banner



How Does Animal Waste Affect The Nitrogen Cycle

  • Journal List
  • Microb Biotechnol
  • v.4(6); 2011 Nov
  • PMC3815407

Microb Biotechnol. 2011 Nov; four(half-dozen): 700–709.

Microbiology of nitrogen bicycle in animate being manure compost

Koki Maeda

1Hokkaido Inquiry Subteam for Waste matter Recycling System, National Agricultural Enquiry Center for Hokkaido Region, National Agricultural and Nutrient Research Arrangement, ane Hitsujigaoka, Sapporo 062‐8555, Japan

iiiEnvironmental Science and Technology, Tokyo Institute of Technology, 4259 Nagatsuta, Midori‐ku, Yokohama 226‐8502, Nihon

Dai Hanajima

1Hokkaido Inquiry Subteam for Waste material Recycling System, National Agronomical Inquiry Heart for Hokkaido Region, National Agricultural and Nutrient Research Arrangement, 1 Hitsujigaoka, Sapporo 062‐8555, Japan

Sakae Toyoda

2Departments of Environmental Chemistry and Engineering

Naohiro Yoshida

3Environmental Science and Technology, Tokyo Institute of Technology, 4259 Nagatsuta, Midori‐ku, Yokohama 226‐8502, Nippon

Riki Morioka

oneHokkaido Research Subteam for Waste matter Recycling System, National Agricultural Research Middle for Hokkaido Region, National Agricultural and Nutrient Research Organization, 1 Hitsujigaoka, Sapporo 062‐8555, Japan

Takashi Osada

4Livestock Research Team on Global Warming, National Institute of Livestock and Grassland Science, National Agricultural and Food Research Organization, 2 Ikenodai, Tsukuba 305‐0901, Japan

Received 2010 Jul 19; Accepted 2010 October 17.

Summary

Composting is the major technology in the treatment of animal manure and is a source of nitrous oxide, a greenhouse gas. Although the microbiological processes of both nitrification and denitrification are involved in composting, the fundamental players in these pathways have not been well identified. Recent molecular microbiological methodologies accept revealed the presence of dominant Bacillus species in the degradation of organic material or betaproteobacterial ammonia‐oxidizing bacteria on nitrification on the surface, and take too revealed the mechanism of nitrous oxide emission in this complicated process to some extent. Some bacteria, archaea or fungi still would be considered potential key players, and the contribution of some pathways, such as nitrifier denitrification or heterotrophic nitrification, might exist involved in composting. This review article discusses these potential microbial players in nitrification–denitrification within the composting pile and highlights the relevant unknowns through recent activities that focus on the nitrogen bike within the animal manure composting process.

Introduction

Composting is the simplest traditional fauna manure management technology that depends on the degradation of organic matter by the microbial community within manure itself (Bernal et al., 2009). Easily degradable organic matter would exist utilized as the free energy source, and CO2, NH3 and wet would be emitted and would generate big amounts of heat; the temperature inside compost piles is nigh 70°C. The mass of the pile decreases significantly, and the process also reduces odorous compounds and pathogens, while killing weed seeds. Because the mature product can be reused as organic fertilizer, composting is a very of import technology from the viewpoint of the apportionment of resources or environmental protection.

Considering the organic or inorganic state of nitrogen contained inside the compost is an important food for crops, the available amount of nitrogen content in composted cloth is a precious component. Through the composting procedure, the organic nitrogen contained within initial fresh manure is degraded into ammonium by a wide diverseness of microorganisms including bacteria and fungi. Part of this nitrogen is lost every bit NH3 by volatilization, or through conversion into gases content such as N2O or N2 through the nitrification/denitrification procedure (Fig. i). The range of nitrogen loss can vary betwixt nineteen% and 77%, which mainly occur through NHiii volatilization and Northward2 emission (Martins and Dewes, 1992; Mahimairaja et al., 1995; Eghball et al., 1997; Tiquia and Tam, 2000; Tiquia et al., 2002). In addition, 0.2–9.ix% of initial nitrogen content can exist emitted as NorthiiO, the intermediate of denitrification or by‐products of nitrification (Kuroda et al., 1996; Hao et al., 2001; 2004; Fukumoto et al., 2003; El Kader et al., 2007; Szanto et al., 2007) (Table 1). The loss of nitrogen during brute manure composting processes is afflicted by various parameters, such as the creature species, nutrition, bulking agents, wet content, turning frequency, carbon/nitrogen ratio and initial nitrogen content.

An external file that holds a picture, illustration, etc.  Object name is mbt0004-0700-f1.jpg

Nitrogen conversion and emission during the composting process.

Tabular array 1

Nitrous oxide emission from manure composting process.

Animal Process type Unit Reference
Dairy 0.582 g‐N2O per kg DM Pattey et al. (2005)
Beefiness 0.162 thou‐North2O per kg DM
Hog Forced aeration one.9–71.ix g‐N2O‐Due north per kiii Osada and Fukumoto (2001)
Cattle static i.1 kg‐N per Mg manure Hao et al. (2001)
Cattle turned 1.nine kg‐N per Mg manure
Cattle Woodchip 0.39 %N Hao et al. (2004)
Cattle Straw 0.68 %Due north
Dairy static ane.2 %Due north El Kader et al. (2007)
Dairy turned i.nine %North
Turkey 0.2–0.4 %N
Squealer turned iii.vii–four.6 %North Fukumoto et al. (2003)
Pig turned 2.five %North Szanto et al. (2007)
Pig static ix.nine %Due north

Nitrous oxide is an important greenhouse gas with strong global warming potential (300 times as that of CO2 (IPCC, 2001). Moreover, because Due north2O is profoundly responsible for ozone depletion, reduction of its emission would be important for environmental protection (Ravishankara et al., 2009). Therefore, an important event in the study of composting is the nitrifier/denitrifier microbial community, which plays a meaning role in nitrogen conversion within the composting pile. In this review article, nosotros bargain with recent inquiry activities that focus on the nitrifier/denitrifier microbial community in composting while referring to like studies in other environments.

Overall microbial and fungal community in the composting process

In that location are many studies virtually microbial community structures in the composting process. Most of them focus on the leaner mainly responsible for the degradation of organic matter. In order to identify microbes present in the compost process, likewise the classical isolation technique, new approaches based on culture‐independent techniques, such as the extraction of Deoxyribonucleic acid from the compost and amplification of 16S rRNA factor by PCR, followed by DNA sequencing are commonly used (Muyzer et al., 1993). The approach based on Dna sequencing provides relevant information on microbes that are hard to culture.

