribution of N; from the second component is d = (15 _ ”ad/(”0, _ ”a.) [3] Absolute flux can be calculated from relative flux (rela- tive contribution) by associating some absolute measure with initial or final quantity of headspace gas: for exam- ple. if final N, concentration is 0.8. and chamber volume is 1 L. then a value of 0.01 for d implies evolution of (0.8)(1 L)(0.01) = 0.008 L of N2. Note that a quadratic expression is used to find ”a, which in turn is used to find d. Arah (1992) used a quadratic to find if and used d to find ‘5a,,. The two systems give identical results. The relevant root in Eq. [2b] can be identified by inspection of Fig. 2 and is necessarily the greater; the other root is identical to a. in this case (cf., Boast et al., 1988 Eq. BERGSMA ET AL: HEURIS‘nC MODEL or nmtooermxs FLUX mom NHROGEN-IS-LABELED SOIL 1711 "Na molecular traction x “N atomic fiction I Fig.1.MMMWMteruyaadCandmmtahwuon-dfialfledNxmmufimm independentiyrandomized,Waatern-ymthbeaesaneoletho-alduplesplottdas’xu‘n. [26]. [31] fl“. ). Note again that all mixtures of equilibrium never past it. The formal proof has been provided by gases necessarily be under the equilibrium curve. This Boast et a1. (1988; see Eq. [19]). is true regardless of the number of equilibrium compo. nents in a mixture. As an informal proof. consider step— . wise additions of many equilibrium components to an Enid?!” “12°: linorbDue accumulating mixture. The first two components create 0 p e 0° a mixture below the curve; subsequent additions dis- The utility of the heuristic model outlined above can place the mixture toward some point in the curve. but be illustrated by revisiting the principle assumption of 23 1712 SOIL scr. soc. AM. 1.. VOL 63. NOVEMBER—DECEMBER 1999 0.5 1' P 0.4 < x F g 0.3J a 3 0.2 3 ii 0.1 A? M A . o i 4. i 4. i t j 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 “Nannietacttona Fig.2. Visaah‘aatioaotthehe-istie-odelfl‘lq-IILIZL-‘Dll-AWhmduwfieqlepa‘a-db‘ www.mmhnMAWMMhMmhflafigadedeI-fie aim-mama“ disequilibrium methods Of N; or N20 flux analysis (Hauck and Bouldin. 1961; Siegel et al., 1982; Mulvaney, 1984; Arab. 1992). Any method that does not measure enrichment of the soil mineral N substrate directly. but infers it from isotOpic data, assumes that soil N; or N20 derives from a single, uniformly labeled pool. The im- portance Of this assumption has been debated (Focht, 1985; Mulvaney and Kurtz. 1985) and evaluated in the lab (Mulvaney, 1988) and field (Mulvaney and Vanden Heuvel. 1988) and tested by simulation (Vanden Heuvel et al., 1988; Arab, 1992). Addition Of ”N-labeled mate- rial to native soil pools creates the possibility Of at least two differently labeled pools. Theory suggests that con- current flow from multiple pools of different enrichment should usually lead to an overestimation Of soil N enrich- ment, and should always lead to an underestimation Of flux (Boast et al.. 1988; Arah. 1992). Arah (1992; see Fig. 2a) simulated the underestima- tion that occurs when the assumption Of pool uniformity fails. Estimated mixing ratio d was compared with actual mixing ratio for 1000 runs in which number of pools was randomized on the Interval [1, 50] (cf., Vanden Heuvel et al.. 1988, for simulations based on two pools) and pool enrichment was randomized on the Interval [7, l] (where 'y is the natural abundance of "N in the atmosphere. or 15a.). The plot Of estimated vs. actual showed very little scatter about a line with a slope of 0.76; thus, for the conditions that were simulated, the isotopic method consistently underestimated N2 flux by about 24%. This non-intuitive result begs for a more fundamental interpretation. We have discovered that Arah’s result can be calcu- lated directly from statistical principles, without re- course tO simulation. It is instructive to begin by visualiz- ing how underestimation occurs (Fig. 3). If a gas sample P'“““ is itself a mixture Of N; that derives from two or more soil N pools. then the character of P will fall somewhere below the equilibrium curve (Fig. 3). The ray approaching P”"" from the direction of A will neces- 24 sarily “overshoot” P, intersecting the curve beyond P. Thus, I5a,, will always be overestimated if 15a, > "a. (as shown by Boast et al., 1988; and Arah, 1992). It is apparent from Fig. 3 (and from Eq. [3]) that d is always underestimated: the segment AW is always longer than the segment AP‘“"‘ (as per Arah, 1992). The underestimation can be quantified if the isotopic character of P‘““" is specified. Regarding Arah’s (1992) simulation. it is appropriate to associate P“ with the statistical expectations of "a and ”x for a mixture of equilibrium gases. whose values of 1’a are randomly and uniformly distributed on the interval [g, h]. With P“ thus specified, the expectation Of underestimation can be calculated. The answer derived here is independent of the number Of contributing gases (pools) because statistical expectation E(x) is independent of the num- ber Of samples. For generality, we take the case where each pool has equal weight. For a random variable uni- formly distributed on the interval [g, h]. 5050) = (g + Ill/2 l4] Var("a) = (h - g)’/12 [5] E("a’) = E2(“a) + Var("a) [6] Considering the dependence of ”x on "a (Eq. [1b]), the expectation of ”x is given by E(”x) = E(2 l5a - 2 "a1) [7a] = ZE("a) - 2E("a’) [7b] Substituting [5] in [6] and [6] in [7b] gives E(”x) = ZE("a) - 2 [Ez("a) + (h — g)’l12] [8] Thus, given a range of enrichments, the expected isoto- pic character Of a random mixture PM, ”1““) can be determined from [4] and [8]. To calculate the resulting underestimation, an atom fraction "a,” is found by substituting ”W and "am for ”x. and l5a... in Eq. [2]. A coefficient of estimation e can be expressed as BERGSMA ET AL: HEURISTIC MODEL OF NITROGEN-GAS FLUX FROM NI'IROGEN—IS-LABELED son. 1713 0.5 T "N: molecular traction x 9 9 .o N W . ‘ l 1r O .a i K 0 t ¢ ¢ ¢ 0 0.1 0.2 0.3 0.4 A 1 4* 4 Y I V 1' fl 0.5 0.6 0.7 0.8 0.9 1 "Natomictctiona FglvmdmmmWMNMdMMHPihMa-n—ed (redthglron-ltiplepoolsnhea WAWBWMAF,M e = doubted/dew!“ [9a] = (1509mm _ ‘50,)/(”ap°’°““‘ _ 150.) [9b] Eq. [9b] can be proven from Eq. [3]. For the case where the range of enrichments is [0.1], ”mm = (V2) by Eq. [4] and ”W = (1/3) by Eq. [8] (Note 2). Inter- estingly. if the lower bound of the interval 3 is equal to y(‘5a.: cf. Arah. 1992). e— = 0.75. regardless Of the range Of enrichment. this value depends neither on the value of 1 nor it (Note 3). In summary, we have provided a more fundamental interpretation of Arah‘s (1992) simulation (Fig. 2a). Our heuristic model for N-gas flux from lsN-labeled soil shows why underestimation occurs. Our equations re- duce the simulation to a relatively straight-forward cal- culation. We show that even under somewhat less restric- tive conditions than Arah’s (i.e., range not specified). the coefficient Of underestimation e evaluates to 0.75. This value agrees well with the slope of 0.76 in Arah’s Fig. 2a, which can also be interpreted as an index Of underes- timation. Simulation is a valuable tool for exploring systems of equations that defy direct solution: the dis- covery Of a direct solution for such a system represents progress. While our solution may not represent any real set of field conditions, it does help predict how field conditions will influence the accuracy Of isotopic meth- ods for N-gas flux measurement. The approach employed above can be used to explore other questions about labeled N; or NO fluxes. For instance, Hauck and Bouldin (1961) state that their sys- tem gives a value for 15N that “represents the average isotope content of the material undergoing denitrifica- tion over a given period of time.” But the concept Of “average” implies that the enrichment is changing with time (e.g., by dilution from concurrent nitrification). and thus the assumption of pool uniformity is violated. In principle. it makes no difference whether the assump- tion is applied to space or time. Is change Of source pool enrichment with time likely to be a significant source of error in field experiments? 25 RWMWMWmMPHMWF,m“g>'~Shafie isalwaysaaderestiaatedts‘ueelluxAMIAP). We sampled N20 over a 3.5-d period in April 1998 in a heavily labeled wheat plot (30 kg ha" as 99% 1(”NO3). Enrichment of the source pool, as inferred from N20 isotope data. dropped gradually from 82 to 72% during this period. Even if this entire drop had occurred during a single incubation, the resulting under- estimation would have been negligible. Let [g. h] be [0.7, 0.8] and let g = am. From Eq. [9b] (which invokes Others), e = 0.999. For comparison, a drop from 80 to 60% during an incubation yields e = 0.993. We conclude that error from temporal changes in enrich- ment for our experiment must have been negligible, and is probably negligible in most cases. DISCUSSION The heuristic model presented above facilitates the design of N-gas flux experiments and the interpretation Of isotopic data for N2 and N20 samples collected over ”N-labeled soil. It is particularly useful for exploring the problem of underestimation that occurs when N; or N20 analyzed by mass spectrometry derives in part from a soil pOOl that is not uniformly labeled: it illustrates how underestimation occurs. We reduced a published simulation of underestimation to a direct calculation based on statistical principles. We showed the general utility of our equations by evaluating a case where en- richment varied over time, rather than in space. Isotopic methods for measuring N2 and N20 flux have general appeal because they are relatively non-disrup- tive Of soil systems and because they represent the only practical direct method for measuring N2. As isotope ratio mass spectrometry (IRMS) becomes more widely available, use of isotopic methods will continue to grow. Although flux equations and analytical methods have been available for decades, there still exists considerable uncertainty regarding the accuracy of the method when applied in the field. Most of the uncertainty pertains to the necessary assumption that empirical methods result in uniformly labeled soil mineral N pools. Underestima- 1714 SOIL SCI. soc. AM. 1.. VOL 63. NOVEMBER—DECEMBER 1999 "N2 molecular fraction II 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1sN atomic fraction .1 Fig.4. Visual'mtionofa methodtoc'ne-ventthe ofpoolnon-nnifornn'ty. Alter-n ”Alibi.” proble- nitnliadian'oa' s-led. thechamberisnmheddM-dwiltedwith'N,AzisdisplaoedrelafivetoApMme-entofm,p'veathehjeaoryofaaecnnlrny pr'ndpalm—ptionis chm-acted thatconvergesonthefilstatP‘flThe tion is expected to result from the failure of the assump- tion, but the magnitude of the underestimation is diffi- cult to predict. Our work does not imply the existence of a theoretical method 10 t hefield. “Sta- tistical expectation", as used in our argument has a precise mathematical definition that is not equivalent to expected error in field measurements, unless field conditions closely match the constraints of the mathe- matical model. The underestimation mlculated by our equations for e will not likely be realized in the field unless (i) the number Of pools is large, (ii) the enrich- ments Of the pools are randomly distributed. and (iii) flux is distributed evenly among the pools. We doubt that any of these conditions is likely to be met in field settings. especially the third. Even if the conditions were met. it would seem impossible to know this a priori. Our experience with NO fluxes shows underestimation to vary within experiments and especially among experi- ments. When pared to estimates made by gas chromatography. agree- ment ranged from 6 to 117% (MS/GC, unpublished data). In a systematic laboratory study of N20 fluxes, Mulvaney (1988) found that differences between MS and GC usually were small (less than 10%) and probably resulted from analytical error. Mulvaney and Vanden Heuvel (1988) found no appreciable difference between MS and GC in the field. unless plots were relabeled. Can pool uniformity ever be assumed? We believe that when added N (labeled) far exceeds preexisting soil pool N. there is initially only one significant pool, which is practically uniform. When uniformity of the soil N pool cannot be as- sumed, it should be demonstrated (e g., Stevens et al., 1997). However. the conceptual model given here (e. g... Fig. 2 and 3) reminds us that it is not critical to know that the soil pool' Is uniform; rather it is critical to know the isotopic character of N gas derived from the soil pool. Uniformity merely makes this easy to calculate ctimatoc are ram 26 “personals-that theevolvingp. (Soda k We tidy.) (by invoking the equilibrium curve as one constraint). The error that attends violation of the assumption of pool uniformity could be avoided if there were altema- tive means of assessing the isotopic character of the N gas derived from the soil. It maybeposs ible ' ‘ ‘ ' L ‘ character of soil-derived gas. Consider two successive incubations of the same unit of soil. Suppose that, after sampling the headspace at the end Of the first incubation, the enclosure is flushed. closed, and spiked with a small amount Of ”N, The spike has the effect of displacing A1. the base of the ray for the second incubation (Fig. 4). The rays for the ’ ‘ c. 5.. at a point identifying the isotopic character Ofw the total evolved gas allowing accurate determinations of flux for both intervals (equations are outlined in Note 4). The principal assumption is that the isotopic character of the evolved gas is constant: probably a more robust assumption than pool uniformity. Another assumption is that change in headspace enrichment due to other processes (e. g., equilibration with soil pore space) is negligible. The amount of gas needed for the spike de- pends on the volume of the headspace and the sensitivity of the mass spectrometer. For N,, such a spike will be relatively expensive until ”N1 costs drop or sensitivity improves such that smaller changes In ”N; abundance become detectable. For N10, however, for which addi- tions of labeled N such as “NO can readily be measured in a normal atmosphere, this approach is already an option. Another approach for independently assessing the isotopic character of soil-derived gas is similar to that above. Two consecutive incubations are conducted, but before the second incubation, the headspace is purged of the gas of' Interest so that the final sample will contain only (mostly) soil-derived gas. This is, again, diffith for N2 because of contamination problems, but is an interesting possibility for N20. We believe the heuristic model described here' is use- BERGSMA ET AL: HEURISTIC MODEL or NITROGEN-GAS FLUX FROM NITROGEN-Is-LABELEO SOIL 1715 ful for clarifying principles. for designing experiments. and for evaluating data related to N: or NO collected over labeled soil. "N approaches to measuring N2 and N20 fluxes are likely to become more common as associ- ated materials and technology become more readily available. Application of conceptual tools to explore the limits and potential of isotopic methods is warranted. Note 1 A Cartesian plot of 2"I: vs. ”a is graphically identical to a ternary plot Of 2‘sx. 2"x. and 30x if the ternary plot is bounded by an equilateral triangle. and if the abscissa is expanded by a factor Of cos(30) " relative to the ordinate (about 15%). Equivalence. but not identity. is preserved even if both conditions are removed. Consider an equilateral triangle Of unit height and horizontal base. Specify three axes that bisect the three vertices and intersect their respective Opposite sides at right angles. Scale these from O to 1. base to vertex. Beginning with the lower left vertex and proceeding clockwise. assign the axis bisecting the vertex to repre- sent ”it. ”x. or ”1:. respectively. Additionally. specify a vertical axis and a horizontal axis (v and h) originating at the lower left vertex and scaled identically to the other three axes. Adopt the term “base“ to represent a line normal to an axis. which passes through its origin. For any given point. what is the relationship between its Cartesian coordinates (h. v) and its ternary coordi- nates (”x. ”x. 3".r)? It is clear by inspection that v=2°x because their bases are collinear. The distance Of any point to the base of the h axis can be divided into two portions falling inside and outside the ternary plot. respectively. Trigonometric analysis shows that h = (”x) cos(30)’l + (”1) tan(30). Multiplying both sides by cos(30). (h) cos(30) = (’° x) + (2" x) sin(30). Since all 15N atoms occur either in the 29x or 30.: frac- tions. and since only half of the atoms in the 29x fraction are ”N, we can write 15a = (”x) + (”x) (0.5). Noting that sin(30) evaluates to 0.5. (h) cos(30) = l5a and h = ('50) cos(30)" Therefore. a plot of (”x) vs. (”a) cos(30)" is identical to a ternary plot of 23x, ”x, and 30x. The factor cos(30)“ merely scales the abscissa. and is largely irrelevant. The plot ”x vs. 1’a also emulates a valid ternary plot. albeit with a compressed base. Note 2 To illustrate this. we calculated the isotopic character of 1000 simulated gas mixtures by mass balance. Number of pools was randomized on the interval [2.50]. Enrich- ment as well as relative weights of pools were random- ized on the interval [0.1]. When plotted. the isotopic 27 characters of the mixtures clustered around the coordi- nates (1/2, 1/3) for (”a. 2"1). Mean coordinates for all values were (0.497. 