Open access

Stable Isotope Methods for the Study of the Nitrogen Cycle

Written By

Evgenia Ryabenko

Submitted: 18 June 2012 Published: 10 July 2013

DOI: 10.5772/56105

From the Edited Volume

Topics in Oceanography

Edited by Enrico Zambianchi

Chapter metrics overview

5,269 Chapter Downloads

View Full Metrics

1. Introduction

Nitrogen, a limiting element for biological productivity, plays a key role in regulating the biogeochemical processes in the ocean. In today’s ocean, all of the major reactions in the N cycle are mediated by assimilatory or dissimilatory functions of marine organisms. Because marine organisms preferentially incorporate lighter stable isotope of nitrogen 14N instead of 15N, each major metabolic reaction in the N cycle involves irreversible kinetic fractionation of nitrogen.

The isotopic composition of a pool of nitrogen can be used to identify the relative importance of sources that are isotopically distinct, or processes that add or remove nitrogen with a characteristic pattern of isotopic discrimination. The strongest isotopic fractionations are associated with dissimilatory processes that mediate the transfer of nitrogen from one inorganic pool to another [1]. In contrast, processes such as primary production, which move nitrogen directly into and through the food web are associated with comparatively weak isotopic fractionations [1]. The extent of nitrogen isotope fractionation also depends upon the kinetics of individual metabolic reactions, concentration of products and reactants, environmental conditions (e.g., oxygen concentrations) and the microbial species involved.

The crucial challenge in using nitrogen isotope methods is the complexity of the marine nitrogen cycle and the potential influence of multiple processes on the isotopic Composition of several biologically active pools of nitrogen. Successful use of nitrogen isotopes in resolving N cycle fluxes and processes requires an understanding of the general distribution of nitrogen isotopes in marine systems, the nature of isotopic fractionation, and a careful consideration of the processes at work in the system of interest. In this chapter I will introduce the reader to the two different approaches of nitrogen isotope analysis: natural abundance or tracer methods (15N-labeling). The natural abundance of the stable isotope of nitrogen can provide critical insights into processes acting on a variety of scales with minimal alteration and manipulation of the system being studied. The isotopic signal collected in this type of approach is an integrated value that incorporates large spatial and temporal variation of the relevant processes. Conversely, isotope tracer methods involve short-term incubations of small, isolated samples of water under in situ or simulated in situ conditions. Tracer methods are primarily used for rate experiments that complement larger-scale natural abundance studies.

I begin the chapter with basic definitions of isotopic fractionation and analytical considerations of the sample measurements. I follow with an overview of the processes of nitrogen sources and sinks. I conclude with a discussion of how isotopic data can contribute to the current debate on the balance of N-inputs and losses in the nitrogen cycle.

Advertisement

2. Fundamentals

2.1. Isotopes and calculation of their ratio

Isotopes are atoms of an element that share the same number of protons but a different number of neutrons. In the scientific nomenclature, isotopes are specified in the form Enm, where “m” indicates the mass number (the sum of protons and neutrons in the nucleus) and “n” refers to the atomic number of an element “E”. There are more than 10 nitrogen isotopes known. Most of these isotopes are radioactive and highly unstable with longest half-life time for N713 of 10 minutes. The only two stable nitrogen isotopes are N714 and N715, which have seven protons each and seven or eight neutrons in their nucleus, respectively. 15N is the less frequent stable isotope, constituting of 0.365% of the global nitrogen pool [2]. Consequently, it is more practical to measure the difference or ratio of two isotopes instead of the absolute quantity of each.

Isotopic compositions are expressed in terms of “delta” (δ) values which are given in parts per thousand or per mil (‰). Nitrogen isotope ratios, for example, are expressed as the ‰-difference to atmospheric N2, which has a constant 14N/15N of 272 ± 0.3 [3]. The δ15N-value in the sample is then calculated by the following equation (1).

δ15N(vs.air)=((15N/14N)sample(15N/14N)air1)×1000E1

The δ-values do not represent absolute isotope abundances but rather the ‰-difference to a widely used reference standard, such as VSMOW (Vienna Standard Mean Ocean Water). The δ-value is then calculated from equation (2), by measuring the isotope ratios (R) for the sample and the reference standard:

δ15N=(RsampleRstandard1)×1000E2

where Rsample and Rstandard represent the isotope ratio δ15N(vs.air) in the sample and in the standard respectively, calculated using equation (1). By convention, R is the ratio of the less abundant isotope over the most abundant isotope (i.e. 15N/14N for nitrogen).

As previously discussed, the δ15N value changes under the influence of chemical and physical processes. If the process is complete the resulting δ15N value in the product is equal to the value of the reagent. Only if the reaction is incomplete the fractionation of isotopes happens, which means that the δ15N values in the product and substrate differ.

2.2. Isotope fractionation effect

There are two different fractionation processes, both of which will be discussed here: equilibrium and kinetic fractionation processes.

Equilibrium fractionation processes are reversible processes. They are mainly driven by changes in the internal energy of a molecule, i.e. vibrations of the atoms within a molecule with respect to each other and rotations around the molecular axes. The equilibrium fractionation factor αeq. is related to the equilibrium constant K as shown in equation (3), where n is the number of exchanged atoms.

αeq.= k1/nE3

During equilibrium reactions, the heavier isotope preferentially accumulates in the compounds with a higher number of bonds. During phase changes, the ratio of heavy to light isotopes in the molecules also changes. For example, as water vapor condenses in rain clouds (a process typically viewed as an equilibrium process), the heavier water isotopes (18O and 2H) become enriched in the liquid phase while the lighter isotopes (16O and 1H) remain in the vapor phase. In addition, the equilibrium isotopic effect decreases as the system temperature increases.

Kinetic fractionation processes are also associated with incomplete processes like evaporation, dissociation reactions, biologically mediated reactions and diffusion [4]. Kinetic isotope fractionation reflects the difference in the bond strengths or motilities of the isotopic species. The degree of isotopic fractionation associated with a reaction is commonly expressed with α, which is the ratio of rate constants for molecules containing the different isotopes:

α =14k/15kE4

where 14k and 15k are the rate constants for molecules containing the light and heavy isotopes, respectively. Most biological reactions discriminate against the heavier isotope, yielding 1.00<α<1.03. This makes it convenient to define an isotopic enrichment factor (ε) that highlights more clearly the range of variation:

ε=(α 1)× 1000E5

An alternative method for describing isotopic enrichment is the Rayleigh relationship. This is an exponential relationship that describes the partitioning of isotopes between two reservoirs as one reservoir decreases in size. The equation can be used if the following conditions are met: 1) material is continuously removed from a mixed system; 2) the fractionation accompanying the removal process at any instance is described by the fractionation factor α, and 3) α does not change during the process. Under these conditions, the evolution of the isotopic composition in the residual (reactant) material is described by:

R = R0f(α1),E6

where R is the isotopic ratio of the product, R0 is the initial ratio of the reactant, f is the fraction of the substrate pool remaining and α is the kinetic fractionation factor.

2.3. Measurements of δ15N

Stable isotope measurements are becoming a routine tool in studies of marine ecosystems due to the increasing availability of mass spectrometers. Biological oceanographers usually using continuous flow systems for 15N measurements that integrate a preparatory system (e.g., an elemental analyzer) with a mass spectrometer [5]. The mass spectrometer typically measures the isotopic composition of N2 generated by combustion and carried in a stream of He gas to an open split interface that introduces a small fraction of the gas stream into the ion source for measurement. This analytical approach requires little or no manipulation of organic samples, though care must be taken to avoid contamination with exogenous nitrogen, particularly when tracer-level 15N experiments are being carried out in the vicinity. Since N2 is typically the analyte, minimizing atmospheric contamination in the preparatory system is also quite important. Alternatively, N2O gas may be used for isotope measurements, which practically eliminates the potential contamination from the atmosphere and gives additional information of oxygen isotope composition of nitrogen compound (e.g nitrate or nitrite). In this case, nitrogen compounds must be chemically converted into N2O prior the isotope analysis [6-11]. Special care has to be taken due to the potential exchange of oxygen isotopes between nitrate and water leading to deviation of initial value (e.g. [10]).

Tracer-level 15N experiments used for rate measurements typically add about 10% of the ambient source concentration of 15N-compound. The initial planning of tracer studies requires consideration of several practical issues. First of all, one has to determine the length of incubation, which is connected to turnover rate of the relevant pools and the detection limits for the analytical methods. The turnover rate (pool size-1×uptake rate) has to be estimated for both source and product pools to ensure that tracer can be detected in the product pool, while dilution of the source pool is either insignificant or small [12, 13]. Increasing incubation time will increase the amount of tracer in the product but it will reduce the potential of isotope dilution affecting the source pool. Given a best-guess turnover rate for the product pool, one can estimate the relative trade-offs between incubation time and changes in isotopic labeling of the source and product pools at a given enrichment level (7).

Turnover rate × incubation time × source enrichment = change in enrichment, e.g., 0.1 d-1× 0.2 d-1× 10% = 0.2%.E7

Generally the shortest incubation possible would resolve this dichotomy for biogeochemical processes such as nitrification or denitrification. However, complications may arise for uptake studies since short incubations primarily represent the transport (movement across the cell membrane) of N-compounds, whereas longer incubations primarily represent assimilation into organic compounds such as amino acids or proteins.

Finally, one has to consider the degree of replication and treatment of the initial, or time zero, concentrations and enrichment of both the product and source pools. To account for small, natural (%) level changes in these pools, studies either assume that particles have constant natural abundance values, or they assess the natural abundance periodically during the study. Alternatively, tracer can be added to the incubation bottle, which is then deployed and then immediately filtered at an initial time point for comparison with the final values. One then needs to decide how to interpret the initial time point, as there is often a substantial time between tracer addition and filtration (often 30 min). During this interval, the population may experience abnormal conditions (e.g., high light exposure or rush uptake due to nutrient perturbation) prior to filtration. It was also shown recently by Mohr et al. that measurements of N2-fixation rates by introducing a tracer in bubble form can lead to significant underestimation of the process [14]. The equilibration, i.e. the isotopic exchange between the 15N2 gas bubble and the surrounding water is controlled primarily by diffusive processes. The major variables that influence the rate of isotopic exchange include the surface area to volume ratio of the bubble, the characteristics of the organic coating on the bubble surface [15], and temperature and the rate of renewal of the water-bubble interface [16]. To avoid this problem, the authors recommend adding an aliquot of 15N2-enriched water to incubators used in rate measurements.

Samples of oceanographic interest often pose additional analytical challenges simply because they are difficult to obtain in large quantities and most mass spectrometry systems require fine-tuning and careful attention to leaks in order to process samples containing less than a few micromoles of nitrogen. In practice, open ocean samples including suspended particles, small zooplankton, and sinking organic matter are often available only in µmole or sub-µmole quantities. These low concentrations create a strong incentive to minimize the mass requirements of the analytical systems used for marine samples, and to develop methods for correcting for the influence of any analytical blank, which will disproportionately affect small samples. Even with substantial care to reduce this source of contamination, a nitrogen blank on the order of 0.05–0.15 µmol N is typical for systems in current use [17]. For analysis of the smallest samples, the influence of this blank can be reduced provided that an appropriate range of mass and isotopic standards is run, allowing estimation of the size and δ15N of the blank.

Operationally, isotope ratio mass spectrometers measure the difference in 15N abundance between a sample and a reference gas calibrated to atmospheric N2. This reference gas calibration may be carried out directly by comparison to atmospheric values [18], or indirectly using any of a number of organic and inorganic secondary standards (e.g., nitrate and ammonium salts, acetanilide, glutamic acid) available from NIST or IAEA. For precise system calibration and correction of blank effects, it is also important to analyze a size series of a working standard that is chemically similar to the samples of interest.