It has been reported that some Bacillus species are important in the composting pile in the thermophilic phase, when agile degradation of organic compounds occurs (Blanc et al., 1997; 1999; Ishii et al., 2000; Peters et al., 2000; Dees and Ghiorse, 2001; Zhang et al., 2002; Ishii and Takii, 2003; Schloss et al., 2003; Iida et al., 2005; Kim et al., 2006; Wang et al., 2007; Yamamoto et al., 2009). These Bacillus can grow and degrade organic compounds nether thermophilic weather condition upwardly to 65°C, and Thermus species are ascendant instead of Bacillus species above 70°C (Beffa et al., 1996). While this is relevant, it should exist noted that the bacterial community structure changes dramatically fifty-fifty in the maturing stage, when active deposition of organic compounds has almost ended. In the maturing phase, mesophilic Proteobacteria or Actinobacteria are known to be dominant (Danon et al., 2008), and these bacterial groups are considered responsible for the maturation process.

In the composting procedure, the temperature in the core zone of the pile reaches 60–70°C, and there are temperature slope effects inside the pile (Fernandes et al., 1994). In addition, there is an oxygen gradient and anoxic conditions deep inside the piles (Hao et al., 2001), particularly in passively aerated composting systems. In these various complicated environments, bacterial communities differ significantly betwixt the surface and the core zone (Maeda et al., 2010a). In the dairy cattle manure composting procedure, nitrite and nitrate accumulate on the surface layer of the pile even in the initial stage of the procedure, when there are still hands degradable organic compounds (Maeda et al., 2010b). In the surface layer, 16S rRNA‐dependent bacterial community assay suggests that some Proteobacteria or Bacteroidetes are ascendant, which is significantly different from the case in thermophilic cadre zones. Some part of these bacterial species are thought to contribute to the nitrification or denitrification that actively occurs in the surface layer.

Although Dees and Ghiorse (2001) reported that they failed to detect archaea in the compost piles, while they institute many fungal species in the compost samples whose temperatures did not exceed 50°C. In this regard, Anastasi and colleagues (2005) reported the isolation of 194 fungal species, the Acremonium, Aspergillus, Cladosporium, Malbranchea, Penicillium, Pseudallescheria and Thermomyces species from compost. In another study, Hultman and colleagues (2009) reported that fungal biomass tin represent between 6.three% and 38.v% of total biomass in municipal waste material compost based on phospholipid fatty acid assay. They as well found that the fungal community suffers dramatic changes during the composting process, as does the bacterial community, and that a fungal community succession differed between a total‐calibration composting facility and a laboratory‐scale small-scale reactor. Studies are needed on the function of the fungal community in the degradation of organic matter in the huminification process, or potential interaction with the bacterial customs and its contribution to the nitrification/denitrification pathway.

Microorganisms relevant to the nitrogen cycle in composting

Nitrifiers

Nitrification is known to exist carried out by bacteria, archaea and fungi (De Boer and Kowalchuk, 2001; Leininger et al., 2006; Laughlin et al., 2008). In the bacterial process, nitrification consists of ii steps, ammonia oxidation and nitrite oxidation, and each of these reactions is performed past an individual microbial group: ammonia‐oxidizing leaner (AOB) (Kowalchuk and Stephen, 2001) and nitrite‐oxidizing bacteria (NOB) respectively. Nitrous oxide is known to exist produced as a by‐product of hydroxylamine oxidation. betaproteobacterial AOB or Thaumarchaeota ammonia‐oxidizing archaea (AOA) are considered of import in ammonia oxidation (Brochier‐Armanet et al., 2008), and the major NOB in the environment are alphaproteobacterial Nitrobacter or Nitrospira. Nitrifiers grow slowly under laboratory weather condition, and their cultivation or isolation is very time‐consuming. In club to speed the tracking of these microbes, a molecular biology approach using primers specific to the 16S rRNA genes and the ammonia monooxygenase gene of betaproteobacterial AOB have been developed and used to runway nitrifiers in the surround (Innerebner et al., 2006; Junier et al., 2010).

In the composting procedure, temporal nitrite accumulation in the middle stage and high accumulation of nitrate in the mature phase were observed (He et al., 2000; 2001; Fukumoto et al., 2006; Fukumoto and Inubushi, 2009; Maeda et al., 2010c), and information technology is evident that nitrification occurs in the compost pile. Nevertheless, it remains unclear which microbes are responsible for this process. Some studies have detected sequences similar to known AOB species Nitrosomonas europaea‐eutropha or Nitrosospira in the composting process (Kowalchuk et al., 1999; Jarvis et al., 2009; Maeda et al., 2010b). Jarvis and colleagues (2009) also detected Nitrosomonas in the theromophilic stage and Nitrosospira in the maturation stage of household waste composting, while Maeda and colleagues (2010b) detected Nitrosomonas throughout the process, especially from the surface layer of a cattle manure composting pile. Some previous studies also detected Nitrosomonas from landfill embrace, an organic‐rich environment similar to the composting process (Mertoglu et al., 2006; Zhu et al., 2007). Betaproteobacterial AOB are chemoautotrophic and generate energy from the hydroxylamine oxidation stride, the ATP produced is used to fix COii as a carbon source. Therefore, the presence of these AOB indicates that these bacteria oxidize ammonia in the composting process. However, the extent of the contribution to net nitrification is still unknown.

To clarify the contribution of AOB to ammonia oxidation in composting, the contribution of AOA must exist studied. Although AOA is known to exist responsible to some extent for nitrification in environments containing less organic matter, such equally soil, ocean or river sediment (Leininger et al., 2006; Santoro et al., 2010), in that location are some reports that AOB'due south contribution is much more important than AOA's for actual nitrification in organic‐rich environments such as wastewater from handling plants (Park et al., 2006; Wells et al., 2009). Another report shows that ammonia oxidation in zinc‐contaminated soil is restored non past AOA but by AOB (Mertens et al., 2009). Many heavy metals are included in livestock manure, especially in swine manure (Nicholson et al., 1999; Ko et al., 2008). AOB might be more than important than AOA under these weather condition. Although a previous report failed to notice the AOA in compost piles (Maeda et al., 2010b), Yamamoto et al. (2010) reported the existence of AOA from cattle manure compost in the later stage of the process. The relative contribution of AOB and AOA on actual nitrification needs to be clarified.

Heterotrophic nitrification is a reaction in which heterotrophic leaner oxidize ammonia or degrade organic thing to nitrate directly (Papen and Von Berg, 1998). Many bacterial species are known to undergo this reaction, and species such equally Paracoccus denitrificans or Pseudomonas putida are known to possess amoA sequences distinct from those of autotrophic nitrifiers (Moir et al., 1996; Daum et al., 1998). Although these bacteria have potential to contribute net nitrification in the compost, its actual contribution is not known at all. This reaction past heterotrophs has non been considered in depth considering these species do not generate free energy from this procedure, nor exercise they accrue high concentrations of nitrite; yet, this futile reaction may exist of relevance in an environmental setting. These heterotrophic nitrifiers assimilate more than ammonium than chemoautotrophic AOB, which leads to higher biomass, and they have been considered not useful for wastewater treatment systems (Podmirseg et al., 2010). Efforts to unveil nitrification process include the development of new culture media for thermophilic nitrifiers in compost under heterotrophic conditions is ongoing (Shimaya and Hashimoto, 2008). In summary, it can be said that to understand the nitrogen bicycle in the composting procedure, we need to acquire more nearly the function of the heterotrophic nitrifiers in the process.