0.334). Mean coordinates for 100 000 simulated mixtures were (0.5004, 0.3336). Note 3 In other words. h can approach arbitrarily close to 7. and underestimation remains unchanged. When h = 7. however, e is undefined. The apparent suggestion is that an almost perfectly uniform pool still leads to significant underestimation, when It is very close to y. The point is moot. since flux estimates could hardly be made from such a poorly labeled pool. Anyway, the reader is cau- tioned that this result only arises under the assumptions stated. The proof follows. What is e for a mixture of equilibrium N, samples whose enrichments are uniformly distributed on the in- terval [7, h]? DEW. 7:150 q=nxa=27-272 k=lsap i= E("a) = ('y + h)/2 j=E(” x)=2i-2[i’+(h -y)2/12] k = l-(m ' 2) I [(m - 2Y- 8(4 M7)]"’]/4 m = (i - (1)/(1' - 7) e = (i - 7)/(k - 7) Substitutions j= 2 x [(y + h)l2] - 2[(y + h)2/4 + (h - y)2/12] = (~4h2 — 472 - 4h‘y + 67 + 6h)/6 i-q=(-4h’-47’-4h7+67+6h)/6-(27- 272) = (-4h’ + 87’ — 4m - tn, + 6h)/6 m=(1"“(I)/[(7+h)f2*-7]=(-4h’+87z -4h7 - 67 + 6h)/l3(h - 7)] m-2=[(-4h2+8y’—4hy—6ry+6h)—2(3h - 37)]/[3(h - 7)] = (-4’!2 + 872 -4h7)/l3(h - 7)] q 7=27 272-l(-4h’+87’- 4h7-67+ 6h)/[3(h- 7)]17 = (- 27’ 272’! + 47h’)/I3(h- 7)] (m - 2)2= l('-4h2 + 87’— 4h7)/[3(h - 7)]!2 = (1611‘ +647‘ -48h’7’+32h’7-64h7‘)/l9(h-7)’] (m-2)’-8(q-M7)=(16h‘+647‘-48h’7’+32h’7 - 64h7’)/[9(h - 7)’] - 8(-27J - 27”! + 47W)/I3(h - 7)] = (16h‘ + 161‘ + — 64th - My’hy [9(h - 7)’] = (411’ + 472 - 8h7)’/l9(h - 7)’l [(m - 2)2 - 8(4 - m7)]"2 = (4'!2 + 47’ - 8h7)/[3(h - 7)] = l-(m - 2) I [(m - 2)2 - 8 (q - m7)ll/2]/4 = l-l(-4h2 + 872 - 4h7)/l3(h - 7)]1 I [(4122 + 472 - 8h7)/[3(h - 7)]11/4 1716 SOIL SCI. soc AM 1.. VOL 63. NOVEMBER-DECEMBER 1999 k. = (4h2 - 81’ + 4127 + 4h2 + 47’ - 8h‘y)/[12(h - 7)] = (219 - 7’ - h7)/[3(h - 7)] = (7 + 2W3 k,=(4h’-873+4h-y-4h2-4-f + 3h7)/[12(h - 7)] = 7 For k,: e = (i - 7)/(k - 7) (undefined) For k.: e = (i - 7)/(k - 7) = [(h + 7 - 27)/2]/l[(7 + 210/3] ‘ ‘Yl = [(h " 7W] x {3’0}! - 27)] = [3(h — 7)]/ [40! - 7)] = 3/4 Note 4 Using the symbology in Fig. 4, fractional contribution of the soil pool to the final mix is calculated as dn = (lsaMn '— ”(Rd/(”0P - 150*") where dis the fractional contribution. n is the incubation number. “a is the enrichment Of the sample. A is the initial chamber headspace. M is the final mix. and P is the soil-derived component. The enrichment of P can be found by solving for the intersection Of the two rays as follows: (50050) + C1 = (Sz)("a) + C2 ("0)(s. ' 52) = C2 ’ Cl ”0 = (C2 ' CI)/(Si ‘ 52) where sis slope, C is intercept (by analogy to Eq. [2a]) and the subscripts reference the two incubations. Slope is calculated as 3n = (293m "' ”rm/("am ‘ 150M) where ”x is the mole fraction of singly-substituted mol- ecules. Intercept is calculated as C = ”x... - 50’0".) = ”1A.. - 50’0“.)- ACKNOWLEDGMENTS We thank S.K. Hamilton and EA. Paul for encouragement and direction during various stages of manuscript preparation. We are grateful to Matt Emmons for technical insights regard- ing N-gas analysis. We are indebted to four anonymous review- ers whose comments greatly improved the manuscript. This work is supported by grants from the USDA NR] and TECO Programs. the NSF RTG and LTER Programs, and the Michi- gan Agricultural Experiment Station. REFERENCES Arah. J.R.M. 1992. New formulae for mass spectrometric analysis of nitrous oxide and dinitrogen emissions. Soil Sci. Soc. Am. J. 56: 795—80). Arah. J.R.M. 1997. Apportioning nitrous oxide fluxes between nitrifi- 28 cation and denitrification using gas-phue mass spectrometry. Soil. Biol. BiochenL 291295-1299. Boast. CW.. R.L Mulvaney. and P. Baveye. 1988. Evaluation of nitrogen-15 tracer techniques fordirectmeasmementof denitrifica- tion in soil: 1. Theory. Soil Sci. Soc. Am. J. 52:1317—1322. Bouwman. A.F. 1990. Exchange of greenhouse gases between terres- trial ecosystems and the atmosphere. p. 61-127. In A.F. Bouwman (ed.) Soils and the greenhouse effect. John Wiley & Sons. New York. Brooks. P.D.. DJ. Herman. GJ. Atkins. SJ. Proser. and A. Barrie. 1993. Rapid. isotopic analysis of selected soil gases at atmospheric concentrations. p. 193-202. In LA. Harper et al. (ed) Agriailtural ecosystem effects on trace gases and global climate change. ASA Spec. Pub]. 55. ASA. Madison. WI. Eichner. MJ. 1990. Nitrous oxide from fertilized soils: summary of available data. J. Environ. Qual. 19:272-2m. Focht. DD. 1985. Differences in nitrogen-15 enrichments of evolved nitrous oxide and dinitrogen and the question of a uniform nitrate- 15 pool. Soil Sci. Soc. Am. J. 49786-787. Hauck. R.D..and D.R. Bouldin. 1961. Distribution of isotopic ' in nitrogen gas during denitrification. Nature (London) 191: 871-872. Hauck. R.D.. SW. Melsted. and RE. Yankwich. 1958. Use of N-iso— rope distribution in nitrogen gas in the study of denitrification. Soil Sci 86287-291. Intergovernmental Panel on Climate Change. 1996. J.T. Houghton et al. (ed) Climate change 1995: The Science of climate change. Cambridge University Press. Cambridge. Mosier. A.R.. W.D. Guenzi. and EE. Schweizcr. 1986. Soil losses of dinitrogen and nitrous oxide from irrigated corps in northeastern Colorado. Soil Sci. Soc. Am J. 50:344-348. Mosier. AR. and L Klemedtsson. 1994. Measuring denitrification in the field. p. 1047—1066. In R.W. Weaver et al. (ed) Methods of soil analysis. Pan 2: Microbiological and biochemical Properties. SSSA Book Series. no. 5.x. SSSA. Madison. WI. Mulvaney. RI. 1984. Determination of "N-labeled dinitrogen and nitrous oxide with triple collector mas spectrometers Soil Sci. Soc. Am. J. 48:690-692. Mulvaney. R.L. 1988. Evaluation of nitrogen.15 tracer techniques for direct measurement of denitrification in soil: Ill. laboratory studies. Soil Sci. Soc Am. J. 52:1327—1332. Mulvaney. R.L. and CW. Boast. 1986. Equations for determination of nitrogen-15 labeled dinitrogen and nitrom oxide by mass spec- trometry. Soil Sci. Soc. Ant J. 50360—363. Mulvaney. R.L.. and LT. Kurt; 1982 A new method for determina- tion Of "N-labeled nitrous oxide. Soil Sci. Soc. Am. J. 46:1178-1184. Mulvaney. RJ.. and LT. Kurt; 1985. Reply to “Differences in nitro- gen-15 enrichments of evolved nitrous oxide and dinitrogen and the question of a uniform nitrate-15 pool”. Soil Sci. Soc. Am. J. 49:787. Mulvaney. KL. and R.M. Vanden Heuvel. I988. Evaltntion of nitro- gen~15 tracer techniques for direct measurement of denitrifiation in soil: IV. Field Studies. Soil Sd. Soc. Am. J. 52:1332-1337. Siegel. R.S., RD. Hauck. and LT. Kurtz. 1982. Determination of "N, and application to measurement of N: evolution during denitrifica- tion. Soil Sci. Soc. Am. J. 46178-74. Stevens. RJ.. RJ. Laughlin. GJ. Atkins. and SJ. Prosser. 1993. Auto- mated deterrnination Of nitrogen-IS-labeled dinitrogen and nitrous oxide by mass spectrometry. Soil Sci. Soc. Am. J. 57:981-988. Stevens. RJ.. RJ. Laughlin. LC. Burns. and J .R.M. Arah. 1997. Men- suring the contributions of nitrification and denitrification to the flux of nitrous oxide from soil. Soil Biol Biochcm 29:139—151. Vanden Heuvel. R.M.. R.L. Mulvaney. and KO. Hoeft. 1988. Evalua- tion of nitrogen-15 tracer techniques for direct measurement of denitrification in soil: 1]. Simulation studies. Soil Sci. Soc Am. J. 52:1322-1326. Weier. K.L.. J .W. Doran. J .F. Power. and D.T. Walters. 1993. Denitrifi- cation and the dinitrogen/niacin oxide ratio as affected by soil water. available carbon. and nitrate. Soil Sci. Soc. Am. J. 57:66-72. Chapter 3 A NOVEL METHOD FOR DIRECT DETERMINATION OF NITROUS OXIDE AND DINITROGEN FLUX FROM 15N-LABELED SOIL Summary Field measurements of denitrification gas products are useful for studying both the contribution Of nitrous oxide to global climate change and the loss Of fertilizer nitrogen as nitrous oxide or dinitrogen. Understanding the relationship between N2 and N20 fluxes may lead to process-based models that improve prediction of N gas fluxes. We report here an analytical method that uses a single conceptual approach for independent analysis Of N2 and N20 for the same field incubation. Sequential samples are removed from a chamber cover placed over 15N-labeled soil. A small amount Of each sample is analyzed for the isotopic composition of N2, and the remainder for N20. Each aliquot passes through an ascarite trap to remove water and a LiOH trap to remove 002; N20 is cryogenically focused in a liquid nitrogen trap. Aliquots are then carried by He through chromatographic columns to further resolve interfering masses. Finally, a triple-collector mass spectrometer detects peaks for mlz 28, 29, and 30 (N2) or 44, 45, and 46 (N20). The peaks are integrated and expressed as ratios relative to a laboratory standard. Flux is determined by the change in isotopic character of chamber N2 or N20 with time. Unlike other published methods for N2, 02 is separated chromatographically rather than removed chemically. For N20, this is the first report Of a method that infers flux 29 directly from the change in isotopic character of headspace N20. Since the method is equivalent for the two denitrification products, a subsequent estimate of N20 mole fraction (N20 1 [N20 + N2] ) may be free of a well-known systematic error that attends such methods when gases derive from non- uniforrnly enriched substrate. Introduction The observed rate of increase in the concentration of N20 in the Earth's atmosphere of 0.25% per year has raised concem over its contribution to global warming (IPCC 1996) and to the destruction of stratospheric ozone (Hahn and Crutzen 1982). A significant proportion of the increase in N20 may derive from terrestrial environments (Bouwman 1990), especially as a consequence of agricultural activity (Eichner 1990, Mosier et al. 1998). Denitrification and nitrification are responsible for most N20 flux from soils (Firestone and Davidson 1989) although their relative importance is still unclear (e.g. Stevens et al. 1997). Microbial denitrification in soils (Paul and Clark 1996, Robertson 1999) results in variable proportions of its end-products, N20 and N2. The relationship between denitrifier N20 production and total denitrification is conveniently defined as the N20 mole fraction: N20 I [N20 + N2]. The N20 mole fraction is highly variable spatially and temporally (Letey et al. 1980, Hutchinson and Davidson 1993, Weier et al. 1993). Single estimates of N20 mole fraction cannot be used reliably to relate denitrification and N20 flux (Vinther 1984, Aulakh et al. 1992, Weier et al. 1993) because of the variability in mole fraction. 30 Improved methods of measuring N2 and N20 emissions from soil can foster the understanding of global N20 emissions, loss of fertilizer N from the plant- soil system during denitrification, and the relative contributions of nitrification and denitrification to soil N20 flux. Direct methods for measuring fluxes of both gases are few. Chamber techniques prevail for N20 (Mosier 1989) because its concentration is readily determined by gas chromatography using an electron-capture detector. Dinitrogen can also be quantified by gas chromatography, but owing to its abundance in the atmosphere its concentration is usually not detectably altered during chamber incubations. The acetylene inhibition technique (AIT) is often employed to measure total denitrification gas flux (e.g. Mosier and Klemedtsson 1994) since acetylene blocks reduction of N20 to N2 (Yoshinari and Knowles 1976). Dinitrogen flux can be determined by difference, if a control chamber or soil core is used to measure N20 alone; however, the potential variability between chambers introduces uncertainty, which must be accommodated by adequate replication. Also, AlT may be ineffective for heavy textured soils (Arah et al. 1993). Furthermore, acetylene blocks nitrification at levels lower than those that inhibit N20 reductase, which can cause denitrification to be underestimated because of substrate depletion (Bollmann and Conrad 1997). Finally, some microorganisms metabolize acetylene (Topp and Germon 1986). An alternative method for determining N2 flux is to label soil with 15N03' or ”NH; and analyze a time series of chamber gas samples by isotope ratio 31 mass spectrometry (IRMS) (Hauck and Bouldin 1961, Siegel et al. 1982). This method, known as the 15N isotope dilution technique, is becoming more common (Mulvaney and Kurtz 1984, Mosier et al. 1986, Stevens et al. 1993). One modification involves reducing N20 to N2 chemically before analysis to measure total denitrification (Mosier et al. 1990, Arah et al. 1993, Mosier and Klemedtsson 1994). If total denitrification is measured, the N20 mole fraction can be determined if N20 flux is independently assessed (e.g. Arah et al. 1993). Dinitrogen flux has also been determined by radioactive (‘3N) isotopic labeling (Firestone et al. 1979, Speir et al. 1995) but the radioactive tracer is short-lived and limited in availability. The procedure described here estimates fluxes of N20 and N; from soil independently by detecting changes in isotopic composition of headspace N20 and N2. Data from a field experiment are provided as an illustration. This is the first demonstration of N20 and N2 flux concurrently measured by equivalent methods. Similarity of method and sampling of the same headspace may reduce error in the estimate of the N20 mole ratio. Materials and Methods Overview The method may be summarized as follows. ‘5N03' is added in aqueous solution to soil. A headspace is confined over the soil surface and N20 and N2 are allowed to accumulate (eg. by denitrification). The headspace is sampled using evacuated glass vessels fitted with stopcocks. Samples are 32 purified in the laboratory, and then analyzed by isotope ratio mass spectrometry. 'lsotopic character is defined as the relative abundances of all isotopic masses of a gas (Bergsma et al. 1999). The shift in isotopic character of a headspace gas (either N; or N20) indicates the relative contributions (to the final mixture) of soil-derived gas and atmospheric gas initially present. Changes in isotopic character also indicate the enrichment of the soil N pool undergoing denitrification. The rate of emission can be calculated from the relative contribution of the soil source as a function of elapsed time, chamber area, headspace volume, and initial or final concentration. The initial concentration of N20 or N2 may be assumed, but for N20 it is more accurate (and not very difficult) to obtain a measurement. Vessels and Sampling Pyrex vessels are used to collect headspace samples. The 500-mL volume of the vessels provides sufficient N20 for analysis at ambient concentrations in air. Analysis of N2 is performed on a subsample from each vessel, prior to N20 analysis. The vessels are oblong, with stopcocks at each end. The stopcocks, constructed of glass barrels and pistons, are sealed with VITON O-rings and terminate in 1/ " (6.4 mm) 0.0. glass tubing. Vessels are pre-evacuated to a pressure of less than 1.3 Pa. In the field, the vessels are fitted to a chamber cover using latex tubing. Samples are collected by rapidly opening the chamber-side stopcock for ~10 s. 33 Analysis For isotopic characterization of N2 and N20, samples are processed using a preparation system interfaced to a mass spectrometer (Figure 3.1). Sample vessels are attached to the terminus of the system using 1/ " ID CAJON Ultratorr unions. The system is then evacuated. For analysis of N2, the sample trap is isolated from vacuum; the vessel stopcock is opened for 10 s and sample diffuses into the sample trap (0.5 m x 1116" large internal diameter nickel tubing, coiled). The remaining sample (the majority) is reserved in the vessel for later N20 analysis. For both gases, the sample passes through a 10 cm column of ascarite (10 mm ID.) for removal of water, then to a 20 cm column (4 mm ID.) of anhydrous lithium hydroxide for removal of C02, which has the same mass spectrum as N20. For N20 analysis, the sample trap remains open to vacuum and the complete contents of the sample vessel pass through the chemical scrubbers at a regulated rate. In contrast to analysis of N2, however, the sample trap is chilled with liquid nitrogen to retain N20. The evacuation of the vessel is regulated by an electronic mass flow controller (25 mL min“) to maintain the efficiency of the sample trap. After trapping of either N2 or N20, the sample trap is isolated by pneumatically operated valves (and then thawed, in the case of N20). A stream of He (50 psi head pressure) then carries the contents of the trap onto the column of a gas chromatograph. For analysis of N20, a J.W. Scientific GS-Q column is used to assure separation from C02. A molecular sieve column (Alltech, 8 m by 1/ " O.D., 5A) is used to separate N2 from 002, Ar, 34 and 02 (oven temperature, 50° C). The column terminates at the inlet of the isotope-ratio mass spectrometer. Prior to arrival of a sample peak, the mass spectrometer analyzes a pulse of reference gas (Note 1). The ratios 2/1 and 3/1 are calculated for both the sample and reference by routines which integrate the signals from beams 1,2, and 3 (masses 44, 45, and 46 for N20 or 28,29, and 30 for N2). For N2 analysis, a pressure regulator adds make-up helium (30 psi) into the flow path just before the inlet of the mass spectrometer (Figure 3.1) to dilute the N; for optimal peak size. Because of the low abundance of N20 and high abundance of N2 in air, typical fluxes from an isotopically enriched