flu- ..x. ‘ . ‘ ri£u.i. u"-~.~n|v ML. . .(I. 7 .. .M u. . éfégxf a. 5?}széfifiéi%awm,§..a.,£m .5 .».r,..... 41.3. .7: .. ..|(A. 1.. . .1! :1 x .3 w..¢.£:fii...$£m§. . ma 5... a. » jar, This is to certify that the dissertation entitled STABLE NITROGEN ISOTOPES AS A TOOL FOR UNDERSTANDING NITROGEN CYCLING PROCESSES AND MECHANISMS OF NITROUS OXIDE PRODUCTION presented by Robin Leslie Sutka has been accepted towards fulfillment of the requirements for Ph.D. degree in Environmental Geosciences Major professor 7am OSWVW CW Date 8/? ’01 MS U is an Affirmative Action/Equal Opportunity Institution 0- 12771 LIBRARY Michigan State University STABLE NITROGEN ISOTOPES AS A TOOL FOR UNDERSTANDING NITROGEN CYCLING PROCESSES AND MECHANISMS RESPONSIBLE FOR NITROUS OXIDE PRODUCTION By Robin Leslie Sutka A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Geological Sciences 2002 UMI Number: 3064317 ® UlVII UMI Microform 3064317 Copyright 2002 by ProQuest lnfon'nation and Learning Company. All rights reserved. This microform edition is protected against unauthorized copying under Title 17, United States Code. ProQuest Information and Learning Company 300 North Zeeb Road PO. Box 1346 Ann Arbor, MI 48106-1346 ABSTRACT STABLE NITROGEN ISOTOPES AS A TOOL FOR UNDERSTANDING NITROGEN CYCLE PROCESSES AND MECHANISMS RESPONSIBLE FOR NITROUS OXIDE PRODUCTION By Robin Leslie Sutka The stable nitrogen isotopic composition of dissolved nitrate was determined at six stations ranging from the oligotrophic North Pacific Subtropical Gyre (NPSG) to the eutrophic Eastern TrOpical North Pacific (ETNP). Dissolved inorganic nitrogen, oxygen and phosphate concentrations were determined in water column samples at all six stations. In productive, oxic waters nitrate isotopically enriched in 15N (maximum 515N- NO3' value of 12.5 0/00) was most likely the result of assimilatory nitrate reduction. In contrast, high OlSN-Nog' values (maximum of 12.3 °/oo), associated with high nitrate deficits and anoxic conditions, supported the interpretation of isotopic fractionation due to denitrification. A one-dimensional vertical advection and diffusion model was used to estimate the fractionation factor for denitrification at two stations in the ETNP. A comparison of modeled to observed 8'5N-NO3' provided estimates of the isotopic enrichment factor (8) of 30 at station 4 and 30 to 35 at station 5. Isotopically light nitrate (1.1 and 3.2 °/oo) was observed in the upper 200 m of the water column at stations in the ETNP. Tracer studies of 15NIL; and biogeochemical indicators of nitrogen fixation, such as N/P ratios, OlsN-POM and T richodesmium abundance, supported the interpretation of nitrification as the most plausible explanation for low 6‘5N-NO3' values observed in upper water column samples. Nitrification rates increased across the transect from a maximum rate of 1 nmol TI (1'1 at station 1 and 23.7 nmol 1'1 (11'1 at station 6. The relative importance of individual microbial pathways to nitrous oxide production is not well known. The intramolecular distribution of 15 N in nitrous oxide provides a basis for distinguishing biological pathways. Concentrated cell suspensions of M. capsulatus Bath and N. europaea were used to investigate the Site preference of nitrous oxide by microbial processes during nitrification. The average site preference of nitrous oxide formed by hydroxylamine oxidation by M. capsulatus Bath (5.5 +/- 3.1 ° 00) and N. europaea (-1.4 +/- 1.7 °/oo) and nitrite reduction by N. europaea (-7.7 +/- 3.1 ° 00) differed significantly (AN OVA, fags): 247.9, p=0). These results demonstrate that the mechanisms for hydroxylamine oxidation are distinct in M. capsulatus Bath and N. europaea. The average SRO-N20 values of nitrous oxide formed during hydroxylamine oxidation by M. capsulatus Bath (53.1 +/— 2.9 °/oo) and N. europaea (~23.4 +/- 7.2) and nitrite reduction by N. europaea (4.6 +/- 1.4) were also significantly different (ANOVA, fags): 279.98, p=0) and suggests although the nitrogen isotope value of the hydroxylamine was similar, the A15 N associated with hydroxylamine oxidation by M. capsulatus Bath and N. europaea (-2.3 and 26.0 °/oo for M. capsulatus Bath and N. europaea respectively) provided evidence that differences in isotopic fractionation were associated with two mechanisms. The site preferences in this study are the first measured values for isolated microbial processes. The differences in site preference are significant and indicate that isotopomers provide a basis for apportioning biological processes of nitrous oxide. Dedicated to the memory of my friend and father, Robert Sutka. iv ACKNOWLEDGEMENTS I would like to thank my mom and sister for supporting me through the last 5 years. My family has often not known what I am doing, but they support me unconditionally. I like to acknowledge the role of Hasand Gandhi as a friend and teacher. He is undoubtedly the most patient and kind person that I have ever known and I will miss him very much. I would like to thank Robyn Engleright for her friendship and support that have made the last year more pleasant. I would like to thank my committee for support and guidance through the last five years. Especially my advisor Peggy Ostrom for her support through the good and bad times. I would also like to thank Nathaniel Ostrom for his creative approaches toward solving problems. I have learned a great deal about mass spectrometry in the laboratory from both Peggy and Nathaniel. In addition, they have become good friends that have been a wonderful source of support. Thanks also to Dr. John Breznak for his advice on experimental design. Through his vast expertise and great skill as a teacher he was able to provide great options, but allow me the freedom of deciding what would meet my experimental needs. I would like to thank Dr. David Hyndman for serving on my committee and providing support with the modeling efforts. My dissertation would not be possible without the financial support that I received throught my teaching assistantships in the Department of Chemistry and the Department of Geological Sciences. I also received additional funding from the Graduate school that made summer research and completion of my dissertation possible. In addition, I would like to thank Dr. David Emerson for providing me some wonderful motivation for finishing my dissertation in order to work on what will certainly be interesting research. TABLE OF CONTENTS LIST OF TABLES .......................................................................... viii LIST OF FIGURES ........................................................................ ix INTRODUCTION ......................................................................... 1 CHAPTER 1 Stable nitrogen isotope dynamics of dissolved nitrate in a transect from the North Pacific Subtropical Gyre to the Eastern Tropical North Pacific. Abstract .............................................................................. 3 Introduction .......................................................................... 4 Materials and methods ............................................................. 6 Results and Discussion ............................................................ 8 General water column characteristics .................................. 8 Nitrogen isotope biogeochemistry ..................................... 9 Conclusion ......................................................................... 19 CHAPTER 2 Nitrogen isotopomer site preference of nitrous oxide produced by Nitrosomonas europaea and Methylococcus capsulatus Bath. Abstract ............................................................................ 37 Introduction ........................................................................ 38 Materials and Methods ........................................................... 42 Culture and media ....................................................... 42 Concentrated cell suspensions ......................................... 43 Mass spectrometric analysis of nitrous oxide and isotope units...45 vi Standard characterization ................................................ 45 Controls ..................................................................... 46 Results and Discussion ............................................................ 47 Standard characterization ................................................ 47 Controls .................................................................... 47 Bulk nitrogen and oxygen isotopic composition of nitrous oxide.48 Nitrogen isotopomeric site preference of nitrous oxide ............. 50 Conclusion ................................................................ 53 Bibliography ................................................................................ 61 vii LIST OF TABLES Table 1. Summary of nitrous oxide data including concentration, site preference, 815N- N20 and O'BO-NZO from concentrated Methylococcus capsulatus Bath cell cultures with hydroxylamine as the substrate. Each letter (A, B, C and D) corresponds to a separate experiment and number indicates the time course sample. Nitrous oxide formation via hydroxylamine oxidation is stimulated in this set of experiments. ‘Anomalous value (37.7 0/00) omitted in the calculation of average and standard deviation of the 231.80-N20smow .......................... page 56. Table 2. Summary of nitrous oxide data including concentration, site preference and ONO-N20 from concentrated Nitrosomonas europaea cell cultures with hydroxylamine as the substrate. Each letter (A, B, C and D) corresponds to a separate experiment and number indicates the time course sample. Nitrous oxide formation by hydroxylamine oxidation is stimulated in this set of experiments. ....................................................................................... page 57. Table 3. Summary of nitrous oxide data including concentration, site preference and 8180-N20 from concentrated Nitrosomonas europaea cell cultures with nitrite as the substrate. Each letter (A, B, C and D) corresponds to a separate experiment and number indicates the time course sample. Nitrous oxide formation via nitrite reduction is stimulated in this experiment ..................................... page 58. Table 4. Summary of istopomeric compositions of nitrous oxide in the literature with an emphasis on biological pathways ............................................ page 59. viii LIST OF FIGURES Figure 1. Locations of sampled stations in a transect from west to east ......... page 22. Figure 2. Dissolved nitrate concentration , oxygen concentration, nitrate deficits, S'SN-NO3' values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 1 ............................... page 23. Figure 3. Dissolved nitrate concentration , oxygen concentration, nitrate deficits, OlsN-NO3’ values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 2 ............................... page 25. Figure 4. Dissolved nitrate concentration , oxygen concentration, nitrate deficits, S'SN-NO3' values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 3 ............................... page 27. Figure 5. Dissolved nitrate concentration , oxygen concentration, nitrate deficits, 815 N-NOg' values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 4 ............................... page 29. Figure 6. Dissolved nitrate concentration , oxygen concentration, nitrate deficits, S'SN-Nog’ values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 5 .............................. page 31. Figure 7. Dissolved nitrate concentration , oxygen concentration, nitrate deficits, SlsN-NO3' values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 6 .............................. page 33. ix Figure 8. Comparison of modeled to observed nitrate concentrations and 5'5N values of dissolved nitrate at station 4. Parameters used at station 4 include a K of 7.65E-5 m2 sec", R0 of 0.85E-7 pmol L" see“, (I) of1.0E-3 m" and t1 ofO.76E-2 m". ...................................................................................... page 35. Figure 9. Comparison of modeled to observed nitrate concentrations and S'SN values of dissolved nitrate at station 5. Parameters used at station 5 include a K = 7.65E-5 m2 sec", R0 = 1.35E-7 nmol L" sec", 0) = 1.0E-3 m'1 and p = 0.76E-2 m'l. ....................................................................................... page 36. Figure 10. Keeling plot approach to characterizing the site preference of the nitrous oxide standard where Qa = moles of 98 % 15N"‘-nitrous oxide and Qm = total moles of nitrous oxide ............................................................ page 55. Figure 11. Proposed pathway for the formation of nitrous oxide from the reduction of nitrite showing involvement of an enzyme-bound intermediates modified from Weeg-Aerssens et al. (1988) ..................................................... page 60. INTRODUCTION Increases in the atmospheric concentration of the important greenhouse gas (nitrous oxide) N20, have stimulated an interest in identifying its sources and sinks and characterizing mechanisms of production of this important greenhouse gas. The ocean environment is a net source of N20, but the relative importance of the oceans compared to other environments is not well known (Yoshinari, 1997). In the oceans, two microbial processes, nitrification and denitrification, produce N20. Isotopic data have suggested that in oxic waters of the Pacific Ocean, nitrification is the dominant source of the trace gas N20 (Kim and Craig, 1990; Dore and Karl, 1996; Ostrom et al., 2000). However, in anoxic environments, denitrification likely plays an important role in the flux ofN20 (J orgenson et al., 1984). In addition, N20 can be produced during methanotrophic nitrification (Yoshinari, 1985). The first objective of this dissertation research was to delineate nitrogen cycling in oceanic environments and develop methods to identify specific nitrogen cycle process that produce N20. The second objective was to define the isotopomeric composition of N20 formed during distinct microbial pathways. The first chapter details research that is part of a collaborative effort between Michigan State University and the University of Hawaii to investigate the controls on the production of N20 in the oceans. Nitrogen isotope biogeochemistry of dissolved nitrate was examined in a west to east transect extending from the NPSG to the ETNP, where a change in environmental conditions from oligotrophy to eutrophy occurs. While N20 was not investigated in this study, delineation of microbial processes in the water column provides insight into the production mechanisms for N20 in related studies. In the second chapter, laboratory studies were done to characterize the isotopomeric composition of N20 produced during hydroxylamine oxidation by N. europaea and M. capsulatus Bath and nitrite reduction by N. europaea. The isotopomeric data were the first reported values for autotrophic nitrification, methanotrophic nitrification, and nitrifier denitrification. Applications of this data include understanding in situ sources of N20 in oceanic and soil environments. CHAPTER 1 STABLE NITROGEN ISOTOPE DYNAMICS OF DISSOLVED NITRATE IN A TRANSECT FROM THE NORTH PACIFIC SUBTROPICAL GYRE TO THE EASTERN TROPICAL NORTH PACIFIC ABSTRACT The stable nitrogen isotopic composition of dissolved nitrate was determined at six stations ranging from the oligotrophic North Pacific Subtropical Gyre (NPSG) to the eutrophic Eastern Tropical North Pacific (ETNP). Dissolved inorganic nitrogen, oxygen and phosphate concentrations were determined in water column samples at all six stations. In productive, oxic waters nitrate isotopically enriched in 15N (maximum OlsN-NO3' value of 12.5 0/00) was most likely the result of assimilatory nitrate reduction. In contrast, high SUN-N03” values (maximum of 12.3 °/oo) in association with high nitrate deficits and anoxic conditions supported the interpretation of isotopic fractionation due to denitrification. A one-dimensional vertical advection and diffusion model was used to estimate the fractionation factor for denitrification at two stations in the ETNP. A comparison of modeled to observed SlsN-NO3' data indicated an isotopic enrichment factor (a) of 30 at station 4 and 30 to 35 at station 5. Isotopically light nitrate (1.1 and 3.2 0/00) was observed in the upper 200 m of the water column at stations in the ETNP. Tracer studies of 1"NFL; and biogeochemical indicators of nitrogen fixation, such as N/P ratios, OlsN-PN and T richodesmium abundance, supported the interpretation of nitrification as the most plausible explanation for low 5'5 N-N03' values observed in water column samples. The nitrification rates increased along the transect from a maximum rate of 1 nmol l'1 d'1 at station 1 to 23.7 nmol l‘1 d'1 at Station 6. INTRODUCTION Oceans account for approximately half of the world’s primary production (Karl, 1999). This productivity is fueled by recharge of nutrients from deep waters and in most of the world’s oceans, the availability of fixed nitrogen limits productivity (Howarth, 1988). The availability of nitrogen is not homogeneous and this results in variations in ecosystem productivity. Two systems with contrasting nitrogen dynamics include subtropical/tropical gyres and coastal upwelling zones. Subtropical/tropical gyre ecosystems are characterized by oligotrophic conditions and were once thought to be marine deserts (Karl, 1999). However, given their large size (40 % of the Earth’s surface area), gyre ecosystems can contribute substantively to global primary productivity (Karl, 1999). In contrast, the fraction of the world’s oceans characterized by coastal upwelling zones is small, but eutrophic conditions support a disproportionately large amount of primary productivity. The North Pacific Subtropical Gyre (NPSG) is the largest open ocean gyre (area of 2 x 107 kmz) (Karl, 1999). In this ecosystem, nitrogen and phosphorus are supplied to the surface by diffusion across the thennocline and horizontal transport fi'om adjacent waters (Karl, 1999; Reid et al., 1978). However, the observation that the amount of primary productivity in the NPSG cannot be accounted for by the supply of nutrients by diffusion from deep waters led to a reevaluation of nitrogen cycling (Letelier and Karl, 1998). The periodic appearance of T richodesmium blooms at station ALOHA and an increase in the N/P ratio of dissolved and particulate matter provided evidence for a significant role of nitrogen fixation in the NPSG (Karl et al., 1997). Quantitative estimates indicate that diazotrophic microbes could fuel up to half of the new production in the NPSG ( Karl et al., 1997). In the ETNP, nutrients are supplied by upwelling off the Central American coast. The large flux of sinking organic carbon from primary production increases oxygen demand. The result of this increased respiration is a large lens of oxygen deficient water in the ETNP. The suboxic conditions and upwelled nitrate provides the necessary environment for water column denitrification. The associated loss of fixed nitrogen is significant in the region (10 x 1012 to 30 x 1012 g N per year, Cline and Kaplan, 1975) and variations in denitrification rates may contribute to glacial-interglacial changes in the atmospheric concentration of carbon dioxide. In addition to effects on past climate change, the production of nitric oxide and N20 at the oxic/anoxic interface in these waters has important consequences on future climate change (Ward, 2000). Studies of ecosystems that differ with respect to nutrient supply provide an important framework for understanding nitrogen cycle controls on primary productivity and climate change (Yoshida et al., 1984 Ganeshram et al., 1995; Dore et al., 1998; Karl, 1999). The current study contributes additional information on nitrogen cycling through the analysis of samples across a west to east transect. The stable nitrogen isotopic composition of dissolved nitrate and l5N-NI-L, oxidation studies are used make interpretations of predominant microbial processes. This suite of data allows the evaluation of nitrogen cycling and primary production in two contrasting systems, the NPSG and ETNP. MATERIALS AND METHODS Water column samples were collected during a May-June 2000 sampling cruise that was part of the Eastern Pacific Redox Experiment (EPREX) aboard the RN Roger Revelle. Six stations were sampled during the cruise along a west to east transect from the oligotrophic North Pacific gyre to the eastern tropical North Pacific (ETNP) (Station 1: 22.7°N, 158°W; Station 2: 16.0°N, 150°W; Station 3: 16.0°N, 136°W; Station 4: 15.8°N, 119.1°W; Station 5: 16.3°N, 107.2°W; Station 6: 15.6°N, 98°W) (Figure 1). The water column was characterized at each station by determining temperature, salinity, dissolved oxygen and chlorophyll fluorescence. Water samples were collected to determine nitrate concentration and stable nitrogen isotope values. Nitrate was extracted from the seawater samples by a distillation method (Ostrom et al., 1997). During the distillation NH] was converted to NH3 by addition of 4 mL of distilled 5 N NaOH to adjust the pH above 10. Finely ground Devarda’s alloy (0.3 g) was added to the sample to reduce N03' to NH3 for a second distillation. Approximately 250 mL of condensate was collected in a flask containing 30 mL of 0.084 N HCl. The sample was bound to 100 mg of zeolite molecular sieve in two successive bindings and dried at 40 °C. Zeolite bound samples were placed in a quartz tube to which excess copper and copper oxide were added, evacuated, sealed and heated to 850 °C for 1 hour. and slowly cooled (Macko, 1981). Nitrogen (N2) was purified cryogenically on a vacuum line and analyzed on a Prism isotope ratio mass spectrometer (Micromass, Manchester, UK). Samples were corrected for background associated with addition of Devarda’s alloy using an isotope mass balance equation (Ostrom et al., 1998). Accuracy and precision of replicate samples for stable nitrogen isotopic analysis of dissolved nitrate are typically better than 1 °/oo. The stable nitrogen isotopic composition is expressed in per mil notation (°/oo): 5‘5N = [(‘5N/‘4N,,m,../ 'SN/‘4N,,,,.d,,d) - 1] x 1000 (Equation 1) The isotopic standard for 6'5 N is atmospheric N2. The concentration of dissolved ammonium in samples was determined during shipboard analyses using a solvent extraction techniques (Brzezinski, 1988). Dissolved oxygen concentrations were determined by Winkler titration (Carpenter, 1965). Analyses of dissolved nitrate, nitrite and phosphate concentrations were determined at the University of Washington (Unesco, 1994; Valderrama, 1981). For determination of nitrification rates via 15N-ammonia oxidation, seawater samples were collected at the depth of intended in situ incubation. Typically, samples were taken at the nitracline, chlorophyll maximum, upper edge of the chlorophyll maximum and at a depth 150 m below the chlorophyll maximum. To determine nitrification rates 4 L of seawater was collected at selected depths below the chlorophyll maximum in the water column. Isotopically enriched (99.9%15N-NH4CI) was added to the seawater that was transferred into glass bottles. The 4 L bottles were lowered back to the depth that they were taken and incubated in situ for 24 or 48 hours. At the end of the incubation period, the bottles were brought to the surface, filtered, transferred to Nalgene bottles and frozen until analysis. For determination of the S‘SN value of the dissolved nitrate in the incubation samples, the nitrate was first extracted by distillation. A sample volume of 1 L was distilled following a modification of the distillation technique described by Ostrom et al. (1997). The modification was necessary to ensure complete removal of residual 15N-NH4 from the seawater and distillation glassware. To remove lSN-NH4, the sample was distilled three times. To remove trace NH4 from the distillation apparatus, the glassware was washed in 1 N HCl and heated to 500°C for 2 hours. Tests of this technique before sample analysis proved that the modifications successfully removed residual ‘5 N from the distillation apparatus. The calculation of nitrification rates from the oxidation of 15NH4 was based on the equations described in Glibert and Capone (1993) Nitrate deficits were calculated using equations of Cline and Richards (1972). The amount of nitrate used as a substrate during denitrification is the difference between the expected nitrate concentration and the observed concentration of nitrate plus nitrite. N03'(¢xp¢c,cd) = 14.8 x P04+ (measured) (Equation 2) N03' deficit = N03'(cxp¢ctcd) — (N 03'(m¢a,u,cd) + N02'(mcasu,cd)) (Equation 3) Nitrogen to phosphorus ratios of inorganic dissolved nutrients were calculated according to Fanning (1992). N/P = ([N02'] + [N0;'] + [NH4+])/ [POX] (Equation 4) RESULTS AND DISCUSSION General Water Column Characteristics Stations along the transect (Figures 2 to 7) capture the gradual broadening, both horizontally and vertically of anoxic water from west to east and concomitant change from oligotrophic conditions in the NPSG to eutrophic conditions in the ETNP (Cline and Kaplan, 1975; Ward and Zafiriou 1988; Karl, 1999). The oxygen minimum zone occurs at an increasingly shallow depth from west to east with a change from 700 m at station 1 to 65 m at station 6 (Figures 2 and 7). Furthermore, minimum oxygen concentrations at station 1 are 30 uM and decrease to <1 pM at Station 6. The decreasing depth of the oxygen minimum zone is associated with a change in depth of the chlorophyll maximum from 140 m at station 1 to 40 m at Station 6 (Figures 2 and 7). The fluorescence increases from west to east with a maximum of 0.25 RFU at Station 1 to 0.95 RFU at station 6 (Figures 2 and 7). Ammonium concentrations in surface waters at station 1 are between 0 and 20 nM and at stations 4, 5 and 6 increase to 100 to 300 nM (Figures 2, 5, 6 and 7). This is expected due to the enhanced productivity and subsequent mineralization of organic matter in the eastern portion of the transect. Nitrogen Isotope Biogeochemistry The primary microbial processes that influence the nitrogen isotopic composition of nitrate include denitrification, assimilatory nitrate reduction, nitrogen fixation and nitrification (Cline and Kaplan, 1975; Mariotti et al., 1981; Liu et al., 1996). During these microbial-driven nitrogen transformations, isotopic discrimination can result due to a slight difference in the enzymatic reaction rate between 15 N and l4N containing compounds (Mariotti et al., 1981). The result is an accumulation of 14N in the product and 15N in the residual substrate (Mariotti et al., 1981). Nitrification and nitrogen fixation can result in nitrate that is depleted in '5 N. The fractionation factor associated with nitrogen fixation is small and produces fixed nitrogen with a nitrogen isotopic composition approximately equal to that of atmospheric nitrogen 0 0/00 (Hoering and Ford, 1960; Minagawa and Wada, 1978). Isotopically light nitrate can be produced by magnification of fixed nitrogen and subsequent nitrification. In Kuroshio waters, l4N enriched nitrate (1 to 3 °/oo) was attributed to the activity of unusually high numbers of a nitrogen fixer, T richodesmium (Liu et a1, 1996). In contrast to nitrogen fixation, laboratory studies reveal that microbial nitrification has a relatively large isotopic enrichment factor (a of -35 °/oo) (Mariotti et al., 1981; Yoshida, 1988), where 8 is defined by: s,,,,=(or-1)103 (Equation 5) Where: a = k('4N)/k(‘5N) k(l4N) = the reaction rate constant for molecules containing 14N k(15 N) = the reaction rate constant for molecules containing 15N In Conception Bay, nitrification was proposed as the likely cause of low 515 N values of dissolved nitrate (515 N-NO3') (average 0.2 °/oo) (Ostrom etal., 1997). Nitrate enriched in 15N has been associated with the processes of denitrification and assimilatory nitrate reduction by phytoplankton (Cline and Kaplan, 1975; Brandes et al., 1998 and Voss et al., 2001). During the reduction of N03' to N2, the residual pool of N03' becomes increasingly enriched in 15 N. This can be substantive due to a large 8 associated with denitrification of approximately -30 in previous studies in the ETNP (V 055 et al., 2001; 8 = -27 +/- 3 °/oo, Brandes et al., 1998). In anoxic regions of the ocean, denitrification has resulted in OISN-NO3' values as high as 18.7 ° 00 (V 085 et al., 2001). In oxic waters, preferential assimilation Of '4NO3' by phytoplankton can also result in high SUN-N03] (Cline and Kaplan, 1975). However, the fractionation due to assimilatory nitrate reduction, 8 of O to -19, is not as large as that associated with denitrification (Wada and Hattori, 1978; Wu et al., 1997; Sigman et al., 1999; Altabet et al., 1999). 10 High 515 N-NO3' values observed in samples from suboxic (oxygen concentrations equal to or less than 1 11M) waters at our stations suggest denitrification (Figures 4 to 7). Maximum B'SN-N03' values of 11.8, 12.3 and 10.6 °/.,0 were observed in samples taken from stations 4, 5 and 6 respectively. Nitrate deficits within anoxic waters are generally high with maximum values of 14.0, 16.1 and 16.0 pM at stations 4, 5 and 6 and indicate active denitrification (Figure 5, 6 and 7). In the ETNP, enrichment in 15’N of nitrate in association with suboxic waters and high nitrate deficits has previously been attributed to denitrification (Cline and Kaplan, 1975; Brandes et al., 1998; Voss et al., 2001). The isotopic enrichment factor for denitrification was estimated using a model that incorporated the vertical advection and diffusion of waters in and out of the suboxic waters at stations 4 and 5. The typical model for estimating the fractionation factor for a single step reaction is derived from the Rayleigh equation (Mariotti et al., 1981; Macko et al., 1986). An important assumption in the Rayleigh model is that after the reaction is initiated, there are no additional inputs of substrate and that the reaction is unidirectional. Thus, estimating the fractionation factor for denitrification from a Rayleigh model in open ocean systems is inappropriate because there is an influx of nitrate from water bodies below the zone of denitrification. In open-systems, studies have employed a one- dimensional vertical advection and diffusion model to describe the concentration and isotopic composition of nitrate as a function of depth in the water column (Cline and Kaplan, 1975). The current study applied a vertical advection-diffusion model to estimate a for denitrification using water column data from 150 m to 700 m at stations 4 and 5 (Cline and Kaplan, 1975; Velinsky et al. 1991). Boundary conditions included a depth interval with a linear temperature and salinity relationship and apparent absence of 11 microbial processes other than denitrification on the nitrogen isotopic composition of dissolved nitrate. The first step is to find an analytical solution of the following differential equation (Equation 6) subject to the boundary conditions [N]z= z] = [N]1, [N]z = 22 = [N12- d2[N]/d22 + wd[N]'/dz — (Roe'”)/K = 0 (Equation 6) Where: 2 = depth (m) [N] = concentration of nitrate (uM) K = vertical eddy diffusion coefficient (m2 sec“) R0 = rate of denitrification (pmol L'l sec") 00 = mixing parameter (Vz/K) (m'l) Vz = vertical velocity (m see") p = describes the attenuation of microbial activity with depth (m") The solution of Equation 6 is used to determine the values for the parameters that provide the best comparison of modeled to observed nitrate concentrations. The parameters evaluated in the first step include K, R0, (0 and u. The second step is to obtain a numerical solution of the isotope portion of our model. Using an implicit finite difference approximation to Equation 7 produces a tridiagonal matrix. The system of equations was inverted and solved using the Thomas algorithm (Roache, 1998). We increased the computational mesh until successive refinements produced no additional change in the results. A typical step size used in the computations is ~ 0.5 m. dZC/dzz + to dC/dz —(1/a)C/[N] (Roo'W/K) + ((Ot-l)/Ot)( Room/K) 1000 = 0, (Equation 7) 12 Where: C = 5 ‘ [N] . o = o‘SN—No; (°/,,) or = isotopic fractionation factor A comparison of modeled to observed nitrate concentrations at stations 1 to 3 indicated that denitrification was not active at the time that we sampled the water column. The parameters used at station 4 that provided a reasonable comparison of modeled to observed nitrate concentrations include K = 7.65E-5 m2 sec", R0 = 0.85E-7 umol L'l sec", 0) = 1.0E-3 m‘1 and p. = 0.76E-2 rn'I (Figure 8). At station 5, the values for K, 00 and u were the same, but a value of R0 of 1.35E-7 pmol L'l sec'l was used (Figure 9). Despite a good comparison of modeled to observed nitrate concentrations at station 6, e was not estimated due to insufficient isotopic data in the modeled depth interval. At station 4, the model estimate was an e of -30. Since estimates of a were very sensitive to changes in 5N-N03' near the boundary, we define our estimates of 8 based on data near the center of the modeled depth interval. A comparison of modeled to observed SISN-NO3' at station 5 suggest an e of -30 to -35 (Figure 9). The non-ideal fit of the modeled to observed data at station 5 is likely due to the low number of observed OlSN-N03' values. A comparison of the results from station 4 to those at station 5 suggest that the isotopic fractionation factor is Similar despite a change in the rate of denitrification (R) from 0.85E-7 umol L'l sec“1 at station 4 to 1.35E-7 umol L'l sec'1 at station 5. Our estimates of the isotopic enrichment factor are within the range of previous studies in the ETNP (a = -27 to -60) (V OSS et al., 2001; Brandes et al., 1998; Cline and Kaplan, 1975) and closely resemble the value for majority of the estimates (a of -30). 13 In the upper water column (<200 m depth), multiple processes influence the distribution of inorganic nitrogen including assimilatory nitrate reduction, nitrogen fixation and nitrification. This leads to the potential for overlap of active processes within similar depth intervals. For example, in an incubation study using 15N substrates, Ward et a1. (1989) demonstrated that nitrification occurs simultaneously with the assimilation of nitrate by phytoplankton. Interactions between microbial-driven nitrogen transformations are further complicated in the ETNP where a shallow oxygen minimum zone promotes denitrification at the base of the euphotic zone. The potential for interaction between nitrification and denitrification at oxic/anoxic interfaces in the environment has been noted by Ward (1996). At the boundary of the interface, a flux of substrates and products can promote a coupling of the oxidation and reduction of different inorganic nitrogen species (Ward, 1996). As a consequence of the coupling and Spatial overlap of processes, interpretations of the isotopic composition of inorganic nitrogen in the upper water column often requires additional geochemical information such as nitrogen to phosphorus ratios, fluorescence and dissolved oxygen concentration. In productive, oxic waters high S'SN-N03‘ values that occur at stations 2, 5 and 6 are suggestive of assimilatory nitrate reduction (Figures 3, 6 and 7). In our data, this is most evident at station 6 where high SUN-N03” values of 8.3 and 10.0 ° 0, at 28 and 44 m, respectively, are coincident with the chlorophyll maximum (as interpreted from fluorescence data) (Figure 7). In addition, localized isotopic maxima in oxic surface waters at stations 2 and 5 likely reflect an influence of assimilatory nitrate reduction (Figures 3 and 6). This is consistent with previous interpretations of high SUN-N05 found in near-surface waters of the ETNP (Cline and Kaplan, 1975). 14 Low S'SN-N03' values (1.1-3.2 °/oo) in the upper 200 m depth are Observed in our samples from stations 3, 4, 5 and 6 (Figures 4 to 7). The processes of nitrogen fixation or nitrification could be invoked as the plausible cause of this isotopically light nitrate. In near—surface waters of the ETNP, low 515 N-N03' values (5 °loo) were attributed to nitrogen fixation (Brandes et al., 1998). Estimates from an isotope mass balance model indicated that as much as 20 % of the nitrate was derived from nitrogen fixation (Brandes et al., 1998). However, isotopic data alone cannot distinguish nitrate derived from nitrification and nitrogen fixation. Other lines of evidence indicative of nitrogen fixation include the 615 N value of particulate organic nitrogen (PON), nitrogen to phosphorus ratios and the abundance of nitrogen fixing organisms (Saino and Hattori, 1987; Liu et al., 1996; Karl, 1999). Studies in the Western Pacific and Sargasso Sea have suggested that isotopically light PON (f)15 N of -3 °/oo to 2 °/oo) is the result of nitrogen fixation activity (Saino and Hattori, 1987; Altabet et al., 1991). These low OISN values of PON occur in waters where T richodesmium, a nitrogen fixing organism, are present (Saino and Hattori, 1987; Altabet et al., 1991). The observation that a culture of T richodesmium had a BISN value of -l.6 °/,,o provides further evidence that low 5'5 N values of PON (SUN-PON) can result from nitrogen fixation (Carpenter et al., 1997). Within the same depth interval where we observed isotopically light nitrate (ETNP), 8'5N-P0N values reported by Voss et al. (2001) between 4.4 and 8.2 ° 00 are not indicative of nitrogen fixation. In addition to the nitrogen isotopic composition of PON, the nitrogen to phosphorus ratio of dissolved and particulate matter is also used to indicate nitrogen fixation. This is based on the observation that cells of nitrogen fixing organisms are 15 characterized by nitrogen to phosphorus ratios in excess of Redfield (16:1 and 44:1 for Redfield and nitrogen fixing cells respectively) (Carpenter, 1983). Nitrogen to phosphorus ratios of dissolved inorganic nitrogen and phosphorus were demonstrated to be significantly higher than Redfield (40:1) in Kuroshio water northeast of Taiwan where unusually abundant populations of nitrogen fixers were found (Liu et al., 1996). In previous studies at station 1, an increase above the Redfield ratio to 20:1 or higher in the nitrogen to phosphorus ratio of the dissolved and particulate pools was interpreted by Karl et al. (1997) as evidence for the importance of nitrogen fixation. Our data indicate that maximum nitrogen to phOsphorus ratios of total dissolved inorganic nitrogen to total phosphorus were never higher than 12.1 :1 in the upper 200 m of all stations (Figures 2 to 7). In the region of our transect where we observed isotopically light nitrate the nitrogen to phosphorus ratios are sub-Redfield and not consistent with a significant input of nitrogen via nitrogen fixation. Station 1 (ALOHA) was demonstrated by Karl et