On the other hand, there have non been many molecular ecological studies on NOB. Because NOB has a diverse taxonomy, including Nitrobacter (α‐Proteobacteria), Nitrosococcus mobilis (γ‐Proteobacteria), Nitrospina gracilis (δ‐Proteobacteria) and Nitrospira (Nitrospira), it is difficult to detect all these strains past methods such as FISH (fluorescent in situ hybridization) or PCR, which depend on 16S rRNA sequences. Few studies have focused on functioning gene sequences of nitrite oxidoreductase of Nitrobacter (Poly et al., 2008; Wertz et al., 2008). Even though these methods may be effective for understanding nitrification in the composting process, they have not been used for studies of nitrite oxidation yet, and NOB in the composting process has not been characterized well.

Denitrifiers

Nitrite or nitrate generated by nitrifiers would ordinarily be reduced by heterotrophic denitrifiers and emitted into the atmosphere equally N2O or Due northtwo. Denitrification by bacteria has been well studied and the details of its molecular mechanisms have been characterized. The reaction consists of 4 reduction steps, namely, NOthree→ NO2→ NO → N2O → Due north2. The genes nar, nir, nor and nos are coding the catalysing enzymes (Rudolf and Kroneck, 2005; Tavares et al., 2006). Denitrifying leaner are known to be phylogenetically diverse, with at least l genera (Zumft, 1997). Therefore, the study of denitrifiers that depend on 16S rRNA factor sequences is very difficult, and functioning genes that code each enzyme catalysing each denitrification step are oft used for studies of environmental denitrifiers (Sharma et al., 2005; Wertz et al., 2009). Considering of the relative abundance of information in public databases, nitrite reductase (nirS and nirK) or nitrous oxide reductase (nosZ) have been used oftentimes.

In that location are two types of nitrite reductase: nirS, cytochrome c nitrite reductase, which has haem iron in its active centre (Einsle et al., 1999; 2000), and nirK, a copper‐containing nitrite reductase (Tater et al., 1997; Antonyuk et al., 2005). It is possible to distinguish these nitrite reductases in denitrifiers by using diethyldithiocarbamate (DDTC), which chelates copper of the nirK denitrifier and prevents the process. DGGE primers targeting nitrite reductase gene sequences have been developed (Throback et al., 2004) and used to study denitrifiers in various environments. It is frequently discussed which types of nitrite reducers would exist dominant in the environment. For case, ane report shows that nirS denitrifiers are dominant in subtropical macrotidal estuaries (Abell et al., 2009). Considering the horizonal transfer of denitrifying genes may occur inside the environment, the incidence of nirK or nirS does not always agrees with the 16S rRNA gene phylogenetic sequences. Heylen and colleagues (2006a) concluded that nir genes may non exist suitable to evaluate microbial diversity of denitrifiers in the surround. Thus, interpretation of biodiversity based on nir sequence analysis need to be interpreted with care.

Nitric oxide reductase (NOR) catalyses the reduction of NO to N2O. Nitric oxide, produced by the reduction of nitrite, is known to be toxic to microorganisms, and they demand to metabolize it to protect themselves. Three kinds of NOR take been reported, namely cNOR, qNOR and qCuNOR (Tavares et al., 2006). The cytochrome c‐dependent nitric oxide reductase (cNOR) of P. denitrificans has been studied well. It is a component of cytochrome bc complex with two non‐haem irons in its active centre. Braker and Tiedje (2003) were the start to study denitrifying communities using norB equally a functional marking, and others accept used it for studies on environmental samples (Dandie et al., 2007) or isolates (Heylen et al., 2006b), just non yet for compost samples.

Nitrous oxide reductase is the final oxidoreductase of denitrification that transforms N2O to Northwardtwo (Brown et al., 2000). Considering this multi‐copper‐containing enzyme prevents the accumulation of a potent greenhouse gas, information technology plays an important office in the nitrogen bike (Zumft, 2005). Nitrous oxide reductase is most sensitive to molecular oxygen amongst the enzymes involved in denitrification, and its function is inhibited under aerobic atmospheric condition. This enzyme can as well exist inhibited by C2Htwo easily (Yoshinari and Knowles, 1976) and is oft used for the report of the denitrification potential of environmental samples (Teissier and Torre, 2002). Moreover, the nosZ gene that codes this enzyme is used every bit a biomarker in molecular ecological studies (Scala and Kerkhof, 1999; Stres et al., 2004).

Some bacterial denitrifiers and fungi are known non to possess nitrous oxide reductase (Takaya, 2009). Although NiiO reduction is thermodynamically favourable and Due northtwoO is suitable for an electron acceptor, some denitrifiers produce N2O as the last product of the denitrification process. This might exist explained by the fact that nitrous oxide is not toxic to microorganisms, whereas NO is toxic to bacterial cells. The lack of N2O reduction makes ∼20% difference to the bioenergetics of the bacterium (Richardson et al., 2009).

Diversity of denitrifiers in the environments

A written report about denitrifier communities in the composting process revealed an initial variation of nirK diversity and stability after that (Maeda et al., 2010a). Hallin and colleagues (2006) also reported that the addition of methanol or ethanol to activated sludge significantly afflicted the diverseness of nirS but not that of nirK. On the other hand, the add-on of mature compost that contains NOii or NO3‐N did non affect nirK variety but significantly affected nosZ diversity, suggesting that denitrifiers possessing the nosZ cistron in the compost would be more than sensitive to environmental conditions. It is necessary to isolate the major denitrifier revealed by molecular methods in club to understand the bodily denitrification occurring in the environment. In a denitrifier community study of rice paddy soil, Ashida and colleagues (2010) successfully isolated a major denitrifier through the enhancement of denitrification activity with succinate amendment and molecular methods such as the 16S rRNA gene clone library arroyo (Ishii et al., 2009) or the stable isotope probing approach (Saito et al., 2008). Moreover, Ishii and colleagues (2011) proved that denitrifiers with different 16S rRNA cistron phylogeny possess same nirS or nirK gene in the same surround (Fig. 2). Their information testify previously unknown complex human relationship between 16S rRNA gene and functional gene possession. To understand the denitrifier community completely, it is necessary to combine independent approaches such every bit molecular and conventional cultivation approaches. The molecular methods used to characterize the unknown and uncultivated denitrifier communities, and the subsequent single‐cell isolation strategy would be constructive for the denitrifiers that are truly functioning for actual denitrification in the environment (Ishii et al., 2010).