source (eg. soil N03) to a chamber headspace cause a marked increase in the 15N enrichment of N20, but only mildly perturb the level of enrichment in N2. Therefore, the potential range of isotope ratios is greater for N20 than for N2. The Micromass Prism mass spectrometer used in this study has a second head amplifier that is readily interchanged with the first. Initially these were both configured such that the two minor ion beams were amplified 100 times more than the major beam (resistor values 5 - 108 Q, 5 - 1010 Q, and 5 - 1010 a). This configuration anticipates that the minor beams will be much smaller than the major, which is usually the case for natural abundance measurements and is still the case for our analysis of N2 in our chamber experiments. We modified our second head amplifier with the result that all three resistors were of equal but intermediate sensitivity (1 - 109 (2). This configuration makes no assumptions about the relative strengths of the major and minor beams and is appropriate for 35 enriched gases (>596). Normally all samples in a set are analyzed for N2 before reconfiguring the mass spectrometer for N20. Data Processing The equations of Arah (1992) were modified by Bergsma et al. (1999) to calculate fluxes and estimate enrichments of the gas source. Application of these equations to N20 is completely analogous to their use for N2. However, because of naturally occurring isotopes of oxygen, the molecular fractions “N20 and 46N20 do not strictly correspond to the molecular fractions [“N‘5N0+‘5N14NO] and 15N‘sNO. We have derived equations that express 29(N.,.)O/"‘(N2)O and 3°(N2)0128(N2)O as functions of “NZO/“Nzo, 46Nzo/“Nzo, "OI“O and 18OI"’O; thus the equations for N; can be used directly for N20 (Note 2). Since the equations estimate d (the fraction of mixed gas derived from the soil source), estimation of absolute flux requires some estimate of absolute abundance. If a represents gas from the atmosphere (pre-existing headspace gas) and p represents gas derived from the soil mineral pool, then d=pl[a+p]. [1] Therefore p = d * [a + p] ; alternatively [2] p = da / [1-d]. [3] Since concentration and therefore actual volume of headspace N2 hardly changes during a typical incubation, [a + p] z a and therefore [5] 36 p =- da. [6] For N20 flux, however, Eq. [3] must be used rather than Eq. [6]; a is the average abundance of N20 in the atmosphere (~3.1 - 10'9 L - L") multiplied by chamber volume. Alternatively, concentration can be measured by gas chromatography at the beginning and end of an incubation, so exact values of a (beginning; Eq. [3]) and [a + p] (end; Eq. [2]) are known. In the field experiment described later, similar results were obtained whether Eq. [3] was used with an estimated a, Eq. [3] was used with a measured a, or Eq. [2] was used with a measured [a + p]. Verification Tests were conducted to assess the performance of the mass spectrometer and gas purification system. To determine whether isotopic analysis is influenced by the size of the sample, replicate samples of N20 reference gas were analyzed using the traditional head amplifier at four different intensities for the major beam, covering the range of valid sample sizes for this system (Table 3.1 ). Our mass spectrometer is designed for analysis of samples having isotope ratios near natural abundance. To test for a memory effect (i.e., whether analysis of highly enriched samples potentially biased the analysis of subsequent samples by contaminating some portion of the system with residual 15N) we analyzed a sample of N20 at natural abundance (45l44 only) immediately following analysis of a highly enriched sample. Enriched samples were prepared by mixing natural abundance N20 with 45N20, as described 37 elsewhere in this study. Samples were injected by gas tight syringe. Measured enrichments are reported in 6 (per mil) notation (Table 3.2) to facilitate interpretation: 5=[(RSAM/RSTD)'1]1OOO where R is the ratio of mlz 45 to 44, RSAM is the sample ratio, and Rsm is the ratio of the standard. Since the laboratory standard is identical to the reference gas, a measurement of 0 %o (0 per mil) was expected for the standard gas, and the subscripts refer to the sample and the standard. We tested the response of the mass spectrometer for linearity over a moderate range of enrichments using the traditional head amplifier. Natural abundance N20 and purified 45N20 (>98% 15NMNO) were mixed on a vacuum line. Natural abundance N20 was measured with a calibrated glass bulb (~20 mL) and enriched N20 was measured with a gas-tight syringe (0-10 uL). Gases were frozen into a 1L flask submerged in liquid nitrogen. After thawing, aliquots of mixture could be removed from the flask and analyzed, or the mixture could be refrozen for further dilution or enrichment. Aliquots (~03 mL) were analyzed using the preparation system described above. Results are reported as a calibration line (Figure 3.2). The response of the mass spectrometer was tested for linearity over a wide range of enrichments using the modified head amplifier. Natural abundance N20, purified 45N20, and/or 46N20 were mixed in a flask in ratios of small whole numbers. In order to prevent previously observed fractionation effects at low pressure, flask pressure was increased to ~1 - 105 Pa (1 atm.) by 38 adding He (99.999%). Replicate aliquots were removed from the flask for analysis, with highly repeatable analytical results (Table 3.3). The N20 analysis system was tested for long-term stability, for an effect of the presence of a LIOH column on measured isotope ratios, and for an effect of water in the LIOH column. A mixture of natural abundance N20, 45N20, and 48N20 was prepared in a ratio of 1:1 :2 and analyzed. Three weeks later, more aliquots were analyzed using each of the following: no LiOH column, a hydrous LIOH column, and a column of LiOH dried by heating on a vacuum line overnight (Table 3.4). The between-date comparison is also a test of the stability of laboratory mixtures. Rapid movement of gas through narrow apertures can cause fractionation. We tested whether significant fractionation occurs during filling of the Pyrex sampling vessels described previously. Three vessels were left open in the laboratory for several days, and three were evacuated and filled rapidly at the time and location that the other three were closed. N20 in the vessels was analyzed as described, using the modified head amplifier (Table 3.5). Laboratory Denitrification The equations used for data processing allow calculation of the mean isotopic enrichment of the soil mineral N pool from which evolved N2 or N20 is derived. We conducted a laboratory experiment to verify the accuracy of calculated enrichments. Fresh soil was collected in July 1998 from a wheat plot at the Kellogg Biological Station, Ml, that had not been fertilized at planting. 39 Water-extractable NO; in this soil was ~2.8 pg - g‘1 dry soil. Ten 9 of fresh sieved soil (8% moisture) were added to each of fourteen 600 mL Erlenmeyer flasks fitted with septa. Each flask also received ~1 g steel wool activated with detergent solution (as a sink for trace levels of 02 - Parker, 1955; cited in Kaspar and Tiedje, 1994) and 10 mL of ~1.0 mM sodium succinate. Flasks were fitted with evacuated sampling vessels using one-holed stoppers, then flushed with high purity nitrogen (via the septa, using a source needle and vent needle) and monitored for N20 production. Quantitative consumption of native soil NO; was suggested by cessation of N20 accumulation and confirmed by selective destructive sampling for soil N031 Stock solutions of 99.93 atom percent 15N-KN03 and natural abundance KN03 were prepared and mixed to give secondary solutions of ~0, 10, 20, 40, or 100% 1SN target enrichment. Each flask was flushed with N2, and then received 1 mL of a secondary solution, to deliver ~5 pg N as N031 When the N20 concentration in the flask headspace reached 1-2 ppm, samples for isotopic analysis were collected by opening the vessel stopcocks. Samples were analyzed within two days, plotted as 1"’x vs. 15a (Figure 3.3) and tabulated (Table 3.6). The process was repeated for analysis of N2 production, using 20 g fresh soil per flask (1.4 pg N03'-N - g dry soil"), 20 mL H20, 1 mL of 0.1 M sodium succinate. and 1 mL of 0.1 M secondary solution. Flasks were flushed thoroughly with high purity N2. Four blank vessels were flushed with N2 as a reference. Evolution and subsequent disappearance of N20 in the headspace suggested active denitrification. Anaerobic conditions were confirmed by monitoring headspace 02 by gas 40 chromatography. After several days, headspaces were sampled by opening vessel stopcocks. Destructive sampling of one vessel showed ~2 ug N as N03' remaining (cs. ~1.4 mg added). N2 in sample vessels was analyzed for isotopic abundances. Average enrichments were calculated by the equations of Bergsma et al. (1999), using the average isotopic character of the reference flasks to represent initial headspace (Table 3.6). Field Demonstration In April 1998 a field of winter wheat at Kellogg Biological Station, MI was fertilized with NH4NO3 at a rate of 30 kg N - ha". Fertilizer was excluded from six microplots (0.25 m2) nested within treatment plots (32 m2), three of which were clipped. The microplots received 99% ‘5N-KN03 at a rate of 30 kg N - ha". Aluminum frames (0.0846 m2) were installed in each microplot as bases for gas sampling chambers. A clipped microplot was selected for intensive study. Rain fell sporadically over the two week period following fertilization. Nine one-hour incubations were distributed around these rain events. A lid (30 cm x 30 cm x 14 cm) was placed over the selected frame (sealed with a moat of water). At the beginning and end of each incubation, gas samples were collected for analysis by gas chromatography (GC), infrared gas analysis (IRGA), and mass spectrometry (IRMS). Samples for CC and IRGA were collected by syringe to 3 mL Vacutainers (4 mL overpressure). For MS analysis, pre-evacuated glass vessels were connected to the chamber lid using latex hose and opened at appropriate intervals. 41 To minimize the pressure artifact at the soil surface associated with opening the evacuated vessels, "barostatic" chamber lids were used. Each barostat consists of a resealable polyethylene bag (44.4 um thick, 0.94 L capacity) and a short length of threaded pipe (ID. 7 mm) which attaches one wall of the bag to an internal wall of the lid (using a washer and 0-ring) and vents the bag to the exterior of the lid. The bag can be opened for assembly and resealed for deployment. The pipe can be stoppered, except during sample collection, to guard against leakage due to bag failure or diffusion. Normally two barostats are used per lid, and are pressure-tested prior to each incubation. For the experiment described above, only one unstoppered bag per lid was used. Results Verification The data in Table 3.1 indicate that the measured isotope ratios for N20 reference gas do not vary appreciably with sample size on our mass spectrometer although, as expected, larger samples provide better precision. \M'ten equipped with the traditional head amplifier, the mass spectrometer demonstrates linearity over a large range of moderate enrichments (Figure 3.2) with only a slight bias in favor of the heavy isotope (slope ~1.05). Apparently our system slightly overestimates the enrichment of a natural abundance sample immediately following analysis of a highly enriched sample (Table 3.2). In our judgment, the overestimation is negligible. 42 Agreement of measured and calculated values was usually within 1% for high levels of enrichment using the modified head amplifier (Table 3.3). We could not find evidence that the presence of LIOH or the presence of water in the LiOH significantly affects measured isotope values of N20 (Table 3.4; P > F for the effect of LIOH condition on 45N20 and 46N20: 0.8306 and 0.8989, respectively). Furthermore, measurements are stable across periods as long as three weeks (P > F for the effect of date on 45N20 and 46N20: 0.4102 and 0.5362, respectively). We found no evidence that sample gas is significantly fractionated upon collection using evacuated vessels (Table 3.5). Repeatability was better when gas was collected quickly by opening pre-evacuated vessels (SE: 5.5- 10‘“, 2.4- 1043 for ion beam ratios 2/1 and 3/1, respectively) than upon collection over long time intervals by diffusion of ambient air (1.4 X 10" and 1.9 x 1043 for 2/1 and 311, respectively). Laboratory Denitrification The measured and predicted enrichments for N20 and N2 produced from labeled soil in flasks agreed well (Table 3.6). The predictions are slightly different from the 'target' enrichments because of the assumption of 0.3663% 15N in natural abundance KNOa (i.e., equal to the 15N abundance in atmospheric N2) and a measured 0.07% 14N in stock 15N-KNO3. Despite evidence of statistically significant differences (Student's ttest), absolute differences are small: in all but one case the means of measured enrichments are within 1% (0.01) of the predicted value. These results indicate the lower limit of the 43 accuracy of our method and eliminate systematic error as a possible explanation for unusual field results. In the case of N20, the laboratory denitrification experiment provides a test of an important assumption: namely, that microbial denitrification of a uniform soil N pool results in an equilibrium mixture of masses (that is, N atoms with masses 14 and 15 are paired in a statistically random fashion, such that the abundances of singly-, doubly- or un-labeled molecules in the product pool can be predicted from the enrichment of the substrate; see Note 3). Since no headspace N20 was initially present, sampled N20 was entirely from a soil source with presumably uniform enrichment. Under these circumstances, the sampled gas should be in equilibrium. Figure 3.3 shows nine samples (three replicates at three enrichments) plotted with the equilibrium curve (see Bergsma et al. 1999). Since the data points lie very near the curve, the equilibrium assumption for microbial denitrification appears to be supported. Previous studies have tested the equilibrium character of N20 by comparing 15N content calculated from 45R and 46R (Stevens et al., 1997; 1998a; 1998b). Here, information from 45R vs. 46R is combined to calculate isotopic character, a two-dimensional quantity, which can be compared to the equilibrium curve, a two-dimensional reference. Field Demonstration The labeled soil experiment (April 1998) is a field application of the analytical methods described above. Nitrous oxide and dinitrogen fluxes showed similar temporal trends; however, during the first two days of the 44 experiment N20 and N2 fluxes generally increased and decreased, respectively (Figure 3.4). The associated rise in N20 mole fraction (~20/[N20 + N2], Figure 4.5) is counter-intuitive: accumulation of soil water due to precipitation should decrease soil redox status and favor production of N2. Enrichment of the soil pool undergoing denitrification was estimated non-destructively by analysis of the shift in isotopic character of headspace gases during each incubation (Hauck and Bouldin 1961, Bergsma et al. 1999). The apparent enrichment, based on N20 isotope data, dropped from 0.82 to 0.72 during the 4 day period, suggesting that significant nitrification was occurring. Soil pool enrichment estimated from N2 data was considerably more variable, perhaps reflecting lower sensitivity for N2 (see precision and detection limits, below). Agreement of mass spectrometric and gas chromatographic flux estimates for N20 was good (Figure 3.4), with 5 values of the MS/GC ratio between 94% and 107%, and three values between 66 and 77%. Differences could be due to analytical error, to non-uniform enrichment of the soil mineral N undergoing denitrification (cf. Mulvaney et al. 1988), or to other sources of N20 production (e.g. Robertson and Tiedje 1987, Stevens et al., 1998b). Precision and Detection Limits There is not as yet a universally recognized method for calculating detection limits when denitrification fluxes are quantified from shifts in isotopic character. However, means and standard deviations of ion ratios 2/1 and 3/1 (the primary output of isotope ratio mass spectrometers) are easily compared (Table 3.7). Precision for 211 (N2) in this study is comparable to that in previous 45 reports, and for 3/1 is less. Precision for N20 and N2 in this study are similar. Mean 3/1 ratios for N2 are substantially higher than those reported by others. This is probably an artifact of high mass 30 background due to the formation of NO in the spectrometer source (Stevens et al 1993). Post hoc correction of data represented in Figure 3.4 using a representative calibration curve lowered the mean N2 flux by ~4%. Detection limit has been defined as three times the standard deviation of the blank (Miller and Miller, 1988, cited in Stevens et al. 1993). Interpreting this to mean that ratios 2/1 and 3/1 in end-of—incubation samples must exceed initial ratios by three standard deviations (as tabulated), we calculated minimum detectable flux independently for 2/1 and for 3/1, and then selected the larger of the two. Detection limit is a function Of analytical precision, chamber volume, chamber area, enrichment of the soil pool, and duration of the incubation. For the field