Correcting for an analytical blank is a multi-step process involving separate regression analyses to evaluate the size of the blank and the dependence of δ15N values on size. This information can then be used to remove the isotopic influence of the blank. Any real sample can be treated as a mixture of sample material and contaminants of various origins. If the blank contribution is constant in composition and magnitude across analyses, then a simple mass balance can be constructed for the mixture analyzed:

Mmixδ15Nmix= Msampleδ15Nsample+ Mblankδ15Nblank,E8

where M is the size of an individual pool (sample, blank, or the mixture). In general, the nitrogen analytical blank is small enough that blank corrections are minor for samples containing more than 1.5–2 µmol of nitrogen.

Advertisement

3. Sampling for δ15N measurements

Modern water column nitrogen isotope analysis is used on broad variety of samples as particles and dissolved nitrogen species, e.g. nitrate, ammonia or N2 and N2O. For each of these methods, special care must be taken during sample collection and preparation.

3.1. Particles

The simplest and most common method to remove particles from incubation water is by filtration. During the filtration special care should be taken to minimize the pressure difference to avoid strong cell rupture [18]. Several studies showed that Teflon and aluminium oxide filters capture up to 60% higher amount of particulate organic nitrogen (PON) than the GF/F filters [19, 20]. GF/F filters have been also found to retain only ~50% of bacteria [21]. In addition to incomplete retention of particles, glass-fiber filters adsorb dissolved organics on the high surface area of the filter [22]. For 15N studies, although mass retention of DON is not a major issue, retention of the highly labeled incubation water on the filter has the potential to affect the isotopic ratio of the retained particulates so that wetting the filter prior to filtration and rinsing afterwards with filtered seawater is a common practice. Filtered seawater is used to prevent osmotic shock of the living particles. Regardless of the type of filter, the particle load needs to be preserved and converted into the gaseous form prior to mass spectroscopic analysis. δ15N of particles can be analysed directly via combustion (conversion into N2) or via persulfate oxidation (conversion into NO3-) [23].

3.2. Nitrate and nitrite

The most abundant dissolved nitrogen specie in the ocean is nitrate (NO3-) followed by nitrite (NO2-). In order to perform the isotopic analysis both have to be converted into forms amenable to enter the mass spectrometer. The methods differ by level of reduction, with conversion from NO3- to NO2-, NO, N2 or complete reduction to ammonium.

Although NO2- concentrations are usually very small compared to NO3-, there are cases when both should be analyzed without an interference of δ15N signals of one another. To do this one has to separate NO3- and NO2- in the sample by excluding one of them from the sample. Olson [24] used sulfamic acid, while Yakushiji and Kanda [25] took advantage of the reaction of sulfanilic acid with NO2-. Rather than forming the final azo dye, the diazonium ion can be destroyed by heating which effectively destroys the initial NO2- present in the sample. Two new techniques remove nitrite by reaction with sodium azide in acetic acid buffer to produce N2O [7] or via reaction with ascorbic acid to produce NO which is then removed bubbling [11].

Nitrate conversion to ammonia commonly uses DeVarda’s alloy approach [26]combined with moderate heating. However, this method can add a significant blank that varies with the batch, and combustion of the alloy failed to reduce the blank while also reducing the reduction efficiency [27]. Tanaka and Saino [28] replaced Devarda’s alloy with an aluminum reagent that reduced the blank and permitted analysis of nitrate at low concentrations.

Biological reduction of NO3- to NO2- or gaseous forms has been used by several investigators. Risgaard-Petersen et al. [29] and [30] used cultures of denitrifying bacteria to respire NO3- and NO2- toN2 which was then stored for analysis by mass spectrometry. Sigman et al [8] changed this method by using mutant strain of a denitrifying bacteria that only produced N2O to avoid the problem with contamination of N2 from the air. The chemical alternative using sodium azide buffer to produce N2O from NO2- was developed by McIlvin and Altabet [7] for natural abundance measurements in seawater samples. It was further improved for samples with low salinities [6] and can be used to analyze nitrate as well as nitrite in the samples.

3.3. Ammonium

Ammonium is not necessarily dominant by mass, but it is often predominant with respect to flux. NH4+ is produced primarily by the breakdown and remineralization of organic forms of nitrogen and consumed by autotrophic assimilation and re-incorporation into organic molecules, chemosynthetic nitrification to NO2-, and anammox oxidation by NO2- to N2. For isotopic analysis ammonia has to be extracted quantitatively from the solution and converted to suitable gaseous species (e.g. N2).

Initial approaches to isolating ammonium for isotopic analysis by steam distillation were derived from the soil literature. In all methods involving NH4+, great care is required to minimize ambient contamination (e.g., cleaning compounds such as Windex), from sample/reaction containers and from reagents.

Distillation of NH4+ under alkaline conditions with subsequent conversion to dinitrogen gas has been used by several groups but is time consuming and has problems with cross-contamination and fractionation [31-34]. In the diffusion method, ammonia passes through a membrane into an acidic medium and is trapped as an NH4+ salt [34-37]. This approach, however, is also time-intensive and not very reliable at low concentrations [34]. For freshwater samples Hg precipitation and cation exchange are reported to show very good results for isotope analysis [32, 38]. An organic reaction with NH3 producing indophenol permits organic or solid-phase extraction works well for 15N-labeled samples [32] but cannot be used for samples with [NH4+] < 5µM [38]. Zhang and Altabet [39] developed a robust method for sea water samples at low concentrations (0.5 µM or lower), which involves oxidation of ammonia into NO2- and subsequent reduction to N2O.

3.4. N2O and N2

In addition to the previously discussed methods where production of N2 or N2O from other forms of nitrogen is an analytical step, there are related methods for analysis of the gases where the gas itself is the form of interest. N2O can be formed during nitrification and denitrification processes, while N2 is the end product of denitrification and anammox. Isotope paring technique (IPT) for analysis of N2 requires analysis of all three isotopic variants of N2 (28N2, 29N2, 30N2) [40]. Either a quadrupole or magnetic sector mass spectrometer can be used for the isotope pairing analysis, but the quadrupole can also simultaneously determine the N2/Ar ratio. Extraction of N2 from water samples, generally stored in a gas-tight container sealed with a septum (e.g., Exetainer), involves partitioning of the gas into a suitable headspace such as helium. The gas is allowed to equilibrate and the headspace is sampled for introduction of the gas phase into the mass spectrometer [41].

Rather than extraction into a headspace, several investigators have employed membrane inlet mass spectrometers (MIMS) to directly sample N2 for denitrification and nitrogen fixation studies [30, 42]. MIMS can also be used directly to measure denitrification by monitoring changes in the N2/Ar ratio over time since the Ar composition reflects changes in temperature and other physical factors while change in the N2 concentration reflects the balance between nitrogen fixation and denitrification [43]. An advantage of combining MIMS with the isotopes is the ability to separate the two contributions to the net flux. MIMS also obviates the need for the generally slow, separate extraction step. The membrane inlet configuration varies but always features a gas-permeable membrane that separates the mass spectrometer from the solution [44, 45]. A probe inlet [44]permits analysis of depth zonation of rate processes in sediments but has only been employed for N2/Ar measurements.

Analysis of 15N2O during nitrification and denitrification rate studies has become increasingly popular. Punshon and Moore [46] further developed the method to extract 15N2O produced from either 15NH4+ or 15NO3- by natural samples. Using purge-trap gas chromatography with a quadrupole mass spectrometer, they were able to determine production rates for N2O from nitrification. The origin of N2O (denitrification or nitrification) can be determined using isotopic signature of oxygen isotope. δ18O isotope signature in N2O molecule is a product of mixing from dissolved O2 and H2O and any fractionation process associated with its production [47]. For instance, production of N2O via hydroxylamine oxidation (nitrification) will only include oxygen atoms from dissolved O2, while via denitrification it will have δ18O isotope of NO2- or NO3-, originating from both dissolved O2 and H2O [47, 48].

The site preference isotopic signature (isotopomers) is also used to identify pathways of N2O production. In the molecule of N2O the 15N can be at a central (α) or outer (β) position. The intermolecular distribution of 15N is expressed as site preference (SP) (eq.9)[49].

SP=δ15Nα- δ15NβE9

In contrast to δ18O signature, SP is not dependent on the substrate’s isotopic composition but reflects the microbial production mechanism [48]. For example, production of N2O via hydroxylamine oxidation results in an SP of 33‰, while via denitrification the SP is 0‰ [50]. Good summery on isotopomers for soil studies is presented in the book of Environmental Isotope Geochemistry [51] showing the variable range of SP values for nitrification. In general, isotopomers are valuable tool for distinguishing between nitrification and denitrification processes of N2O production.

Advertisement

4. Main processes of nitrogen cycle and their influence on isotope balance

Most nitrogen in marine environments is present in five forms: N2, a quite stable molecule that requires specialized enzymatic systems to break and use it; NO3-, the most oxidized form of nitrogen and the dominant form of N within oxic environments; NH4+, the most reduced natural form of N and the dominant biologically available form found in anoxic environments; particulate nitrogen, predominant within sediments and primarily in the form of organic N, and dissolved organic N (DON), which is a complex mixture of compounds with a wide range of compositions [52-55]. Nitrate, nitrite, ammonium, and organic nitrogen are typically grouped together as “fixed N” in discussions about the availability of nitrogen. Although each form has a different level of reactivity, a complex web of reactions links these different compounds. The most important processes discussed here are: N2 fixation, nitrification, assimilation and N-loss (conversion of fixed N to N2). As it was mentioned already earlier, all biologically mediated processes lead to nitrogen isotope fractionation. For example, denitrification converts NO3- to N2, which has the net effect of depleting the ocean’s supply of combined nitrogen and doing so with strong discrimination against 15N. This results in an increase in the δ15N of the major pool of oceanic combined nitrogen (NO3-) concomitant with a reduction in its overall size. In contrast, N2-fixation adds combined nitrogen with a low δ15N to the ocean, counteracting both the mass and isotopic effects of denitrification (figure 1). In the subsequent parts of this chapter I will discuss all these processes in detail and their influence on the nitrogen isotope budget in the ocean.

Figure 1.

The impact of different processes on the δ15N of oceanic NO3- [56]. Axes show deviation of δ15N signal from oceanic average and loss/input of nitrogen due to different processes.

4.1. Nitrogen fixation in the ocean

Microbial N2 fixation in the ocean inputs “new” N into the surface waters by reducing N2 gas into two molecules of ammonia. The reaction is logically consumes energy and is catalyzed by nitrogenase enzyme complex, which is extremely oxygen sensitive (req. 10).

N2+ 8 H++ 16 ATP + 8 e 2 NH3+ H2+ 16 ADP + 16 PiE10

The nitrogenase proteins are highly similar among diazotrophs, and the well-conserved nifH gene is commonly used for phylogenetic and ecological studies [57]. In the open ocean N2 fixation research has focused most intensively on the cyanobacterium, Trichodesmium [58, 59], which often occur as aggregates (often referred to as colonies) visible to the naked eye (‘sea sawdust’). It can also occur, however, as individual filaments. In early reports N2 fixation was observed in the pelagic zone and associated with the cyanobacterial epiphytes (Dichothrix fucicola) of the brown macroalga, Sargassum [60] and cyanobacterial endosymbionts of certain oceanic diatoms [61, 62]. Trichodesmium has a unique physiology that employs spatial and temporal segregation and increased oxygen consumption to allow it to simultaneously fix N2 and CO2 (and thus evolve O2)[63].

Evidence of N2 fixation began to first accumulate in the late 1970s with several pioneering N isotope studies examining the natural abundance of the ratio of δ15N in the western Pacific Ocean [59, 64]. The average isotopic signature of nitrate in the deep ocean lay about 5‰ [27, 65, 66]. Nitrogen fixation is the process by which N2 is incorporated from the air into the cell. As the δ15Nair is set to 0‰, the kinetic isotope effect of the process is small (-2‰ to +2‰). Surface particulate and dissolved inorganic N pools in the Kuroshio often had a 15N-depleted isotopic signature, indicative of a N2 fixation source [67]. Using nifH gene studies, N2-fixing was found also in the picoplankton, as well as in heterotrophic bacteria from the guts of copepods [68]. Heterotrophic N2 fixation does not consume CO2 during the process as diazotrophs do but instead produces it. However, the mechanism of heterotrophic N2 fixation is not yet understood and it is still unclear and the potential contribution of these organisms to marine N2 fixation is uncertain.