al. (1997) to have seasonal (July to October) and ENSO period increases in the significance of nitrogen fixation. However, we sampled station 1 in May and an El Nino event did not take place in 2000. The abundance of T richodesmium is potentially indicative of the significance of nitrogen fixation in marine environments. In the western Pacific, isotopically light nitrate (1 to 3 0Am) is observed in waters characterized by a cell density of T richodesmium as high as 4.9 x 106 trichomes m'3 (Liu et al., 1996). At station 1, Karl et al. (1997) noted increases in T richodesmium abundance in association with seasonal increases in N to P ratios of particulate and dissolved matter. East of 130 °W in the North Pacific T richodesmium has not been observed (Carpenter, 1983). Thus, the locations where 16 isotopically light nitrate were observed in our current study fall outside the easterly range of T richodesmium. Although the influence of nitrogen fixation from other genera (e.g. Synechococcus, Cyanothece and Erythrosphaera) in these waters cannot be discounted, T richodesmium is the dominant diazotroph in tropical and subtropical oceans (Carpenter et al., 1997 Zehr et al., 2000). Thus we conclude that aside from isotopically light nitrate, OISN values of PON, nitrogen to phosphorus ratios and T richodesmium abundance do not suggest nitrogen fixation as a source of isotopically light nitrate. Thus, we consider nitrification as a likely mechanism. Previous incubation studies have found nitrification rates of 1 to 137.4 nmol L'1 d'1 at ALOHA (station 1) and a maximum of 20 nmol L'l d'l in the ETNP (Dore and Karl, 1996; Ward and Zafiriou, 1988). Our incubation studies document that nitrification occurs at every station in this study (Figures 2 to 7). At station 1, maximum nitrification rates of 1 nmol L'l (II are within the range of previous studies (1 - 137.4 nmol L'1 d") (Dore and Karl, 1996) (Figure 2). The large variation in nitrification rates at station 1 could be due to differences in sampling depth or possible temporal dynamics of nitrification rates. The data for station 6 (maximum of 23.7 nmol L'1 d") are similar to the rate of 20 nmol L'1 d'1 from a study in the ETNP by Ward and Zafiriou (1988) (Figure 7). There was an increase in nitrification rates from west to east in the transect with a maximum rate at station 6 about 20 times greater than station 1. High ammonium concentrations (100 toi300 nM) at stations 5 and 6 indicate ample substrate for nitrification. The data also indicate that the highest rates of nitrification generally occur at the base of the chlorophyll maximum (Figures 2 to 7). The depth distribution of rate measurements could be explained by light inhibition of 17 nitrifiers and competition for ammonium with phytoplankton. Culture studies have indicated that nitrifiers are inhibited by high light (Horrigan et al., 1981). At the depth that phytoplankton are typically most active high light levels and are inhibitory to nitrification. As light intensity decreases with depth, nitrifiers are less inhibited and phytoplankton becomes less active. The data demonstrate that minima in SlsN-N03' occurs where nitrification was active. Most notably at station 4, the maximum nitrification rate of 15.8 nmol L’1 d'1 and the minimum SUN-N05 of 3.2 °/oo coincide (Figure 5). However, maximum nitrification rates do not always coincide with isotopic minima. For example at station 5, the nitrification rate is highest at a depth of 65 m, but the 515N-N03’ of approximately 9 °/00 at a depth of 75 m suggests a predominance of assimilatory nitrate reduction (Figure 6). This is consistent with a study by Ward et al., (1989) that demonstrated that nitrification and assimilatory nitrate reduction activity at the same depths. The influence of assimilatory nitrate reduction can obscure the isotopic signature of nitrification. The decrease in the 6'5N-N03' with depth at stations 5 and 6 above 150 suggests that the influence of assimilatory nitrate reduction on the isotopic composition of nitrate decreases with depth and the importance of nitrification increases. Thus, isotopically light nitrate observed in the data at stations 5 and 6 (128 and 104 m, respectively) reflects the predominant influence of nitrification processes despite maximum nitrification rates at a depth of 65 m (Figures 6 and 7). In combination with knowledge of isotopic fractionation during nitrification, the incubation studies support the interpretation that low SlSN-N03 (1.1-3.2 °/oo) in the data reflect the influence of nitrification rather than nitrogen fixation. 18 CONCLUSIONS Variations in the relative importance of microbial processes of the nitrogen cycle as a function of ecosystem conditions have implications for trace gas production and primary productivity. Low O‘SN-N03' values in the ETNP have been interpreted to be indicative of nitrogen fixation (Brandes et al., 1998). In the past, inputs of nitrogen via fixation and the abundance of diazotTOphic organisms in the open ocean were considered insignificant (Ward, 2002; Paerl and Zehr, 2000). More recently, studies of the NPSG have revealed a significant contribution of nitrogen fixation to primary production (Karl et al., 1997 and Capone er al., 1997). The ETNP is a major site for the loss of fixed nitrogen via denitrification and a recent study suggests that nitrogen fixation has the potential to balance the nitrogen losses (Brandes et al., 1998). However, our results indicate that nitrification is an alternative explanation for isotopically light nitrate in the euphotic zone of the NPSG and ETNP during the period of our study. In contrast to nitrogen fixation, nitrification does not increase the pool size of available nitrogen. Furthermore, the oxidation of ammonia to nitrate results in coupling of nitrification and denitrification that contributes to a loss of fixed nitrogen to the system. If the loss of nitrogen via coupled nitrification/denitrification in the euphotic zone is found to be significant, it has the potential to impact estimates of productivity and organic matter export. The distribution of nitrification activity in the water column is inconsistent with the new production paradigm of Dugdale and Goering (1967). A central component of this paradigm is that increases in the productivity result from delivery of new sources of nitrogen such as nitrate from deep water or nitrogen fixation (Dugdale and Goering, 19 1967). Thus, nitrate delivered to the euphotic zone via upwelling from deep waters and nitrogen derived from fixation are termed “new nitrogen” (Dugdale and Goering, 1967; Eppley and Peterson, 1979). In contrast, ammonia is termed “regenerated nitrogen” since it is the product of Short-term recycling processes in the euphotic zone (Dugdale and Goering, 1967). In order to maintain steady state levels in productivity, imports of new nitrogen and exports from the system in the form of sinking particles and fisheries harvest must be balanced (Eppley and Peterson, 1979). An important assumption is that nitrification occurs solely in deep waters. It is now recognized that nitrification in the euphotic zone is significant and contributes to the phytoplankton nitrate demand (Dore and Karl, 1996; Ward, et al., 1989). Our results support the interpretation of these studies (Dore and Karl, 1996; Ward, et al., 1989) and for the first time provide corroborating stable nitrogen isotopic evidence for euphotic zone nitrification. It is clear that euphotic zone nitrification must be accounted for in models of primary productivity and studies of nitrogen isotope biogeochemistry. The production of isotopically light nitrate in the euphotic zone has implications for interpretation of nitrogen isotope signals in modern and ancient environments. Variations in the 5'5 N of sedimentary organic matter are thought to be indicative of past changes in nutrient inventories or utilization (Ganeshram et al.,1995 and Farrell et al., - 1995). Ganeshram et al., (1995) suggested that nitrate concentrations in oxygen deficient waters are regulated by variations in the rate of denitrification (Ganeshram et al., 1995). They assumed that the isotopic composition of organic matter produced in near surface waters and deposited in sediments is a function of fractionation during phytoplankton uptake and the isotopic composition of the source nitrate (Ganeshram et al., 1995). A 20 second assumption is that nutrients in near surface waters of the ETNP are completely utilized by phytoplankton. Thus, the nitrogen isotopic composition of the organic matter should be identical to that of its source (nitrate) and denitrification controls the 6'5 N of nitrate (Ganeshram er al., 1995). Increases in denitrification would be reflected as an increase in the nitrogen isotopic composition of nitrate and associated organic matter. However, euphotic zone nitrification could obscure the isotopic signature of denitrification by contributing MN enriched nitrate to the nutrient pool. Therefore, the expectation that the nitrogen isotopic composition of organic matter is a function of the extent of denitrification may not be entirely correct. An alternative explanation for fluctuation in the nitrogen isotopic composition of sedimentary organic matter is variation in the importance of euphotic zone nitrification. This could result from increases in upwelling, which would stinrulate primary productivity and subsequently increase the supply of regenerated NHa to support euphotic zone nitrification. At present time, the relative importance of increases in the rate of denitrification versus euphotic zone nitrification as mechanisms to influence the isotopic composition of nitrate during glacial periods is unknown. However, our results definitively demonstrate that it is not appropriate to conclude that changes in the rate of denitrification and fractionation during assimilatory nitrate reduction are the only processes driving sedimentary nitrogen isotopic signals. 21 >>o om we on . ..... 853cm Zo om ammo 9 «we; E0: Bowen: m :_ 25sz 8.953 Co 20383 .F 959“. 22 Figure 2. Dissolved nitrate concentration, oxygen concentration, nitrate deficits, OlsN-NO3' values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 1. 23 =25 E:_coEE< 9 com com 2:. $3... 355 2a.. cow—355: » d D _ gum—mu 29:: o 9. 9. up a m n o q and o wan—m: 3:33.02“? a “I 9.. «F a v 220 cod 88 .8 as. P :8an c 3338 me . T pppppppp cinema Asécogxoo ecu 2:. a co: _. - q ’ o be 09%0090000 q S. a.” cw 2 £328.? . -oma ficam eomw .cco .cmh .065 .cm0 -cco -cmm .ccm .cmfi .ocv -cmm -ccn .cmN -OGN .omw -cOv an ac Depth (m) .N 959“. 24 Figure 3. Dissolved nitrate concentration, oxygen concentration, nitrate deficits, S'SN-NO3' values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 2. 25 =25 E:_:OEE< com SN 2:. c A A 4. It cc... and Sum: 3:33.03.“— EL .25: 28 cozmuEbE A mNOvacvm 111..” L 1 m 8... or «r a v =88 28.... . 23¢ a m n o q q a « jl u q a «room: n=2 0 oocw .omtwm >32 .nOzizmvm A N .5sz ASE—.0903 0 cow 2:. c j.- _ . . o] w W g 0 u a £322.... .. v a.” ow. 3 a -oocw .cma -ccm -omm .ccw tank .005 -cmo -cco -cmm -com tome ccv rem” -ccn -cmN .ocN -cmw car an c Depth (m) .m 2:2“. 26 Figure 4. Dissolved nitrate concentration, oxygen concentration, nitrate deficits, OlsN-N03' values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 3. 27 gone“. 28:: A :23 5998 o 2. 2. up a o m a cow 2: c x . x . 1. .. H 61.82. 5mm .23 . ficmm -com .8» -2: .93 .30 -cmm .23 .93 63 . ._ m6 :3 . . . . . u e -ccn _ .cmw L , -Su . . - ago. -92 o r o , o om . 2: O. 0 e 0 an m p a d1 A r r . i » > rt » s t L i i ii A q p — c a A 8.. on... 8... 2 2 a e o N. a e a o 8 cw cu 2 o ADH—mv $0C00MQLOH=KI l\z O NOON .VIN 0CD“ tMOZIZer O Aziv wuflhgc I .9 mun—mum m cozmum N I @ Depth (m) is: see 28 cozmoztgco mu cm 3. 3. m o I 0000000000 :25 E:_:OEE< O can con 2:. c 28 Figure 5. Dissolved nitrate concentration, oxygen concentration, nitrate deficits, OlsN-N03' values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at Station 4. 29 :25 52:08:29 can ecu 2:. a L _ _ ”I. 8... . 8.5. 5...”: oucoumocoafi. < I 852 Nm m e c :ocmu 29:: o :23 59680 3. 3. 5. m o m a can 8.. c i. A A ”II: II A .F a)! - .93 .25 . .3» .25 $2 42:. 5mm .25 -omm scan $3 62‘ vcmn m8... womu .53 63. . , o . % o 2: o . o I Q on pl- - .m . t . . . . o in... L .25.. 9a.. cozmoEbE A macaw—.2...“ o III... I + 0 I annnnnnno O s I I I @fim Man I 0 9o O i m . t . _ t . n i ‘ N. a o n c ”02.2% A 8 on 8 S a n=z . £3 28:: . Doom .mtm 0:2. - v Caz—3m cco —. Depth (m) .m 3:9“. 30 Figure 6. Dissolved nitrate concentration, oxygen concentration, nitrate deficits, SUN-N03; values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at station 5. 31 :25 £229.53 0 8.” cow 3.. c q _ d cc... 3an 3:83.02“? :oeou 22:: . owns 53.8 o a 3 2 N. m o n a _ . . . :fiITeII 1 a . O62: o . - .