demonstration above, headspace was ~14 L, area was 0.0846 m2, enrichment averaged 0.77 15N, and duration was ~1 h; consequently our estimated detection limit for N2 is 216 g - ha'1 - d". This is larger than reported estimates of 5 g - ha'1 - cl'1 (Siegel et al. 1982) and 12 9 ~ ha'1 - of1 (Stevens et al. 1993), but not directly comparable because sampling configurations were different. For N20, our detection limit was 2.72 - 10“ 9 N20- N - ha'1 - d", which is equivalent to a minimum detectable change in headspace concentration of 6 - 10'12 L - L". Stevens et al. (1993) report a minimum detectable change in headspace concentration of 2.1 - 10'6 L - L'1 for N20. For N20 analysis by gas chromatography, assuming a nominal CV of 1% 46 (perhaps optimistic), minimum detectable concentration change is ~ 1 - 10‘8 L- L". Discussion Our results provide evidence that flux of N20 and N2 from 15N-labeled soil can be reliably measured by isotopic analysis of chamber headspace gases using mass spectrometry to directly analyze N20 and N2 after chemical, cryogenic, and chromatographic purification. We verified that our analytical technique gives reasonably accurate, precise, and unbiased results. The field data show that our analytical technique can be an integral part of a complete experimental system. Although similar methods exist for N2, flux of N20 has never been analyzed in this way before. Usually N20 is reduced to N2 and analyzed in a mixture with sample N2 (Mosier et al. 1990) or laboratory standard N2 (Mulvaney and Kurtz 1982). Our method for N20 gives lower detection limits and coordinates well with the 15N dilution technique for N2 because it involves the same assumptions regarding homogeneity of the soil mineral pool. An isotopic method for measuring total N; flux from soil has been available for three decades (Hauck and Bouldin 1961) but it has only recently been exploited (e.g. Siegel et al. 1982, Mulvaney and Kurtz 1984, Stevens et al. 1993). Our technique for estimating N2 flux differs in some details from those published. The conventional approach for purifying N2 involves removing condensibles (e.g. C02) in a cold trap and removing oxygen with chemical traps or hot copper (e.g. Siegel et al. 1982, Mosier and Klemedtsson 1994). Boyd et al. (1994) claim superior convenience and efficiency for a mixture of 47 CaO granules and Cu for purifying nanomole quantities of N2 (mainly removing C02 and 02, respectively). We remove water and the majority of C02 using chemical and cold traps, and separate 0;, Ar, and trace 002 and CO from N2 on a molecular sieve column during continuous flow mass spectrometry. Our purification system is efficient and inexpensive, and requires very little maintenance. The conceptual approach to the flux calculation is in principal the same as originally proposed (Hauck and Bouldin 1961) and uses the equations and notation of Arah (1992) as modified by Bergsma et al. (1999). In early studies of N2 flux, 15N-IaoeIed fertilizer was added to soil and evolved gas was evaluated for mass ratio 29/28 (e.g. Rolston et al. 1978, 1982). One limitation of this approach is that only the N2 flux derived from fertilizer is determined, and not any flux derived from native soil mineral N (see Mosier and Klemedtsson 1994). Another way of describing this limitation is that the isotope ratio 29128 does not completely characterize N2 when the sample is a mixture of atmospheric N2 and soil-derived N2 of different enrichment; the ratio 30128 must also be measured (Hauck et al. 1958) since such mixtures are not in isotopic equilibrium. Hauck and Bouldin (1961) showed that measurement of the ratio 30I28 allows calculation of the average enrichment of the N pool experiencing denitrification, which in turn allows a calculation of total N2 flux (whether from native soil N, label, or both). The principle assumption of this approach is that flux derives from a single, uniformly labeled pool. 48 The N20 molecule, like the N2 molecule, can be singly or doubly substituted with respect to 15N (giving masses 44, 45, and 46 instead of 28, 29, and 30). Therefore, it is possible, as we have shown, to estimate flux of N20 by measuring shift in isotopic character (as for N2). This has not been performed until now, probably because isotope ratio mass spectrometers, although more precise than gas chromatographs, may not have been sufficiently sensitive. In the past, as much as 0.1 mg N was needed for analysis, as compared to our 0.2 pg; see Mulvaney and Kurtz, 1982; 1985). When measuring soil denitrification by chamber methods, it is usually of interest to determine both N2 flux and N20 flux. Five methods have been previously reported for determining N20 flux from soil otherwise labeled for N2 determination. Some researchers measure N2 flux using the shift in isotopic character ("Hauck technique") and N20 flux by gas chromatography (Mosier et al. 1986). Others measure N20 by GC and [ N20 + N2 ] by the Hauck technique, reducing N20 over hot copper and thereby letting it mix with sample N2 before analysis (Mosier et al. 1990, Arah et al. 1993). Third, some researchers calibrate their isotope ratio mass spectrometer so that the concentration of N20 can be derived from the sum of ion currents above baseline for masses 44, 45, and 46 (Stevens et al. 1993). A fourth approach is to trap N20 from the sample, mix it with a known quantity of standard N2, and then reduce the N20 to N2 (Mulvaney and Kurtz 1982, Mulvaney and Kurtz 1984, Mulvaney and Vanden Heuvel 1988). Finally, a linear mixing equation can estimate N20 flux if the 49 enrichments of the soil mineral N, the label, and the evolved N20 have been independently determined (Brooks et al. 1993). Our method for quantifying flux of N20 has advantages relative to the five methods outlined above. The first three methods of flux estimation depend on a change in headspace concentration, a measure of net flux, whereas isotope methods generally measure gross fluxes (see discussion in Hart et al. 1994). Net positive flux of N20 is only the same as total flux if significant quantities of N20 are not consumed during incubation; consumption of N20 has not been widely tested, and may be especially important under soil covers, within which N20 concentration (and therefore likelihood of consumption) is increasing. The fourth and fifth methods are isotopic approaches, but assume negligible background N20 in the chamber headspace. The atmospheric background is indeed small (~31 - 10'9 L - L“) but negligible only if concentration due to flux is several orders of magnitude greater. Clearly a method that does not assume negligible background is potentially more sensitive. Our method requires larger samples (500 mL) and longer analysis times (40 min/sample) than those reported for automated methods (Brooks et al. 1993, Stevens et al. 1993) but sensitivity is greatly enhanced, no destructive sampling of soil is required (in contrast to the fifth method above), and gross flux rather than net flux is estimated. In addition to the general advantages above, our method for determination of N2 and N20 carries specific advantages for determination of the N20 mole ratio during denitrification. (1) Because both N2 and N20 are 50 measured by shifts in isotope ratios rather than changing concentrations, both flux estimates represent gross rather than not flow from the soil surface, resulting in an internally-consistent ratio. (2) When the soil mineral pool is not uniformly enriched, flux of N2 is underestimated (see Boast et al. 1988, Vanden Heuvel ef al. 1988, Arah 1992, Bergsma et al. 1999). To the extent that the same mineral N pool is the source for both N20 and N2, proportional underestimation of both will be similar when equivalent methods are used. Therefore the calculated ratio of the two fluxes will be relatively independent of this source of error. (3) Unlike the acetylene inhibition technique, the method described here and the others outlined above allow fluxes of both N2 and N20 (and therefore the N20 mole ratio) to be determined from a single experimental unit, thus reducing statistical uncertainty due to natural variability among control and experimental units. The statistical need for replication is reduced. Mile any given method for determining N20 flux may have its advantages, important gains are often made by coordinating multiple methods. Mulvaney (1988) measured N20 by GC and by MS to test the assumption that N20 was derived from a uniformly labeled pool of soil N. Arah et al. (1993) used both GC and MS methods to test the suitability of acetylene inhibition for measuring denitrification in heavy-textured soils. Stevens et al. (1997) employed concentration and isotope distribution data from mass spectrometry to examine relative contributions of nitrification and denitrification to N20 flux. As noted above, GC values that are lower than MS values may constitute evidence of concurrent production and consumption of N20 by soil. As the 51 precision and convenience of these methods improve, so does the potential for characterizing fundamental controls on denitrification dynamics. Such improvements should lead to a more comprehensive perspective on regional and global N budgets, in addition to local insight regarding soil N cycling. Notes Note 1. For N20 analysis, a well-characterized laboratory standard is used. For N2 analysis, ~150 pL of 3°N2 are mixed with ~20 mL of laboratory standard to improve the stability of 30l28, since natural abundance N2 has very little mass 30. The absolute abundances of m/z 28, 29, and 30 are measured manually: voltage shifts are used to sequentially place each mass in the same collector for three replicate cycles. After correction for background readings, the ratios of the means are calculated and used later to interpret the analytical data for the samples analyzed with this reference. Note 2. Let x, y, and 2 represent the fractional abundances of 2E’(N2)O, ”(N2)O, and 30(N2)O. Let r, s, and t represent the fractional abundances of “N20, 45N20, and 46N20. Let c, d, and e represent the fractional abundances of 160,170, and 180. That the mass spectrometer measures slr and tlr, although ylx and zlx are of interest. Now, r=xc s = yc + xd and t = 20 + yd + xe. Then s/r = [yc + xd] /[xc] = ylx + d/c and 52 t/r = [20 + yd + xe]/[xc] = zlx + [yd]l[xc] + elc. Thus, ylx = slr - d/c and zlx=tlr-y/x*dlc-elc. In terms of simple mass ratios, (”mac/”(Nam = [‘5N20/“N201 - I1 701‘ 601 and I°°(N2)0/28(N2)Ol = Dime/“N201 - [29(N2)0/28(N2)OI * ["01‘601 - [mo/”0]- ln shorthand, 29R = 45R - ”R and NR = 46R _29R17R _ 18R Literature values for ['7OI'GO] and [180/160] are used. Note 3. Use of the term “equilibrium” in discussions of the 15N isotope dilution technique is related to its classical use in discussions of chemical reactions. The reaction ”N2 +3°N2 <-> 2I”N21 proceeds spontaneously in both directions, but at a negligible rate because of high activation energies. Thus, the different molecular-mass fractions of an N2 sample do not normally equilibrate, except perhaps at geologic time scales. The reaction rates increase, of course, if the sample is heated (> 1000 ° C) or in the presence of some other form of energy (e.g. microwave). At chemical equilibrium, the relative concentrations of products and reactants no longer change. The relative proportions of the three mass fractions is then approximately that which is predicted from the composite 15N abundance of the 53 whole system (neglecting fractionation effects). Any sample meeting this criterion may be said to be in equilibrium, even if no “equilibration” has occurred. 54 Table 3.1. Test for stability of measured ratios with varying sample size. Variations were simulated by adjusting the strength of the major beam for analysis of N20 laboratory standard gas. 211 and 3/1 refer to mlz ratios 45l44 and 46/44, respectively. SE is standard error, n is number of samples. major beam (Amps) 2/1 mean SE 311 mean SE n 2.60 . 10'” 7.858 - 10“ 3.2 - 10'5 2.109 - 10“ 2.9 . 10"?— 4.43 - 10'9 7.853 . 10'3 2.0. 10'6 2.102 . 108 2.0- 10‘ 4 1.3010:8 7857-103 3010'7 2106-103 8.5- 10'7 3 1.79 - 10“ 7.858 . 10*3 4010'7 2.106 - 10*3 4.2 - 10'7 3 55 Table 3.2. Test for a memory effect during analysis. Highly enriched N20 was prepared by mixing various quantities of labeled and unlabeled N20. Analysis of laboratory standard N20 immediately followed each analysis of enriched gas. Since the laboratory standard is the same as the reference gas, a value of 0 %o is expected if there is no memory effect. date enriched sample (960) standard sample (%o) 1 223-103 18 1 532- 103 81 1 577 - 1o3 59 2 540.103 78 2 517 . 103 43 3 414- 103 17 3 279.103 37 56 Table 3.3. Summary of analyses of high-enrichment laboratory mixtures of N20. 44:45:46 represents the mixing ratio of natural abundance N20, 45N20, and “N20. Predictions '58 and 45N20 adjust for gas purity. 15a is the atom fraction of 15N in the sample, consistent with the notation of Arah et al. (1992). 45N20 is the molecular fraction of mass 45 in the sample. Measured 15a and “N20 are means for all samples where number of samples n is more than 1. 44:45:46 predicted 15a measured 15a predicted 45N20 measured “N20 1 :0: 1 0.5022 0.5057 0.0040 0.0045 1 :1 :0 0.2499 0.2658 0.4937 0.5085 1:1 :1 0.5013 0.5029 0.3283 0.3278 1 :1 :2 0.6265 0.6267 0.2460 0.2532 57 Table 3.4. Test for effects of the LiOH water trap, water in LiOH, and time on measured enrichment of a 1:1 :2 mixture of natural abundance N20, 45N20, and 46N20. 'Hydrous' refers to hydrous LiOH used as a chemical trap, 'dried' refers to hydrous LiOH dried on a vacuum line overnight. 45N20 and “N20 represent the mean calculated molecular fraction for N20 of masses 45 and 46. SE is standard error. Date LiOH 45N20 SE 46N20 SE 1 dried 0.25318 0.0001 1 0.5001 1 0.00009 2 dried 0.25326 0.00026 0.49961 0.00048 2 hydrous 0.25320 0.00018 0.49980 0.00038 2 none 0.25310 0.0001 1 0.49998 0.00028 58 Table 3.5. Test for an effect of collecting sample gas quickly using pre- evacuated vessels. Air samples were collected by diffusion into open vessels or by rapid filling of pre-evacuated vessels when stopcocks were opened. Samples were analyzed for N20 using the unmodified head amplifier. 2/1 and 311 are mlz ratios of 45l44 and 46/44, respectively. Means 1: standard errors are reported; n is 3. P > N is the significance level for Student's ttest of differences in means. Least significant number (LSN) is the smallest number of samples needed to demonstrate significant differences in means at a confidence level of or = 0.05. Ratio Diffusion Pre—evac P >|t| LSN 2/1 7.835 - 10‘3 :l: 5.5- 10‘6 7.830 - 10‘3 i 1.4- 165 0.3965 28 3/1 2133108124- 10‘ 2129- 10311910“5 0.2602 16 59 Table 3.6. 15N enrichment of soil NOg‘ pools for laboratory denitrification experiments, calculated and measured . Calculations are based on the enrichments and mixing ratios of natural abundance KNOa' and highly enriched KNOa' (99.93 atom %). Measurements use the equations of Bergsma et al. (1999). 'n' is number of samples. Means :l: standard errors are reported for separate experiments measuring N2 and N20. P>|t| is the significance level for Student's ttest. calculated N2: measured P>|t| N20: measured P>|t| n 0.1032 0.1072 :1: 0.0053 0.531 0.1026 :l: 0.0003 0.210 3 0.2028 0.1969 :l: 0.0031 0.199 0.2007 :l: 0.0013 0.0336 3 0.4019 0.3827 :l: 0.0007 0.001 0.3976 i 0.0012 0.068 3 0.9993 0.9692 :l: 0.0020 0.042 (ratio out of range) 2 60 Table 3.7. Comparison of spectrometer precision by study. Precision is reported for analysis N2 or N20. 2/1 refers to mlz 29l28 or 45l44. 3/1 refers to 30/28 or 46/44, except as noted. STD is standard deviation and n is number of samples. Citation gas ratio mean STD CV n Siegel et al. 1982 N2 2/1 7.35 - 10" 9.3 - 10" 1.2- 10“ 15 Siegel et al. 1982 N2 3/1 *1.34- 10'5 2 7 - 10‘7 2.0- 10‘2 15 Stevens of al. 1993 N2 211 3.50 - 10'3 5.3- 10" 1.5- 10*3 7 Stevens et al. 1993 N2 3/1 1.01 - 10'5 5.3 - 10” 5.3 - 10“ 7 this study N2 2/1 7.22 - 10'3 4.8- 10'7 8.7- 10“ 12 this study N2 3/1 1.55 - 10'4 1.2- 106 7.7- 10*3 12 this study N20 211 7.83 - 10'3 2.4 - 1043 3.1 - 10“ 3 this study N20 311 2.13 - 10'3 3.3- 10‘6 1.6- 10" 3 *Siegel et al. measured 30/[29 + 28]. 61 Figure 3.1. Gas purification system connected to mass spectrometer. Valves are configured such that during purification, He bypasses the cold trap and travels onto the 60 column, while sample passes into the cold trap (N2) or through the cold trap and then to vent via the mass flow controller (N20). During analysis, helium passes through the cold trap and onto the GC column, while the sample vessel remains open to vacuum via the mass flow controller. During N2 analysis the sample trap is not chilled. Different columns for N2 and N20 are used. Interface helium is used only for N2 analysis. The Penning valve passes about 10% of the gas stream to the spectrometer. 62 :58? 0% 52.50 30.38.: MERE: ostousw 75:98on $9: . oz? 559... 5.8 96.» .35 5:28 838.9395 .3 85?. 3.0.652 :o 9:393 a...“ lots; ) a _cmm¢> .l: :o: £88 288 P 05>Mw 0 ‘l K 63 Figure 3.2. Measured vs. calculated molecular fraction of 45N20 for moderate enrichments of N20, prepared volumetrically in the laboratory from enriched and unenriched standards; analyzed with the unmodified head amplifier. 64 an ego: c2895. 5300.9: ooze. 62930.8 mod mod cod mod No.0 5.0 oo.o * u 4_ + 4. cod i Pod i No.0 w 6 B S n m D. . 9 seamen“: zN . - . .. O 880 x887) reed w a. O m B I 3.5 u. m m. m .