An external file that holds a picture, illustration, etc.  Object name is mbt0004-0700-f2.jpg

The relationship of 16S rRNA gene and functional denitrifier genes (nirS, nirK) possession of denitrifiers isolated from rice paddy soil (Redrawn from Ishii et al., 2011). Isolated strains were written in assuming, and the possession human relationship was connected by dotted lines.

The relationship between denitrifying cistron diversity or abundance and potential denitrification activity in soil has been well studied. Potential denitrification action, Due north2O/(N2O + North2) and the denitrifier community would be affected by pH (Palmer et al., 2010). Cuhel and colleagues (2010) reported that nirS diverseness correlates with soil pH. On the other hand, Hallin and colleagues (2009) reported that denitrification activity did not correlate with denitrifier cistron composition, only did correlate with the size of the total bacterial customs or nosZ affluence. Another study reports that the nosZ ratio to total bacterial community is much more than important than denitrifying gene abundance for potential N2O production (Philippot et al., 2009). Although much effort has been fabricated in these environmental studies, it is still difficult to explain denitrification and nitrous oxide product past denitrifying factor abundance or diversity. Moreover, a previous paper reported that AOA possesses a novel nirK sequence (Bartossek et al., 2010), which had not been covered by previous denitrifier studies, and much effort should be made to acquire about this unknown archaean denitrification.

It is also known that some autotrophic nitrifiers accept the ability to denitrify (Wrage et al., 2001; Wrage et al., 2004; Shaw et al., 2006). These autotrophic nitrifiers possess nirK‐type nitrite reductase with distinct DNA sequences from those of heterotrophic denitrifiers. It is non well understood yet how these autotrophic nitrifiers acquired the nirK gene, which might have occurred past horizontal cistron transfer; and how they became tolerant to nitrite produced by themselves (Casciotti and Ward, 2001). Because nitrifiers produce nitrite, they have many advantages for the utilization of nitrite every bit the substrate for denitrification. Therefore, nitrifier denitrification may contribute much more than heterotrophic denitrifiers, only it is difficult to distinguish these pathways with the current analytic techniques.

Who is responsible for nitrous oxide emission?

It is hard to understand how to share NO in nitrification and denitrification procedures. Equally stated above, nitrification would be performed by AOB, AOA, heterotrophic nitrifiers or fungi, whereas denitrification would be performed by heterotrophic denitrifiers, denitrifying fungi and autotrophic/heterotrophic nitrifiers. However, the relative contribution to net nitrification or denitrification of each grouping is not notwithstanding articulate. Studies that focus on nitrifying genes are but almost AOB or AOA, and those for denitrifying gene analysis are just most leaner. The development of a tool for such study will exist needed.

Stable isotope analysis of N2O is an culling approach to studying its product processes because the relative abundance of stable isotopes is a function of their abundance in source materials and the isotope fractionation factor of each concrete/chemical process. In detail, intramolecular 15Due north distribution inside the Northward2O molecule (site preference, SP) has been found to depend only on enzymatic reaction processes and not on substrates (Toyoda and Yoshida, 1999; Yoshida and Toyoda, 2000; Toyoda et al., 2005; Sutka et al., 2006). Nitrous oxide, which originates from bacterial nitrification (hydroxylamine oxidation) and denitrification (nitrite reduction), tin exist distinguished past using SP. Still, SP cannot distinguish betwixt nitrifier denitrification and heterotrophic denitrification, and a contempo study showed that fungal denitrification produces Northward2O with SP similar to that of bacterial nitrification (Sutka et al., 2008). In addition, isotope affluence is afflicted by nitrous oxide reduction (Ostrom et al., 2007; Jinuntuya‐Nortman et al., 2008). Although this analytical technique has some limitations every bit stated to a higher place, it would be a powerful tool by using all available isotopic data (Northward and O isotope ratios and SP) in a complementary style (east.g. Koba et al., 2009) or by combining other belittling approaches, such as a wide range of molecular methods. Isotopomer analysis of N2O straight collected from a composting pile by the dynamic chamber method (Osada and Fukumoto, 2001) revealed that bacterial denitrification is the most important and responsible nitrous oxide production pathway (Maeda et al., 2010b). This report relied on the isotopic characteristics of N2O produced by isolates not from compost but from other environments (such as soil). Therefore, in future studies, we need to isolate the major nitrifier, denitrifier, denitrifying fungi and isotope signature of their producing Due northiiO.

Future perspective

Because of the ease with which it is managed, composting will continue to exist a major engineering for treating animal manure. However, analysing the techniques developed previously cannot explain the nitrogen cycle and nitrous oxide emission yet. In the future, combining distinct approaches such every bit molecular methods, stable isotope analysis and classical isolation techniques will help us to sympathize the nitrogen bicycle during the composting process in detail. The results should lead to the development of relevant mitigation strategies, which will include: identification of the master players in the nitrification–denitrification process in the composting piles; isolation of these key players; and analysis of their physiological, biochemical and ecological properties. It would be also of interest to identify the nitrous oxide reducers and to study their office in the composting piles.

Acknowledgments

This piece of work was supported by the National Agronomics and Food Research Organization (NARO), Nippon. The authors would like to acknowledge Dr. Ishii for providing Effigy ii.