4.1.1. Controlling factors for N2 fixation and its distribution

Numerous factors, physical, chemical and biotic, can affect the extent of N2 fixation in an ecosystem [69-72]. Many factors that affect nitrogenase activity co-vary, such as light, temperature, oxygen and turbulence, making it difficult to determine the relative impact of each. Species-level responses to these variables also differ. The heterocystous diazotrophs, for instance, are typically restricted to special marine environments. Metabolically active populations of Trichodesmium have been observed at 18.3 °C [54], but activity was low, and substantial growth is typically not seen until water temperature exceeds 20 °C [73]. In the cooler Baltic waters Nodularia spp. and Aphanizomenon spp. Are observed, which are restricted to 16–22 °C and 25–28 °C, respectively [74]. In warmer waters of the tropics and subtropics, heterocystous diazotrophs (such as Richelia intracellularis) are only encountered as symbionts of diatoms like Hemiaulus spp., and Rhizosolenia spp. [75].

For photoautrophic diazotrophs, nitrogenase activity is intimately linked to photosynthesis [76]. Thus, light is an obvious and important factor potentially regulating or constraining this process. Carpenter [73] has summarized much of the early marine work largely relating to Trichodesmium spp. with respect to their relationship to light. Whereas many non-heterocystous cyanobacteria fix nitrogen during the night, and thereby uncouple N2 fixation from photosynthesis, Trichodesmium fixes nitrogen exclusively during the light period and shows a strong diel pattern of activity with maxima during midday [59, 63, 69, 77]. Natural populations of Trichodesmium, often found in the upper layers of the euphotic zone, appear to be adapted to high light with a relatively shallow compensation depth (typically 100–200 mmol quanta m-2 s-1) for photosynthesis [78].

To calculate how much nitrate is produced via N2-fixation, Redfield stoichiometry can be used [79]. With the given NO3-exp and the measured concentration of the actual dissolved inorganic nitrogen (DIN = NO3-+ NO2-+ NH4+), the amount of NO3-, which is removed by denitrification (NO3-def), can be calculated as follows:

NO3-def= NO3-exp– DINE11

An alternative stoichiometric method involves the quasi-conservative tracer N* [80, 81]. Gruber and Sarmiento used high quality data from JGOFS and GEOSECS databases to develop a general relationship between fixed inorganic nitrogen (DIN) and phosphate (PO43-) for the world’s oceans:

N* = DIN – 16 [PO43-] + 2.90E12

The value of 2.90 is the deviation from the amount of DIN predicted by the Redfield stoichiometry (N:P =16) and the world-ocean N:P regression relationship. Negative values of N* are interpreted to show the net denitrification whereas positive values show the net nitrogen fixation.

Figure 2.

N*= [NO43-] – 16*[ PO43-] + 2.9 on the 26.5 sigma theta surface (World Ocean Atlas 2005). Note the large negative N* values in the Tropical Pacific and Arabian Sea and positive N* values in the Tropical Atlantic Ocean.

Strong positive N* values have been observed in the tropical and subtropical North Atlantic (fig.2) and are proposed to be a result of the input into these areas of diazotrophic (i.e. nitrogen fixer) biomass with a higher N:P content than that typical of eukaryotic phytoplankton of the upper ocean [80-82]. The observation of relatively high concentrations of DON in surface waters of regions of the tropical oceans have also been attributed to nitrogen fixation [83-86]. At the Hawaiian Ocean Time (HOT) series station, pools of DON increased during a period in which microbial nitrogen fixation also became more prominent [87, 88].

Iron limitation may be important for cyanobacterial N2 fixation as Fe is a cofactor in photosynthesis and nitrogenase, leading to is ~5 times higher Fe requirement for diazotrophy than that for ammonia assimilation [77]. Even though the concentration of dissolved Fe (dFe) is low in many of the environments where N2 fixation occurs [89-92], a number of studies have speculated on diazotroph response to mineral dust Fe fertilization [93, 94]. A 1999 Saharan dust event coincided with increases in dissolved Fe concentrations on the west Florida shelf and a 100-fold increase in Trichodesmium biomass [95]. N2 fixation rates were not measured, but DON concentrations doubled, presumably due to exudation by N2 fixers [95]. A Saharan dust addition experiment to surface water samples collected along a west African cruise transect (35°W–17°W), found a minimal increase in CO2 fixation and a large stimulation in N2 fixation, suggesting that diazotrophs were co-limited by both P and Fe [96]. It should be noted however, the analysis of aerosol dust shows that while providing Fe, the dust also supplies P and combined N [96-98]. Work on Trichodesmium spp. culture has identified mechanisms to deal with P limitation. For instance, under P stress conditions it uptakes a greater amount of DIP and dissolved organic P (DOP), decreasing the requirement for cellular P.

N2 fixation is an energetically costly process in comparison to ammonia assimilation, thus it is logical to assume that presence of high concentrations of fixed N play a role of an inhibitor. Indeed, under mM concentrations of ammonia (concentrations essentially never observed in the real surface ocean), N2 fixation is inhibited in most diazotrophs [99]. However, mM concentrations of NO3- stopped N2 fixation in some strains [100] but have little to no effect in others [101]. The higher energy requirement for the nitrate reduction compared to ‘direct’ assimilation of ammonia might explain these differences. Addition of ammonia to natural samples of Trichodesmium spp. can shut down nitrogenase activity [69]. However, pure culture studies have shown mixed results, with some cultures indicating no reduction of their activity [102] and others showing a reduction in activity under short-term additions of fixed N sources [103]. Interestingly, indication for N2 fixation has been found also in the nitrate-rich upwelling areas off Chile and Peru [104, 105], indicating that N2 fixation may not be as strongly regulated by fixed N as previously presumed.

4.1.2. Nitrogen fixation in sediments

Within the marine realm, N2 fixation in sediments has been mostly recognized in pelagic environments or in benthic microbial mats, sea grass sediments, and coral reef systems [106]. A recent study by Bertics et al. [107] revealed that N2-fixation, coupled to sulfate reduction, is stimulated by bio-irrigation in organic-rich coastal sediments. The authors suggested that N2-fixation has been largely overlooked in sediments, and that its occurrence might be significant, contradicting the general belief that sediments are areas of net nitrogen loss through denitrification and anaerobic ammonium oxidation [108, 109].

4.2. Nitrification

Nitrification process oxidizes NH4+ into NO2- and further to NO3-. Ammonium rarely occurs at high concentrations in the surface ocean as it is rapidly recycled between heterotrophic and N2 fixing organisms and many heterotrophic and photosynthetic plankton species. N2 fixing organisms play a role of ammonia source as they excrete NH4+ directly or release organic N that is microbialy degraded to NH4+, while plankton utilize ammonia as a N source for building up its biomass. Similarly, accumulation of NO2- is rarely observed in oxygenated habitats, although NO2- is an essential intermediate in several oxidation and reduction processes in the N cycle.

Nitrification is carried out through a combination of two microbial processes: ammonia oxidation to NO2- and nitrite oxidation to NO3-. The first step of the process is enabled by nitrifying organisms, ammonium-oxidizing bacteria and archaea (AOB and AOA), while the second step is carried out by nitrite-oxidizing bacteria (NOB). No known organism can carry out both reactions. The overall reaction of NH4+ oxidation (eq. 13) for AOB shows that the process consumes molecular oxygen and produces NO2- and hydrogen ions. Molecular oxygen is required in the first step of the reaction, which is catalyzed by a monooxygenase (NH3 monooxygenase, AMO). Oxygen is also consumed by the terminal oxidase as a result of electron transport generating adenosine-5’-triphophate (ATP) for cellular metabolism. The immediate product of AMO is hydroxylamine, which is further oxidized by hydroxylamine oxidoreductase (HAO) to NO2-. In contrast, AOA apparently do not possess the hydroxylamine reductase gene, so the pathway of ammonia oxidation in these organisms must be quite different.

NH4+ + O2+ H++ 2 e-AMONH2OH + H2OHAONO2-+ 5 H++ 4 e-E13

Depending on the conditions, NO, N2O and even N2 have been reported as secondary products in autotrophic NH4+ oxidation by both marine and terrestrial strains [110, 111]. Although N2O and NO can be produced in vitro by HAO from hydroxylamine [112], the reduction of NO2- appears to be the dominant pathway in all cells [113-115]. One of the first studies of kinetic isotopic effect of nitrification in pure cultures of Nitrosomonas europaea was made by the Mariotti group [116], where a fractionation factor of 34.7 ± 2.5‰ was determined for conversion of NH4+ into NO2-. The later work of Casciotti [117] identified that the fractionation factor for different Nitrosomonas cultures of ammonia-oxidizing bacteria was linked to amino acid sequences. The isotopic effect in these cultures ranged from 14.2‰ to 38.2‰.

N2O is produced by AOB through two separate pathways: hydroxylamine decomposition and nitrite reduction, so-called “nitrifier denitrification” [114]. The isotopic compositions (δ15Nbulk, δ18O, δ15Nα, δ15Nβ, and site preference = δ15Nα - δ15Nβ) of the produced N2O can provide insight into the mechanisms of N2O production under different conditions [48, 50, 118]. For example, N2O production through nitrifier denitrification has low δ15Nbulk and low site preferences relative to that produced by hydroxylamine decomposition [119]. As discussed earlier, the N2O site preference (SP) is determined by the enzymatic mechanism [49, 120]. The site preference of N2O produced during nitrification is +30‰ to +38‰, while N2O production via denitrification and nitrifier-denitrification has a SP of -10‰ to +5‰ [48, 118, 121, 122]. The large difference between the SP values of these two mechanisms of N2O production provides a good tool with which to distinguish their contributions.

The biochemistry of the NO2- oxidation is simpler than that of NH4+ oxidation because it is only involves a transfer of two electrons without intermediates (eq. 14). The additional oxygen atom in NO3- is derived from water, and the molecular oxygen that is involved in the net reaction results from electron transport involving cytochrome oxidase.

NO2- + H2Onitrite oxydoreductaseNO3-+ 2 H+ + 2 e- + 0.5 O2cytochrome oxidaseNO3- + H2OE14

A recent study of isotopic fractionation during nitrite oxidation to nitrate [1] showed a unique negative isotopic effect of -12.8‰. Because of the inverse isotope effect, when nitrite oxidation is taking place, the δ15NNO2 (and δ18ONO2) values are expected to be lower than the NO2- initially produced by ammonia oxidation or nitrate reduction. In most parts of the ocean, however, NO2- does not accumulate and the isotope effects of different processes complicate the interpretation. In the euphotic zone NO2- is either oxidized to NO3- or assimilated into particulate nitrogen (PN) and the heavy isotope can be preferentially shunted in one direction vs. the other [119].

4.2.1. Controlling factors of nitrification and its distribution

Nitrifying bacteria are traditionally considered to be obligate aerobes; they require molecular oxygen for reactions in the N oxidation pathways and for respiration. A significant positive correlation between apparent oxygen utilization (AOU) and N2O accumulation is often observed in marine systems [123-125]. The relationship implies that nitrification is responsible for N2O accumulation in oxic waters where it is released as a byproduct. The relationship breaks down at very low oxygen concentrations (~6 mM, [126]), where N2O is consumed by denitrification.

There is abundant evidence from culture studies that both AOB and NOB are photosensitive. Several studies of nitrification rates in surface seawaters from various geographical regions show profiles that are consistent with light inhibition of both NH4+ and NO2- oxidation [127-129]. The highest nitrification rates occur in a region near the bottom of the euphotic zone. It is in this interval where nitrifying bacteria can compete with phytoplankton for NH4+, as the rates of nutrient assimilation are reduced due to light limitation. A sharp peak of nitrification rate is often observed at water column depth where the light intensity is 5–10% of surface light intensity [118, 127-129].

In both deep and shallow sediments, nitrification can be one of the main sinks for oxygen [130, 131]. In continental shelf sediments, nitrification and denitrification are often closely linked. Coupled nitrification/denitrification is invoked to explain the observation that the rate of N2 flux out of sediments can greatly exceed the diffusive flux of NO3- into the sediments [132]. Ammonium is produced during aerobic and anaerobic remineralization of organic matter. It is then oxidized to NO3- and subsequently reduced to N2. Anaerobic oxidation of NH4+ can be a factor supporting the imbalance between supply and consumption of NO3-. Nitrification can supply up to 100% of the NO3-, which is later consumed by denitrification [133, 134].

Although oxygen and NH4+ conditions likely differ between planktonic and sediment environments, there is no clear evidence from clone libraries that water column and sediment nitrifying communities are significantly different in composition and regulation.

To measure nitrification rates, isotope dilution can be used. In this method, the isotope dilution of the NO2- pool (in the case of NH4+ oxidation) or the NO3- pool (in the case of NO2- oxidation) is monitored over time [135]. Alternatively, in the 15N method 15NH4+ or 15NO2- is added to a sample and then the appearance of 15N label in the NO2- or NO3- pools is measured over time [24]. One disadvantage of the method is that the increase in NH4+ or NO2- concentrations caused by the tracer addition can result in enhanced NH4+ or NO2- oxidation rates [129]. When small additions of 15N are used, the isotope dilution of the substrate pool must be monitored over time.

4.3. Nitrogen assimilation

Nitrification and nitrogen assimilation are usually found at similar depths in the ocean as they compete for the same reactants (NH4+). At the same time, nitrification, producing NO2- and NO3-, provides assimilation with additional reactants. Nitrogen assimilation is the process of incorporation of reactive nitrogen species (NO3-, NO2- and NH4+) into the bacterial cell. Ammonium, which is produced in the photic zone by heterotrophic processes, is assimilated immediately by phytoplankton and heterotrophic bacteria before it can be nitrified. Ammonium is often the dominant form of dissolved N assimilated in a variety of marine and estuarine systems. This form of N is energetically efficient for cells to use, because it is already reduced and is a common cellular transient in N metabolism, requiring little additional energy for assimilation. Despite the low NH4+ concentrations in oceanic systems, uptake and regeneration of NH4+ are tightly coupled and result in rapid turnover times [136].

Nitrate concentrations in the surface ocean are usually maintained at low levels as a) phytoplankton assimilate NO3- rapidly and b) nitrate can be supplied by mixing or diffusion from the deep NO3- reservoir only at the regions with strong upwelling. The assimilation of NO3- is more energetically demanding than NH4+, because it requires a synthesis of NO3- and NO2- reductases, associated active transport systems, and the turnover of cellular ATP and nicotinamide adenine dinucleotide phosphate (NADPH) [136]. In order to assimilate NO3- for growth, phytoplankton must first possess the genetic capacity to synthesize the necessary enzymes and transport systems- a capacity which not all phytoplankton have. Further, the supply of NO3- is limited by nitrification and vertical mixing.

Since nitrite reductase (NiR) is required for NO3- assimilation, organisms that assimilate NO3- can by default assimilate NO2-. Consequently, some organisms that cannot use NO3- can still use NO2- (e.g., some clones of Prochlorococcus; [137]). It has been suggested that this rather unique N physiology has arisen from the evolutionary loss of genes, which are necessary for the assimilation of NO3- [138, 139]. The resultant evolution of many ecotypes specifically adapted for unique environments where NO2- production can be high.

Nitrogen isotope fractionation can take place at different stages of the processes. Wada and Hattori [140] first argued that N isotope fractionation by phytoplankton occurs during nitrate reduction, while Montoya and McCarthy [141] ultimately favored fractionation associated with nitrate transport into the cell. The N isotope data presented in Granger at al. [142] suggest that an isotope effect is driven solely by the reductase, and not by uptake at the cell surface and that this effect results in a fractionation factor of 5 – 10‰. Ammonium assimilation is shown to have higher fractionation factor of +14 – +27‰ [143, 144].

In most parts of the ocean, however, assimilation and nitrification occur concurrently, which additionally complicates the isotope analysis. In general, nitrite oxidation, as discussed above, transfers NO2- with an elevated δ15N to the NO3- pool, while nitrite assimilation transfers the residual NO2- with a lower 15N:14N ratio into the PN pool. If the fractionation factor of nitrite assimilation is 1‰ [144] and fractionation factor of nitrite oxidation is -15‰ [1], then the δ15NNO2 at steady state will be lower than the source of NO2-, unless nitrite assimilation is >95% of the NO2- sink. This has the opposite effect of the ammonia oxidation/assimilation branching where ammonia oxidation transfers low 15N:14N material into the NO2- and NO3- pools and higher 15N:14N material into the PN pool [119].

4.3.1. Factors controlling nitrogen assimilation and its distribution

Most nitrogen assimilation takes place in the surface waters of the ocean. The most important factors controlling assimilation are the presence of oxygen for respiration and/or light for photosynthesis and enzymes responsible for metabolic processes. Most of the enzymes involved in the uptake and assimilation of N are bound to energy sources and thus are affected by light, presence of oxygen, and the supply of enzyme co-factors and metabolic substrates. For example, uptake and reduction of NO3-, NO2-, and urea have been linked to the light supply in phytoplankton due to the necessity for ATP and NADPH from photo-phosphorylation. While uptake and reduction of these compounds is thought to proceed at maximum rates only in the light under nutrient replete conditions, active uptake of these compounds may occur even in the dark [145].

The accumulation of intracellular product pools, on the other hand, can result in the inhibition of uptake and assimilation. The posttranslational modification of enzymes can regulate uptake and assimilation by modulating the number of active sites available for catalysation of specific reactions. Studies examining the transcriptional activation of enzymes involved in the process have begun to demonstrate clear functional relationships between genomes and ecological capabilities. For example, the presence of NH4+ represses protein expression (i.e., enzymes) involved in the assimilation of alternative N sources (e.g., NO3- and N2) in some organisms [146]. Overall, most factors controlling assimilation are directly connected with enzymes, which require substrate compounds or energy and oxygen supply for their metabolic activity.

4.4. Nitrogen loss (denitrification and anammox) in the ocean

The term “nitrogen loss” is commonly used for the processes in which fixed nitrogen (nitrate, nitrite or ammonia) is converted into non-bioavailable N2 in the absence of oxygen. This includes denitrification process, in which combined nitrogen is reduced to gaseous end products, as well as newly discovered process of anammox (anaerobic ammonium oxidation), which converts nitrite and ammonia into N2 [147]. Another important process of reduction is DNRA (dissimilatory nitrate reduction to ammonia) will be also discussed in this section as it is an important source of ammonia in the ocean. Even though the reduction of nitrate to ammonia is not a N-loss as such, but it is also taking place in suboxic waters, and the ammonia produced can diffuses up to the anoxic waters and may well be oxidized to N2 by anammox bacteria [148].

4.4.1. Canonical denitrification

Canonical denitrification is a heterotrophic process in which nitrogen oxides serve as the terminal electron acceptor for organic carbon metabolism [149]. The nitrogen oxides are reduced mainly to molecular nitrogen and some nitrous oxide may be formed as a side product. The sequential reduction is enabled by four well-studied enzyme systems: nitrate reductase, nitrite reductase, nitric oxide reductase and nitrous oxide reductase [150] (eq. 15).

NO3-nitratereductaseNO2-nitritereductaseNOnitricoxidereductaseN2OnitrousoxidereductaseN2E15

The intermediate NO2- is known to escape the cell and is frequently found in denitrifying environments; likewise, N2O also can accumulate externally [151]. This leads to local maxima of both intermediates, which tend to occur near the boundaries of the suboxic zone. Another intermediate, NO, is actually a free radical and, thus, very reactive and highly toxic to most bacteria including denitrifiers [152]. In order to minimize the accumulation of NO within the cell, nitrite reductase and nitric oxide reductase are controlled interdependently at both the transcriptional and enzyme activity levels [152, 153].

4.4.2. Anammox

Anammox bacteria are chemoautotrophic bacteria that fix CO2 using NO2- as the electron donor [154] and are thought to be strict anaerobes. Recent publication on the tolerance of anammox bacteria to oxygen levels, however, showed that they can perform a significant activity even under ~20 µmol/l O2 concentrations [155]. All anammox species found have evolved a membrane-bound intra-cytoplasmic compartment called the anammoxasome. The anammoxasome membrane is made up of high-density lipids, called ladderanes because of their ladder-like structure, which are thought to be specific to anammox bacteria. The proposed mechanism of anammox involves a hydrazine hydrolase, which catalyzes the combination of hydroxyl ammine and ammonium to form hydrazine. The hydrazine-oxidizing enzyme subsequently oxidized it to N2 [156, 157]. The ladderane-lipid membrane is thought to act as a barrier to diffusion thus isolating the toxic intermediates of the anammox reaction within the anammoxasome. Interestingly, for every 1 mol of N2 produced, 0.3 mol of NO3- is also produced alongside from NO2-. This reaction is believed to be important for replenishing electrons for the acetyl-CoA carbon fixation pathway [158]. At the same time, these bacteria can perform dissimilatory NO3- reduction to NH4+ (DNRA), which is subsequently combined with NO2- to produce N2, thus mimicking denitrification [159].

Figure 3.

Morphology of the anammox cell and proposed model for the anammox process. HH, hydrazine (N2H4) hydrolase; HZO, hydrazine oxidizing enzyme; NR, nitrite reducing enzyme [160].

Anammox bacteria have been identified in many naturally occurring anoxic environments including marine, freshwater, and estuarine sediments [161, 162], oxygen minimum zones [163, 164], soils [165-167], wetlands [168], and anoxic tropical freshwater lakes [169]. These bacteria are not only widespread but may also be very active (e.g. ~67% fixed N-loss [162]). Anammox organisms may live in low abundance but are capable of significant growth and metabolism when a shift in environmental parameters permits anammox communities to grow [170]. The anammox bacteria are slow growing, with a doubling time of ∼11 days, even under optimal growth conditions in bioreactors [171]. The state-of-the art method to identify anammox is currently based on isotopic labeling of NH4+ or NOx-. It relies on the fact that N2 produced by anammox combines 1:1 atoms from NH4+ and NO2- [172]. The selective formation of 15N14N (29N2) rather than 15N15N (30N2) is a clear evidence of anammox activity[162].

4.4.3. Dissimilatory nitrate reduction to ammonium or DNRA

The dissimilatory reduction of nitrate to ammonium is a bacteria-mediated heterotrophic process occurring in anaerobic conditions. Nitrate is first reduced to nitrite and then to NH4+. Bock et al. [173] also showed that N. europaea and N. eutropha were able to nitrify and denitrify at the same time when grown under oxygen limitation ([O2] ~ 0.2 - 0.4 µmol/l). Under these conditions, oxygen and NO2- served simultaneously as electron acceptors and both N2 and N2O were produced, whereas under anaerobic conditions N2 was the predominant end product. This type of nitrification–denitrification pathway may help to explain why ammonium oxidizers remain active in nearly suboxic environments and enhanced N2O production is observed [174, 175].

The reaction has been reported for sediments underlying anoxic waters [176, 177] and sediments with substantial free sulfide, possibly due to sulfide inhibition of nitrification and denitrification [42]. Suboxic NO3- reduction to NH4+ appears to occur in environments where there is excess available carbon, such as estuarine sediments [178]. Perhaps more important are the NO3-fermenting organisms such as Thioploca and Thiomargarita, that are found in sediments underlying the major suboxic-denitrifying water columns of the Arabian sea, eastern Tropical Pacific and Namibia [179-181]. These organisms are able to couple the reduction of NO3- to NH4+ with the oxidation of reduced sulfur compounds. Both Thioploca and Thiomargarita are able to concentrate NO3- at up to 500 mM levels in large vacuoles within their cells for subsequent sulfide oxidation [180, 181].

4.4.4. Controlling factors for N-loss processes and its distribution

The controlling factor for all N-loss processes is molecular oxygen. When the concentration of the denitrification intermediate NO2- is plotted against dissolved oxygen, we see that NO2- does not appear in the water columns of the Eastern Tropical Pacific until oxygen concentrations are reduced below about 2 µM (see figure 4 [182, 183]).

Figure 4.

Dissolved oxygen versus nitrite for OMZs. Presented data are taken from publically available cruises I07-N, P19 (http://woce.nodc.noaa.gov/woce_v3/wocedata_2/bathymetry/default.htm) and publications of Cline & Richards [182] and Morisson [183].

Another indication for the denitrification process is consumption of N2O within the water column, where oxygen concentrations are reduced levels less than c. 5 µM [65]. In oceanic systems the relative importance of heterotrophic denitrification, compared to autotrophic anammox has been debated and believed to depend on the amount of organic matter (OM) available [184, 185]. However, in most of the cases N-loss processes co-exist and it is rather a question which one is playing the first fiddle [186, 187].

About less than half of the current marine denitrification of the global ocean is thought to occur in the three main pelagic Oxygen Minimum Zones (OMZs): Eastern Tropical North Pacific (ETNP), the Eastern Tropical South Pacific (ETSP) and the Northern Arabian Sea (Fig. 4). These zones occur in intermediate waters (~150–1000 m) in locations where the ventilation rate is insufficient to meet the oxygen demand. To calculate nitrogen loss in the water column, the amount of substrate that is consumed in N-loss processes or the amount of N2 produced should be determined. Similar to estimations of the N-fixation, the Redfield stoichiometry is used to calculate N-deficit or N*. Negative values of N* are interpreted to show the net N-loss whereas positive values show the net nitrogen fixation (see above).

Devol et al. [188] used N2:Ar ratios in the Arabian Sea OMZ to determine the amount of N2 produced during denitrification. The amount of N2 excess in the OMZ was computed from the increase in the N2:Ar ratio over that present in the source waters. Measurements of N2 excess predicted a larger nitrogen anomaly than that estimated by nitrate deficit. The discrepancy is said to be due to incorrect assumptions of the Redfield stoichiometry [189]. Inputs of new nitrogen through N-fixation, N2 contributions from sedimentary denitrification along continental margins, the anammox reaction, or metal catalyzed denitrification reactions all lead to a shift of the N:P ratio from the Redfield stoichiometry.

Due to N-loss processes, i.e. denitrification and anammox, the δ15N values in the water column increase to 15‰ or more. For the eastern tropical North Pacific OMZ Brandes et al. [190] calculated the fractionation factors between 22‰ (Arabian Sea closed system model) and 30‰ (eastern tropical North Pacific open-system model). Several authors [190-193] have observed high nitrogen isotopic values and nitrate deficits in the Pacific and Arabian Sea OMZ’s and estimated fractionation factor, which all cluster around a value of +25‰. The fractionation factor of N-loss processes in the sediments was shown to be significantly lower (~3 ‰) [194]. Thus, in the waters with strong upwelling, the water column signal can be altered via mixing with water originating from the sediments.

4.4.5. Sedimentary denitrification

Another location where oxygen is typically depleted and denitrification takes place is in marine sediments, especially the continental margin and hemipelagic sediments. Respiratory processes within the sediments act as sinks of oxygen, while overlying water deliver nitrate to the sediment. In most continental shelf sediments, oxygen penetrates to less than 1 cm below the sediment–water interface and even in most deep-sea sediments oxygen penetration is restricted to the approximately upper 10 cm. Thus, ample environments exist for N-loss in marine sediments. Due to the fact that marine sediments are usually rich with organic matter, denitrification and DNRA are the prevailing mechanisms in the sediments [176].

Advertisement

5. Modern ocean nitrogen budget

The discussion of N-budget is usually based on balance or imbalance of nitrogen sources and sinks. Nitrogen sources in this case are N2-fixation, atmospheric wet and dry deposition, and riverine input. Nitrogen sinks include denitrification and anammox in the water column, sedimentary denitrification, and sedimentation of organic matter. Nitrification and DNRA processes “recycle” nitrogen from ammonia into nitrate and vice versa. Although the primary sources and sinks are relatively well known, the quantification of their fluxes is associated with considerable uncertainty [81, 195-199] (see figure 5). The major problem in calculating global budgets of N sources and sinks is extrapolating measured fluxes from a limited number of oceanic regions to the global ocean.

Figure 5.

Nitrogen budget estimations in the literature.

Additional complications in calculating N budgets are caused by interplay between and potential coexistence of N cycle processes at the same depth. A schematic water column profile with corresponding processes over O2, NO2-, and N2O gradients is presented in the figure 6 (based on water column profiles in the South Pacific ocean [65]). In the surface layer, nitrogen assimilation and rapid recycling is taking place, producing nitrite where oxygen and bioavailable N are plentiful enough. Ammonia oxidation (first step of nitrification) lead to efflux of N2O [200] in the surface waters. Ammonium release and its oxidation in the well-oxygenated waters lead to high production of NO2- (primary nitrite maximum). At oxygen concentration ~ 5 µmol/l, denitrification process become more preferential producing to “secondary nitrite maximum” in the OMZ. This NO2- can also be a source for anammox process at the top of OMZ, while DNRA process can provide a source of ammonia [148]. N2O consumption in the water column provides a strong evidence for denitrification producing N2. At the borders of OMZ nitrification and denitrification processes may coexist [196], leading to production of N2O at these water depths.

Figure 6.

Schematic water column profile with corresponding processes over O2, NO2-, and N2O gradients.

Nitrogen isotope ratios in the water column and in sedimentary records provide additional constraints for nitrogen budget calculation. Average δ15N of fixed nitrogen in the deep ocean is close to 5‰. The sedimentary records show that this global average increased during deglaciation periods, when stronger denitrification was taking place. However, the global average δ15N range for the last 30.000 years was in the range 4-6‰ [201], which suggests that N sources and sinks should be balanced, at least the millennial time scale. In this case, the processes increasing δ15N (assimilation, N-loss) has to be balanced with those, which are decreasing it (N2-fixation).

Figure 7.

Nitrogen isotope range and processes influencing on the 15N budget. Dashed line at 5‰ represent an oceanic average value [197, 202, 203].

The conceptual model of the nitrogen isotope budget is initially based on two processes: N2-fixation and denitrification. Sedimentary denitrification was believed not to change the isotopic signature (εsed) of nitrogen due to N-limitation in the sediments, while water column denitrification (εwc) has a fractionation factor of ε between 25 and 35 ‰. The N2- fixation fractionation factor εfix is close to 0‰ as it incorporates nitrogen from the atmosphere. An early attempt to construct a one-dimensional model was undertaken by Brandes and Devol [197], where they set fractionation factor to εfix and εsed to 1.5‰; and εwc to 25‰. On the figure 8 scale of nitrogen isotope is illustrated. In the Brandes and Devol model, sedimentary N-loss does not change the 15N budget substantially (εsed in the range of 0 to 1.5‰). Thus, a large amount of sedimentary denitrification is necessary to balance water column denitrification and bring overall isotopic values of the losses into the range of sources, leading to the value of Sediment to Water column denitrification (W:WC) ratio of 3.8. The authors noticed that with fluxes estimations of total N-loss of 400 Tg N/yr and fixation of 100 Tg N/yr, “the combined budget would be out of balance by over -200 Tg N/yr” [197]. This implies that fixation rates should be 2–3 times higher than 100 Tg N/yr, which was an accepted value at that time. Recent observations that N2- fixation is indeed underestimated by up to two fold support these calculations.

More recent calculation of the nitrogen isotope budget in year 2007 included so-called “consumed-N” into the model, which represents nitrogen utilization in the euphotic zone [202]. During N utilization, nitrogen fractionation takes place, shifting the “initial” value of the ambient NO3- from +5‰ to 17 ‰. Water column N-loss leads to production of N2 gas with δ15N values 25‰ lighter than the “initial”. This process shifts the nitrogen isotope budget significantly and even leads to balance of water column and sedimentary denitrification (S:CW =1). A recent model of the pre-industrial N-budget includes a sedimentary fractionation factor εsed of 4‰ and εfix of 1.5‰ [203]. The value of “consumed-N” was set to +12‰, taking into account the fractionation factor of utilization of 7‰. In this model, the S:CW denitrification ratio varies from 2 to 3.5. Current budget estimates can, however, differ from these prior estimates due to an increase of riverine and atmospheric input of nitrogen into the ocean [198].

Advertisement

6. Conclusions

Of the two approaches whereby nitrogen isotope measurements are used to study marine N cycle processes, stable isotope natural abundance measurements have the great advantage of providing information on integrated processes with little or no manipulation of the system. In contrast tracer-based rate measurements are better used for rate measurements. The natural abundance and tracer approaches are strongly complementary, providing insight into the nitrogen cycle on very different spatial and temporal scales. The magnitude of processes that have distinctive isotopic fractionation patterns can potentially be quantified if the fractionations are known accurately. Thus, there is a significant ongoing need for careful measurement of isotopic fractionation patterns under controlled laboratory conditions as well as under field conditions, which in the best case should be combined. In order to interpret isotopic distributions and trends and to construct meaningful isotopic budgets, the openness of the system (ventilation rates) and the nature of N-exchanges should be known.

The growing set of δ15N measurements makes it increasingly feasible to incorporate isotopic constraints into ecosystem and biogeochemical models of nitrogen in the ocean. A robust isotope budget for the ocean will require extension of these sampling efforts. Time-series measurements from a variety of locations will also greatly enhance the utility of nitrogen isotope measurements, particularly in quantifying the flux of nitrogen through the biota during blooms and other transient events. Such transient events are effectively natural tracer experiments created by isotopic fractionation during phytoplankton consumption of inorganic nitrogen and could provide an ecosystem-level analog to the short-term tracer experiments typically carried out in small volumes aboard ship. Complementing nitrogen stable isotope measurements with focused rate measurements may ultimately provide the best strategy for studying the dynamics of the nitrogen cycle across a variety of temporal and spatial scales.

References

  1. 1. CasciottiK. LInverse kinetic isotope fractionation during bacterial nitrite oxidation. Geochimica et Cosmochimica Acta, 200920612076
  2. 2. NierA. OA redetermination of the relative abundances of the isotopes of carbon, nitrogen, oxygen, argon, and potassium. Physical Review, 1950789793
  3. 3. CoplenT. BH. RKrouseand J. KBöhlkeReporting of nitrogen isotope abundances- (Technical report). Pure and Applied Chemistry, 1992907908
  4. 4. HoefsJStable Isotope Geochemistry 6ed2009Berlin: Springer Verlag.
  5. 5. OwensN. J. Pand A. PReesDetermination of nitrogen-15 at sub-microgram levels of nitrogen using automated continuous flow isotope ratio mass spectrometry. Analyst, 198916551657
  6. 6. RyabenkoEM. AAltabetand D. W. RWallaceEffect of chloride on the chemical conversion of nitrate to nitrous oxide for δ 15N analysis. Limnology and Oceanography-Methods, 2009545552
  7. 7. McilvinM. Rand M. AAltabetChemical conversion of nitrate and nitrite to nitrous oxide for nitrogen and oxygen isotopic analysis in freshwater and seawater. Analytical Chemistry, 200555895595
  8. 8. SigmanD. Met alA bacterial method for the nitrogen isotopic analysis of nitrate in seawater and freshwater. Analytical Chemistry 200141454153
  9. 9. CasciottiK. Land M. RMcilvinIsotopic analyses of nitrate and nitrite from reference mixtures and application to Eastern Tropical North Pacific waters. Marine Chemistry, 2007184201
  10. 10. CasciottiK. Let alMeasurement of the oxygen isotopic composition of nitrate in seawater and freshwater using the denitrifier method. Analytical Chemistry, 200249054912
  11. 11. GrangerJet alA method for nitrite removal in nitrate N and O isotope analyses. Limnology and Oceanogrophy-Methods, 2006July): 205212
  12. 12. GlibertP. Met alIsotope-dilution models of uptake and remineralization of ammonium by marine plankton. Limnology and Oceanography, 1982639650
  13. 13. HarrisonW. Gand L. RHarrisIsotope-dilution and its effects on measurements of nitrogen and phosphorus uptake by oceanic microplankton. Marine Ecology-Progress Series, 1986253261
  14. 14. MohrWet alMethodological underestimation of oceanic nitrogen fixation rates. Plos One, 2010e12583
  15. 15. FrewNThe role of organic films in air-sea gas exchange, in The sea surface and globale change, R.A. Duce and P.S. Liss, Editors. 1997Cambridge University Press: Cambridge, UK. 121172
  16. 16. AsherW. Eand J. FPankowPrediction of gas/water mass transport coefficients by a surface renewal model Environmental Science & Technology, 199112941300
  17. 17. MariottiAAtmospheric nitrogen is a reliable standard for natural 15N abundance measurements. Nature, 1983685687
  18. 18. Montoya, J. P.: Nitrogen stable isotopes in marine environments, in: Nitrogen in Marine Environment, edited by: Capone, D. G., Bronk, D. A., Mulholland, M. R., and Carpenter, E. J., Elsevier Inc, Amsterdam, 1277-1302, 2008.
  19. 19. WilkersonF. Pand GGrunseichFormation of blooms by the symbiotic ciliate Mesodinium rubrum- The significance of nitrogen uptake Journal of Plankton Research, 1990973989
  20. 20. RaimbaultPGSlawykand NGarciaComparison between chemical and isotopic measurements of biological nitrate utilization: further evidence of low new-production levels in the equatorial Pacific. Marine Biology, 200011471155
  21. 21. LibbyS. Pand P. AWheelerParticulate and dissolved organic nitrogen in the central and eastern equatorial Pacific. Deep Sea Research Part I: Oceanographic Research Papers, 1997345361
  22. 22. TaguchiSand E. ALawsOn the microparticles which pass through glass-fiber filter type GF/F in coastal and open waters. Journal of Plankton Research, 19889991008
  23. 23. MaskeHand EGarcia-mendozaAdsorption of dissolved organic matter to the inorganic filter substrate and its implications for 14C uptake measurements. Applied and Environmental Microbiology, 199438873889
  24. 24. RaimbaultPand GSlawyk A Semiautomatic wet-oxidation method for the determination of particlulate organic nitrogen collected on filters. Limnology and Oceanography, 1991405408
  25. 25. OlsonR. JDifferential photoinhibition of marine nitrifying bacteria: A possible mechanism for the formation of the primary nitrite maximum. Journal of Marine Research, 1981227238
  26. 26. YakushijiHand JKandaDetermination of experimentally enriched 15N in nitrate nitrogen based on an improved method of azo dye formation. Journal of Oceanography, 1998337342
  27. 27. BronkD. Aand B. BWardGross and net nitrogen uptake and DON release in the euphotic zone of Monterey Bay, California. Limnology and Oceanography, 1999573585
  28. 28. SigmanD. Met alNatural abundance-level measurement of the nitrogen isotopic composition of oceanic nitrate: an adaptation of the ammonia diffusion method. Marine Chemistry, 1997227242
  29. 29. TanakaTand TSainoModified method for the analysis of nitrogen isotopic composition of oceanic nitrate at low concentration. Journal of Oceanography, 2002539546
  30. 30. Risgaard-petersenNet alApplication of the isotope pairing technique in sediments where anammox and denitrification coexist. Limnology and Oceanography-Methods, 20036373
  31. 31. JensenK. MM. HJensenand R. PCoxMembrane inlet mass spectrometric analysis of N-isotope labelling for aquatic denitrification studies. Fems Microbiology Ecology, 1996101109
  32. 32. FeastN. Aand P. FDennisA comparison of methods for nitrogen isotope analysis of groundwater. Chemical Geology, 1996167171
  33. 33. PrestonTet alIsotope dilution analysis of combined nitrogen in natural waters.1. Ammonium. Rapid Communications in Mass Spectrometry, 1996959964
  34. 34. LehmannM. FS. MBernasconiand J. AMckenzieA method for the extraction of ammonium from freshwaters for nitrogen isotope analysis. Analytical Chemistry, 200147174721
  35. 35. ODeenW. A. a. n. d L. KPorterDevarda’s alloy reduction of nitrate and tube diffusion of the reduced nitrogen for indophenol ammonium and nitrogen-15 determinations. Analytical Chemistry, 198011641166
  36. 36. HolmesR. Met alMeasuring 15N-NH4+ in marine, estuarine and fresh waters: An adaptation of the ammonia diffusion method for samples with low ammonium concentrations. Marine Chemistry, 1998235243
  37. 37. Risgaard-petersenNN. PRevsbechand SRysgaardCombined microdiffusion-hypobromite oxidation method for determining nitrogen-15 isotope in ammonium. Soil Science Society of America Journal 199510771080
  38. 38. SebiloMet alThe use of the ‘Ammonium diffusion’ method for δ15N-NH4+ and δ15N-3-measurementsComparison with other techniques. Environmental Chemistry, 200499103
  39. 39. DudekNM. ABrzezinskiand P. AWheelerRecovery of ammonium nitrogen by solvent extraction for the determination of relative 15N abundance in regeneration experiments. Marine Chemistry, 19865969
  40. 40. ZhangLet alSensitive measurement of NH4+ 15N/14N (δ15NH4+) at natural abundance levels in fresh and saltwaters. Analytical Chemistry, 2007
  41. 41. SteingruberS. Met alMeasurement of denitrification in sediments with the N-15 isotope pairing technique. Applied and Environmental Microbiology, 200137713778
  42. 42. CaponeD. Gand J. PMontoyaNitrogen fixation and denitrification, in Methods in Marine Microbiology, J. Paul, Editor 2001Academic Press: New York. 501515
  43. 43. AnS. MW. SGardnerand TKanaSimultaneous measurement of denitrification and nitrogen fixation using isotope pairing with membrane inlet mass spectrometry analysis. Applied and Environmental Microbiology, 200111711178
  44. 44. KanaT. Met alDenitrification in estuarine sediments determined by membrane inlet mass spectrometry. Limnology and Oceanography, 1998334339
  45. 45. HartnettH. Eand S. PSeitzingerHigh-resolution nitrogen gas profiles in sediment porewaters using a new membrane probe for membrane-inlet mass spectrometry. Marine Chemistry, 20032330
  46. 46. KanaT. Met alMembrane Inlet Mass Spectrometer for rapid high-precision determination of N2, O2, and Ar in environmental water samples. Analytical Chemistry, 199441664170
  47. 47. PunshonSand R. MMooreNitrous oxide production and consumption in a eutrophic coastal embayment. Marine Chemistry, 20043751
  48. 48. CasciottiK. LMMcilvinand CBuchwaldOxygen isotopic exchange and fractionation during bacterial ammonia oxidation Limnology and Oceanography, 2010753762
  49. 49. FrameC. Hand K. LCasciottiBiogeochemical controls and isotopic signatures of nitrous oxide production by a marine ammonia-oxidizing bacterium. Biogeosciences 201030193059
  50. 50. ToyodaSand NYoshidaDetermination of nitrogen isotopomers of nitrous oxide on a modified isotope ratio mass spectrometer. Analytical Chemistry, 199947114718
  51. 51. SutkaR. Let alDistinguishing nitrous oxide production from nitrification and denitrification on the basis of isotopomer abundances. Applied and Environmental Microbiology, 2006638644
  52. 52. OstromNand POstromThe isotopomers of nitrous oxide: Analytical considerations and application to resolution of microbial production pathways, in Handbook of Environmental Isotope Geochemistry, M. Baskaran, Editor 2011Springer Berlin Heidelberg. 453476
  53. 53. WardB. Band D. ABronkNet nitrogen uptake and DON release in surface waters: importance of trophic interactions implied from size fractionation experiments. Marine Ecology-Progress Series, 20011124
  54. 54. BrandesJ. AA. HDevoland CDeutschNew developments in the marine nitrogen cycle. Chemical Reviews, 2007577589
  55. 55. MccarthyJ. Jand E. JCarpenterOscillatoria (Trichodesmium) thiebautii (Cyanophyta) in the central north Atlantic Ocean. Journal of Phycology, 19797582
  56. 56. SeitzingerS. PR. WSandersand RStylesBioavailability of DON from natural and anthropogenic sources to estuarine plankton. Limnology and Oceanography, 2002353366
  57. 57. MontoyaJ. PE. JCarpenterand D. GCaponeNitrogen fixation and nitrogen isotope abundances in zooplankton of the oligotrophic North Atlantic. Limnology and Oceanography, 200216171628
  58. 58. ZehrJand HPaerlBiological nitrogen fixation in the marine environment., in Microbial Ecology of the Oceans., D.L. Kirchman, Editor 2008Wiley-Liss, Inc.: New York. 481525
  59. 59. CarpenterE. JNitrogen fixation by Oscillatoria (Trichodesmium) thiebautii in the southwestern Sargasso Sea. Deep-Sea Research, 1973285288
  60. 60. SainoTand AHattoriDiel variation in nitrogen fixation by a marine blue-green alga, Trichodesmium thiebautii. Deep-Sea Research, 197812591263
  61. 61. CarpenterE. JNitrogen fixation by a blue-green epiphyte on pelagic Sargassum. Science, 197212071209
  62. 62. MagueT. HF. CMagueand OHolmhansenPhysiology and chemical composition of nitrogen-fixing phytoplankton in the central north Pacific ocean. Marine Biology, 1977213227
  63. 63. VillarealT. AWidespread occurrence of the Hemiaulus-cyanobacterial symbiosis in the southwest north Atlantic ocean. Bulletin of Marine Science, 199417
  64. 64. Berman-frankIet alSegregation of nitrogen fixation and oxygenic photosynthesis in the marine cyanobacterium Trichodesmium. Science, 200115341537
  65. 65. WadaEand AHattoriNatural abundance of 15N in particulate organic matter in the North Pacific Ocean. Geochimica et Cosmochimica Acta, 1976249251
  66. 66. RyabenkoEet alContrasting biogeochemistry of nitrogen in the Atlantic and Pacific oxygen minimum zones. Biogeosciences 2012203215
  67. 67. LehmannM. Fet alOrigin of the deep Bering Sea nitrate deficit: Constraints from the nitrogen and oxygen isotopic composition of water column nitrate and benthic nitrate fluxes. Global Biogeochemical Cycles, 2005GB4005
  68. 68. DelwicheC. Cet alNitrogen isotope distribution as a presumptive indicator of nitrogen-fixation. Botanical Gazette, 1979S65S69
  69. 69. ZehrJ. PE. JCarpenterand T. AVillarealNew perspectives on nitrogen-fixing microorganisms in tropical and subtropical oceans. Trends in Microbiology, 20006873
  70. 70. CaponeD. Get alBasis for diel variation in nitrogenase activity in the marine planktonic cyanobacterium Trichodesmium thiebautii. Applied and Environmental Microbiology, 199035323536
  71. 71. HowarthR. WRMarinoand J. JColeNitrogen-fixation in fresh-water, estuarine, and marine ecosystems.2. Biogeochemical controls. Limnology and Oceanography, 1988688701
  72. 72. HowarthR. Wet alNitrogen fixation in freshwater, estuarine, and marine ecoystems. 1. rates and importance. Limnology and Oceanography, 1988469687
  73. 73. KarlDet alDinitrogen fixation in the world’s oceans. Biogeochemistry, 20024798
  74. 74. CarpenterE. JPhysiology and ecology of marine Oscillatoria (Trichodesmium). Marine Biology Letters, 19836985
  75. 75. LehtimakiJet alGrowth, nitrogen fixation, and nodularin production by two Baltic sea cyanobacteria. Applied and Environmental Microbiology, 199716471656
  76. 76. SohmJ. AE. AWebband D. GCaponeEmerging patterns of marine nitrogen fixation. Nature Reviews Microbiology, 2011499508
  77. 77. GallonJ. RN2 fixation in phototrophs: adaptation to a specialized way of life. Plant and Soil, 20013948
  78. 78. Berman-frankIet alIron availability, cellular iron quotas, and nitrogen fixation in Trichodesmium. Limnology and Oceanography, 200112491260
  79. 79. LaRocheJ. and E. Breitbarth, Importance of the diazotrophs as a source of new nitrogen in the ocean. Journal of Sea Research, 20056791
  80. 80. CodispotiL. Aand F. ARichardsAn analysis of the horizontal regime of denitrification in the eastern tropical North Pacific. Limnology and Oceanography, 1976379388
  81. 81. DeutschCet alDenitrification and N2 fixation in the Pacific Ocean. Global Biogeochemical Cycles, 2001483506
  82. 82. GruberNand J. LSarmientoGlobal patterns of marine nitrogen fixation and denitrification. Global Biogeochemical Cycles, 1997235266
  83. 83. MichaelsA. Fet alInputs, losses and transformations of nitrogen and phosphorus in the pelagic North Atlantic Ocean. Biogeochemistry, 1996181226
  84. 84. HansellD. Aand R. AFeelyAtmospheric intertropical convergence impacts surface ocean carbon and nitrogen biogeochemistry in the western tropical Pacific. Geophysical Research Letters, 200010131016
  85. 85. HansellD. Aand T. YWaterhouseControls on the distributions of organic carbon and nitrogen in the eastern Pacific Ocean. Deep-Sea Research Part I-Oceanographic Research Papers, 1997843857
  86. 86. VidalMC. MDuarteand SAgustiDissolved organic nitrogen and phosphorus pools and fluxes in the central Atlantic Ocean. Limnology and Oceanography, 1999106115
  87. 87. AbellJSEmersonand PRenaudDistributions of TOP, TON and TOC in the North Pacific subtropical gyre: Implications for nutrient supply in the surface ocean and remineralization in the upper thermocline. Journal of Marine Research, 2000203222
  88. 88. KarlDet alThe role of nitrogen fixation in biogeochemical cycling in the subtropical North Pacific Ocean. Nature, 1997533538
  89. 89. KarlD. Met alEcosystem changes in the North Pacific subtropical gyre attributed to the 1991-92 El-Nino. Nature, 1995230234
  90. 90. BergquistB. Aand E. ABoyleDissolved iron in the tropical and subtropical Atlantic Ocean. Global Biogeochemical Cycles, 2006
  91. 91. BlainSSBonnetand CGuieuDissolved iron distribution in the tropical and sub tropical South Eastern Pacific. Biogeosciences, 2008269280
  92. 92. BoyleE. Aet alIron, manganese, and lead at Hawaii Ocean Time-series station ALOHA: Temporal variability and an intermediate water hydrothermal plume. Geochimica et Cosmochimica Acta, 2005933952
  93. 93. WuJet alSoluble and colloidal iron in the oligotrophic North Atlantic and North Pacific. Science, 2001847849
  94. 94. MahaffeyCet alBiogeochemical signatures of nitrogen fixation in the eastern North Atlantic. Geophysical Research Letters, 20031300
  95. 95. JohnsonK. Set alSurface ocean-lower atmosphere interactions in the Northeast Pacific Ocean Gyre: Aerosols, iron, and the ecosystem response. Global Biogeochemical Cycles, 20031063
  96. 96. LenesJ. Met alIron fertilization and the Trichodesmium response on the West Florida shelf. Limnology and Oceanography, 200112611277
  97. 97. MillsM. Met alIron and phosphorus co-limit nitrogen fixation in the eastern tropical North Atlantic. Nature, 2004292294
  98. 98. BakerA. Ret alAtmospheric deposition of nutrients to the Atlantic Ocean. Geophysical Research Letters, 200322962300
  99. 99. RidameCand CGuieuSaharan input of phosphate to the oligotrophic water of the open western Mediterranean Sea. Limnology and Oceanography, 2002856869
  100. 100. RamosJ. LFMaduenoand M. GGuerreroRegulation of nitrogenase levels in Anabaena sp. ATCC 33047 and other filamentous cyanobacteria Archives of Microbiology, 1985105111
  101. 101. Martín-nietoJAHerreroand EFloresControl of nitrogenase mRNA levels by products of nitrate assimilation in the cyanobacterium Anabaena sp. Strain PCC 7120. Plant Physiology, 1991825828
  102. 102. SanzalferezSand F. FDelcampoRelationship between nitrogen fixation and nitrate metabolism in the Nodularia strains M1 and M2. Planta, 1994339345
  103. 103. MulhollandM. RKOhkiand D. GCaponeNutrient controls on nitrogen uptake and metabolism by natural populations and cultures of Trichodesmium (Cyanobacteria). Journal of Phycology, 200110011009
  104. 104. HollC. Mand J. PMontoyaInteractions between nitrate uptake and nitrogen fixation in continuous cultures of the marine diazotroph Trichodesmium (Cyanobacteria). Journal of Phycology, 200511781183
  105. 105. RaimbaultPand NGarciaEvidence for efficient regenerated production and dinitrogen fixation in nitrogen-deficient waters of the South Pacific Ocean: impact on new and export production estimates. Biogeosciences, 2008323338
  106. 106. FernandezCLFariasand OUlloaNitrogen fixation in denitrified marine waters. Plos One, 2011e20539
  107. 107. CarpenterE. Jand D. GCaponeNitrogen fixation in the marine environment, in Nitrogen in the marine environment, D.G. Capone, et al., Editors. 2008Elsevier. 141198
  108. 108. BerticsV. Jand WZiebisBioturbation and the role of microniches for sulfate reduction in coastal marine sediments. Environmental Microbiology, 201030223034
  109. 109. GilbertFet alHydrocarbon influence on denitrification in bioturbated Mediterranean coastal sediments. Hydrobiologia, 19976777
  110. 110. HerbertR. ANitrogen cycling in coastal marine ecosystems. FEMS Microbiology Reviews, 1999563590
  111. 111. SchmidtIR. J. MVan Spanningand M. S. MJettenDenitrification and ammonia oxidation by Nitrosomonas europaea wild-type, and NirK- and NorB-deficient mutants. Microbiology-Sgm, 200441074114
  112. 112. ZartDand EBockHigh rate of aerobic nitrification and denitrification by Nitrosomonas eutropha grown in a fermentor with complete biomass retention in the presence of gaseous 2or NO. Archives of Microbiology, 1998282286
  113. 113. HooperA. Band K. RTerryHydroxylamine oxidoreductase of Nitrosomonas: Production of nitric oxide from hydroxylamine. Biochimica et Biophysica Acta (BBA)- Enzymology, 19791220
  114. 114. HooperA. Bet alEnzymology of the oxidation of ammonia to nitrite by bacteria. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology, 19975967
  115. 115. PothMand D. DFochtN kinetic analysis of N2O production by Nitrosomonas europaea: an examination of nitrifier denitrification. Applied and Environmental Microbiology, 198511341141
  116. 116. RemdeAand RConradProduction of nitric oxide in Nitrosomonas europaea by reduction of nitrite. Archives of Microbiology, 1990187191
  117. 117. MariottiAet alExperimental determination of nitrogen kinetic isotope fractionation- Some principles- Illustration for the denitrification and nitrification processes. Plant and Soil, 1981413430
  118. 118. CasciottiK. LD. MSigmanand B. BWardLinking diversity and stable isotope fractionation in ammonia-oxidizing bacteria. Geomicrobiology Journal, 2003335353
  119. 119. SutkaR. Let alStable nitrogen isotope dynamics of dissolved nitrate in a transect from the North Pacific Subtropical Gyre to the Eastern Tropical North Pacific. Geochimica et Cosmochimica Acta, 2004517527
  120. 120. CasciottiK. Land CBuchwaldInsights on the marine microbial nitrogen cycle from isotopic approaches to nitrification. Frontiers in Microbiology, 2012
  121. 121. SchmidtH. Let alIs the isotopic composition of nitrous oxide an indicator for its origin from nitrification or denitrification? A theoretical approach from referred data and microbiological and enzyme kinetic aspects. Rapid Communications in Mass Spectrometry, 200420362040
  122. 122. ToyodaSet alFractionation of N2O isotopomers during production by denitrifier. Soil Biology & Biochemistry, 200515351545
  123. 123. SutkaR. Let alNitrogen isotopomer site preference of N2O produced by Nitrosomonas europaea and Methylococcus capsulatus Bath. Rapid Communications in Mass Spectrometry, 2003738745
  124. 124. NevisonC. Det alInterannual and seasonal variability in atmospheric N2O. Global Biogeochemical Cycles, 2007GB3017
  125. 125. CohenYand L. IGordonNitrous oxide in the oxygen minimum of the eastern tropical North Pacific: evidence for its consumption during denitrification and possible mechanisms for its production. Deep Sea Research, 1978509524
  126. 126. YoshinariTNitrous oxide in the sea. Marine Chemistry, 1976189202
  127. 127. NevisonCJ. HButlerand J. WElkinsGlobal distribution of N2O and the delta N2O-AOU yield in the subsurface ocean. Global Biogeochemical Cycles, 20031119
  128. 128. LipschultzFet alBacterial transformations of inorganic nitrogen in the oxygen-deficient waters of the Eastern Tropical South Pacific Ocean. Deep Sea Research Part A. Oceanographic Research Papers, 199015131541
  129. 129. WardB. BNitrogen transformations in the Southern California Bight. Deep-Sea Research Part A-Oceanographic Research Papers, 1987785805
  130. 130. WardB. BM. CTalbotand M. JPerryContributions of phytoplankton and nitrifying bacteria to ammonium and nitrite dynamics in coastal waters. Continental Shelf Research, 1984383398
  131. 131. BlackburnT. Hand N. DBlackburnCoupling of cycles and global significance of sediment diagenesis Marine Geology, 1993101110
  132. 132. GrundmanisVand J. WMurrayNitrification and denitrification in marine sediments from Puget Sound. Limnology and Oceanography, 1977804813
  133. 133. DevolA. Hand J. PChristensenBenthic fluxes and nitrogen cycling in sediments of the continental margin of the eastern North Pacific Journal of Marine Research, 1993345372
  134. 134. LaursenA. Eand S. PSeitzingerThe role of denitrification in nitrogen removal and carbon mineralization in Mid-Atlantic Bight sediments. Continental Shelf Research, 200213971416
  135. 135. LehmannM. FD. MSigmanand W. MBerelsonCoupling the 15N/14N and 18O/16O of nitrate as a constraint on benthic nitrogen cycling. Marine Chemistry, 2004120
  136. 136. ClarkD. RA. PReesand IJointA method for the determination of nitrification rates in oligotrophic marine seawater by gas chromatography/mass spectrometry. Marine Chemistry, 20078496
  137. 137. MulhollandM. Rand M. WLomasNitogen uptake and assimilation, in Nitrogen in the Marine Environment, D.G. Capone, et al., Editors. 2008Elsevier Inc.: Amsterdam et al. 303384
  138. 138. MooreJ. Ket alAn intermediate complexity marine ecosystem model for the global domain. Deep-Sea Research Part II-Topical Studies in Oceanography, 2002403462
  139. 139. Garcia-fernandezJ. MN. TDe Marsacand JDiezStreamlined regulation and gene loss as adaptive mechanisms in Prochlorococcus for optimized nitrogen utilization in oligotrophic environments. Microbiology and Molecular Biology Reviews, 2004630638
  140. 140. Garcia-fernandezJ. Mand JDiezAdaptive mechanisms of nitrogen and carbon assimilatory pathways in the marine cyanobacteria Prochlorococcus. Research in Microbiology, 2004795802
  141. 141. WadaEand AHattoriNitrogen isotope effect in the assimilation of inorganic nitrogenous compounds by marine diatoms. Geomicrobiology Journal, 197885101
  142. 142. MontoyaJ. Pand J. JMccarthyIsotopic fractionation during nitrate uptake by phytoplankton grown in continuous culture. Journal of Plankton Research, 1995439464
  143. 143. GrangerJet alCoupled nitrogen and oxygen isotope fractionation of nitrate during assimilation by cultures of marine phytoplankton. Limnology and Oceanography, 200417631773
  144. 144. HochM. PM. LFogeland D. LKirchmanIsotope fractionation associated with ammonium uptake by a marine bacterium. Limnology and Oceanography, 199214471459
  145. 145. WaserN. Aet alNitrogen isotope fractionation during nitrate, ammonium and urea uptake by marine diatoms and coccolithophores under various conditions of N availability. Marine Ecology-Progress Series, 19982941
  146. 146. AntiaN. JP. JHarrisonand LOliveiraThe role of dissolved organic nitrogen in phytoplankton nutrition, cell biology and ecology. Phycologia, 1991189
  147. 147. EppleyR. Wand W. HThomasComparison of half-saturation constants for growth and nitrate uptake of marine phytoplankron Journal of Phycology, 1969375379
  148. 148. StrousMJ. GKuenenand M. S. MJettenKey physiology of anaerobic ammonium oxidation. Applied and Environmental Microbiology, 199932483250
  149. 149. LamPet alRevising the nitrogen cycle in the Peruvian oxygen minimum zone. Proceedings of the National Academy of Sciences of the United States of America, 200947524757
  150. 150. CodispotiL. Aet alThe oceanic fixed nitrogen and nitrous oxide budgets: Moving targets as we enter the anthropocene? Scientia Marina, 200185105
  151. 151. ZumftW. Gand HKornerEnzyme diversity and mosaic gene organization in denitrification. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology, 19974358
  152. 152. CodispotiL. ATYoshinariand A. HDevolSuboxic respiration in the oceanic water column., in Respiration in Aquatic Ecosystems, P.A. del Giorgio and P.J.L.B. Williams, Editors. 2005Blackwell Scientific: Oxford. 225247
  153. 153. ZumftW. GCell biology and molecular basis of denitrification. Microbiology and Molecular Biology Reviews, 1997533616
  154. 154. FergusonS. JDenitrification and its control. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology, 199489110
  155. 155. GüvenDet alPropionate oxidation by and methanol inhibition of anaerobic ammonium oxidizing bacteria Applied and Environmental Microbiology, 200510661071
  156. 156. KalvelageTet alOxygen sensitivity of anammox and coupled N-cycle processes in oxygen minimum zones. Plos One, 2011e29299
  157. 157. JettenM. S. Met alAnaerobic ammonium oxidation by marine and freshwater planctomycete-like bacteria. Applied Microbiology and Biotechnology, 2003107114
  158. 158. Van NiftrikLet alThe anammoxasome: an intracytoplasmic compartment in anammox bacteria. FEMS Microbiology Letters, 2004713
  159. 159. StrousMet alDeciphering the evolution and metabolism of an anammox bacterium from a community genome. Nature, 2006790794
  160. 160. KartalBet alAnammox bacteria disguised as denitrifiers: nitrate reduction to dinitrogen gas via nitrite and ammonium. Environmental Microbiology, 2007635642
  161. 161. KuypersM. M. Met alAnaerobic ammonium oxidation by anammox bacteria in the Black Sea. Nature, 2003608611
  162. 162. DaleO. RC. RTobiasand BSongBiogeographical distribution of diverse anaerobic ammonium oxidizing (anammox) bacteria in Cape Fear River estuary. Environmental Microbiology, 200911941207
  163. 163. ThamdrupBand TDalsgaardProduction of N2 through anaerobic ammonium oxidation coupled to nitrate reduction in marine sediments. Applied and Environmental Microbiology, 200213121318
  164. 164. DalsgaardTet alN2 production by the anammox reaction in the anoxic water column of Golfo Dulce, Costa Rica. Nature, 2003606608
  165. 165. PodlaskaAet alMicrobial community structure and productivity in the oxygen minimum zone of the eastern tropical North Pacific. Deep Sea Research Part I: Oceanographic Research Papers, 20127789
  166. 166. HuBl., et alNew anaerobic, ammonium-oxidizing community enriched from peat soil. Applied and Environmental Microbiology, 2011966971
  167. 167. HumbertSet alMolecular detection of anammox bacteria in terrestrial ecosystems: distribution and diversity. ISME J, 2009450454
  168. 168. ZhuGet alAnaerobic ammonia oxidation in a fertilized paddy soil. ISME J, 201119051912
  169. 169. ErlerD. VB. DEyreand LDavisonThe contribution of anammox and denitrification to sediment N2 production in a surface flow constructed Wetland. Environmental Science & Technology, 200891449150
  170. 170. SchubertC. Jet alAnaerobic ammonium oxidation in a tropical freshwater system (Lake Tanganyika). Environmental Microbiology, 200618571863
  171. 171. HannigMet alShift from denitrification to anammox after inflow events in the central Baltic Sea. Limnology and Oceanography, 200713361345
  172. 172. StrousMet alThe sequencing batch reactor as a powerful tool for the study of slowly growing anaerobic ammonium-oxidizing microorganisms. Applied Microbiology and Biotechnology, 1998589596
  173. 173. van de GraafA.A., et al., Autotrophic growth of anaerobic ammonium-oxidizing micro-organisms in a fluidized bed reactor. Microbiology, 199621872196
  174. 174. BockEet alNitrogen loss caused by denitrifyling Nitrosomonas cells using ammonium or hydrogen as electron-donors and nitrite as electron-acceptor. Archives of Microbiology, 19951620
  175. 175. BlackmerA. MJ. MBremnerand E. LSchmidtProduction of nitrous oxide by ammonia oxidizing chemoautotrophic microorganisms in soil. Applied and Environtal Microbiology, 198010601066
  176. 176. GoreauT. Jet alProduction of 2-andN2O by nitrifying bacteria at reduced concentrations of oxygen. Applied and Environmental Microbiology, 1980526532
  177. 177. BohlenLet alBenthic nitrogen cycling traversing the Peruvian oxygen minimum zone. Geochimica et Cosmochimica Acta, 201160946111
  178. 178. JensenM. Met alIntensive nitrogen loss over the Omani Shelf due to anammox coupled with dissimilatory nitrite reduction to ammonium. Isme Journal, 201116601670
  179. 179. RysgaardSNRisgaardPetersen, and N.P. Sloth, Nitrification, denitrification, and nitrate ammonification in sediments of two coastal lagoons in Southern France. Hydrobiologia, 1996p. 133141
  180. 180. FaríasLPotential role of bacterial mats in the nitrogen budget of marine sediments: the case of Thioploca spp. Marine ecology progress series, 1998291292
  181. 181. FossingHet alConcentration and transport of nitrate by the mat-forming sulphur bacterium Thioploca. Nature, 1995713715
  182. 182. SchulzH. Net alDense populations of a giant sulfur bacterium in Namibian shelf sediments. Science, 1999493495
  183. 183. ClineJ. Dand F. ARichardsOxygen deficient conditions and nitrate reduction in Eastern Tropical North-Pacific Ocean. Limnology and Oceanography, 1972885900
  184. 184. MorrisonJ. Met alThe oxygen minimum zone in the Arabian Sea during 1995. Deep-Sea Research Part II-Topical Studies in Oceanography, 199919031931
  185. 185. KoeveWand PKählerHeterotrophic denitrification vs. autotrophic anammox- quantifying collateral effects on the oceanic carbon cycle. Biogeosciences, 201023272337
  186. 186. VossMand J. PMontoyaNitrogen cycle. Oceans apart. Nature, 20094950
  187. 187. BulowS. Eet alDenitrification exceeds anammox as a nitrogen loss pathway in the Arabian Sea oxygen minimum zone. Deep-Sea Research Part I-Oceanographic Research Papers, 2010384393
  188. 188. GalanAet alAnammox bacteria and the anaerobic oxidation of ammonium in the oxygen minimum zone off northern Chile. Deep-Sea Research Part II-Topical Studies in Oceanography, 200911251135
  189. 189. DevolA. Het alDenitrification rates and excess nitrogen gas concentrations in the Arabian Sea oxygen deficient zone. Deep-Sea Research Part I-Oceanographic Research Papers, 200615331547
  190. 190. Van MooyB. A. SR. GKeiland A. HDevolImpact of suboxia on sinking particulate organic carbon: Enhanced carbon flux and preferential degradation of amino acids via denitrification. Geochimica et Cosmochimica Acta, 2002457465
  191. 191. BrandesJ. Aet alIsotopic composition of nitrate in the central Arabian Sea and eastern tropical North Pacific: A tracer for mixing and nitrogen cycles. Limnology and Oceanography, 199816801689
  192. 192. SigmanD. Met alThe δ15N of nitrate in the Southern Ocean: Consumption of nitrate in surface waters. Global Biogeochemical Cycles, 199911491166
  193. 193. VossMJ. WDippnerand J. PMontoyaNitrogen isotope patterns in the oxygen-deficient waters of the Eastern Tropical North Pacific Ocean. Deep-Sea Research Part I-Oceanographic Research Papers, 200119051921
  194. 194. ClineJ. Dand I. RKaplanIsotopic fractionation of dissolved nitrate during denitrification in the eastern tropical north Pacific Ocean. Marine Chemistry, 1975271299
  195. 195. LehmannM. Fet alThe distribution of nitrate 15N/14N in marine sediments and the impact of benthic nitrogen loss on the isotopic composition of oceanic nitrate. Geochimica et Cosmochimica Acta, 200753845404
  196. 196. CodispotiL. AAn oceanic fixed nitrogen sink exceeding 400 Tg Na-1 vs the concept of homeostasis in the fixed-nitrogen inventory. Biogeosciences, 2007233253
  197. 197. CodispotiL. Aand J. PChristensenNitrification, denitrification and nitrous-oxide cycling in the eastern tropical South-Pacific Ocean. Marine Chemistry, 1985277300
  198. 198. BrandesJ. Aand A. HDevolA global marine-fixed nitrogen isotopic budget: Implications for Holocene nitrogen cycling. Global Biogeochemical Cycles, 200211201134
  199. 199. GallowayJ. Net alNitrogen Cycles: Past, Present, and Future. Biogeochemistry, 2004153226
  200. 200. GruberNThe marine nitrogen cycle: Overview and challanges, in Nitrogen in Marine Environment, D.G. Capone, et al., Editors. 2008Elsevier Inc. 151
  201. 201. CodispotiL. AInteresting times for marine N2O. Science, 201013391340
  202. 202. GalbraithE. Det alGlobal nitrogen isotopic constraints on the acceleration of oceanic denitrification during deglacial warming. Nature Geosciences, 2012submitted.
  203. 203. AltabetM. AConstraints on oceanic N balance/imbalance from sedimentary 15N records. Biogeosciences, 20077586
  204. 204. SomesCNitrogen Isotopes in the Global Ocean: Constraints on the pre-industrial fixed-N budget, in Biogeochemical modelling 2012Christian-Albrechts-Universität zu Kiel: Kiel.

Written By

Evgenia Ryabenko

Submitted: 18 June 2012 Published: 10 July 2013