- 8.5 I mod 65 Figure 3.3. Isotopic character of nine N20 samples from laboratory denitrification of prepared ‘5N03' label, plotted against the equilibrium curve. As expected, these samples are in equilibrium. Vertical bars intersect the curve at predicted enrichments, open circles represent measured values. 66 omd ovd cozom: Eoum me. and 0N6 L o_,.o .1- 3 can: cod cod qt— — A i 3.6 r omd J i 00.0 1 r ovd uonoerj JBII’IOOIOLU Xe: l omd cod 67 Figure 3.4. N2 and N20 fluxes with cumulative precipitation for nine field incubations over the same highly labeled plot. 68 (uro) uouettdgoeld eAltejnwno so 8:9“. 98 6.2.4 as} 855 of 938 So _ _ _ 58.0 ms. so oaz 5295605 8N - Sod m 8.... u m m w. 8.0 - .z - 5.0 Mn N Se - ms. 3 z 00.0? rd 69 Figure 3. 5. Estimated soil enrichment, comparison of MS and GC, and N20 mole fraction for nine incubations represented in Figure 3.4. 70 no 26:. 28 85:. 829 >23 >22; 518 w 4 n \9 -914 o 0. ............................... o. .\0....¢.i° 22 + osz m oaz -r No B m I to m. a m... 0 -- 90 w oaz .3 2s =8 m i w . 9 11 m 0 W1 N2 E 29 =8 o -- a 71 Chapter 4 NITROUS OXIDE MOLE FRACTION DURING DENITRIFICATION IN SOIL: RESPONSE TO RECENT MOISTURE HISTORY VARIES AMONG ECOSYSTEMS Summary Very little is known concerning the effects of recent moisture history (antecedent moisture regime) and of ecosystem differences on the relative proportion of N20 and N2 produced during denitrification (N20 mole fraction). We conducted laboratory incubations of sieved soil from cropped and successional ecosystems under two moisture histories. The soils were pedogenically identical but had been managed differently for the past decade. Fresh soils were air-dried, re-packed, and amended with nitrate, glucose, and sufficient water (about 85% water-filled pore space) to stimulate denitrification. One set of incubations received 80% of prescribed water 2 d before incubation and the remaining water at the start of the incubation; the other set of incubations received all water at the start of the incubation. Production of nitrous oxide and dinitrogen was estimated using acetylene inhibition (measuring resultant N20 by gas chromatography) and also by 15N isotope dilution (characterizing headspace samples by isotope ratio mass spectrometry). The response of N20 mole fraction to recent moisture history varied by ecosystem. Mean N20 mole fractions (N20 l[ N20 + N2] ) measured using acetylene inhibition were 0.36 and 0.90 for cropped pre-wet and control soils, respectively, and were 0.34 and 0.33 for successional pre-wet and 72 control soils. Isotope data for N20 showed that in most cases, the soil N03' pool undergoing denitrification was nearly uniform in its N isotopic composition. 15N isotope dilution consistently gave estimates of N2 production that were about one-third of estimates by acetylene inhibition, suggesting that most of the N2 came from an unlabeled source. Explicit recognition of ecosystem differences in response of N20 mole fraction to recent moisture history may improve modeled estimates of global N20 flux. Introduction The proportion of denitrification end product that is nitrous oxide (N20 mole fraction) is an important aspect of the global budget of N20, a significant greenhouse gas (IPCC 1996) and regulator of stratospheric ozone (Hahn and Crutzen 1982). A major source of N20 is microbial denitrification in soil, which produces dinitrogen and nitrous oxide in proportions that vary widely (Tiedje 1988, Robertson 1999). Many factors are recognized as influencing the N20 mole fraction, including soil moisture, nitrate or nitrite concentration, pH, aeration, temperature, carbon availability, enzyme status, and moisture history (Colbourn and Dowdell 1984, Sahrawat and Keeney 1986, Firestone and Davidson 1989, Arah and Smith 1990, Bouwman 1990, Aulakh et al. 1992, Hutchinson and Davidson 1993). However, few experimental studies have considered the influence of moisture history on the nitrous oxide mole fraction (e.g. Dendooven and Anderson 1995, Dendooven et al. 1996) and we know of none that has looked for an interaction of moisture history and ecosystem management history. Since most N is probably lost from soils during brief 73 periods following irrigation or rainfall (Smith and Tiedje 1979, Sextone et al. 1985, Rolston et al. 1982, Mummey et al. 1994, see also Davidson 1991 and references therein) dependency of the N20 mole fraction on short-term soil moisture history could have large consequences for the relationship between nitrous oxide production and total denitrification. The study of N20 mole fraction is hampered primarily by the difficulty of analyzing N2 flux from soil. In this study, we estimated nitrous oxide mole fraction for incubations of soil from two ecosystems (row crop agriculture and early native succession field) and for two recent moisture histories in a factorial design. Our primary objectives were to determine the effect of recent moisture history on N20 mole fraction and to determine whether the effect can vary among ecosystems. A secondary objective was to compare the use of the 1"SN isotope dilution method with the acetylene inhibition method for estimating N20 mole fraction. Materials and Methods Soil collection and processing Soils (Kalamazoo/Oshtemo soil series; Austin 1979) were collected from the Long-term Ecological Research site at the W. K. Kellogg Biological Station, Hickory Corners, Michigan, 42° 24' N, 85° 24' W. The soils at this site are Typic Hapludalfs (fine-loamy, mixed, mesic) derived from glacial till that was deposited about 10,000 years ago. The ongoing LTER experiment at KBS is a randomized complete block design with 6 replicate blocks and 7 treatments on the main site, for a total of 42 1-ha plots. We sampled three replicates of two treatments: a high-input corn-wheat-soybean rotation and a 74 native succession treatment last plowed in 1988. The annual cropping system is tilled and receives conventional applications of fertilizer and pesticides. The successional treatment is managed only by occasional burning and/or removal of woody biomass. Soil was collected in December 1999 from blocks 4-6 on the LTER site. For each of 6 plots (two treatments x 3 replicate blocks), 4 soil cores (2 cm diameter by 16 cm depth) were collected at each of 5 semi-permanent sampling stations. Soil was bulked by plot, sieved (4 mm mesh), air-dried for several weeks (to about 1% gravimetric moisture), and stored in bags at room temperature (“stock soil”) until the start of the experiment. Due to analytical limitations, soil from the two ecosystems was tested on separate dates, four weeks apart. Stock soil was tested on both dates for nitrate and ammonium availability by KCI extraction (1M) followed by analysis using an Alpkem auto- analyzer. Experiment and treatments We incubated soil from each ecosystem for 24 hours in 1 L glass mason jars. Each jar received 150 g dry soil from one of three field-level replicates, packed to a volume of 125 mL (:l: ~5%) for a target bulk density of 1.2 g dry soil - cm‘a. Each jar within a replicate set was assigned to one of two moisture histories (pre-wet or control) and one of 4 sampling strategies ('5N- labeled soil, unlabeled soil, acetylene-amended soil, or soil for mineral N analysis). Two additional jars were established without soil to serve as blanks for gas analysis, for a total of 26 jars per ecosystem. 75 All soils received 9.75 mg of KN03 (about 9 pg N03'-N - g dry soil“), 20 mg glucose (about 53 pg C - g dry soil ‘1), and 56.6 mL deionized water (for a target water-filled pore space of ~85%). “Pre-wet” soils received 80% of their prescribed water 48 hours before the start of the incubation, with remaining water reserved as a vector for nitrate and glucose. “Control” soils received all of their water and nutrients as a single solution at the start of the incubation. Blank jars received only 56.6 mL water (no soil). All soils received nitrate and glucose immediately prior to the start of the incubation. Solutions were delivered as a slow trickle down the edge of a tipped jar to minimize soil disturbance and air entrapment. The delivery method produced a wetting front that moved laterally across the soil within about 15 minutes. The labeled soils received 9.84 mg K15N03, the molar equivalent of the 9.75 mg K“N03 received under the other three strategies. The acetylene jars received 80 mL C2H2 at the start of the incubation for a 10% headspace concentration, known to inhibit nitrous oxide reductase in these soils (Robertson and Tiedje 1987). All jars were fitted with air-tight lids; rubber septa and CAJON UltraTorr unions (custom o-ring seal) were added as necessary for syringe sampling and sampling to Pyrex vessels (0.5 L, pre-evacuated, with stopcocks) for 15N analysis. Sampling and analysis The mineral-N soils were sampled destructively for analysis of nitrate and ammonium concentrations about 2 hours after the start of the incubation (10 g soil, dry weight equivalent, extracted in 100 mL 1M KCI). N20 76 concentrations in other jars were measured by gas chromatography at 0, 6, 12, and 24 hours after the start of the incubation. At the close of the incubation (24 h) gas samples were collected for analysis by isotope ratio mass spectrometry (”N-labeled and unlabeled treatments). The vessel stopcocks were opened for about 10 s and then sealed. Analysis was performed within two weeks, using methods described elsewhere (Bergsma et al. submitted). For N20, mlz ratios 46I44 and 45l44 were measured. For N2, ratios 30/28 and 29l28 were measured. Equations for estimating the 15N enrichment of the soil mineral N pool and the fraction of headspace gas derived from the soil mineral pool (d) require initial and final measurements of isotopic character (Arah 1992, Bergsma et al. 1999). Constraints of the experiment allowed only a final sampling. Therefore, each labeled sample was paired with its corresponding unlabeled sample to represent final and initial conditions, respectively. An advantage of this pairing is that it controls for (slight) biological and mechanical artifacts that could influence isotopic character under the experimental conditions described above. To guard against bias due to a label effect, N2 flux was calculated as [NZOmtyhm - Nzohbejed] for comparison to isotope data, but as [Ngomjybm - NgOummw] for all tests of treatment effects. Differences among treatment means were tested for statistical significance by ANOVA, using JMPIN software version 3.1.5 (Sall and Lehman 1996). 77 Results Experimental design We tested whether mineral N availability in stock successional soil changed during the interval between the date on which cropped soil was incubated and the date on which successional soil was incubated. Mean extractable nitrate in stock soil from the successional ecosystem dropped slightly between the two experimental dates, from 2.2 1 0.4 pg N - g dry soil '1 to 1.7 1 0.4 pg N - g dry soil " (statistically significant at the 0.05 level). Mean extractable ammonium was unchanged (15.2 1 3.2 pg N - g dry soil '1 and 15.2 1 2.0 pg N - g dry soil "). We tested whether pre-wet and control soils differed in available mineral N two hours after the start of incubation. Mean extractable nitrate in mineral-N jars was 2.3 pg N - g dry soil '1 lower for pre-wet cropped soils than for controls (19.5 1 2.5 and 21.8 1 2.2 pg N - g dry soil ’1 respectively). For successional soils, pre-wet soils were only 0.8 pg N - g dry soil '1 lower (9.0 1 3.3 vs. 9.8 1 0.1). For both ecosystems, mean extractable ammonium was sharply higher in pre-wet soils relative to controls: 3.1 pg N - g dry soil '1 higher for cropped soils (4.8 1 0.4 vs. 1.7 1 0.3, respectively) and 12.7 pg N - g dry soil '1 higher for successional soils (21.4 1 2.3 and 8.7 1 0.7). We tested whether N20 production in labeled soils differed significantly from N20 production in unlabeled soils. Final N20 concentration for each labeled jar was divided by final N20 concentration in the corresponding unlabeled jar. On average, the labeledzunlabeled ratio was 0.98. However, the 78 average differs by ecosystem: 1.12 for cropped soil and 0.85 for successional soil. For this reason, production of N; by acetylene inhibition was calculated from concentration data for unlabeled jars for tests of treatment effects, but from labeled jars for comparison of acetylene inhibition with 15N isotope dilution. N20 mole fraction The time-course of change in concentration of N20 in all jars (except blanks, where change was negligible) is summarized in Figure 4.1. N20 accumulated more rapidly in the acetylene-amended jars than in the others. Since acetylene inhibits the reduction of N20 to N2, production of N2 by the un- amended soils may be inferred by difference (Yoshinari and Knowles, 1976). Initially, total denitrification (represented by N20 produced in acetylene jars) was less for control soil than for pre-wet soil, but was indistinguishable by the end of the incubation. For both ecosystems, pre-wet soil produced N20 immediately and steadily throughout the incubation. Successional control soils did not respond differently than successional pre-wet soils. However, the response of the cropped control soils was almost identical to the response of the cropped acetylene-amended soils, diverging only slightly by the end of the incubation. The near-identity suggests that cropped control soils produced significantly less N2 than soils for the other three combinations of ecosystem and recent moisture history. Final N20 concentrations were used to calculate total denitrification, N20 production, and N2 production (by difference) as pg N - g dry soil’1 (Figure 4.2). 79 Results were analyzed by ANOVA. When the block effect is included in the model, successional soils (grouped) had significantly higher total denitrification than cropped soils, at the 0.05 confidence level. Cropped controls had significantly higher N20 production and lower N2 production (p = 0.02 and 0.05, respectively). Although the difference in total N2 production for successional pre-wet and successional control soils was not significantly different, review of time-series data for the entire incubation period shows that pre-wet soils consistently led their corresponding controls in N2 production by a small margin (data not shown). Nitrous oxide mole fraction was calculated as N20 I [ N20 + N2 ] for each combination of ecosystem, recent moisture history, and block. Analysis of the results (ANOVA) is summarized in Table 4.1. Mean N20 mole fractions were 0.36 and 0.90 for cropped pre-wet and control soils, respectively, and were 0.34 and 0.33 for successional pre-wet and control soils. In the analysis of variance, cropped soils showed a strong effect of recent moisture history, while successional soils did not. The difference accounts for the highly significant interaction term (ecosystem x history: p = 0.012) that results when the entire model is considered. IsotOpic data For N; or N20, isotopic character has been defined as the relative proportions of the three (primary) molecular fractions (2°N2, 29N2, 30N2 or 2a(N2)O, ”(N2)O, 3‘°(N2)O; see Bergsma et al., 1999). Isotopic character is conveniently represented by a plot of 29x vs. 15a: that is, the 29(N2) molecular fraction vs. the 80 composite 15N atomic fraction. Figure 4.3 gives the isotopic character of N; at the end of the 24 hour incubation for unlabeled (lower left) and labeled (all other) jars. Each unlabeled/labeled pair (one line segment) represents one incubation unit. There is strong tendency toward colinearity for pairs of line segments representing paired pre-wet and control incubations, indicating similar soil nitrate enrichment (consistency within replicate field plots). Lengths of segments representing cropped controls are generally shorter than those for cropped pre-wet treatments, indicating smaller flux. Successional control and pre-wet segment lengths (and therefore fluxes) are similar for replicates 4 and 5, but not 6. Figure 4.4 shows the final isotopic character of N20 for unlabeled and labeled jars, paired to represent incubation units. Most of the labeled incubations resulted in equilibrium mixtures of N20 masses. Differences between lengths of pre-wet and control segments were small. If the soil mineral pool undergoing denitrification is not uniformly labeled, the enrichment of the soil pool is an overestimate (Boast et al. 1988, Arah 1992, Bergsma et al. 1999). Since uniformity was not assumed in this study, we adopt the convention of referring to the apparent enrichment of the soil mineral pool. Apparent enrichment for each incubation is shown in Table 4.2. For both cropped and successional soils, apparent enrichment calculated from N2 isotope data agrees strongly with that calculated from N20 isotope data. Enrichment of the soil mineral pool was predicted from the amount of label 81 added and the amount of nitrate initially present (Table 4.2). Predicted enrichments agree well with the calculated enrichments. Production of N2 estimated by isotope methods shows the same patterns as production of N2 estimated by acetylene inhibition (Figure 4.5). However, the isotope method gives values consistently lower than the acetylene method. For each combination of ecosystem, moisture history, and replicate, the ratio MSIGC was calculated, where MS refers to N2 flux by mass spectrometry (‘SN isotope dilution method) and GC refers to N2 flux by gas chromatography using acetylene inhibition. Analysis of variance showed no effect of block, ecosystem, or moisture history, and no interaction of ecosystem and moisture history (P > 0.4 for all effects). WIth one outlier removed (from 12 total values), mean and standard error for MSIGC is 0.34 1 0.04. Discussion Experimental Design Our experiment was designed to test for an effect of recent moisture history (antecedent moisture regime) and for an effect of ecosystem differences on N20 mole fraction during denitrification in soil. Validation of the design requires that (1) tests performed on different soils were equivalent in other respects, (2) the moisture-history treatments were equivalent in other respects, and (3) the N20 observed was the product of denitrification. Successional soils were tested four weeks after the cropped soils. The very small (less than 0.5 pg N - g dry soil ’1) drop in extractable NO; -N in stock soil during that interval, and the absence of change in NHf-N, suggest that no 82 1'32“?