References

  • Abell G., Revill A., Smith C., Bissett A., Volkman J., Robert South. Archaeal ammonia oxidizers and nirS‐type denitrifiers dominate sediment nitrifying and denitrifying populations in a subtropical macrotidal estuary. ISME J. 2009;4:286–300. [PubMed] [Google Scholar]
  • Anastasi A., Varese Thousand., Marchisio V.F. Isolation and identification of fungal communities in compost and vermicompost. Mycologia. 2005;97:33–44. [PubMed] [Google Scholar]
  • Antonyuk South., Strange R., Sawers G., Eady R., Hasnain S. Diminutive resolution structures of resting‐state, substrate‐and product‐complexed Cu‐nitrite reductase provide insight into catalytic mechanism. Proc Natl Acad Sci Usa. 2005;102:12041–12046. [PMC gratis article] [PubMed] [Google Scholar]
  • Ashida Northward., Ishii Southward., Hayano S., Tago K., Tsuji T., Yoshimura Y. Isolation of functional unmarried cells from environments using a micromanipulator: awarding to study denitrifying bacteria. Appl Microbiol Biotechnol. 2010;85:1211–1217. et al. [PubMed] [Google Scholar]
  • Bartossek R., Nicol G., Lanzen A., Klenk H., Schleper C. Homologues of nitrite reductases in ammonia‐oxidizing archaea: variety and genomic context. Environ Microbiol. 2010;12:1075–1088. [PubMed] [Google Scholar]
  • Beffa T., Blanc M., Lyon P., Vogt Chiliad., Marchiani Grand., Fischer J., Aragno M. Isolation of Thermus strains from hot composts (60 to lxxx degrees C) Appl Environ Microbiol. 1996;62:1723–1727. [PMC gratuitous article] [PubMed] [Google Scholar]
  • Bernal Yard., Alburquerque J., Moral R. Composting of creature manures and chemical criteria for compost maturity cess. A review. Bioresour Technol. 2009;100:5444–5453. [PubMed] [Google Scholar]
  • Blanc M., Marilley 50., Beffa T., Aragno Thou. Notes: rapid identification of heterotrophic, thermophilic, spore‐forming bacteria isolated from hot composts. Int J Syst Evol Microbiol. 1997;47:1246–1248. [PubMed] [Google Scholar]
  • Blanc M., Marilley L., Beffa T., Aragno One thousand. Thermophilic bacterial communities in hot composts as revealed by most probable number counts and molecular (16S rDNA) methods. FEMS Microbiol Ecol. 1999;28:141–149. [Google Scholar]
  • Braker G., Tiedje J.M. Nitric oxide reductase (norB) genes from pure cultures and environmental samples. Appl Environ Microbiol. 2003;69:3476–3483. [PMC gratis article] [PubMed] [Google Scholar]
  • Brochier‐Armanet C., Boussau B., Gribaldo Southward., Forterre P. Mesophilic Crenarchaeota: proposal for a 3rd archaeal phylum, the Thaumarchaeota. Nat Rev Microbiol. 2008;6:245–252. [PubMed] [Google Scholar]
  • Brown Yard., Tegoni M., Prudêncio K., Pereira A., Besson South., Moura J. A novel blazon of catalytic copper cluster in nitrous oxide reductase. Nat Struct Mol Biol. 2000;7:191–195. et al. [PubMed] [Google Scholar]
  • Casciotti K., Ward B. Dissimilatory nitrite reductase genes from autotrophic ammonia‐oxidizing leaner. Appl Environ Microbiol. 2001;67:2213–2221. [PMC free article] [PubMed] [Google Scholar]
  • Cuhel J., Simek K., Laughlin R., Bru D., Cheneby D., Watson C., Philippot L. Insights into the effect of soil pH on N2O and Nii emissions and denitrifier community size and activity. Appl Environ Microbiol. 2010;76:1870–1878. [PMC free article] [PubMed] [Google Scholar]
  • Dandie C., Miller M., Burton D., Zebarth B., Trevors J., Goyer C. Nitric oxide reductase‐targeted real‐fourth dimension PCR quantification of denitrifier populations in soil. Appl Environ Microbiol. 2007;73:4250–4258. [PMC gratuitous article] [PubMed] [Google Scholar]
  • Danon G., Franke‐Whittle I., Insam H., Chen Y., Hadar Y. Molecular analysis of bacterial community succession during prolonged compost curing. FEMS Microbiol Ecol. 2008;65:133–144. [PubMed] [Google Scholar]
  • Daum Thou., Zimmer W., Papen H., Kloos K., Nawrath K., Bothe H. Physiological and molecular biological characterization of ammonia oxidation of the heterotrophic nitrifier Pseudomonas putida. Curr Microbiol. 1998;37:281–288. [PubMed] [Google Scholar]
  • De Boer Westward., Kowalchuk G. Nitrification in acid soils: micro‐organisms and mechanisms. Soil Biol Biochem. 2001;33:853–866. [Google Scholar]
  • Dees P.M., Ghiorse W.C. Microbial diversity in hot synthetic compost every bit revealed past PCR‐amplified rRNA sequences from cultivated isolates and extracted Deoxyribonucleic acid. FEMS Microbiol Ecol. 2001;35:207–216. [PubMed] [Google Scholar]
  • Eghball B., Power J., Gilley J., Doran J.West. Nutrient, carbon, and mass loss during composting of beef cattle feedlot manure. J Environ Qual. 1997;26:189–193. [Google Scholar]
  • Einsle O., Messerschmidt A., Stach P., Bourenkov G., Bartunik H., Huber R., Kroneck P. Structure of cytochrome c nitrite reductase. Nature. 1999;400:476–480. [PubMed] [Google Scholar]
  • Einsle O., Stach P., Messerschmidt A., Simon J., Kröger A., Huber R., Kroneck P. Cytochrome c nitrite reductase from Wolinella succinogenes. J Biol Chem. 2000;275:39608–39616. [PubMed] [Google Scholar]
  • El Kader N., Robin P., Paillat J., Leterme P. Turning, compacting and the addition of water as factors affecting gaseous emissions in farm manure composting. Bioresour Technol. 2007;98:2619–2628. [PubMed] [Google Scholar]
  • Fernandes Fifty., Zhan W., Patni North., Jui P. Temperature distribution and variation in passively aerated static compost piles. Bioresour Technol. 1994;48:257–264. [Google Scholar]
  • Fukumoto Y., Inubushi Chiliad. Consequence of nitrite accumulation on nitrous oxide emission and total nitrogen loss during swine manure composting. Soil Sci Found Nutr. 2009;55:428–434. [Google Scholar]
  • Fukumoto Y., Osada T., Hanajima D., Haga K. Patterns and quantities of NHthree, N2O and CH4 emissions during swine manure composting without forced aeration – consequence of compost pile scale. Bioresour Technol. 2003;89:109–114. [PubMed] [Google Scholar]
  • Fukumoto Y., Suzuki Yard., Osada T., Kuroda Chiliad., Hanajima D., Yasuda T., Haga K. Reduction of nitrous oxide emission from pig manure composting by addition of nitrite‐oxidizing bacteria. Environ Sci Technol. 2006;40:6787–6791. [PubMed] [Google Scholar]