-ccm .38 L 355 2a.. cos—35.x: A N 3 2 2 m a 11'! . L A and 2.6 9. N_. a Q 9 NF m o m c ..u. 0 r! w. . . _ . n 3. an an 2. o E2 oooow 3-9 0:2,. Oz.zo.wA 22$ 29:: u .o 2:9". mcozfiw 4 4 — i Depth (m) 32 Figure 7. Dissolved nitrate concentration, oxygen concentration, nitrate deficits, 615N-N03’ values, nitrification rates, N/P ratios, ammonium concentration and fluorescence of samples collected at Station 6. 33 .221. .8ng Or m —.=ON_"0—umoumm=cm O OON OOF O L a. I sell. I t - L . 83. - r Omm I I otOOO . . Omw . I 9 OOO - fiOmh i I # CON - r OmO . I 2 OOO t I .YOmm I 3 06m t I AT OM? a I OOv . I Own r I . OOn . . SN OON . II Om —. - I OO—. - A I I o o o Om r .. t .3.w +. O at _ q 1) 55 :8 32:5 . .. 82:25.3 9 92 5335.32 A OOm OON OO_. O mN ON mv O—. m O _ I d _ T. _ N 4 _ O , _|J.|I 4 a . a. t . L océ and cod 2. up w v c NF m o m a 3 on cm 2. o 3mm: 3:33.02"... E2 o ooom 5-2. ocasmozizmpmA 22322:: e 8:3 .N 2:5 Depth (m) 34 n -oz-z.=w :23 22:2 3 2 N_. 3 3 m N we 8 on an N ON 2:. . . . . 2: 33250 I .822 II 58 - $8 -8» - .8... G G u a SS. m . $3 m... m m. L8.” - 58 .SN - .SN _ _ F _ . OOF _ _ . L OOF 95822: 8\o m E 8\o CV. 98 cm. 5253 c332: mos—m.» 5:QO ._-E Wmcfio .«o 1 98 78 Two.— mo 8 .708 74 SEE Wmmmdmo em .708 NE Wmmofi CO V— N “.6205 v :ousm “a mum: BBoEENm .v 5:86 “a 38:: 32036 HO mos—g 22w 98 3235588 88:: @20on NO 533800 .w 053m 35 :23 3.3.2 €2.22» 3. 2. up 2. 2. m a mv 3. mm an mm cu . . . . . 2:. . . . . can 302030 I .822 l. I I com I I can r I can I I can 0 G a a . I Be m - - as. m M m. I I com I I can .. I com I I can _ p p p p COP _ _ _ h OOF 03080005 8\o m E 8? o7- was cm. 5053 030608 002? :0:QO .7E «-mobd u E 98 78 TB: u 8 .708 7..— BE: sum—mm.— " om .700m NE mImmcs n v— N 00205 m .593 “N v00: E80883” .m scram In 88:: “002830 «0 mos—N.» 22m 28 82358080 035: @2002: we 838800 .a 053m 36 CHAPTER 2 NITROGEN ISOTOPOMER SITE PREFERENCE OF N20 PRODUCED BY NITROSOMONAS E UROPAEA AND ME T H YLOCOCC US CAPSULA T US BATH ABSTRACT The relative importance of individual microbial pathways to nitrous oxide (N20) production is not well known. The intramolecular distribution of ‘5N in N20 provides a basis for distinguishing biological pathways. Concentrated cell suspensions of M. capsulatus Bath and N. europaea were used to investigate the site preference of N20 by microbial processes during nitrification. The average site preference of N20 formed by hydroxylamine oxidation by M. capsulatus Bath (5.5 +/- 3.1 ° 00) and N. europaea (-1.4 +/- 1.7 ° 00) and nitrite reduction by N. europaea (-7.7 +/- 3.1 ° 00) differed significantly (AN OVA, fags): 247.9, p=0). These results demonstrate that the mechanisms for hydroxylamine oxidation are distinct in M. capsulatus Bath and N. europaea. The average filsO-NZO values of N20 formed during hydroxylamine oxidation for M. capsulatus Bath (53.1 +/- 2.9 ° 00) and N. europaea (-23.4 +/- 7.2) and nitrite reduction by N. europaea (4.6 +/- 1.4) were significantly different (ANOVA, 11235): 279.98, p=0) . Although the nitrogen isotope value of the source hydroxylamine was similar, the A15 N associated with hydroxylamine oxidation by M. capsulatus Bath and N. europaea (-2.3 and 26.0 °/oo for M. capsulatus Bath and N. europaea respectively) provided evidence that differences in isotopic fractionation were associated with two mechanisms. The site preferences in this study are the first measured values for isolated microbial processes. The differences in site preference are significant and indicate that isotopomers provide a basis for apportioning biological processes of N20. 37 INTRODUCTION N20 is a trace gas that on a molecule for molecule basis is 200 times more effective than carbon dioxide as a greenhouse gas (Lacis et al., 1981). In addition, concentrations of N20 in the atmosphere have been increasing at a rate of 0.3% per year (Prinn et al., 1990; Rasmussen and Khalil, 1986). The rise in atmospheric concentrations of N20 is thought to result predominantly from microbial activity (J iang and Bakken, 1999) and not anthropogenic emissions. In certain environments, such as agricultural fields, management practices could mitigate emissions of carbon dioxide and trace gases by stimulating or inhibiting microbial processes (Robertson et al., 2000). Mitigation of future emissions by management in such environments depends on knowledge of the microbial processes responsible for trace gas production. To date, many studies have focused on defining the relative importance of nitrification and denitrification, but have not addressed the possibility of methanotrophic nitrification as a mechanism for the production of N20 (Blackmer et al., 1980; Kim and Craig, 1990; Dore et al, 1998, Naqvi et al., 1998; Perez et al, 2001; Ostrom et al., 2001). The possible contribution of methane oxidizing organisms to N20 production was demonstrated in a culture study by Yoshinari (1987). Defining the relative importance of methane oxidizers and autotrophic nitrifiers will be an important step in managing N20 emissions from terrestrial environments (Robertson et al., 2000). The microbial processes in soil and water that produce N20 include nitrification, denitrification, methanotrophic nitrification and heterotrophic nitrification (J iang and Bakken, 1999). Recently, the importance of autotrophic nitrification as a source of N20 in oceanic environments has been emphasized (Dore etal., 1998; Naqvi et al., 1998). 38 During nitrification, N20 is formed by two mechanisms. In the first mechanism, N20 is a by-product of the oxidation of hydroxylamine to nitrite by the enzyme hydroxylamine oxidoreductase (Jiang and Bakken, 1999). In the second, nitrite is reduced to N20 by the enzyme nitrite reductase (Ritchie and Nicholas, 1972; Poth and Focht, 1985). N20 production via nitrite reduction is stimulated by oxygen limitation (Ritchie and Nicholas, 1972; Poth and F ocht, 1985). This pathway was first documented in studies using the organism, Nitrosomonas europaea and is referred to as nitrifier denitrification (Ritchie and Nicolas, 1972; Poth and Focht, 1985). This term distinguishes it from denitrification by heterotrophic denitrifiers, although the enzymes involved are similar (Ritchie and Nicholas, 1974; Poth and Focht, 1985). Like autotrophic nitrifiers, methanotrophic bacteria are capable of oxidizing ammonia and hydroxylamine. An important consequence of ammonia oxidation by methanotrophs is the production of N20 (Hanson and Hanson, 1996). While the production of N20 by methane oxidizers has been evaluated in culture studies (Yoshinari, 1985), N20 production by methanotrophs in field-based studies has not been unequivocally demonstrated. Two studies provide important information on the role of methanotrophs in nitrification (Bodelier and Frenzel, 1999; Mandemack et al., 2000). Using competitive inhibition, Bodelier and Frenzel (1999) showed that methanotrophs were major contributors to nitrification in the rice rhizosphere. The results of Mandemack et al., (2000) suggest the involvement of methane oxidizers in N20 production in landfill soils. In incubations of these soils, the production of N20 was stimulated by pre-treatment with 1 % methane and rates were proportional to methane concentration (Mandemack et al. 39 2000). In addition, fatty acids isolated within the zone of maximum N20 flux were indicative of type H methanotrophs (Mandemack et al., 2000). While these studies provide evidence for a role of methane oxidizers as potential contributors to N20 production, they do not definitively demonstrate production of N20 by methanotrophs. The intramolecular distribution (isotopomers) of 15 N in N20 is emerging as a new tool for defining the relative importance of microbial sources of N20 (Toyoda and Yoshida, 2000; Perez et al, 2001; Popp et al., in press; Toyoda et al., in press). In the N20 molecule, the central nitrogen atom is preferred to as the alpha (or) and the terminal or end nitrogen atom is referred to as beta (B) (Toyoda and Yoshida, 1999). Quantifying the distribution of 15N in N20 is possible due to the fragmentation of N20+ to NO+ in the ion source of the mass spectrometer. The isotopic composition of 15N in the a position is determined from the fragment ion, NO+, after correction for rearrangement that occurs in the ion source (Toyoda and Yoshida, 1999). The relative abundance of 15 N in the molecule N20 is commonly expressed in 5 notation. 6 = [(Rsamplc/Rsmdard) — 1] * 1000 (Equation 8) Where: Rsamplc = 15’N/ 14N and ‘SO/wO for sample Rsmdard = 15N/MN and 18O/mO for standard The standards for nitrogen and oxygen isotope measurements are atmospheric N2 and VSMOW, respectively. The site preference is equal to the difference between S'SN“ and 5‘5NB. Site preference = 515N“ — S'SNB (Equation 9) Where: 4O 5'5N“ = (‘5R“(,,m,I.,/‘5R“(..d, — 1) *1000 B'SNB = (‘5R5(,,mpI.I/'5R”IIIII) — 1) *1000 15Ror = [l4NlSNl60/14N14NIOO] ISRB =[15N14N16O/MNMNI6O] Studies by Toyoda and Yoshida (2000) and Breninkenmeijer and Rockman (2000) were a critical step in suggesting that isotopomers could provide a basis to apportion sources of N20. Toyoda and Yoshida (2000) demonstrated that the site preference of N20 varied throughout the atmosphere, but the general trend was a lower site preference in the troposphere samples and a greater site preference in the stratospheric air samples. The low site preference for the troposphere samples is likely a result of local soil emissions and fossil fuel combustion sources (Toyoda and Yoshida, 2000). An important research goal since the Toyoda and Yoshida (2000) study is source apportionment. Thus far, efforts have been directed at determining the isotopomer compositions of N20 in environments that have been implicated as sources, such as oceans and soils, to determine which source is most significant (Perez et al., 2001; Popp et al., in press; Toyoda et al., in press). While this work has increased our understanding of the role of biological mechanisms in determining the isotopomeric composition of ' N20, the isotopomer signature of distinct microbial pathways is still undefined. In order to determine the isotopomer signatures associated with individual microbial pathways, it is important to utilize experimental conditions that isolate distinct pathways. This study examines the isotopomer site preference of N20 produced via hydroxylamine oxidation by cultures of a methane oxidizer and autotrophic nitrifier and 41 reduction of nitrite via nitrifier denitrification. Nitrosomonas europaea was selected as a candidate autotrophic nitrifier because the physiology and biochemistry of ammonia oxidation is best understood in this species (Poth and Focht, 1985; Bothe et al., 2000). Among the methane oxidizing organisms, we chose Methylococcus capsulatus Bath owing to its ability to oxidize ammonia (Dalton, 1977). The objective of this study is to define the site preference distribution for the N20 produced during methanotrophic nitrification, autotrophic nitrification and nitrifier denitrification. This is an important first step in the use of isotopomers as a tool for apportioning microbial sources and in particular, evaluating the importance of methanotrophic nitrification as a source of N20 in certain environments. MATERIALS AND NIETHODS Culture and Media A culture of Methylococcus capsulatus Bath grown on solidified nitrate mineral salts (N MS) medium was provided by A. DiSpirito from Iowa State University. The cultures were maintained in modified nitrate mineral salts medium (NMS)-Whittenbury medium amended with 5 uM CuSO4 in 25 mL test tubes with a headspace of 3:7 (vzv) methane:air (Whittenbury and Dalton, 1981). Prior to inoculation, 5 mL of medium containing all components except the phosphate and vitamin solution were dispensed and autoclaved. The medium was cooled and amended with the phosphate and vitamin solution. The final pH was adjusted to 6.9-7.1 using 1 M Na2C03. The complete NMS medium was inoculated with 0.1 mL of stock culture and incubated horizontally at 40 °C. Stock cultures were stored in medium containing 5 % v/v DMSO held in 2.5 mL cryogenic vials (N algene) and frozen at -70°C. 42 A frozen stock of Nitrosomonas europaea ATCC 19718 was purchased and shipped on dry ice from the American Type Culture Collection, Manassas, VA. The recipe for the medium used to grow N. europaea was provided by A. Hooper, University of Minnesota. The medium consisted of 7.91 g NH4804, 0.76 g KzHPO4' 3HzO, 1 mL 0.5 % w/v Phenol red solution, 1 mL of trace element solution A and 3 mL of trace element solution B per 1.0 L of E-Pure water (A. Hooper, pers.comm). Trace element solution A is an autoclaved mixture containing 7.91 g CuSO4, FeSO4, and 5.89 g EDTA in 1.0 L of E-Pure water. Trace element solution B is an autoclaved mixture containing 68 g MgClz ' 6H20 and 3.08 g CaClz ' 2HzO per 1.0 L of E-Pure water. Approximately 100 mL of sterile media was dispensed into 300 mL tissue culture flasks and the final pH was adjusted to 7.5-7.7 with 5 % (w/v) K2CO3. At least 10 mL of culture was added to inoculate fresh media in tissue culture flasks. The cultures were incubated horizontally at 27 °C. The cultures were checked for purity by inoculating 10 