- strong artifacts were introduced by the delay. The delay itself is small compared to the initial air-dry storage time for both soils (> 8 weeks). To test whether the two patterns of moisture addition (pre-wet vs. control) had similar consequences for soil N status, we sampled destructively for extractable NO; and NH4’ in the mineral-N jars as soon as possible (~2 h) after the start of the incubation. Mean NOa'-N concentrations were 1-2 pg N - g dry soil '1 lower for the pre-wet soils, 8 small difference relative to the amount of N added (9 pg N - g dry soil ") and the amount of N03‘-N extracted (9-21 pg N - g dry soil"). Extractable ammonium was sharply higher in pre-wet soils than in control soils. However, the difference is due more to a reduction of control [NHX], relative to background, than to the enhancement of pre-wet [NHX] (data not shown). Apparently significant NH; was lost (possibly through volatilization or assimilation) when water first contacted the dry soil, but was replaced thereafter via mineralization. However, NH: is not the substrate for denitrification. Furthermore, since the pattern was similar across ecosystems, this artifact does not explain the interaction of ecosystem and moisture history in our results. Was the observed N20 a product of denitrification? Nitrous oxide can be produced by nitrification and dissimilatory nitrate reduction to ammonium as well as by denitrification (T iedje 1988). However, nitrification is an aerobic process, for which optimum moisture ranges between 30% and 70% water- filled pore space (Davidson 1991). Many studies address the partitioning of N20 production by source, including nitrification and denitrification (e.g. 83 Mulvaney and Kurtz 1984, Robertson and Tiedje 1987, Klemedtsson et al. 1988, Skiba et al. 1993, Mummey et al. 1994, Stevens et al. 1997, Stevens of al. 1998b, Hutsch et al. 1999, Panek et al. 2000, Wolf and Russow 2000) and also dissimilatory nitrate reduction to ammonium (e.g. Stevens et al. 1998a). VVIth notable exceptions (e.g. Hutchinson et al. 1993) denitrification is usually the major source of nitrous oxide in saturated and nearly-saturated soils. Because WFPS was about 85% for this study, we assume that N20 was a product of denitrification. Furthermore, for a majority of the incubations, N20 derived from soil was in isotopic equilibrium, which implicates a single, uniformly labeled soil mineral pool (Figure 4.4; see Bergsma et al. 1999) and strongly supports denitrification as the only important source of N20. Also, there was good agreement between calculated and estimated enrichments for the soil NO; pool contributing to flux (Table 4.2) suggesting that N20 derived predominantly from N032 N20 mole fraction For the successional soil, the N20 mole fraction was about one-third. The pre-wet treatment brought mean soil moisture to approximately 68% WFPS, but apparently did not greatly enhance denitrification enzyme status relative to the controls. We conclude that denitrifying enzymes, especially nitrous oxide reductase, persisted well in the successional soil during several months of air-dry conditions (< 1% gravimetric moisture). Enzyme persistence in dry soils has been observed by others (e.g. Smith and Parsons, 1985). 84 For the cropped soils, however, the pre-wet treatment apparently enhanced the activity of nitrous oxide reductase relative to the controls. N20 mole fraction was also about one-third for the pre-wet soils, but about 0.9 for the (previously dry) control soils. Although total denitrification was Similar for the two moisture histories, a much greater fraction of N20 was further reduced to N2 in the pre-wet soils. We conclude that nitrous oxide reductase did not persist well in the cropped soil when air-dry, but that its activity was significantly enhanced by 48 hours of high soil moisture, achieving levels similar to those for successional soils. Since total denitrification was only slightly less for cropped soils than for successional soils, precursor enzymes (such as nitrite reductase and nitrate reductase) may have been less affected by drying than was nitrous oxide reductase. Our observation that - under at least some conditions - N20 mole fraction may be higher for the cropped soils than for the successional soils may help explain field data showing threefold greater annual flux of N20 from the cropped system (3.5 1 0.21 g NZO-N - ha'1 - d") than from the successional system (1.1 1 0.05 g N20-N - ha'1 - d", Robertson et al. 2000) To the best of our knowledge, no other published study of N20 mole fraction has tested for a potential interaction between ecosystem effects and moisture history effects. However, there are some reports of effects of either ecosystem or moisture history on relative proportions of N2 and N20. Merrill and Zak ( 1992) reported an N20 mole fraction of 0.7 to 0.9 for well-drained sugar maple forests in northern lower Michigan; in contrast, the N20 mole 85 fraction in a silver maple - red maple swamp was 0.25. Dendooven et al. (1996) found an effect of moisture history on relative production of nitrous oxide and dinitrogen (N202N2) for pasture soil, but the difference was small: 0.54 for soil cores previously submerged for 96 hours, and 0.4 for cores submerged for 6 hours. Conversion to nitrous oxide mole fraction yields values of 0.35 and 0.29: similar to the values presented here for successional soils and pre-wet cropped soil. Mulvaney and Kurtz (1984) studied N20 and N2 flux for three 15N- amended soils subjected to wetting and drying cycles. We calculate from their Table 4.1 an average and standard error of 0.33 1 0.02 (n = 12), similar to the result for our successional soils: 0.33 1 0.04 (n = 6). In a study of three N- amended soils, Jacinthe et al. (2000) found that N20 mole fraction was initially 0.68, increased to 0.95 with imposition of a water table at a depth of 10 cm, and decreased to 0.35 within one week. Our results show that the dependency of nitrous oxide mole fraction on recent moisture history can vary among ecosystems, even when the ecosystems are pedogenically identical. The large difference in response between the successional soils and the cropped soils may be related to differences in soil physical properties, soil carbon patterns, and microbial community characteristics resulting from 10 years of contrasting soil management regimes. First, the cropped soil is plowed regularly and has poorer aggregation than the successional soil. Because the soil used here was sieved (4 mm), effect of aggregate structure would have been restricted to smaller size classes of aggregates. Physical differences may have influenced 86 N20 mole fraction by altering the distribution of water and anaerobic microsites, where most denitrification may occur (McConnaughey and Bouldin 1985). Second, the response difference between the two soil types may be related to differing soil carbon patterns. Although net primary productivity in the cropped ecosystem is about double that of the successional ecosystem (mean 1 standard error = 9.24 1 1.41 vs. 4.24 :l: 0.37 MT - ha'1 - y") the successional system is accumulating soil organic carbon while the cropped system is not (Robertson et al. 2000). After ten years, soil organic carbon content (0 to 7.5 cm depth) was unchanged for the cropped system (1.00 1 0.05 %) but had risen significantly for the successional system (1.63 1 0.06 %). In addition to greater absolute carbon content, the successional soil may have a greater variety of substrates for microbial heterotrophs because of greater plant species diversity. Differences in soil carbon can influence factors controlling N20 mole fraction. Menyailo and Huwe (1999) found that 26 years of soil development under six species of trees caused changes not only in soil chemistry, but also in the persistence and dynamics of denitrifying enzymes. Carbon quality (C:N ratio) was the most important soil chemical factor in explaining differences in N20 emission among soil types. Finally, differences in the microbial communities between the cropped and successional ecosystems may account for the different responses to soil moisture history. It is possible, perhaps even likely, that differing soil properties caused functionally significant divergence in microbial community composition. For instance, nitrate availability is typically much lower for the 87 successional soil (0.63 1 0.04 pg N03'—N - g'1) than for the cropped soil (6.54 1 0.53 pg NOa'-N - g"; Robertson et al. 2000). In the successional soil, a hypothetical sub-group of denitrifiers with the ability to maintain enzyme status (especially NOS) during dry periods could have a competitive advantage in exploiting the flush of carbon that occurs on soil wet-up (e.g. Groffman and Tiedje 1988), since they could use N20 as well as N03' as a terminal electron acceptor if oxygen were limiting. ln cropped soils, the incentive for N08 maintenance would be less, because of the abundance of the more energetically-favorable electron acceptor N031 Thus, if variation exists among denitrifier taxa in their ability to maintain NOS status during soil drying, then a putative mechanism of natural selection exists that could explain our results, in terms of differences in microbial community composition. Variation may indeed exist among denitrifiers in their ability to maintain NOS status during soil drying. Cavigelli and Robertson (2000b) isolated 31 denitrifier taxa from two ecosystems: the cropped ecosystem studied here and a nearby never-tilled successional field. They showed that considerable variability exists among taxa for sensitivity of the NOS enzyme to varying levels of oxygen, a parameter clearly related to soil drying. Furthermore, Cavigelli and Robertson (2000a) found differences in denitrifying ability for whole soil microbial communities (slurry assay) for the cropped ecosystem and the never- tilled successional field. Denitrifying enzymes were more sensitive to oxygen levels in the agricultural soil, and nitrous oxide reductase was more active in the successional soil. Their results are consistent with our suggestion that the 88 microbial community in the successional soils may have experienced selection for denitrifiers with the ability to maintain the status of denitrification enzymes, especially NOS. The story of microbial community for the KBS LTER treatments is, however, complicated. Cavigelli and Robertson (20008, 2000b) compared the conventionally-tilled agricultural treatment (our “cropped” system) and a never- tilled successional treatment. Our study compared the conventionally-tilled system to a historically-tilled successional treatment. Buckley and Schmidt (2000) used biochemical techniques to characterize relative abundance of seven broad taxonomic groups in the microbial communities of KBS LTER treatments. They determined that the communities from the conventionally- tilled system and the historically-tilled system were much more similar to each other than to the community from the never-tilled system. Therefore, comparisons between our results and those of Cavigelli and Robertson should be made with appropriate reserve. Still, there is a strong possibility that at finer taxonomic levels, functionally significant differences exist between the microbial communities from the conventionally-tilled cropped ecosystem and the historically-tilled successional ecosystem (D. Buckley, personal communication). The existence of these differences and their importance for denitrification remain to be demonstrated. Questions of mechanism notwithstanding, evidence of a role for moisture history in controlling N20 mole fraction has an important place in the biogeochemistry of nitrogen. Many studies suggest that most N is lost from 89 soils during brief periods following irrigation or rainfall (Smith and Tiedje 1979, Sextone etal. 1985, Rolston et al. 1982, Mummey et al. 1994, Davidson, 1991). Dependency of the N20 mole fraction on short-tem soil moisture history could have large consequences for the relationship between nitrous oxide production and total denitrification. Since N20 flux has been modeled for intervals as short as one day (e.g. Li et al. 1992a, 1992b), the time scale implied by “short-term” in our study (48 hours) is relevant for efforts to constrain the global N20 budget (e.g. Bouwman 1990, Eichner 1990). Isotope data In principle, isotope data for N20 allow an independent estimate of N20 production. However, the estimate depends on a two-member mixing model (Bergsma et al. 1999) in which absolute contribution from both members is significant. Our concentration data (Figure 4.1) show that N20 concentration changed by almost three orders of magnitude (0.3 pg N - g dry soil " to ~200 pg N - g dry soil '1) even in the least productive jars. We conclude that the final isotopic character of N20 in the headspace of labeled jars essentially represents the isotopic character of soil-derived N20. Thus, the 15N-NZO data in this study are suitable for tests of equilibrium and estimates of soil enrichment, but not for independent estimates of N20 production. Estimates of N2 production by mass spectrometry were typically only one-third of estimates by acetylene inhibition. One of the two methods may have been biased, or the methods may have reflected qualitatively different aspects of the experimental system. The acetylene method could have been 90 biased if denitrification was enhanced by the presence of acetylene. There is some evidence in the literature for the stimulation of denitrification by acetylene addition (e.g. Klemedtsson et al. 1988), but usually for longer incubations and to a lesser extent than would be the case here. Furthermore, the soils studied here were amended with glucose, which argues against one putative mechanism of denitrification enhancement by acetylene: namely, release from carbon limitation (see also Topp and Germon 1986). In oxic soils acetylene may cause denitrification to be underestimated due to scavenging of intermediate nitric oxide (Bollman and Conrad 1997). Our soils, however, were largely anaerobic, and the putative error (if any) is overestimation not underestimation. Few studies have explicitly compared the acetylene inhibition technique with the 15N isotope dilution technique for estimation of N2 flux. In their seminal paper reviving interest in the 15N isotope dilution method for N2 flux, Siegel et al. (1982) confirm accuracy of enrichment estimates for their method, but not necessarily of flux estimates. Rolston et al. (1982), measuring only N-gas derived from fertilizer, found reasonable agreement between the acetylene method and a 15N-accumulation method. Mosier et al. (1986) compared total denitrification under acetylene inhibition (N20 only) with total denitrification by 15N mass spectrometry (N20 + N2; N20 catalytically reduced to N2 prior to analysis). Acetylene-amended plots consistently gave higher fluxes (nominal MSIGC ~0.75): this was attributable to acetylene treatment itself, rather than analytical bias between the two methods, since no difference in methods was 91 found when both were applied to samples from the acetylene-amended plots (see Table 1, last column, in Mosier et al. 1986). Aulakh et al. (1991) reported very similar total denitrification for two acetylene methods and for the 15N method of Mosier et al. (1986). Arah et al. (1993) found that acetylene inhibition consistently gave lower N2 flux estimates than 15N isotope dilution, and concluded that for heavy textured soils the acetylene block was incomplete. Mulvaney (1988) and Mulvaney and Vanden Heuvel (1988) compared fluxes of N20 as measured by mass spectrometry (reducing N20 catalytically to N2) and by gas chromatography. Both studies found only modest differences, and no strong bias towards underestimation or overestimation of N20 flux. For our study, there is no strong justification for questioning the accuracy the N2 measurements made by acetylene inhibition. The difference between the two methods is perhaps explained as bias in the 15N-based measurements — an explanation that is not entirely satisfying. Underestimation is expected from the 15N-dilution technique when the soil mineral N pool undergoing denitrification is not uniform (Boast et al. 1988, Arah 1992, Bergsma et al. 1999). But how much underestimation? Simulation (Arah 1992) and direct calculation (Bergsma et al. 1999) show that if flux derives from a very large number of pools with enrichments randomly distributed from natural abundance levels to 100%, the central tendency of the underestimation is 0.75. Albeit theoretical, the value 0.75 serves as a convenient null hypothesis. It does not explain our nominal MSIGC ratio of 0.34. Also, Figure 4.4 shows that for most incubations, N20 derived from soil 92 was probably in equilibrium, implying a well-mixed soil source. Since N20 is the direct precursor of N2 (Payne 1981), one would expect N; from soil also to be in equilibrium (Focht 1985) and therefore free of the underestimation ascribed to non—uniform pools. Furthermore, estimates of enrichment (of the soil mineral pool undergoing denitrification) based on N2 in our study agreed well with estimates based on N20. Others have found similar results (Mulvaney and Kurtz 1984, Mosier et al. 1986). Given the above considerations, the most satisfying explanation for the differences between the two methods of calculating N2 production is that the methods gave qualitatively different information about the observed system: the acetylene method reported gross N2 production while the 15N dilution method reported only production from a highly enriched, uniformly labeled pool, i.e. the enriched N20 or its substrate. A second, unenriched soil mineral N pool was also a source of N2 (indeed the major source), but not a net source of N20. Under these circumstances, N2 production would have been underestimated without affecting the estimate of enrichment for the labeled pool (see Focht 1985) which would explain the agreement of N2 and N20 data for estimates of pool enrichment (Table 4.2). Estimates of enrichment (of the soil mineral pool) based on isotopic data also agreed well with estimates predicted by mass balance (based on knowledge of extractable NO;{ levels for stock soil and knowledge of the magnitude of the K15N03 addition; Table 4.2). The agreement of the mass balance estimates and the 15N estimates suggests further that the putative unlabeled source of N2 is not a static, extractable NOg’ pool. Perhaps 93 the unlabeled source is a dynamic result of mineralization and nitrification occurring as a consequence of rapid soil wet-up, tightly-coupled to denitrification. This scenario implies nitrification rates (a few pg N - g dry soil'1 - d") that are about an order of magnitude greater than typical potential nitrification rates for these soils in the field (about 0.1 pg N - g dry soil’1 - d"; Robertson et al. 2000). More work is needed to account for the interesting differences between the acetylene inhibition method and the 15N-dilution method for estimating N2 flux as used in this study. Conclusions The design of the study as implemented was suitable to test for effects of recent moisture history (antecedent moisture regime) and ecosystem differences on N20 mole fraction during denitrification. N20 mole fraction in successional soils was not affected by moisture history, but in cropped soils it was sharply lower when soil moisture had been high for 48 hours prior to incubation. We suggest that persistence of nitrous oxide reductase in the successional soils was less sensitive to water stress during soil drying, perhaps because the lower level of native soil nitrate selects (in the successional microbial community) for denitrifier taxa with enhanced capacity for enzyme maintenance. Explicit recognition of ecosystem differences in response of N20 mole fraction to recent moisture history may improve modeled estimates of global N20 flux. Furthermore, understanding the impact of soil management regimes on mole fraction dynamics within ecosystems may lead to strategies that minimize flux of N20 to the atmosphere. 94 The isotope data show that the mineral N pool undergoing denitrification was isotopically uniform in most cases. There is no conclusive explanation for the strong differences between estimates of N2 production by acetylene inhibition and by 15N isotope dilution. Perhaps, under the experimental conditions described, there existed an alternative substrate for production of N2, but not for N20; the 15N method may then have reported only production of labeled N2, while acetylene inhibition would have reflected gross production of N2. Factors likely contributing to the difference between methods are the complexity of the soil environment and the dynamic nature of N transformations during rapid re-wetting of soil. 95 Table 4.1. Nitrous oxide mole fraction ( N20 I [ N20 + N2] ) analyzed by ecosystem and recent soil moisture history. Estimates are mean 1 standard error, in pg N - g dry soil". Effects are P values (Prob. > F). “Ecosystem by history” is the interaction term. 96 «N50 mmood 83.0 mvvod mod N mvd - - uoEnEoo - - SE6 ommod - no.0 a and 9.0 a wad .m:o_mmoooam - - R85 88.0 - a; a 8.0 mod n 8.0 8:88 {on to ice be foo be Ex“: $31 $de $21 929: 929: 929: boE: 3 Soto Soto Soto come :36 :moE 823388 82338 >632 3.83 605988 .228 103.05 Ecum>moom ._..v 2an 97 Table 4.2. Comparison, by ecosystem and replicate, of predicted soil N03' enrichment (atom fraction 15N) with apparent enrichment of the soil pool undergoing denitrification. “Predicted” is calculated by mass balance from the extractable NO3' levels in stock soil and the known addition of KNOa. Apparent enrichments are from N2 isotope data (“by N2”), from N20 isotope data (“by N20”), or “average” of N2 and N20 estimates. Ecosystem Rep. predicted by N2 by N20 average cropped 4 0.42 0.57 0.53 0.55 cropped 5 0.35 0.31 0.38 0.34 cropped 6 0.42 0.71 0.58 0.64 successional 4 0.90 0.87 0.87 0.87 successional 5 0.78 0.83 0.83 0.83 successional 6 0.84 0.84 0.86 0.85 98 Figure 4.1. Change in headspace concentrations (volume/volume) of nitrous oxide during incubation, with selected (for clarity) standard error bars. Dashed lines = successional soils, solid lines = cropped soils. Open symbols = control, filled symbols = pre-wet. Gray squares = + acetylene, black diamonds = - acetylene. Control and pre-wet curves diverge strongly for cropped soils (solid lines with open or filled diamonds) but not for successional soils (dashed lines with open or filled diamonds). 99 .3. 2:9: 5:335 co Ego: vm o 4 919° -- Two... .. v-moN H. -- Eon. w .m -- "two.“‘ 0 . w 61:8 .8390 IT Tmom W go‘s-ca condo-.0 IT . W . _o._Eoo ._m:o_wmooo:m :6.-- -- Ymoo x .. woes-ca ._m:o_mmooo:m :6.-- m 8:038"... + 63:8 .8380 limit -- Ymos ... a . .. 05.38.... + “oz-ma condo-6 .......I....... 0:238er .9250 ._m:o_mmooo:m :-E.-.. -. Twod .. 820581-6305 ._m:o_mmoooam ----l. H. -- 1mg Wmoé 100 Figure 4.2. Summary of production of N2 and N20. “Total denitrification” is N20 + N;- by acetylene inhibition, “N20” is production of N20 in the absence of acetylene, and “N2” is the difference. Even though total denitrification did not differ significantly between moisture histories for cropped soils (bars with solid borders), the control incubations produced significantly more N20 and correspondingly less N2. Bars with dashed borders represent soils from successional plots. 101 Nz ONZ cosmoEono .20... .958 ._mco_mmo8:m F 83.08 ._m:o_mmm83m l 61:8 cacao-6 W 83-05 .8808 I as 2:9“. c'>. v nos Mp 6 - N 6n 102 Figure 4.3. Isotopic character of N2 for labeled and unlabeled jars at the end of the incubation, paired (by line segment) to represent incubation units. Dashed lines = successional soils, solid lines = cropped soils. Open symbols = controls, filled symbols = pre-wet. Circles = Replicate 4, squares = Replicate 5, triangles = Replicate 6. Due to the ovenrvhelming abundance of unlabeled N2 from the atmosphere relative to labeled soil—derived N2, displacement of isotopic character during the incubation (i.e., length of the line segments) is much smaller than for N20 (Figure 4.4). Only a small portion of the equilibrium curve (downward-opening parabola in Figure 4.4) is visible at this scale. For cropped soils, control segments (solid lines with open symbols) are much shorter than pre-wet segments (solid lines with closed symbols) indicating less N2 production for controls. For successional soils (dashed lines) the opposite effect or no effect is seen. 103 :ozomt Eofi 2.... m-mmm.m m-mom.m mummmd m-wow.m m-mmk .m m-mond .2. 28:. Wmmod _ H A _ d cos-ma ”285.?- 85... 63:8 ”£88? :80 .mofioiom 2 8888 88% .856 )\ tine X». \ 0 “\“flst C 0“”9“ . Q s‘!“‘ .0“ 8 Q‘ ‘ C S 8. ‘ . \ ‘ ’ a. \5 - . 9‘ ‘0‘ \ \\ $9 ‘ “ N O 8‘ ‘ . ‘ \ 8 ‘ ‘- 9‘ \‘ ‘8‘ \ t \s .1 s \.\ \. s Q ‘ ‘ s I. \.w1 \ I‘ .9 9 g C. o \ ’ ‘ - Q ‘ \ . .. 1... l . 8 ‘ 5‘ ‘ Q ‘n' 8‘ ‘ t \ \ o I Q E \( 5‘ \‘ 3‘ ‘0 ‘ .1 9 S \\ 5‘ 0 00‘. 1 \‘ ‘ \ 3 3‘ ‘ ’ b.\ ‘ 9. \ ‘ $3 \ \ \ OI \\ 5 \ Q- s ‘ I -- 8890 .moE. 9.8 ‘ I I Q ‘ \ ‘ ‘ \ \ ‘ ‘ Q Q ‘ ‘ ‘ .m:o_mmm8:m “mos. 88% l l 1 l m-momN - mummmk . m-mva 104 - m-moms uonoeu Jemoejour ‘Nez - m-mmmN - mimovfi m-mmvN Figure 4.4. Isotopic character of N20 for labeled and unlabeled jars at the end of the incubation, paired (by line segment) to represent incubation units. Dashed lines = successional soils, solid lines = cropped soils. Open symbols = controls, filled symbols = pre-wet. Circles = Replicate 4, squares = Replicate 5, triangles = Replicate 6. Due to the overwhelming abundance of N20 from soil relative to unlabeled atmospheric N20, enriched samples represent essentially the isotopic character of N20 derived from soil. Samples falling on or near the equilibrium curve (downward-opening parabola) indicate a source in isotopic equilibrium, implying a single, uniformly labeled substrate pool (e.g. a homogenous mixture of native N03' and 15N03' label). 105 00; 828.30”. 2 0:888 wedge 6956 5.8:: 2:85 29 .3. 239.... 00.0 00.0 0v.0 0N.0 00.0 p p — H _ _ cos-ca 8.8:...» 8...... hams-«ea 61:8 ”2853 :80 .- ‘ I! ‘0 I I “. Is. 8‘ 8 I. 9 C \ ‘0 I... 1‘ ‘I {.6 S ‘ ‘ 0 1 1. ‘ ‘ ‘ “ .I‘ ‘n- ’ “‘ 0‘ ‘ ’fl‘ ‘ C \ ‘ ‘ ‘ “ I 0 1 19 “‘ “ ‘8 E‘ .‘ ‘. “ “. 0 \ 6 I ‘ ‘ “ 9 6 6 I ‘0 C \ C O I. .1 1 5 O “ 0‘ ‘n‘. 0 ‘ 11 t 1 1 O 0 .1 o. 0 ‘ “ C.‘ ‘ . 1 10 9‘ .1 c. 5 I 11 1 1 O I. a. 9 s O | s 1. O 1 1 t 10 ‘ 0‘ ‘ $ 6 .I Q \s 1.. 1 c \ i b. . 1 1 t c \1 s. u. 0 ‘ " ‘ 11 t 1 v 1 $6 ‘9‘. vs 1 o 1 1 c 1 1 o .1 1 c s O 9 C .u U 0‘ " ‘ ‘ i . \ . .1 i .1 ’ I. .V . . 3 ‘ 53.5...38 .028 L -0_..0 63 -- 09.0) g. N 0 NW I I m N o 1 T o ‘8 o -- 00.0 uoltoerj rajnoejour l - 0Y0 -- 0Y0 00.0 106 Figure 4.5. N: production calculated independently by the acetylene inhibition technique (AIT) and by mass spectrometry (M8), for both cropped and successional soils. In every case but one, MS gave a much smaller estimate of N2 production than did AIT. 107 .3 2:9“. w: ._mco_mmooo:m t< __m:o_mmmoo:m ms. 62520 ._._< .3390 it 0. Fl v.0. 1' 0.0 l o._‘ a, N l_|go's Mp 6 - N 6n l C! 00 65:8 cm I 6235 my. I .9550 mm l 83.95 mm E 6:50 5. a 63.05 vm n 108 OVERALL CONCLUSION A general conclusion that emerges from my research is that progress in understanding the relationship between denitrification and N20 flux from soil is still methods-limited. The problem arises from the position that denitrification occupies in the nitrogen cycle: right on the brink of a sea of molecular N that constantly scours the beaches of inquiry. Acetylene inhibition and 15N dilution are primarily devices to increase the sensitivity of measuring N2 production, the missing piece of the puzzle. Unfortunately, both interact with the system they were designed to explore, thereby compromising the interpretation. I do not mean to suggest that accurate measurements of N2 are rare or impossible, only that they are difficult and far from routine. The good news is that when the tide goes out, the stories we find on the beach tend to be consistent. Variation in N2 and N20 production seems to abide proximally by principles of differential enzyme induction, and more distally by consistently-recognized ecological controls. The main question for the biogeochemistry community is, “How much needs to be known about ecological controls -- as distributed across ecosystems - in order to adequately predict regional and global N20 flux?” When the practical threshold for data gathering meets that need, a major breakthrough will be within reach. I hope that in some way my brief dashes among the waves will have made a contribution. 109 APPENDIX 110 APPENDIX 111 Table A1. Tests of significance (of differences among means) for Figure 4.2: probabilities of finding the observed differences in means by chance alone. Each row represents a single model of “Response”, incorporating the effects for which there are column entries. In most cases, models returning P > 0.05 for an effect were re-run without that effect (“-”). Dropping the effect (“-") is a notational convenience only; statistically it is no different from grouping the responses across levels (“grouped”). Levels of block are [4, 5, 6]. Levels of ecosystem are [cropped, successional]. Levels of history are [pre-wet, control]. “NA” means the effect (column) is not meaningful for testing the response (row). 112 .7... x " 0'4. mugs}? 44.4 <2 whovd noaaoco $3.65 ”2983005 ouz - m8; - 8:385 ouz 82.0 - - 8:885 owz 88.0 88d - 8:385 onz 88.0 983 88.0 8:885 ouz - <2 83.0 .mco_mmmoo:m ”cozoaooa ouz 89.0 <2 <83 385883 “8.6385 omz - <2 55.0 8&20 ”88:85 onz 88.0 <2 88d 8&90 ”83:85 omz - - £86 58.? 8:852ch :29 - 59.0 88.0 686 8:852ch .92 83.0 moomd 88.0 88.0 8:85:38 .92 >552 x x003 >8sz Emumxmoom xooE omcoamom .3. 2an 113 E 2 v00 F .0 3.00.0 08.5.0 803.0 03.0.0 3.00.0 nmmod vmmod 05:30 mmood 0000.0 00.0.0 0050.0 53.20 ”20003020 Nz $3.20 ”2283005 Nz $223808 ”00003090 Nz 5228803 ”20003020 ANz 08090 E26390 ANz cozozpoa Nz :26:an ANz 2.835580 2 2%.; 114 LITERATURE CITED 115 Literature Cited Arah, J. R. M. 1992. New formulae for mass spectrometric analysis of nitrous oxide and dinitrogen emissions. Soil Science Society of America Journal 56: 795-800. Arah, J. R. M. 1997. Apportioning nitrous oxide fluxes between nitrification and denitrification using gas-phase mass spectrometry. Soil Biology and Biochemistry 29: 1295-1299. Arah, J. R. M. and K. A. Smith. 1990. Factors influencing the fraction of the gaseous products of soil denitrification evolved to the atmosphere as nitrous oxide. Pages 475-480 in A F. Bouwman (ed.) Soils and the Greenhouse Effect. John Wiley and Sons, Ltd. Arah, J. R. M., I. J Crichton, and K A. Smith. 1993. Denitrification measured directly using a single-inlet mass spectrometer and by acetylene inhibition. Soil Biology and Biochemistry 25: 233-238. Arnold, P. W. 1954. Losses of nitrous oxide from soil. Journal of Soil Science 5: 116-128. Aulakh, M. S., J. W. Doran, and A. R. Mosier. 1992. Soil denitrification - significance, measurement, and effects of management. Advances in Soil Science 18: 1-57. Aulakh, M.S., J. W. Doran, and A. R. Mosier. 1991. Field evaluation offour methods for measuring denitrification. Soil Science Society of America Journal 55: 1332-1 338. 116 is"! Austin, Franklin R. 1979. Soil Survey of Kalamazoo County, Michigan. United States Department of Agriculture, Soil Conservation Service, in cooperation with Michigan Agricultural Experiment Station. The Soil Conservation Service: Washington. Bergsma, T. T., N. E. Ostrom, M. Emmons, and G. P. Robertson. (submitted). High-sensitivity determination of nitrous oxide flux from nitrogen-15- labeled soil by isotope ratio mass spectrometry. Soil Science Society of America Journal. Bergsma, T. T., Q. C. Bergsma, N. E. Ostrom, and G. P. Robertson. 1999. A heuristic model for the calculation of dinitrogen and nitrous oxide flux from nitrogen-15-Iabeled soil. Soil Science Society of America Journal 63: 1709-1716. Betlach, M. R. and J. M. Tiedje. 1981. Kinetic explanation for accumulation of nitrite, nitric oxide, and nitrous oxide during bacterial denitrification. Applied and Environmental Microbiology 42: 1074-1084. Blackmer, A. M. and J. M Bremner. 1978. Inhibitory effect of nitrate on reduction of N20 to N2 by soil microorganisms. Soil Biology and Biochemistry 10: 187-191. Blackmer, A. M. and J. M. Bremner. 1979. Stimulatory effect of nitrate on reduction of N20 to N2 by soil microorganisms. Soil Biology and Biochemistry 1 1: 313-315. 117 Boast, C. W., R. L. Mulvaney, and P. Baveye. 1988. Evaluation of nitrogen-15 tracer techniques for direct measurement of denitrification in soil: 1. Theory. Soil Science Society of America Journal 52: 1317-1322. Bollman, A. and R. Conrad. 1997. Acetylene blockage technique leads to underestimation of denitrification rates in oxic soils due to scavenging of intermediate nitric oxide. Soil Biology and Biochemistry 29: 1067-1077. Bouwman, A. F. 1990. Exchange of greenhouse gases between terrestrial _ ecosystems and the atmosphere. Pages 61-127 in A. F. Bouwman (ed.) Soils and the greenhouse effect. John Wiley 8. Sons, New York. Boyd, S. R., A. Rejou-Michel, and M Javoy. 1994. Non-cryogenic purification of nanomole quantities of nitrogen gas for isotopic analysis. Analytical Chemistry 66: 1396-1402. Brooks, P. 0., D. J. Herman, G. J. Atkins, S. J. Prosser, and A. Barrie. 1993. Rapid, isotopic analysis of selected soil gases at atmospheric concentrations. Pages 193-202 in Agricultural Ecosystem Effects on Trace Gases and Global Climate Change. ASA Special Publication no. 55. American Society of Agronomy, Madison, WI, USA. Buckley, 0. H. and T. M. Schmidt. 2000. The structure of microbial communities in soil and the lasting impact of cultivation. Microbial Ecology (in press). Cavigelli, M. A. and G. P. Robertson. 2000a. The functional significance of denitrifier community composition in a terrestrial ecosystem. Ecology 81: 1402-1414. 118 .P Cavigelli, M. A and G. P. Robertson. 2000b. The role of denitrifier diversity in rates of nitrous oxide consumption in a terrestrial ecosystem. Soil Biology and Biochemistry (in press). Colboum P. and R. J. Dowdell. 1984. Denitrification in field soils. Plant and Soil 76: 213-226. Cooper, G. S. and R. L. Smith. 1963. Sequence of products formed during denitrification in some diverse western soils. Soil Science Society of America Proceedings 27: 659-662. Davidson, E. A. 1991. Fluxes of nitrous oxide and nitric oxide from terrestrial ecosystems. Pages 219-235 in John E. Rogers, William B. Wnitman (eds.) Microbial production and consumption of greenhouse gases: methane, nitrous oxides, and halomethanes. American Society for Microbiology, Washington, D. 0, USA. Dendooven, L. and J. M. Anderson. 1994. Dynamics of reduction enzymes involved in the denitrification process in pasture soil. Soil Biology and Biochemistry 26: 1501 -1 506. Dendooven, L. and J. M. Anderson. 1995. Maintenance of denitrification potential in pasture soil following anaerobic events. Soil Biology and Biochemistry 27: 1251 -1 260. Dendooven, L., L. Duchateau, and J. M. Anderson. 1996b. Gaseous products of the denitrification process as affected by the antecedent water regime of the soil. Soil Biology and Biochemistry 28: 239-245. 119 EA? - I 1.2 _ Dendooven, L., P. Splatt, and J. M. Anderson. 1996a. Denitrification in permanent pasture soil as affected by different forms of C substrate. Soil Biology and Biochemistry 28: 141-149. Dendooven, L., P. Splatt, J. M. Anderson, and D. Scholefield. 1994. Kinetics of the denitrification process in a soil under permanent pasture. Soil Biology and Biochemistry 26: 361 -370. Eichner, M. J. 1990. Nitrous oxide from fertilized soils: summary of available data. Journal of Environmental Quality 19: 272-280. Firestone, M. K. and E. A Davidson. 1989. Microbiological basis of NO and N20 production and consumption in soil. Pages 7-21 in M. 0. Andreae and D. S. Schimel (eds) Exchange of Trace Gases between Terrestrial Ecosystems and the Atmosphere. John Wiley and Sons: Chichester; New York. Firestone, M. K. and J. M. Tiedje. 1979. Temporal change in nitrous oxide and dinitrogen from denitrification following onset of anaerobiosis. Applied and Environmental Microbiology 38: 673-679. Firestone, M. K., M. S. Smith, R. B. Firestone, and J. M. Tiedje. 1979. The influence of nitrate, nitrite, and oxygen on the composition of the gaseous products of denitrification in soil. Soil Science Society of America Journal 43: 1140-1144. Focht, D. D. 1974. The effect of temperature, pH, and aeration on the production of nitrous oxide and gaseous nitrogen - a zero-order model. Soil Science 1 18: 173-179. 120 Focht, D. D. 1985. Differences in nitrogen-15 enrichments of evolved nitrous oxide and dinitrogen and the question of a uniform nitrate-15 pool. Soil Science Society of America Journal 49: 786-787. Gaskell, J. F., A. M. Blackmer and J. M. Bremner. 1981. Comparison of effects of nitrate, nitrite and nitric oxide on reduction of nitrous oxide to dinitrogen by soil microorganisms. Soil Science Society of America Journal 45: 1 124-1 127. Groffman, P. M. and J. M. Tiedje. 1988. Denitrification hysteresis during wetting and drying cycles in soil. Soil Science Society of America Journal 52: 1626-1629. Hahn, J. and P. J. Crutzen. 1982. The role of fixed nitrogen in atmospheric photochemistry. Philosophical Transactions of the Royal Society of London 8296: 521-541. Hallmark, S. L. and R. E. Terry. 1985. Field measurement of denitrification in irrigated soils. Soil Science 140: 35-44. Hart, S. C., J. M. Stark, E. A. Davidson, and M. K Firestone. 1994. Nitrogen mineralization, immobilization, and nitrification. Pages 985-1018 in R. W. Weaver et al. (eds) Methods of Soil Analysis, Part 2: Microbiological and Biochemical Properties-SSSA Book Series, no. 5. Hauck, R. D., and D. R. Bouldin. 1961. Distribution of isotopic nitrogen in nitrogen gas during denitrification. Nature (London) 191: 871-872. 121 Hauck, R. D., S. W. Melsted, and P. E. Yankwich. 1958. Use of N-isotope distribution in nitrogen gas in the study of denitrification. Soil Science 86: 287-291. Healy, W. R., R. G. Striegl, T. F. Russell, G. L. Hutchinson, and G. P. Livingston. 1996. Numerical evaluation of static-chamber measurements of soil- atmosphere gas exchange: identification of physical processes. Soil Science Society of America Journal 60: 740-747. Howarth R. W, G. Billen, D. Swaney, A. Townsend, N. Jaworski, K. Lajtha, J. A. Downing, R. N. Elmgren, T. Jordan, F. Berendse, J. Freney, V. Kudeyarov, P. Murdoch, Z. L. Zhu. 1996. Regional nitrogen budgets and riverine N and P fluxes for the drainages to the North-Atlantic Ocean - natural and human influences. Biogeochemistry 35: 75-139. Hutchinson, G. L. and E. A. Davidson. 1993. Processes for production and consumption of gaseous nitrous oxide in soil. Pages 79-94 in G. A. Peterson, P. S. Baenziger (eds.-in-chief) Agricultural Ecosystem effects on Trace Gases and Global Climate Change. ASA Special Publication no. 55. Hutchinson, G. L., W. D. Guenzi, and G. P. Livingston. 1993. Soil water controls on aerobic soil emission of gaseous nitrogen oxides. Soil Biology and Biochemistry 25: 1-9. Hutsch, B. W., X. Z. Wang, K. Feng, F. Yan, and S. Schubert. 1999. Zeitschn'ft fur Pflanzenernahrung und Bodenkunde 162: 607-613. 122 lntergovemmental Panel on Climate Change. 1996. J. T. Houghton, L. G. Meira Filho, B. A. Callander, N. Harris, A. Kattenberg, and K. Maskell (ed.), Climate Change 1995: The Science of Climate Change. Cambridge University Press, Cambridge. Jacinthe, P. A., W. A. Dick, L. C. Brown. 2000. Bioremediation of nitrate- contaminated shallow soils and waters via water table management techniques: evolution and release of nitrous oxide. Soil Biology and Biochemistry 32: 371 -382. Kaspar, H. F. and J. M. Tiedje. 1994. Anaerobic bacteria and processes. Pages 223-243 in R. W. Weaver et al. (eds) Methods of Soil Analysis, Part 2: Microbiological and Biochemical Properties-SSSA Book Series, no. 5. Klemedtsson, L., B. H Svensson, and T. Rosswall. 1988. Relationships between soil moisture content and nitrous oxide production during nitrification and denitrification. Biology and Fertility of Soils 6: 106-111. Koskinen, W. C. and D. R. Keeney. 1982. Effect of pH on the rate of gaseous production of denitrification in a silt loam soil. Soil Science Society of America Journal 46: 1 165-1 167. Kroeze, C., A. Mosier, and L. Bouwman. 1999. Closing the global N20 budget: A retrospective analysis 1500-1994. Global Biogeochemical Cycles 13: 1-8. 123 Lensi, R. and A. Chalamet. 1982. Denitrification in waterlogged soils: in situ temperature-dependent variations. Soil Biology and Biochemistry 14: 1 51 -1 55. Letey, J., N. Valoras, A. Hadas, and D. D. Focht. 1980a. Effect of air-filled porosity, nitrate concentration, and time on the ratio of ~20/N2 evolution during denitrification. Journal of Environmental Quality 9: 227-231. Letey, J., W. A. Jury, A. Hadas, and N. Valoras. 1980b. Gas diffusion as a factor in laboratory incubation studies on denitrification. Journal of Environmental Quality 9: 223-227. Li, C., S. Frolking, and T. A. Frolking. 1992a. A model of nitrous oxide evolution from soil driven by rainfall events: 1. Model structure and sensitivity. Journal of Geophysical Research 97: 9757-9776. Li, C., S. Frolking, and T. A. Frolking. 1992b. A model of nitrous oxide evolution from soil driven by rainfall events: 2. Model applications. Journal of Geophysical Research 97: 9777-9783. Linn, D. M. and J. W. Doran. 1984. Effect of water-filled pore space on carbon dioxide and nitrous oxide production in tilled and non-tilled soils. Soil Science Society of America Journal 48: 1267-1272. Matthews, E. 1994. Nitrogenous fertilizers - global distribution of consumption and associated emissions of nitrous oxide and ammonia. Global Biogeochemical Cycles 8: 411-439. 124 McConnaughey, P. K. and D. R. Bouldin. 1985. Transient Microsite Models of Denitrification: I. Model Development. Soil Science Society of America Journal 49: 886-891. McKenney, D. J., C. F. Drury, W. I. Findlay, B. Mutus, T. McDonnell and C. Gajda. 1994. Kinetics of denitrification by Pseudomonas fluorescens: oxygen effects. Soil Biology and Biochemistry 26: 901 -908. Menyailo, O. V. and B. Huwe. 1999. Activity of denitrification and dynamics of N20 release in soils under six tree species and grassland in central Siberia. Zeitschrift fur Pflanzenernahrung und Bodenkunde 162: 533- 538. Merrill, A. G. and D. R. Zak. 1992. Factors controlling denitrification rates in upland and swamp forests. Canadian Journal of Forest Research 22: 1 597-1604. Mosier, A. R. 1989. Chamber and isotope techniques. Pages 175-187 in M. O. Andreae and D. S. Schimel (eds) Exchange of Trace Gases between Terrestrial Ecosystems and the Atmosphere. John Wiley and Sons: Chichester; New York. Mosier, A. R. and L. Klemedtsson. 1994. Measuring denitrification in the field. Pages 1047-1066 in R. W. Weaver et al. (eds) Methods of Soil Analysis, Part 2: Microbiological and Biochemical Properties-SSSA Book Series, no. 5. 125 Mosier, A. R., S. K Mohanty, A. Bhadrachalam, and S. P. Chakravorti. 1990. Evolution of dinitrogen and nitrous oxide from the soil to the atmosphere through rice plants. Biology and Fertility of Soils 9: 61-67. Mosier, A. R., W. D. Guenzi, and E. E. Schweizer. 1986a. Soil losses of dinitrogen and nitrous oxide from irrigated crops in northeastern Colorado. Soil Science Society of America Journal 50: 344-348. Mosier, A. R., W. D. Guenzi, and E. E. Schweizer. 1986b. Field denitrification estimation by nitrogen-15 and acetylene inhibition techniques. Soil Science Society of America Journal 50: 831-833. Mosier, A., C. Kroeze, C. Nevison, O. Oenema, S. Seitzinger, 0. van Cleemput. 1998. Closing the global N20 budget: nitrous oxide emissions through the agricultural nitrogen cycle - OECD/IPCCIIEA phase II development of IPCC guidelines for national greenhouse gas inventory methodology. Nutrient Cycling in Agroecosystems 52: 225-248. Mulvaney, R. L. 1984. Determination of 1SN-Iabeled dinitrogen and nitrous oxide with triple collector mass spectrometers. Soil Science Society of America Journal 48: 690-692. Mulvaney, R. L. 1988. Evaluation of nitrogen-15 tracer techniques for direct measurement of denitrification in soil: lll. Laboratory studies. Soil Science Society of America Journal 52: 1327-1332. Mulvaney, R. L. and C. W. Boast. 1986. Equations for determination of nitrogen—15 labeled dinitrogen and nitrous oxide by mass spectrometry. Soil Science Society of America Journal 50: 360-363. 126 Mulvaney, R. L. and L. T. Kurtz. 1982. A new method for determination of 15N- labeled nitrous oxide. Soil Science Society of America Journal 46: 1178- 1 184. Mulvaney, R. L. and L. T. Kurtz. 1984. Evolution of dinitrogen and nitrous oxide from nitrogen-15 fertilized soil cores subjected to wetting and drying cycles. Soil Science Society of America Journal 48: 596-602. Mulvaney, R. L. and L. T. Kurtz. 1985. Reply to “Differences in nitrogen-15 enrichments of evolved nitrous oxide and dinitrogen and the question of a uniform nitrate-15 pool”. Soil Science Society of America Journal 49: 787. Mulvaney, R. L. and R. M. Vanden Heuvel. 1988. Evaluation of nitrogen-15 tracer techniques for direct measurement of denitrification in soil: IV. Field studies. Soil Science Society of America Journal 52: 1332-1337. Mummey, D. L., J. L. Smith, and H. Bolton. 1994. Nitrous oxide flux from a shrub-steppe ecosystem: sources and regulation. Soil Biology and Biochemistry 26: 279-286. Nommik, H. 1956. Investigation on denitrification in soil. Acata Agriculturae Scandinavica 6: 195-228. Panek, J. A., P. A Matson, l. Ortiz-Monasterio, and P. Brooks. 2000. Distinguishing nitrification and denitrification sources of N20 in a Mexican wheat system using N-15. Ecological Applications 10: 506-514. Paul, E. A. and F. E. Clark. 1996. Soil Microbiology and Biochemistry. Academic Press: San Diego. 127 Payne, W. J. 1981. The status of nitric oxide and nitrous oxide as intermediates in denitrification. Pages 85-103 in C. C. Delwiche (ed.) Denitrification, nitrification and atmospheric nitrous oxide. Wiley- lnterscience, New York. Robertson, G. P. 1999. Denitrification. Pages C181-190 in M. E. Sumner et al. (eds) Handbook of Soil Science. CRC Press: Boca Raton, FL. Robertson, G. P. and J. M. Tiedje. 1987. Nitrous oxide sources in aerobic soils: nitrification, denitrification, and other biological processes. Soil Biology and Biochemistry 19: 187-193. Robertson, G. P. and J. M. Tiedje. 1988. Deforestation alters denitrification in a lowland tropical rain forest. Nature 336: 756-759. Robertson, G. P., E. A. Paul, and R. R. Harwood. 2000. Nitrous oxide production, methane oxidation, and soil carbon sequestration in intensive agriculture: Composite effects on the atmosphere’s global warming potential. Science (in press). Rolston D. E., D. L. Hoffman, and D. W. Toy. 1978. Field measurement of denitrification I. Flux of N2 and N20. Soil Science Society of America Journal 42: 863-869. Rolston, D. E., A. N. Sharpley, D. W. Toy, and F. E. Broadbent. 1982. Field measurement of denitrification: lll. Rates during irrigation cycles. Soil Science Society of America Journal 46: 289-296. Sahrawat, K. L. and D. R. Keeney. 1986. Nitrous oxide emission from soils. Advances in Soil Science 4: 103-148. 128 Sall, J. and A. Lehman. 1996. JMP Start Statistics. Belmont, CA: Duxbury Press. Sextone, A. J., T. B. Parkin, and J. M. Tiedje. 1985. Temporal response of soil denitrification rates to rainfall and irrigation. Soil Science Society of America Journal 49: 99-103. Siegel, R. S., R. D. Hauck, and L. T. Kurtz. 1982. Determination of ”N; and application to measurement of N2 evolution during denitrification. Soil Science Society of America Journal 46: 68-74. Skiba, U., K A. Smith, and D. Fowler. 1993. Nitrification and denitrification as sources of nitric oxide and nitrous oxide in a sandy loam soil. Soil Biology and Biochemistry 25: 1527-1336. Smith, C. J. and W. H. Patrick, Jr. 1983. Nitrous oxide emission as affected by alternate anaerobic and aerobic conditions from soil suspensions enriched with ammonium sulfate. Soil Biology and Biochemistry 15: 693-697. Smith, M. S. and J. M. Tiedje. 1979. Phases of denitrification following oxygen depletion in soil. Soil Biology and Biochemistry 11: 261 -267. Smith, M. S. and L. L. Parsons. 1985. Persistence of denitrifying enzyme activity in dried soils. Applied and Environmental Microbiology 49: 316- 320. Speir, T. W., H. A. Kettles, and R. D. More. 1995. Aerobic emissions of N20 and N; from soil cores: measurement procedures using 13N-labeled N03' and NH4’. Soil Biology and Biochemistry 27: 1289-1298. 129 Stevens, R. J., R. J. Laughlin, and J. P. Malone. 1998a. Soil pH affects the processes reducing nitrate to nitrous oxide and di-nitrogen. Soil Biology and Biochemistry 30: 1119-1126. Stevens, R. J., R. J. Laughlin, and J. P. Malone. J. P. 1998b. Measuring the mole fraction and source of nitrous oxide in the field. Soil Biology and Biochemistry 30: 541-543. Stevens, R. J., R. J. Laughlin, G. J. Atkins, and S. J. Prosser. 1993. Automated determination of nitrogen-15-Iabeled dinitrogen and nitrous oxide by mass spectrometry. Soil Science Society of America Journal 57: 981 - 988. Stevens, R. J., R. J. Laughlin, L. C. Burns, J. R. M. Arah, and R. C. Hood. 1997. Measuring the contributions of nitrification and denitrification to the flux of nitrous oxide from soil. Soil Biology and Biochemistry 29: 139-151. Tiedje, J. M. 1988. Ecology of denitrification and dissimilatory nitrate reduction to ammonium. Pages 179-244 in A. J. B. Zehnder (ed.), Biology of Anaerobic Microorganisms. John Wiley and Sons, New York. Topp, E. and J. C. Germon. 1986. Acetylene metabolism and stimulation of denitrification in an agricultural soil. Applied and Environmental Microbiology 52: 802-806. Vanden Heuvel, R. M., R. L. Mulvaney, and R. G. Hoeft. 1988. Evaluation of nitrogen-15 tracer techniques for direct measurement of denitrification in soil: ll. Simulation studies. Soil Science Society of America Journal 52: 1 322-1 326. 130 Vinther, F. P. 1984. Total denitrification and the ratio between N20 and N2 during the growth of spring barley. Plant and Soil 76: 227-232. Vitousek, P. M., H. A. Mooney, J. Lubchenco, and J. M. Melillo. 1997. Human domination of Earth’s Ecosystems. Science 277: 494-499. Weier, K L. and J. W. Gilliam. 1986. Effect of acidity on denitrification and nitrous oxide evolution from Atlantic Coastal Plain soils. Soil Science Society of America Journal 50: 1202-1205. Weier, K L., J. W. Doran, J. F. Power, and D. T. Walters. 1993. Denitrification and the dinitrogen/nitrous oxide ratio as affected by soil water, available carbon, and nitrate. Soil Science Society of America Journal 57: 66-72. Wijler J. and C. C. Delwiche. 1954. Investigations on the denitrifying process in soil. Plant and Soil 5: 155-169. Wolf, l. and R. Russow. 2000. Different pathways of formation of N20, N2 and NO in black earth soil. Soil Biology and Biochemistry 32: 229-239. Yoshinari, T. and R. Knowles. 1976. Acetylene inhibition and nitrous oxide reduction by denitrifying bacteria. Biochemical and Biophysical Research Communications 69: 705-710. 131 lrill)jiiliiiiilil