  • Hallin South., Throback I.N., Dicksved J., Pell Thou. Metabolic profiles and genetic variety of denitrifying communities in activated sludge later improver of methanol or ethanol. Appl Environ Microbiol. 2006;72:5445–5452. [PMC free article] [PubMed] [Google Scholar]
  • Hallin S., Jones C., Schloter M., Philippot L. Relationship between N‐cycling communities and ecosystem performance in a 50‐twelvemonth‐former fertilization experiment. ISME J. 2009;three:597–605. [PubMed] [Google Scholar]
  • Hao X., Chang C., Larney F.J., Travis K.R. Greenhouse gas emissions during cattle feedlot manure composting. J Environ Qual. 2001;xxx:376–386. [PubMed] [Google Scholar]
  • Hao X., Chang C., Larney F.J. Carbon, nitrogen balances and greenhouse gas emission during cattle feedlot manure composting. J Environ Qual. 2004;33:37–44. [PubMed] [Google Scholar]
  • He Y., Inamori Y., Mizuochi M., Kong H., Iwami N., Sun T. Nitrous oxide emissions from aerated composting of organic waste. Environ Sci Technol. 2001;35:2347–2351. [PubMed] [Google Scholar]
  • He Y.W., Inamori Y., Mizuochi G., Kong H.N., Iwami Northward., Sunday T.H. Measurements of N2O and CH4 from the aerated composting of food waste product. Sci Full Environ. 2000;254:65–74. [PubMed] [Google Scholar]
  • Heylen Yard., Gevers D., Vanparys B., Wittebolle L., Geets J., Boon N., De Vos P. The incidence of nirS and nirK and their genetic heterogeneity in cultivated denitrifiers. Environ Microbiol. 2006a;eight:2012–2021. [PubMed] [Google Scholar]
  • Heylen One thousand., Vanparys B., Gevers D., Wittebolle L., Benefaction Due north., De Vos P. Nitric oxide reductase (norB) gene sequence analysis reveals discrepancies with nitrite reductase (nir) gene phylogeny in cultivated denitrifiers. Environ Microbiol. 2006b;9:1072–1077. [PubMed] [Google Scholar]
  • Hultman J., Vasara T., Partanen P., Kurola J., Kontro M., Paulin L. Determination of fungal succession during municipal solid waste matter composting using a cloning‐based analysis. J Appl Microbiol. 2009;108:472–487. et al. [PubMed] [Google Scholar]
  • Iida 1000., Ueda Y., Kawamura Y., Ezaki T., Takade A., Yoshida South., Amako G. Paenibacillus motobuensis sp. nov., isolated from a composting machine utilizing soil from Motobu‐town, Okinawa, Japan. Int J Syst Evol Microbiol. 2005;55:1811–1816. [PubMed] [Google Scholar]
  • Innerebner Thou., Knapp B., Vasara T., Romantschuk M., Insam H. Traceability of ammonia‐oxidizing leaner in compost‐treated soils. Soil Biol Biochem. 2006;38:1092–1100. [Google Scholar]
  • Intergovernmental Console on Climatic change. Cambridge Academy Printing; 2001. [Google Scholar]
  • Ishii K., Takii Due south. Comparison of microbial communities in four different composting processes as evaluated by denaturing gradient gel electrophoresis analysis. J Appl Microbiol. 2003;95:109–119. [PubMed] [Google Scholar]
  • Ishii K., Fukui M., Takii South. Microbial succession during a composting process as evaluated by denaturing gradient gel electrophoresis analysis. J Appl Microbiol. 2000;89:768–777. [PubMed] [Google Scholar]
  • Ishii S., Yamamoto M., Kikuchi Thou., Oshima K., Hattori M., Otsuka S., Senoo K. Microbial populations responsive to denitrification‐inducing atmospheric condition in rice paddy soil, as revealed past comparative 16S rRNA gene analysis. Appl Environ Microbiol. 2009;75:7070–7078. [PMC gratis article] [PubMed] [Google Scholar]
  • Ishii S., Tago K., Senoo K. Single‐prison cell analysis and isolation for microbiology and biotechnology: methods and applications. Appl Microbiol Biotechnol. 2010;86:1281–1292. [PubMed] [Google Scholar]
  • Ishii S., Ashida N., Otsuka S., Senoo K. 2011. , and ) Isolation of oligotrophic denitrifiers carrying previously uncharacterized functional gene sequences. Appl Environ Microbiol (in press): doi: ten.1128/AEM.0218910.
  • Jarvis A., Sundberg C., Milenkovski S., Pell Grand., Smars S., Lindgren P., Hallin S. Activity and composition of ammonia oxidizing bacterial communities and emission dynamics of NH3 and Due north2O in a compost reactor treating organic household waste. J Appl Microbiol. 2009;106:1502–1511. [PubMed] [Google Scholar]
  • Jinuntuya‐Nortman M., Sutka R., Ostrom P., Gandhi H., Ostrom N. Isotopologue fractionation during microbial reduction of NiiO within soil mesocosms every bit a function of water‐filled pore infinite. Soil Biol Biochem. 2008;40:2273–2280. [Google Scholar]
  • Junier P., Molina V., Dorador C., Hadas O., Kim O., Junier T. Phylogenetic and functional marker genes to study ammonia‐oxidizing microorganisms (AOM) in the environment. Appl Microbiol Biotechnol. 2010;85:425–440. et al. [PMC complimentary article] [PubMed] [Google Scholar]
  • Kim B., Lee Southward., Weon H., Kwon S., Go S., Park Y. Ureibacillus suwonensis sp. nov., isolated from cotton waste product composts. Int J Syst Evol Microbiol. 2006;56:663–666. et al. [PubMed] [Google Scholar]
  • Ko H., Kim K., Kim H., Kim C., Umeda M. Evaluation of maturity parameters and heavy metal contents in composts fabricated from animal manure. Waste Manag. 2008;28:813–820. [PubMed] [Google Scholar]
  • Koba K., Osaka Grand., Tobari Y., Toyoda Due south., Ohte N., Katsuyama G., Suzuki N. Biogeochemistry of nitrous oxide in groundwater in a forested ecosystem elucidated past nitrous oxide isotopomer measurements. Geoch Cosmoch Acta. 2009;73:3115–3133. [Google Scholar]
  • Kowalchuk Yard., Naoumenko Z., Derikx P., Felske A., Stephen J., Arkhipchenko I. Molecular analysis of ammonia‐oxidizing leaner of the β subdivision of the grade Proteobacteria in compost and composted materials. Appl Environ Microbiol. 1999;65:396–403. [PMC complimentary article] [PubMed] [Google Scholar]
  • Kowalchuk G., Stephen J. Ammonia‐oxidizing leaner: a model for molecular microbial ecology. Ann Rev Microbiol. 2001;55:485–529. [PubMed] [Google Scholar]
  • Kuroda K., Osada T., Yonaga M., Kanematu A., Nitta T., Mouri S., Kojima T. Emissions of malodorous compounds and greenhouse gases from composting swine feces. Bioresour Technol. 1996;56:265–271. [Google Scholar]
  • Laughlin R., Stevens R., Müller C., Watson C. Evidence that fungi tin oxidize NH4 + to NOiii in a grassland soil. Eur J Soil Sci. 2008;59:285–291. [Google Scholar]
  • Leininger South., Urich T., Schloter M., Schwark L., Qi J., Nicol G. Archaea predominate among ammonia‐oxidizing prokaryotes in soils. Nature. 2006;442:806–809. et al. [PubMed] [Google Scholar]