mL of culture in a 20 mL volume of tryptic soy broth (TSB) in a 60 mL serum bottle. Growth was not observed in these tests after 30 days. Stock cultures were stored in a 5 % v/v DMSO at -70 °C. Concentrated cell suspensions Concentrated cell suspensions were used to obtain sufficient concentrations of N20 for subsequent mass spectrometric analysis. Approximately 1 mL of Methylococcus capsulatus Bath culture was used to inoculate each of ten separate 160 mL serum bottles with approximately 30 mL of NMS media and 30 % methane in the headspace. After 40 hours of incubation at 40 0C, the ten cultures were combined in two 250 mL centrifuge bottles (Nal gene). Cells were concentrated by centrifugation (8,000G for 10 minutes at 5 ° C). After the supernatant was decanted, the concentrated cells were combined in one 43 centrifuge cup and resuspended with fresh NMS medium to an approximate volume of 20 mL. For each experiment, four 25 mL butyl stoppered serum test tubes (Bellco) were prepared with 3 mL of concentrated cells and 300 11L of a 0.01 M hydroxylamine. The tubes were stoppered with air in the headspace. The tubes were incubated horizontally at 40 0C. Prior to concentration, Nitrosomonas europaea was maintained in 300 mL tissue culture flasks that contained 100 mL of media in each bottle. The tissue culture flasks were incubated horizontally at 27 0C for three days. Prior to concentration, a subsample was used to inoculate 20 mL of TSB in a 60 mL serum bottle to test for heterotrophic contamination. In all cases, there was no growth in the bottles with TSB following 30 days of incubation. For sufficient cell densities, ten flasks were concentrated via centrifugation (8,000G for 10 minutes at 5 ° C). The cell pellets were washed once and resuspended with 20 mL of 0.1 M phosphate buffer (pH 7.5) containing 171 5.1M CaClz and 78.7 uM MgClz (Poth and F ocht, 1985). For experiments involving hydroxylamine oxidation, four tubes with 2 mL of cell suspension and 300 uL of 0.01 M hydroxylamine solution were stoppered with a headspace of air. To test the isotopomer composition of N20 produced during the nitrifier denitrification pathway, experiments were conducted with nitrite. For each nitrifier-denitrification experiment, at least four serum test tubes were prepared with 2 mL of cell suspension and 300 pL of 0.01 M NaNOz solution. The headspace of the nitrite reduction experiments was purged with N2 and stoppered. 44 Mass spectrometric analysis of N 20 and isotope units Nitrous oxide was sampled from the headspace of the concentrated cell suspensions using a 500 11L or 5 mL gas tight syringe. The gas sample was injected immediately into a Trace Gas system (Micromass) interfaced to an Isoprime (Micromass) mass spectrometer. Using helium as the carrier gas, NzO was purified and concentrated by cryogenic trapping followed by chromatographic separation on a Porapak Q column. Prior to introduction into the mass spectrometer, carbon dioxide and water were removed from the headspace sample in the Trace Gas system with Carbosorb. The nitrogen and oxygen isotopic composition of N20 and the fragment ion NO+ is determined in a single analysis. Triplicate analyses of a N20 sample demonstrate reproducibility of +/- 0.9 for SlsNa-NZO, +/- 0.3 for s‘SN-Nzo and 5'80-Nzo. Calculations of s'SNa-Nzo and BISNB-NZO were done by equations in Toyoda and Yoshida (1999). We observed a rearrangement factor of 18 % versus previous literature estimates of 8 % (Toyoda and Yoshida, 1999; Breninkenmeijer and Rockman, 2000). This difference suggests that the rearrangement factor must be evaluated for each instrument. The correction for 17O for the 5'5N-NZO value was made according to method described by Brand (1995). The concentration of N20 was determined from the area under the mass 44 trace. Standard characterization A Keeling plot approach was used to determine the BISN“ and site preference of an in-house pure (>99.9%) N20 tank standard (Flanagan and Ehleringer, 1998). A series mixtures of 98 % + Na-N20 (Cambridge Isotope Laboratories, MA) with the tank standard NzO (SlsN-NZO = 0.83 °/oo, SIBO-NZO = 37.68 ° 00) were performed to provide an expected range in SUN“ of 10 to 170 °/oo. Mixtures were performed by injecting 45 known volumes of the Cambridge Isotope standard using a gas-tight syringe and injecting into a glass 150 mL vessel (exact volume known) containing the pure N20 tank standard (both the syringe and glass vessel were at atmospheric pressure). To obtain appropriate quantity of N20 for analysis on the Trace Gas system 500 11L of the resulting mixture was injected into a second glass vessel (500 mL) containing pure N2 at atmospheric pressure. A volume of 200 uL from each of the N20-N2 mixtures was injected into the Trace Gas system. Each isotopic analysis was completed in triplicate. On an xy plot Qa/Qm vs. SlsNa-NZO of the mixture, the y—intercept of the best fit line for the 8 mixtures yielded the S‘SNa-NZO of the laboratory N20 tank standard (Q. = moles of 9s % + N°-Nzo and Qm = total moles of N20). Determinations of the ”N‘-N20 of the tank standard were based on the intercept of the plot of Qa/Qm vs. 515Na-NZO based on the following mass balance relationships. 5QO = 5aQa + 5oQo (Equation 10) 5m = Qa(5a — 5o)/Qm + 5o Where: Qm = Qa + Qb Qa = moles of N20 (Cambridge Isotope Lab standard) Qm = total number of moles of N20 (mixture) Q, = moles of N20 (tank standard) 53 = 515Na-N20 (Cambridge Isotope Lab standard) on, = 5‘5N°-Nzo (mixture) 5b = SISNQ-NZO (tank standard) Controls 46 Control experiments were necessary to verify that the stimulated microbial process was responsible for N20 rather than undesired abiological or biological mechanisms. The first control tested whether hydroxylamine can produce N20 abiologically during the time course of a typical incubation (8 hours). In this experiment, 2 mL of phosphate buffer and 300 uL of hydroxylamine were incubated with a headspace of 100 % air at 27 °C. After 8 hours, a 1 mL of a headspace gas was introduced in the mass spectrometer. A second control experiment with 2 mL of concentrated N. europaea cells, 300 11L sodium nitrite and a headspace of air was conducted to test whether N20 was formed via nitrifier denitrification. After 8 hours of incubation, a 1 mL headspace sample was analyzed for N20 concentration. To verify that nitrite reduction does not occur in concentrated cell cultures of M. capsulatus Bath, a third control experiment was conducted. A volume of 2 mL of cell suspension, 300 1.1L of 0.01 M hydroxylamine and 100 uL of 99 % If’N-sodium nitrite were incubated in a test tube with air in the headspace. A 1 mL headspace sample was analyzed to determine N20 concentration RESULTS AND DISCUSSION Standard Characterization The 515 Nat of the standard N20 was obtained from a plot of QalQm vs. 515 Na-NZO of the mixture (Figure 10). The y-intercept of a best-fit line (y = l38520x + 3.61; R2 = 0.9862) yields the 5'5N° oflaboratory N20 tank standard of 3.61 °/,, (Figure 10). Controls Abiological or anoxic microsite N20 production was not detected in the control experiments. In a control experiment to test abiological production of N20 from hydroxylamine, MO was not detected after 8 hours of incubation. In a second control 47 experiment with a concentrated cell suspension of N. europaea, a headspace of air (21 0/o 02) and nitrite as the substrate, the possibility of denitlification within anoxic microsites was examined. The results of the experiment demonstrated N2O was not produced after 8 hours. Since anoxic conditions stimulate the nitrifier denitrification pathway for N20, the result of this experiment suggests that anoxic microsites, if present, were not a site of active denitrification in the concentrated N. europaea cell cultures. To evaluate whether nitrite reduction was active in M. capsulatus Bath, 3 third control was established using concentrated cells, nitrite as the substrate and a headspace of N2 (0 % 02). Using an isotopic mass balance equation, the fraction of N20 calculated to be a product of 15NO2’ in the third control experiment was less than 1 %. Therefore, in the studies of hydroxylamine oxidation by M capsulatus Bath, further reduction of nitrite is not of concern. Bulk nitrogen and oxygen isotopic composition of N 20 The amount of N20 in the headspace increased in the time course experiments indicating production of N20 (Tables 1, 2 and 3). The average 8‘5N values of N20 formed by hydroxylamine oxidation by M. capsulatus Bath (0.0 +/- 1.2 °/oo), hydroxylamine oxidation by N. europaea (-28.3 +/- 6.7) and nitrifier denitrification by N. europaea (-34.8 +/- 2.7) were significantly different (ANOVA, f(2,35)= 247.9, p=0). In addition, the average 5'80 values of N20 formed by hydroxylamine oxidation for M. capsulatus Bath (53.1 +/- 2.9 °/oo), hydroxylamine oxidation by N. europaea (-23.4 +/- 7.2) and nitrifier denitrification by N. europaea (4.6 +/- 1.4) were significantly different (ANOVA, f(2,35)= 279.98, p=0). The average stable oxygen and nitrogen isotopic values of N20 varied in the hydroxylamine oxidation experiments despite the fact that a single 48 substrate solution was used (Table 1 and 2). The isotopic disparities among the hydroxylamine experiments likely result from a difference in the enzymes responsible for hydroxylamine oxidation between M. capsulatus and N. europaea. Whereas, cytochrome P-460 is the enzyme responsible for hydroxylamine oxidation in M. capsulatus Bath (Zahn et al., 1994), the primary enzyme used for hydroxylamine oxidation in N. europaea is hydroxylamine oxidoreductase. Furthermore, the cytochrome P-460 enzyme purified from M. capsulatus Bath differs in terms of activity from that of N. europaea (Zahn etal., 1994). Thus, the enzymes used for oxidation of hydroxylamine in the two organisms are distinct. This could invoke a difference in the fractionation factor associated with the nitrification reaction and, thus, account for the disparities in the bulk nitrogen and oxygen isotope values of N20 produced by the two organisms studied here. The Al5 N for the cultures demonstrates that the two hydroxylamine experiments differ in the degree of fractionation (-2.3 and 26.0 °/oo for M. capsulatus Bath and N. europaea respectively). A15 N = 815N(subsImIe) — BlstroducI) (Equation 11) Where: 8'5 N of hydroxylamine = -2.3 °/oo 515Nofnitrite= 1.1 0/00 1 The A15 N for the nitrifier-denitrification experiment is 35.9 °/oo. This is similar to an estimate of the fractionation factor during N20 production by nitrifiers in a laboratory study (8 = 30 0Am) and suggests that hydroxylamine oxidation may be the rate limiting step in nitrification(Yoshida, 1988). The study by Yoshida (1988) and the current study demonstrate that depletions in 15 N of N20 could result from hydroxylamine oxidation and 49 nitrite reduction by nitrifiers. While bulk stable nitrogen and oxygen isotopic values provide insight into isotopic fractionation, a basis for distinguishing N2O formed by hydroxylamine oxidation and nitrite reduction by nitrifiers still remains. Nitrogen isotopomeric site preference of N 20 Average site preferences of N20 formed by hydroxylamine oxidation by M. capsulatus Bath (5.5 +/- 3.1 °/oo), hydroxylamine oxidation by N. europaea (-1.4 +/- 1.7 0Am) and nitrite reduction by N. europaea (-7.7 +/- 3.1 04,0) are significantly different (ANOVA, F234 = 72.09, p=0). This result indicates that the site preferences of the N20 formed by the three pathways in this study are distinct. This preliminary data set provides the first reported values of the isotopomeric site preference of individual microbial pathways. A negative site preference resulted from the processes of hydroxylamine oxidation and nitrite reduction by N. europaea (Tables 2 and 3). Equilibrium conditions were shown to impose a positive site preference (45 °/oo) (Y ung and Miller, 1997). During kinetic isotope effects, such as the three pathways in this study, the formation of N20 involving a hyponitrite intermediate can result in enrichment in 15N at N“ (Popp et al., in press; Toyoda et al., in press). However, measurements with a negative site preference are reported by Yamulki et al. (2001) and the current study. In the proposed mechanism for the production of N20 via dissimilatory nitrite reduction by Weeg- Aerssens et al., (1988), the N-N bond of N20 is formed by nucleophilic attack of a second nitrite on a Fe-coordinated nitrosyl species (Figure 11). In terms of kinetic fi'actionation, the key reactions in the sequential mechanism that could produce a site preference are steps 1 and 3. This is the case because nitrite is abundant in culture and, 50 therefore, not rate limiting. The first N added in step 1 ultimately occupies the B position and fractionation could result in the process of nitrite binding to the ferrous heme (Figure 11). Nucleophilic attack of nitrite on the coordinated nitrosyl species (step 3) resulting in the formation of the N-N bond can also lead to isotopic discrimination (Figure 11). The relative difference in the magnitude of the kinetic isotope effect (degree of discrimination against 'SN) resulting from steps 1 and 3 will determine the site preference. In order for a negative site preference to result, the NB of N20 must be more enriched in 15N than N“. We propose that the isotopic fractionation involved in nucleophilic attack on the coordinated nitrosyl group is greater than the isotope effect associated with binding of nitrite to ferrous heme. Alternatively, step 3 may be rate limited to a greater extent than step 1 thereby favoring depletion of lsN in N“ (Figure 11). This occurs because as a reaction approaches completion, the isotopic composition of the product approaches that of the initial substrate. Thus, the observed discrimination against 15 N will be greater for rate-limited reactions. While the mechanism associated with N20 formation by hydroxylamine oxidation is unknown, it is possible that a negative site preference could result from a similar difference in fractionation factors or rate limitation associated with yet undescribed pathways in the formation of the N-N bond of N20. Field studies have indicated that N20 formed via biological mechanisms is characterized by low isotopomeric site preferences (Table 4). The average site preference of N20 in tropospheric samples is approximately 19 ° 00 (Y oshida and Toyoda, 2000). The 19 °/(,0 site preference is thought to result from mixing of biologically and combustion derived N20 of low site preference with N20 of higher site preferences such as that produced by photolysis in the stratosphere (Yoshida and Toyoda, 2000). Using an 51 isotope mass balance equation, Yoshida and Toyoda (2000) estimated that the range of site preferences from soil and oceanic sources was -O.5 to +15 °/oo. This is consistent with other field studies that demonstrate low site preferences in areas of the water column with active nitrification or soils with nitrogen additions (Table 4). In a study of N20 production in the North Pacific, Popp et al.(in press) estimated that the site preference of the in situ source of N20 was likely between 0 and 8 °/oo, Although the measured value of the site preference of N20 formed during hydroxylamine oxidation by autotrophic nitrifiers (-1.4 +/- 1.7 °/00) was slightly lower than the estimated range in the field study (0 and 8 0Am), given the possible errors in estimation and analytical measurement they could be considered similar. This provides further evidence that autotrophic nitrification was the dominant in situ source of N20 at ALOHA (Popp et al., in press). The site . preference we observed for the nitrifier-denitrification pathway (-7.7 +/- 3.1 oloo) is much lower than that estimated for oceanic and soil sources (05 to +15 °/oo; Yoshida and Toyoda, 2000). This may suggest that the nitrifier-denitrification pathway is not a dominant mechanism for N20 production in soil and oceanic sources, however, the number of site preference measurements in the literature are few. This is consistent with Ostrom et al. (2001) who proposed that oxidation of hydroxylamine was the predominant pathway of N20 production in the central North Pacific, but also proposed that nitrifier denitrification was significant within a narrow depth interval. The isotopomeric composition of N20 formed via hydroxylamine oxidation by M. capsulatus Bath (5.5 +/- 3.1 °/oo) is similar to the range of estimated values for soil and oceanic sources. This is suggestive of an underappreciated role of methanotrophic nitrification to N20 production in soil and oceanic environments. 52 CONCLUSIONS The A15 N of hydroxylamine oxidation and nitrifier denitrification by N. europaea were, 26.0 and 35.9 °/oo, respectively, and similar to the fractionation previously reported of 8 = 30 °/oo (Y oshida, 1988). However, discrimination against ”N during N20 production by hydroxylamine oxidation by M capsulatus Bath was small (-2.3 oAm). This suggests that B'SN-N2O could be used to distinguish N2O production by autotrophic nitrification and methanotrophic nitrification. Isotopic fractionation of oxygen isotopes during hydroxylamine oxidation by M. capsulatus Bath and N. europaea were significantly different. This is consistent with a study demonstrating that different enzymes are involved in hydroxylamine oxidation in M. capsulatus Bath and N. europaea. These results suggest that, in addition to nitrogen isotope values, oxygen isotope values can be used to distinguish N20 produced by methanotrophic nitrification and autotrophic nitrification. The site preference of N20 formed by hydroxylamine oxidation by M. capsulatus Bath and N europaea and nitrite reduction by N. europaea were significantly different. Therefore, site preference can distinguish N20 formed by hydroxylamine oxidation by M. capsulatus Bath and N. europaea. In addition to differentiating hydroxylamine oxidation mechanisms, site preference can discern N2O produced by nitrifier denitrification from hydroxylamine oxidation. The disparity in site preference is likely a consequence of differences in the enzymes responsible for N20 production between the two organisms. In sequential mechanisms for the production of N20, the degree of discrimination or rate limitation in individual steps of the reaction pathway could impose a distinct site preference. The results of this study indicate that site preference provides a 53 basis for resolving specific microbial pathways of N20 production. This information could be critical for predicting and managing N2O emissions from some ecosystems, such as agricultural systems, where activities to mitigate carbon dioxide emissions may enhance N2O production. 54 59.0 «Sod Sod 88d 886 385 885 o . . _ . _ o -om .9» Ice Iow O O N P O O (amrxgw) aplxo snomu-mmg . - 9; ~83 u o . am . o? 5 m + xommmfl u > o? .025 98:: mo 838 :38 n so . 02x0 msobEIoZ e\e ma .«o 83:. U ad 80:3 @3983 02x0 355:. on“ .«o 85.5.“ng 87. on. wcfitouofiano 2 5889? BE wczoov— .3 «Burn 55 Table 1. Summary of nitrous oxide data including concentration, site preference, S‘SN- N20 and 6'80-N2O from concentrated Methylococcus capsulatus Bath cell cultures with hydroxylamine as the substrate. Each letter (A, B, C and D) corresponds to a separate experiment and number indicates the time course sample. Nitrous oxide formation via hydroxylamine oxidation is stimulated in this set of experiments. 'Anomalous value (37.7 °/oo) omitted in the calculation of average and standard deviation of the 5'80- NZOsmow~ Time site 6‘5N-Nzo... 5‘80- preference (°Ioo) N2Ovsuow Culture N20 (nM) Cloo) (°Ioo) A-M.cap 1 20:09 282 7.2 -0.3 50.4 A-M.cap 2 20:28 451 8.5 -0.2 53.4 A-M.cap 3 21 :24 563 8.1 -3.3 56.2 B.M.cap 1 17:40 305 1 .5 1.7 37.7 B-M.cap 2 18:35 644 6.2 0.4 55.8 B-M.cap 3 19:18 676 7.5 0.7 57.1 C-M.cap 1 20:28 191 0.0 -0.2 49.8 C-M.cap 2 21 :05 260 10.0 -0.2 50.4 C-M.cap 3 21 :47 323 3.8 -0.3 51.3 D-M.cap 1 18:45 317 3.1 1.0 50.8 D-M.cap 2 19:20 387 3.2 0.2 51.9 D-M.cap 3 23:11 576 6.9 0.7 57.1 Average 5.5 0.0 53.1: Standard deviation 3.1 1.2 2.9 56 Table 2. Summary of nitrous oxide data including concentration, site preference and, 6180-N2O and 8'5N-N2O from concentrated Nitrosomonas europaea cell cultures with hydroxylamine as the substrate. Each letter (A, B, C and D) corresponds to a separate experiment and number indicates the time course sample. Nitrous oxide formation by hydroxylamine oxidation is stimulated in this set of experiments. Site 8‘5N- 51°0- preference N20... Nzovsuow Culture Time N 20 (nM (°Ioo) (°Ioo) (°loo) A-N.eur 1 15:28 89 -0.7 -20.4 33.5 A-N.eur 2 15:47 143 1.2 -21.4 31.9 A-N.eur 3 16:05 216 1.2 -24.7 28.1 B-N.eur 1 17:46 284 -1.4 -35.5 16.5 B-N.eur 2 18:29 417 -2.4 -37.5 15.3 B-N.eur 3 19:08 466 -1.8 -38.1 15.1 C-N.eur 1 19:58 158 -2.6 -30.2 20.5 C-N.eur 2 20:38 179 -1.5 -30.6 20.3 C-N.eur 3 21 :18 246 -0.9 -30.3 20.2 D-N.eur 1 14:03 246 -4.6 -27.2 21.4 D-N.eur 2 15:41 249 -2.7 -26.3 22.3 Average -1 .4 -28.3 23.4 Standard deviation 1.7 6.7 7.2 57 Table 3. Summary of nitrous oxide data including concentration, site preference, 5180- N20 and SlSN-N2O from concentrated Nitrosomonas europaea cell cultures with nitrite as the substrate. Each letter (A, B, C and D) corresponds to a separate experiment and number indicates the time course sample. Nitrous oxide formation via nitrite reduction is stimulated in this experiment. Site 5‘5N- 51°0- preference N203" Nzovsmow Culture Time N20 (nM) (°loo) (°Ioo) (°Ioo) A-Nit-denit 1 14:08 89 -10.1 -34.0 4.3 A-Nit-denit 2 14:58 119 -11.1 -33.1 4.1 A-Nit-denit 3 17:38 189 -6.4 -33.3 4.1 A-Nit-denit 4 19:00 204 -6.9 -33.1 4.2 A-Nit-denit 5 19:41 218 -10.8 -33.2 4.1 B-Nit-denit 1 20:18 51 -5.4 -36.0 3.4 B-Nit-denit 2 21 :36 74 -6.0 -36.7 3.5 B-Nit-denit 3 22:16 73 -7.5 -36.8 3.1 C-Nit-denit 1 19:47 98 -5.9 -38.2 8.0 C-Nit-denit 2 20:27 116 -6.0 -38.7 6.5 C-Nit-denit 3 20:47 122 -2.4 -39.1 6.5 D-Nit-denit 1 20:18 214 -11.8 -31.0 5.1 D-Nit-denit 2 21 :36 291 -11.4 -31.5 4.4 D-Nit-denit 3 22:16 559 -3.2 -33.1 3.8 Average -7.7 -34.8 4.6 Standard deviation 3.1 2.7 1.4 58 Table 4. Summary of isotopomeric compositions of nitrous oxide in the literature with an emphasis on biological pathways. Site preference (°Ioo) Study site Comments Authors Tropospheric average trophospheric Yoshida and 19 sample value Toyoda, 2000 Hydroxylamine Methylococcus capsulatus 5.5 ’I. 3.1 oxidation Bath — Laboratory study Current study Hydroxylamine Nitrosomonas europaea - -1.4 *l. 1.7 oxidation Laboratory study Current study Nitrosomonas europaea - -7.7 +I. 3.1 Nitrite reduction Laboratory study Current study North Pacific minimum between 100 Popp et al., in ~8 waters and 300 m press Western Pacific isotopomeric minimum in Toyoda et al., ~11 waters water column in press agricultural Range of isotopomeric Perez et al., 4.9 to 14.2 fields composition in study 2001 urine amended Range of measured Yamulki et al., -2 to +9 grassland values 2001 Yoshida and -0.5 to +15 estimated value isotope mass balance Toyoda, 2000 59 AZ 33:". - t -O i z o z// // _ .1? .z _ z 1... .l. .2 I... l... __ _ ll -O\ .O _ O I- -Z l+w®u Osz _ o _ s m .IN 8% of- \ O- Ouz .6 $0.. . -O i Z 8388 0.4.2-3928 9 mats—283-90 do 5:96.360 . // ‘ 8:93:66 ucm cozoanom . Z I +~mu_ cozoanom . \ .02 865208 65 :0 SE: 283 m do xomzm 2:38.032 . O _ :5 8388 385:1 32.8 828mm: m 2 5:93:60 . mm H 8 888; .0598. m 2 «6:5 6:52 . V - + \\O- m N _. 0 II ZAI/ dz — ‘ +INI .ONI+ — INOZI — a2 I. O m Z i i N i i a... a... I.l a... Tie. a... .l. a... 02+ X N N O \ _ - _ O I- .IIN+ _ - 02+ _ £me .3 3 mcommuoB>E wage—a BEE mo 8:268 65 Set 028 £55: mo cog—88 2t .8.“ .3353 womoaoi .: new: 60 BIBLIOGRAPHY 61 BIBLIOGRAPHY Altabet M.A., Deuser W.G., Honjo S. and Stienen C. (1991) Seasonal and depth-related changes in the source of sinking particles in the North Atlantic. Nature 354, 218-219. Altabet M.A., Pilskaln C., Thunell R., Pride C., Sigman D., Chavez F. and Francois, R. (1999) The nitrogen isotope biogeochemistry of sinking particles from the margin of the eastern North Pacific. Deep-Sea Research 46, 655-679. Blackmer A.M., Bremner J .M. and Schmidt EL. (1980) Production of nitrous oxide by ammonia-oxidizing chemoautotrophic microorganisms in soil. Applied and Environmental Microbiology 40, 1060- 1 066. Bodelier P.L.E. and Frenzel P. (1999) Contribution of methanotrophic and nitrifying bacteria to CH4+ and NIH oxidation in the rhizosphere of rice plants as determined by new methods of discrimination. Applied and Environmental Microbiology 65, 1826- 1833. Bothe H., Jost G., Schloter M., Ward BB. and Witzel K-P. (2000) Molecular analysis of ammonia oxidation and denitrification in natural environments. FEMS Microbiology Reviews 24, 673-690. Brand WA. (1995) PRECON: A fully automated interface for the preGC concentration of trace gases in air for isotopic analysis. Isotopes in Environmental Health Studies 31, 277-284. Brandes J .A., Devol A.H., Yoshinari T., Jayakumar D.A., and Naqvi S.W.A. (1998) Isotopic composition of nitrate in the central Arabian Sea and eastern tropical North Pacific: A tracer for mixing and nitrogen cycles. Limnology and Oceanography 43, 1680-1689. Breninkenmeijer C.A.M and Rockman T. (2000) Mass spectrometry of the intramolecular nitrogen isotope distribution of environmental nitrous oxide using fragment-ion analysis. Rapid Communications in Mass Spectrometry 13, 2028-2033. Brzezinski MA. (198 8) Vertical distribution of ammonium in stratified oligotrophic waters. Limnology and Oceanography 33, 1 176-1 182. Capone D.G., Zehr 1., Paerl H., Bergma B., and Carpenter, E]. (1997) T richodesmium: A globally significant marine cyanobacterium. Science 276, 1221-1229. Carpenter E]. (1983) Nitrogen fixation by marine Oscillatoria (T richodesmium) in the world’s oceans. In Nitrogen in the Marine Environment (eds. E.J. Carpenter and DC. Capone). Academic Press, New York. pp. 65-103. 62 Carpenter E.J., Harvey H.R., Fry B., and Capone DC. (1997) Biogeochemical tracers of the marine cyanobacterium T richodesmium. Deep-Sea Research 44, 27-38. Carpenter J.H. (1965) The Chesapeake Bay Institute technique for the Winkler dissolved oxygen method. Limnology and Oceanography 10; 141-143. Cline JD. and Kaplan LR. (1975) Isotopic fractionation of dissolved nitrate during denitrification in the eastern tropical North Pacific Ocean. Marine Chemistry 3; 271-299. Cline J .D., and Richards RA. (1972) Oxygen deficient conditions and nitrate reduction in the eastern tropical North Pacific Ocean. Limnology and Oceanography 17, 885-900. Dalton H. (1977) Ammonia oxidation by the methane oxidising bacterium Methylococcus capsulatus Strain Bath. Archives of Microbiology 114, 273-279. Delwiche CC, and Steyn PL. (1970) Nitrogen isotope fractionation in soils and microbial reactions. Environmental Science and Technology 4, 929-935. Dore J .E., Popp B.N., Karl D.M., and Sansone RI. (1998) A large source of atmospheric nitrous oxide from subtropical North Pacific surface waters. Nature 396; 63-66. Dore J .E. and Karl D.M. (1996) Nitrification in the euphotic zone as a source of nitrite, nitrate, and nitrous oxide at Station ALOHA. Limnolog) and Oceanography 41, 1619- 1628. Dugdale RC, and Goering 1.]. (1967) Uptake of new and regenerated forms of nitrogen in primary productivity. Limnology and Oceanography 12, 196-206. Eppley R.W. and Peterson PJ. (1979) Particulate organic flux and planktonic new production in the deep ocean. Nature 282, 677-680. Fanning K.A. (1992) Nutrient provinces in the sea: concentration ratios, reaction rate ratios, and ideal covariation. Journal of Geophysical Research 97, 5693-5712. Farrell J.W., Pedersen T.F., Calvert SE, and Nielsen B. (1995) Glacial-interglacial changes in nutrient utilization in the equatorial Pacific Ocean. Nature 377, 514-516. Flanagan LB. and Ehleringer JR. (1998) Ecosystem-atrnosphere CO2 exchange: interpreting signals of change using stable isotope ratios. Trends in Ecology and Evolution 13, 10-14. Ganeshram R.S., Pedersen T.F., Calvert SB, and Murray J.W. (1995) Large changes in oceanic nutrient inventories from glacial to interglacial periods. Nature 376; 755-758. 63 Glibert, RM. and Capone DC. (1993) Mineralization and assimilation in aquatic, sediment, and wetland systems, In Nitrogen Isotope Techniques (eds. R. Knowles and T.H.Blackbum). Academic Press, San Diego, pp. 243-273. Goreau T.J., Kaplan W.A., Wofsky S.C., McElroy M.B., Valois F .W., and Watson SW. (1980) Production of NO2' and N20 by nitrifying bacteria at reduced concentrations of oxygen. Applied and Environmental Microbiology 40, 526-532. Hanson RS. and Hanson TE. (1996) Methanotrophic bacteria. Microbiological Reviews 60, 439-471. Hoering, T. and Ford HT. (1960) The isotope effect in the fixation of nitrogen by Azotobacter. Journal of the American Chemical Society 82, 376-378. Horrigan S.G., Carlucci A.F., and Williams RM. (1981) Light inhibition of nitrification in sea-surface films. Journal of Marine Research 39, 557-565. Howarth R.W. (1988) Nutrient limitation of net primary production in marine ecosystems. Annual Reviews in Ecology 19, 89-110. Jiang Q-Q. and Bakken LR. (1999) Nitrous oxide production and methane oxidation by different ammonia-oxidizing bacteria. Applied and Environmental Microbiology 65, 2679-2684. Karl D.M., Letelier R., Hebel D.V., Bird DR, and Winn CD. (1995) Ecosystem changes in the North Pacific subtropical gyre attributed to the 1991-92 El Nino. Nature, 373, 230-233. Karl D., Letelier R., Tupas L., Dore J., Christian J ., and Hebel. D. (1997) The role of nitrogen fixation in biogeochemical cycling in the subtropical North Pacific ocean, Nature 386, 533-538. Karl D.M. (1999) A Sea of Change: Biogeochemical Variability in the North Pacific Subtropical Gyre. Ecosystems 2, 181-214. Kim K.R.and Craig H. (1990) Two-isotope characterization of N20 in the Pacific Ocean and constraints on its origin in deep water. Nature 347, 58-61. Lacis A.A., Hansen J ., Lee R, Mitchell T., and Lebedeff S. (1981) Greenhouse effect of trace gases. Geophysical Research Letters 8, 1035-1038. Letelier RM, and Karl D.M. (1998) T richodesmium spp. Physiology and nutrient fluxes in the North Pacific subtropical gyre. Aquatic Microbial Ecology 15, 265-276. Liu K-K., Su M-J., Hsueh C-R. and Gong G-C. (1996) The nitrogen isotopic composition of nitrate in the Kuroshio Water northeast of Taiwan: evidence for nitrogen fixation as a source of isotopically nitrate. Marine Chemistry 54, 273-292. Macko, SA. (1981) Stable nitrogen isotope ratios as tracers of organic geochemical processes. Ph. D. thesis, University of Texas, Austin. Macko S.A., Fogel Estep M.L., Engel M.H. and Hare RE. (1986) Kinetic fractionation of stable nitrogen isotopes during amino acid transamination. Geochimica et Cosmochimica Acta 50, 2143-2146. Mandemack K.W., Kinney C., Coleman D., Huang Y-S., Freeman K.H., and Bogner J. (2000) The biogeochemical controls of N20 production and emission in landfill cover soils: the role of methanotrophs in the nitrogen cycle. Environmental Microbiology 2, 298-309. Mariotti A., Germon J.C., Hubert P., Kaiser P., Letolle R., Tardieux P. and Tardieux P. (1981) Experimental determination of nitrogen kinetic isotope fractionation: some principles; illustration for the denitrification and nitrification processes. Plant and Soil, 62, 413-430. Minagawa M. and Wada. E. (1978) Nitrogen isotope fractionation in the biological N2- fixation and its significance in natural ecosystems, Abstracts of the 1978 Annual meeting of the Geochemical Society of Japan, 219. Naqvi S.W.A., Yoshinari T., Jayakumar D.A., Altabet M.A., Narvekar P.V., Devol A.H., Brandes J .A. and Codispoti LA. (1998) Budgetary and biogeochemical implications of N20 isotope signatures in the Arabian Sea. Nature 394,462-464. Ostrom N.E., Macko S.A., Deibel D., and Thompson R]. (1997) Seasonal variation in the stable carbon and nitrogen isotope biogeochemistry of a coastal cold ocean environment. Geochimica et Cosmochimica Acta 61, 2929-2942. Ostrom N.E., Knoke K.E., Hedin L.O., Robertson GR, and Smucker A.J.M (1998) Temporal trendes in nitrogen isotope values of nitrate leaching fiom an agricultural soil. Chemical Geology 146, 219-227. Ostrom N.E., Russ M.B., Popp B., Rust TM. and Karl D.M. (2000) Mechanisms of N20 production in the subtropical North Pacific based on determinations of the isotopic abundances of N20 and O2. Chemosphere — Global Change Science 2, 281-290. Paerl H.W. and Zehr JP. (2000) Marine nitrogen fixation. In Microbial Ecology of the Oceans (ed. D.L. Kirchman). Wiley-Liss, New York. pp. 387-425. 65 Perez T., Trumbore S.E., Tyler S.C., Matson P.A. Ortiz-Monasterio 1., Rahn T. and Griffith D.W.T. (2001) Identifying the agricultural imprint on the global N20 budget using stable isotopes. Journal of Geophysical Research 106, 9869-9878. Popp B.N., Westley M.B., Toyoda S., Miwa T., Dore J.E., Yoshida N., Rust T.M., Sansone F.J., Russ M.B., Ostrom NE, and Ostrom P. in press. Nitrogen and oxygen isotopomeric constraints on the origins and sea-to-air flux of N20 in the oligotrophic subtropical North Pacific gyre. Global Biogeochemical Cycles. Poth M. and Focht, DD. (1985) 15N Kinetic analysis of N20 production by Nitrosomonas europaea an examination of nitrifier denitrification. Applied and Environmental Microbiology 49, 1 134-1 141. Prinn R., Cunnold D., Rasmussen R., Simmonds P., Alear F ., Fraser P., and Rosen R., (1990) Atmospheric emissions and trends of nitrous oxide deduced from 10 years of ALE-GAGE data. Journal of Geophysical Research 99, 5285-5294. Rasmussen, R.A. and Khalil M.A.K. (1986) Atmospheric trace gases: Trends and distributions over the last decade. Science 232, 1623-1624. Reid J.L., Brinton E., Fleminger A., Venrick EL. and McGowan IL. (1978) Ocean circulation and marine life. In Advances in oceanography (eds. H. Chamock and G. Deacon, G). New York. Plenum Press. pp. 65-130. Ritchie G.A.F. and Nicolas D.J.D. (1972) The partial characterization of purified nitrite reductase and hydroxylamine oxidase from Nitrosomonas europaea. Biochemistry Journal 126, 1181-1191. Roache, P (1998) Fundamentals of computational fluid dynamics. Herrnosa. New York. 648 pages. Robertson G.P., Paul EA. and Harwood RR. (2000) Greenhouse gases in intensive agriculture: Contributions of individual gases to the radiative forcing of the atmosphere. Science 289, 1922-1925. Saino T. and Hattori A. (1987) Geographic variation of the water column distribution of suspended particulate nitrogen and its 5N natural abundance in the Pacific and its marginal seas. Deep-Sea Research 34, 807-827. Sigman D.M., Altabet M.A., Francois R., McCorkle DC. and Fischer, G. (1999). The 615N of nitrate in the Southern Ocean: consumption of nitrate in surface waters. Global Biogeochemical Cycles 13, 1149-1166. Toyoda S. and Yoshida N. (1999) Determination of nitrogen isotopomers of nitrous oxide on a modified isotope ratio mass spectrometer. Analytical Chemistry 71, 4711-4718. 66 Toyoda S.N., Yoshida T., Miwa Y., Matsui H., Yamagishi H. and Tsunogai U. (in press) Production mechanism and global budget of N20 inferred from its isotopomers in the western North Pacific. Geophysical Research Letters. UNESCO (1994) Protocols for the Joint Global Ocean Flux Study (J GOFS) Core Measurements. IOC Manual and Guides 29. Valderrama J .C. (1981) The simultaneous analysis of total nitrogen and total phosphorus on natural waters. Marine Chemistry 10, 109-122. Velinsky D.J., Fogel M.L., Todd J .F . and Tebo BM. (1991) Isotopic fractionation of dissolved ammonium at the oxygen-hydrogen sufide interface in anoxic waters. Geophysical Research Letters 18, 649-652. Voss M., Dippner J.W. and Montoya, JP. (2001) Nitrogen isotope patterns in the oxygen-deficient waters of the Eastern Tropical North Pacific Ocean. Deep-Sea Research 148,1905-1921. Wada E. and Hattori A. (1978). Nitrogen isotope effects in the assimilation of inorganic nitrogenous compounds by marine diatoms. Geomicrobiology Journal 1, 85-101. Ward BB. and Zafiriou CC. (1988) Nitrification and nitric oxide in the oxygen minimum of the eastern tropical North Pacific. Deep-Sea Research 35, 1127-1142. Ward B.B., Kilpatrick K.A., Renger E.H., and Eppley R.W. (1989) Biological nitrogen cycling in the nitracline. Limnology and Oceanography 34, 493-513. Ward BB. (1996) Nitrification and denitrification: probing the nitrogen cycle in aquatic environments. Microbial Ecology 32, 247-261. Ward BB. (2000) Nitrification and the marine nitrogen cycle. In Microbial Ecology of the Oceans (ed. D.L Kirchman). Wiley-Liss. New York, pp. 427-453. Weeg—Aerssens E.J., Tiedje J .M. and Averill BA. (1988) Evidence from isotope labeling studies for a sequential mechanism for dissimilatory nitrite reduction. Journal of the American Chemical Society 110, 6851-6856. Whittenbury R. and Dalton H. (1981) The methylotrophic bacteria, p. 894-902. In M.P. Starr, H.G. Truper, A. Balows, and HG. Schlegel (ed), The prokaryotes, vol I. Springer- Verlag, New York. Wu J ., Calvert SE. and Wong CS. (1997) Nitrogen isotope variations in the subarctic Pacific northeast Pacific: relationships to nitrate utilization and tr0phic structure. Deep- Sea Research Part I 44, 287-314. 67 Yamulki S., Toyoda S., Yoshida N., Veldkamp E., Grant B. and Jarvis SC. (2001) Diurnal fluxes and the isotopomer ratios of N20 in a temperate grassland following urine amendment. Rapid Communications in Mass Spectrometry 15, 1263-1269. Yoshida N., Hattori A., Saino T., Matsuo S. and Wada E. (1984) 15N/MN ratio of dissolved N20 in the eastern tropical Pacific Ocean. Nature 307; 442-444. Yoshida N. (1988) 15N-depleted N20 as a product of nitrification. Nature 335, 528-529. Yoshida N. and Toyoda S. (2000) Constraining the atmospheric N2O budget from intramolecular site preference in N20 isotopomers. Nature 405, 330-334. Yoshinari T. (1985) Nitrite and nitrous oxide production by Methylosinus trichosporium. Canadian Journal of Microbiology 31, 139-144. Yoshinari T., Altabet M.A., Naqvi S.W.A.,Codispoti L., Jayakumar A., Kuhland M., Kuhland M., and Devol A. (1997) Nitrogen and oxygen isotopic composition of N20 from suboxic waters of the eastern tropical North Pacific and the Arabian Sea- measurements by continuous-flow isotope-ratio monitoring. Marine Chemistry 56, 253- 264. Yung Y.L. and Miller CE. (1997) Isotopic fractionation of stratospheric nitrous oxide. Science 278, 1778-1780. Zahn J .A., Duncan C. and DiSpirito AA. (1994) Oxidation of hydroxylamine by cytochrome p-460 of the obligate methylotroph Methylococcus capsulatus Bath Journal of Bacteriology 176, 5879-5887. Zehr J .P., Carpenter E]. and Villareal TA. (2000) New perspectives on nitrogen-fixing microorganisms in tropical and subtropical oceans. Trends in Microbiology 8, 68-73. Zehr JP. and Ward BB. (2002) Nitrogen cycling in the ocean: New perspectives on process and paradigms. Applied and Environmental Microbiology 68, 1015-1024. 68 I IIIIIIIIIIIIIIIIIIIIIIIIIIIIIII I ill/illllill/llill/lllllllllill/ill}ill/l 3 1293 02356 2196