  • Maeda K., Hanajima D., Morioka R., Osada T. Label and spatial distribution of bacterial communities within passively aerated cattle manure composting piles. Bioresour Technol. 2010a;101:9631–9637. [PubMed] [Google Scholar]
  • Maeda K., Toyoda S., Shimojima R., Osada T., Hanajima D., Morioka R., Yoshida North. The source of nitrous oxide emission from cattle manure composting procedure revealed by isotopomer analysis and amoA abundance of beta‐proteobacterial ammonia oxidizing bacteria. Appl Environ Microbiol. 2010b;76:1555–1562. [PMC gratuitous article] [PubMed] [Google Scholar]
  • Maeda Yard., Morioka R., Hanajima D., Osada T. The bear on of using mature compost on nitrous oxide emission and the denitrifier community in the cattle manure composting process. Microb Ecol. 2010c;59:25–36. [PubMed] [Google Scholar]
  • Mahimairaja Southward., Bolan N.S., Hedley M.J. Denitrification losses of N from fresh and composted manures. Soil Biol Biochem. 1995;27:1223–1225. [Google Scholar]
  • Martins O., Dewes T. Loss of nitrogenous compounds during composting of brute wastes. Bioresour Technol. 1992;42:103–111. [Google Scholar]
  • Mertens J., Broos K., Wakelin S., Kowalchuk Chiliad., Springael D., Smolders E. Bacteria, non archaea, restore nitrification in a zinc‐contaminated soil. ISME J. 2009;3:916–923. [PubMed] [Google Scholar]
  • Mertoglu B., Calli B., Inanc B., Ozturk I. Evaluation of in situ ammonia removal in an aerated landfill bioreactor. Proc Biochem. 2006;41:2359–2366. [Google Scholar]
  • Moir J., Crossman L., Spiro Due south., Richardson D. The purification of ammonia monooxygenase from Paracoccus denitrificans. FEBS Lett. 1996;387:71–74. [PubMed] [Google Scholar]
  • Murphy Thousand., Turley Southward., Adman E. Construction of nitrite bound to copper‐containing nitrite reductase from Alcaligenes faecalis. J Biol Chem. 1997;272:28455–28460. [PubMed] [Google Scholar]
  • Muyzer G., De Waal E., Uitterlinden A. Profiling of complex microbial populations past denaturing gradient gel electrophoresis analysis of polymerase chain reaction‐amplified genes coding for 16S rRNA. Appl Environ Microbiol. 1993;59:695–700. [PMC complimentary article] [PubMed] [Google Scholar]
  • Nicholson F., Chambers B., Williams J., Unwin R. Heavy metallic contents of livestock feeds and creature manures in England and Wales. Bioresour Technol. 1999;seventy:23–31. [Google Scholar]
  • Osada T., Fukumoto Y. Evolution of a new dynamic bedroom organization for measuring harmful gas emissions from composting livestock waste. Wat Sci Technol. 2001;44:79–86. [PubMed] [Google Scholar]
  • Ostrom N., Pitt A., Sutka R., Ostrom P., Grandy A., Huizinga G., Robertson Chiliad. Isotopologue effects during NiiO reduction in soils and in pure cultures of denitrifiers. J Geophys Res. 2007;112:G02005. [Google Scholar]
  • Palmer K., Drake H., Horn Thou. Clan of novel and highly diverse acid‐tolerant denitrifiers with N2O fluxes of an acidic fen. Appl Environ Microbiol. 2010;76:1125–1134. [PMC gratuitous commodity] [PubMed] [Google Scholar]
  • Papen H., Von Berg R. A most probable number method (MPN) for the estimation of cell numbers of heterotrophic nitrifying bacteria in soil. Plant Soil. 1998;199:123–130. [Google Scholar]
  • Park H., Wells G., Bae H., Criddle C., Francis C. Occurrence of ammonia‐oxidizing archaea in wastewater treatment plant bioreactors. Appl Environ Microbiol. 2006;72:5643–5647. [PMC gratis article] [PubMed] [Google Scholar]
  • Pattey E., Trzcinski 1000.K., Desjardins R.L. Quantifying the reduction of greenhouse gas emissions as a result of composting dairy and beef cattle manure. Nutr Cycl Agroecosys. 2005;72:173–187. [Google Scholar]
  • Peters Due south., Koschinsky Due south., Schwieger F., Tebbe C. Succession of microbial communities during hot composting as detected by PCR‐single‐strand‐conformation polymorphism‐based genetic profiles of small-scale‐subunit rRNA genes. Appl Environ Microbiol. 2000;66:930–936. [PMC free article] [PubMed] [Google Scholar]
  • Philippot 50., Čuhel J., Saby North., Chèneby D., Chronáková A., Bru D. Mapping field‐scale spatial patterns of size and activity of the denitrifier community. Environ Microbiol. 2009;11:1518–1526. et al. [PubMed] [Google Scholar]
  • Podmirseg S., Schoen Thou., Murthy Southward., Insam H., Wett B. Quantative and qualitative effects of bioaugmentation on ammonia oxidizers at a two‐step WWTP. Wat Sci Technol. 2010;61:1003–1009. [PubMed] [Google Scholar]
  • Poly F., Wertz S., Brothier E., Degrange V. First exploration of Nitrobacter diversity in soils by a PCR cloning‐sequencing arroyo targeting functional gene nxrA. FEMS Microbiol Ecol. 2008;63:132–140. [PubMed] [Google Scholar]
  • Ravishankara A., Daniel J., Portmann R. Nitrous oxide (N2O): the dominant ozone‐depleting substance emitted in the 21st century. Science. 2009;326:123–125. [PubMed] [Google Scholar]
  • Richardson D., Felgate H., Watmough N., Thomson A., Baggs E. Mitigating release of the stiff greenhouse gas NtwoO from the nitrogen cycle – could enzymic regulation hold the fundamental? Trends Biotechnol. 2009;27:388–397. [PubMed] [Google Scholar]
  • Rudolf G., Kroneck P. The nitrogen cycle: its biology. Met Ions Biol Syst. 2005;43:75–104. [PubMed] [Google Scholar]
  • Saito T., Ishii Due south., Otsuka South., Nishiyama Grand., Senoo K. Identification of novel betaproteobacteria in a succinate‐assimilating population in denitrifying rice paddy soil past using stable isotope probing. Microb Environ. 2008;23:192–200. [PubMed] [Google Scholar]
  • Santoro A., Casciotti K., Francis C. Activity, abundance and diversity of nitrifying archaea and leaner in the fundamental California Electric current. Environ Microbiol. 2010;12:1989–2006. [PubMed] [Google Scholar]
  • Scala D., Kerkhof L. Diversity of nitrous oxide reductase (nosZ) genes in continental shelf sediments. Appl Environ Microbiol. 1999;65:1681–1687. [PMC free article] [PubMed] [Google Scholar]
  • Schloss P., Hay A., Wilson D., Walker L. Tracking temporal changes of bacterial community fingerprints during the initial stages of composting. FEMS Microbiol Ecol. 2003;46:1–9. [PubMed] [Google Scholar]
  • Sharma S., Aneja G.K., Mayer J., Munch J.C., Schloter G. Diversity of transcripts of nitrite reductase genes (nirK and nirS) in rhizospheres of grain legumes. Appl Environ Microbiol. 2005;71:2001–2007. [PMC free article] [PubMed] [Google Scholar]
  • Shaw L., Nicol G., Smith Z., Fright J., Prosser J., Baggs Eastward. Nitrosospira spp. can produce nitrous oxide via a nitrifier denitrification pathway. Environ Microbiol. 2006;viii:214–222. [PubMed] [Google Scholar]
  • Shimaya C., Hashimoto T. Improvement of media for thermophilic ammonia‐oxidizing bacteria in compost. Soil Sci Plant Nutr. 2008;54:529–533. [Google Scholar]
  • Stres B., Mahne I., Avgustin M., Tiedje J.K. Nitrous oxide reductase (nosZ) gene fragments differ between native and cultivated Michigan soils. Appl Environ Microbiol. 2004;70:301–309. [PMC free article] [PubMed] [Google Scholar]
  • Sutka R.50., Ostrom North.East., Ostrom P.H., Breznak J.A., Gandhi H., Pitt A.J., Li F. Distinguishing nitrous oxide production from nitrification and denitrification on the basis of isotopomer abundances. Appl Environ Microbiol. 2006;72:638–644. [PMC free article] [PubMed] [Google Scholar]
  • Sutka R., Adams G., Ostrom Due north., Ostrom P. Isotopologue fractionation during N2O production by fungal denitrification. Rapid Commun Mass Spectrom. 2008;22:3989–3996. [PubMed] [Google Scholar]
  • Szanto G.Fifty., Hamelers H.V.K., Rulkens Due west.H., Veeken A.H.Grand. NH3, N2O and CH4 emissions during passively aerated composting of harbinger‐rich squealer manure. Bioresour Technol. 2007;98:2659–2670. [PubMed] [Google Scholar]
  • Takaya Northward. Response to hypoxia, reduction of electron acceptors, and subsequent survival by filamentous fungi. Biosci Biotechnol Biochem. 2009;73:i–8. [PubMed] [Google Scholar]
  • Tavares P., Pereira A., Moura J., Moura I. Metalloenzymes of the denitrification pathway. J Inorg Biochem. 2006;100:2087–2100. [PubMed] [Google Scholar]
  • Teissier S., Torre M. Simultaneous assessment of nitrification and denitrification on freshwater epilithic biofilms by acetylene block method. Wat Res. 2002;36:3803–3811. [PubMed] [Google Scholar]
  • Throback I.N., Enwall 1000., Jarvis A., Hallin South. Reassessing PCR primers targeting nirS, nirK and nosZ genes for community surveys of denitrifying bacteria with DGGE. FEMS Microbiol Ecol. 2004;49:401–417. [PubMed] [Google Scholar]
  • Tiquia S.Yard., Tam N.F.Y. Fate of nitrogen during composting of chicken litter. Environ Pollut. 2000;110:535–541. [PubMed] [Google Scholar]
  • Tiquia S., Richard T., Honeyman 1000. Carbon, food, and mass loss during composting. Nutr Cycl Agroecosyst. 2002;62:15–24. [Google Scholar]
  • Toyoda S., Yoshida Due north. Conclusion of nitrogen isotopomers of nitrous oxide on a modified isotope ratio mass spectrometer. Anal Chem. 1999;71:4711–4718. [Google Scholar]
  • Toyoda S., Mutobe H., Yamagishi H., Yoshida N., Tanji Y. Fractionation of N2O isotopomers during production by denitrifier. Soil Biol Biochem. 2005;37:1535–1545. [Google Scholar]
  • Wang C., Shyu C., Ho S., Chiou S. Species variety and substrate utilization patterns of thermophilic bacterial communities in hot aerobic poultry and cattle manure composts. Microb Ecol. 2007;54:1–9. [PubMed] [Google Scholar]
  • Wells G., Park H., Yeung C., Eggleston B., Francis C., Criddle C. Ammonia‐oxidizing communities in a highly aerated full‐scale activated sludge bioreactor: betaproteobacterial dynamics and depression relative abundance of Crenarchaea. Environ Microbiol. 2009;11:2310–2328. [PubMed] [Google Scholar]
  • Wertz S., Poly F., Le Roux X., Degrange V. Development and application of a PCR‐denaturing gradient gel electrophoresis tool to written report the diversity of Nitrobacter‐like nxrA sequences in soil. FEMS Microbiol Ecol. 2008;63:261–271. [PubMed] [Google Scholar]
  • Wertz Southward., Dandie C.E., Goyer C., Trevors J.T., Patten C.L. Diversity of nirK denitrifying genes and transcripts in an agricultural soil. Appl Environ Microbiol. 2009;75:7365–7377. [PMC costless article] [PubMed] [Google Scholar]
  • Wrage Due north., Velthof G., Van Beusichem G., Oenema O. Office of nitrifier denitrification in the production of nitrous oxide. Soil Biol Biochem. 2001;33:1723–1732. [Google Scholar]
  • Wrage N., Velthof M., Laanbroek H., Oenema O. Nitrous oxide production in grassland soils: assessing the contribution of nitrifier denitrification. Soil Biol Biochem. 2004;36:229–236. [Google Scholar]
  • Yamamoto N., Otawa K., Nakai Y. Bacterial communities developing during composting processes in animal manure handling facilities. Asian Austral J Anim Sci. 2009;22:900–905. [Google Scholar]
  • Yamamoto Northward., Otawa Chiliad., Nakai Y. Diversity and affluence of ammonia‐oxidizing bacteria and ammonia‐oxidizing archaea during cattle manure composting. Microb Ecol. 2010;threescore:807–815. [PubMed] [Google Scholar]
  • Yoshida N., Toyoda S. Constraining the atmospheric N2O upkeep from intramolecular site preference in Due northiiO isotopomers. Nature. 2000;405:330–334. [PubMed] [Google Scholar]
  • Yoshinari T., Knowles R. Acetylene inhibition of nitrous oxide reduction by denitrifying bacteria. Biochem Biophys Res Commun. 1976;69:705–710. [PubMed] [Google Scholar]
  • Zhang Y., Ronimus R., Turner Due north., Zhang Y., Morgan H. Enumeration of thermophilic Bacillus species in composts and identification with a random amplification polymorphic Dna (RAPD) protocol. Syst Appl Microbiol. 2002;25:618–626. [PubMed] [Google Scholar]
  • Zhu Southward., Chan G., Cai K., Qu 50., Huang Fifty. Leachates from municipal solid waste product disposal sites harbor similar, novel nitrogen‐cycling bacterial communities. FEMS Microbiol Lett. 2007;267:236–242. [PubMed] [Google Scholar]
  • Zumft Westward.Grand. Cell biology and molecular basis of denitrification. Microbiol Mol Biol Rev. 1997;61:533–616. [PMC free article] [PubMed] [Google Scholar]
  • Zumft W.G. Biogenesis of the bacterial respiratory Cu. J Mol Microbiol Biotechnol. 2005;10:154–166. [PubMed] [Google Scholar]

Articles from Microbial Biotechnology are provided here courtesy of Wiley-Blackwell


Source: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3815407/

Posted by: galazmagentleed80.blogspot.com

0 Response to "How Does Animal Waste Affect The Nitrogen Cycle"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel