Anammox

A bioreactor contaning the anammox bacterium Kuenenia stuttgartiensis, Radboud University Nijmegen

Anammox, an abbreviation for ANaerobic AMMonium OXidation, is a globally important microbial process of the nitrogen cycle.[1] The bacteria mediating this process were identified in 1999, and at the time were a great surprise for the scientific community.[2] It takes place in many natural environments and anammox is also the trademarked name for the anammox-based ammonium removal technology that was developed[3] by the Delft University of Technology.

Process background

C17-C20 ladderane lipids from anammox bacteria containing three linearly concatenated cyclobutane rings and one cyclohexane or five cyclobutane rings. Fatty acids are esterified with methanol or the glycerol backbone, and the ladderane alcohols are ether-linked with glycerol, all in different combinations.[4]

In this biological process, nitrite and ammonium ions are converted directly into diatomic nitrogen and water.[5]

NH4+ + NO2 → N2 + 2H2O.

Globally, this process may be responsible for 30-50% of the N2 gas produced in the oceans.[6] It is thus a major sink for fixed nitrogen and so limits oceanic primary productivity.

The bacteria that perform the anammox process belong to the bacterial phylum Planctomycetes. Currently, five anammox genera have been discovered: Brocadia, Kuenenia, Anammoxoglobus, Jettenia (all fresh water species), and Scalindua (marine species).[7] The anammox bacteria are characterized by several striking properties: they all possess one anammoxosome, a membrane bound compartment inside the cytoplasm which is the locus of anammox catabolism. Further, the membranes of these bacteria mainly consist of ladderane lipids so far unique in biology.[8] Of special interest is the conversion to hydrazine (normally used as a high-energy rocket fuel, and poisonous to most living organisms) as an intermediate.[9] A final striking feature of the organism is the extremely slow growth rate. The doubling time is anywhere from 7–22 days.[4] The anammox bacteria are geared towards converting their substrates at very low concentrations; in other words, they have a very high affinity to their substrates ammonium and nitrite (sub-micromolar range).[10][11] Anammox cells are packed with cytochrome c type proteins (~30% of the protein complement), including the enzymes that perform the key catabolic reactions of the anammox process, making the cells remarkably red.[12] The anammox process was originally found to occur only from 20 °C to 43 °C[10] but more recently, anammox has been observed at temperatures from 36 °C to 52 °C in hot springs[13] and 60 °C to 85 °C at hydrothermal vents located along the Mid-Atlantic Ridge.[14]

History

Figure 2. The biological N cycle. DNRA, dissimilatory nitrate reduction to ammonium.

In 1932, it was reported that dinitrogen gas was generated via an unknown mechanism during fermentation in the sediments of Lake Mendota, Wisconsin, USA.[15] More than 40 years ago, Richards[16] noticed that most of the ammonium that should be produced during the anaerobic remineralization of organic matter was unaccounted for. As there was no known biological pathway for this transformation, biological anaerobic oxidation of ammonium received little further attention.[17] Thirty years ago, the existence of two chemolithoautotrophic microorganisms capable of oxidizing ammonium to dinitrogen gas was predicted on the basis of thermodynamic calculations.[18] It was thought that anaerobic oxidation of ammonium would not be feasible, assuming that the predecessors had tried and failed to establish a biological basis for those reactions. By the 1990s, Arnold Mulder's observations were just consistent with Richard's suggestion.[19] In their anoxic denitrifying pilot reactor, ammonium disappeared at the expense of nitrite with a clear nitrogen production. The reactor used the effluent from a methanogenic pilot reactor, which contained ammonium, sulphide and other compounds, and nitrate from a nitrifying plant as the influent. The process was named "anammox," and was realized to have great significance in the removal of unwanted ammonium. The discovery of the anammox process was first publicly presented at the 5th European congress on biotechnology.[20] By the mid-1990s, the discovery of anammox in the fluidized bed reactor was published.[21] A maximum ammonium removal rate of 0.4 kg N/m3/d was achieved. It was shown that for every mole of ammonium consumed, 0.6 mol of nitrate was required, resulting in the formation of 0.8 mol of N2 gas. In the same year, the biological nature of anammox was identified.[22] Labeling experiments with 15NH4+ in combination with 14NO3 showed that 14-15N2 was the dominant product making up 98.2% of the total labeled N2. It was realized that, instead of nitrate, nitrite was assumed as the oxidizing agent of ammonium in anammox reaction. Based on a previous study, Strous et al.[23] calculated the stoichiometry of anammox process by mass balancing, which is widely accepted by other groups. Later, anammox bacteria were identified as planctomycetes,[24] and the first identified anammox organism was named Candidatus "Brocadia anammoxidans."[25] Before 2002, anammox was assumed to be a minor player in the nitrogen cycle within natural ecosystems.[26] In 2002 however, anammox was found to play an important part in the biological nitrogen cycle, accounting for 24-67% of the total N2 production in the continental shelf sediments that were studied.[27] The discovery of anammox process modified the concept of biological nitrogen cycle, as depicted in Figure 2.

Possible reaction mechanisms

Figure 3. Possible biochemical pathway and cellular localization of the enzyme systems involved in anammox reaction.
Figure 4. Hypothetical metabolic pathways and reversed electron transport in the anammoxosome. (a) Anammox catabolism that uses nitrite as the electron acceptor for the creation of a proton motive force over the anammoxosomal membrane. (b) Proton motive force- driven reversed electron transport combines central catabolism with nitrate reductase (NAR) to generate ferredoxin for carbon dioxide reduction in the acetyl-CoA pathway. HAO, hydrazine oxidoreductase; HD, hydrazine dehydrogenase; HH, hydrazine hydrolase; NIR, nitrite oxidoreductase; Q, quinine. Light blue diamonds, cytochromes; blue arrows, reductions; pink arrows, oxidations.

According to 15N labeling experiments carried out in 1997, ammonium is biologically oxidized by hydroxylamine, most likely derived from nitrite, as the probable electron acceptor.[28] The conversion of hydrazine to dinitrogen gas is hypothesized to be the reaction that generates the electron equivalents for the reduction of nitrite to hydroxylamine.[29] In general, two possible reaction mechanisms are addressed.[30] One mechanism hypothesizes that a membrane-bound enzyme complex converts ammonium and hydroxylamine to hydrazine first, followed by the oxidation of hydrazine to dinitrogen gas in the periplasm. At the same time, nitrite is reduced to hydroxylamine at the cytoplasmic site of the same enzyme complex responsible for hydrazine oxidation with an internal electron transport (Figure 3a). The other mechanism postulates the following: ammonium and hydroxylamine are converted to hydrazine by a membrane-bound enzyme complex, hydrazine is oxidized in the periplasm to dinitrogen gas, and the generated electrons are transferred via an electron transport chain to nitrite reducing enzyme in the cytoplasm where nitrite is reduced to hydroxylamine (Figure 3b). Whether the reduction of nitrite and the oxidation of hydrazine occur at different sites of the same enzyme or the reactions are catalyzed by different enzyme systems connected via an electron transport chain remains to be investigated.[29] In microbial nitrogen metabolism, the occurrence of hydrazine as an intermediate is rare.[31] Hydrazine has been proposed as an enzyme-bound intermediate in the nitrogenase reaction.[32] Recently, using detailed molecular analyses and combining complementary methods, Kartal and coworkers published strong evidence supporting the latter mechanism.[12][33] Furthermore, the enzyme producing hydrazine, hydrazine synthase was purified and shown to produce hydrazine from NO and ammonium.[12] The production of hydrazine from ammonium and NO was also supported by the resolution of the crystal structure of the enzyme hydrazine sythase.[34]

A possible role of nitric oxide (NO) or nitroxyl (HNO) in anammox was proposed by Hooper et al.[35] by way of condensation of NO or HNO and ammonium on an enzyme related to the ammonium monooxygenase family. The formed hydrazine or imine could subsequently be converted by the enzyme hydroxylamine oxidase to dinitrogen gas, and the reducing equivalents produced in the reaction are required to combine NO or HNO and ammonium or to reduce nitrite to NO. Environmental genomics analysis of the species Candidatus Kuenenia stuttgartiensis, through a slightly different and complementary metabolism mechanism, suggested NO to be the intermediate instead of hydroxylamine (Figure 4).[36] However, this hypothesis also agreed that hydrazine was an important intermediate in the process. In this pathway (Figure 4), there are two enzymes unique to anammox bacteria: hydrazine hydrolase (hh) and hydrazine dehydrogenase (hd). The hh produces hydrazine from nitric oxide and ammonium, and hd transfer the electrons from hydrazine to ferredoxin. Few new genes, such as some known fatty acid biosynthesis and S-adenosylmethionine radical enzyme genes, containing domains involved in electron transfer and catalysis have been detected.[36]

Species diversity

Until now, ten anammox species have been described, including seven that are available in laboratory enrichment cultures.[4] All have the taxonomical status of Candidatus, as none were obtained as classical pure cultures. Known species are divided over five genera: (1) Kuenenia, represented by Kuenenia stuttgartiensis,[36] (2) Brocadia (three species: B. anammoxidans, B. fulgida, and B. sinica),[24][37][38] (3) Anammoxoglobus (one species: A. propionicus,[39] (4) Jettenia (one species: J. asiatica,[40][41] and (5) Scalindua (four species: S. brodae, S. sorokinii, S. wagneri, and S. profunda[42][43][44] Representatives of the first four genera were enriched from sludge from wastewater treatment plants; K. stuttgartiensis, B. anammoxidans, B. fulgida, and A. propionicus were even obtained from the same inoculum. Scalindua dominates the marine environment, but is also found in some freshwater ecosystems and wastewater treatment plants.[42][45][46][47] Together, these 10 species likely only represent a minute fraction of anammox biodiversity. For instance, there are currently over 2000 16S rRNA gene sequences affiliated with anammox bacteria that have been deposited to the Genbank (http://www.ncbi.nlm.nih.gov/genbank/), representing an overlooked continuum of species, subspecies, and strains, each apparently having found its specific niche in the wide variety of habitats where anammox bacteria are encountered. Species microdiversity is particularly impressive for the marine representative Scalindua.[43][48][49][50][51][52] A question that remains to be investigated is which environmental factors determine species differentiation among anammox bacteria.

The sequence identities of the anammox 16S rRNA genes range from 87 to 99%, and phylogenetic analysis places them all within the phylum Planctomycetes,[53] which form the PVC superphylum together with Verrucomicrobia and Chlamydiae.[54] Within the Planctomycetes, anammox bacteria deeply branch as a monophyletic clade. Their phylogenetic position together with a broad range of specific physiological, cellular, and molecular traits give anammox bacteria their own order Brocadiales.[55]

Application

The application of the anammox process lies in the removal of ammonium in wastewater treatment and consists of two separate processes. The first step is partial nitrification (nitritation) of half of the ammonium to nitrite by ammonia oxidizing bacteria:

2NH4+ + 3O2 2NO2 + 4H+ + 2H2O

The resulting ammonium and nitrite are converted in the anammox process to dinitrogen gas and circa 15% nitrate (not shown) by anammox bacteria:

NH4+ + NO2 N2 + 2 H2O

Both processes can take place in 1 reactor where two guilds of bacteria form compact granules.[56][57]

For the enrichment of the anammox organisms a granular biomass or biofilm system seems to be especially suited in which the necessary sludge age of more than 20 days can be ensured. Possible reactors are sequencing batch reactors (SBR), moving bed reactors or gas-lift-loop reactors. The cost reduction compared to conventional nitrogen removal is considerable; the technique is still young but proven in several fullscale installations. The first full scale reactor intended for the application of anammox bacteria was built in the Netherlands in 2002.[58] In other wastewater treatment plants, such as the one in Germany (Hattingen), anammox activity is coincidentally observed though were not built for that purpose. As of 2006, there are three full scale processes in The Netherlands: one in a municipal wastewater treatment plant (in Rotterdam), and two on industrial effluent. One is a tannery, the other a potato processing plant.

Advantages

Conventional nitrogen removal from ammonium-rich wastewater is accomplished in two separate steps: nitrification, which is mediated by aerobic ammonia- and nitrite-oxidizing bacteria and denitrification carried out by denitrifiers, which reduce nitrate to N2 with the input of suitable electron donors. Aeration and input of organic substrates (typically methanol) show that these two processes are: (1) highly energy consuming, (2) associated with the production of excess sludge and (3) produce significant amounts of green-house gases such as CO2 and N2O and ozone-depleting NO.[59] Because anammox bacteria convert ammonium and nitrite directly to N2 anaerobically, this process does not require aeration and other electron donors. Nevertheless, oxygen is still required for the production of nitrite by ammonia-oxiding bacteria. However, in partial nitritation/anammox systems, oxygen demand is greatly reduced because only half of the ammonium needs to be oxidized to nitrite instead of full conversion to nitrate. The autotrophic nature of anammox bacteria and ammonia-oxidizing bacteria guarantee a low yield and thus less sludge production.[59] Additionally, anammox bacteria easily form stable self-aggregated biofilm (granules) allowing reliable operation of compact systems characterized by high biomass concentration and conversion rate up to 5–10 kg N m−3.[60] Overall, it has been shown that efficient application of the anammox process in wastewater treatment results in a cost reduction of up to 60%[61][62] as well as lower CO2 emissions.[59]

References

  1. Arrigo, R. A. Marine microorganisms and global nutrient cycles. Nature 437, 349–355 (2005)
  2. Strous, M. et al. Missing lithotroph identified as new planctomycete. Nature 400, 446–449 (1999)
  3. Jetten Michael Silvester Maria, Van Loosdrecht Marinus Corneli; Technische Universiteit Delft, patent WO9807664
  4. 1 2 3 Kartal, B. et al. How to make a living from anaerobic ammonium oxidation. FEMS Microbiology Reviews 37, 428-461 (2013)
  5. Reimann, Joachim; Jetten, Mike S.M.; Keltjens, Jan T. (2015). "Chapter 7 Metal Enzymes in "Impossible" Microorganisms Catalyzing the Anaerobic Oxidation of Ammonium and Methane". In Peter M.H. Kroneck and Martha E. Sosa Torres. Sustaining Life on Planet Earth: Metalloenzymes Mastering Dioxygen and Other Chewy Gases. Metal Ions in Life Sciences. 15. Springer. pp. 257–313. doi:10.1007/978-3-319-12415-5_7.
  6. Devol, A. H. et al. Nitrogen cycle: solution to a marine mystery. Nature 422(6932), 575-576 (2003)
  7. Jetten, M. S. M. et al. Biochemistry and molecular biology of anammox bacteria. Critical Reviews in Biochemistry and Molecular Biology 44(2-3), 65-84 (2009)
  8. Boumann H. A. et al. Biophysical properties of membrane lipids of anammox bacteria: I. Ladderane phospholipids form highly organized fluid membranes. Biochim Biophys Acta 1788(7), 1444-1451 (2009)
  9. "Pee power: Urine-loving bug churns out space fuel". Agence France Press. 2011-10-02. Retrieved 2011-10-03.
  10. 1 2 Strous, M., Kuenen, J.G., Jetten, M.S. 1999. Key Physiology of Anaerobic Ammonium Oxidation. App. Environ. Microb. (3248-3250)
  11. Yan J, Haaijer SCM, Op den Camp HJM, van Niftrik L, Stahl DA, Ko ̈nneke M, Rush D, Sinninghe Damste ́ JS, Hu YY, Jetten MSM (2012) Mimicking the oxygen minimum zones: stimulating interaction of aerobic archaeal and anaerobic bacterial ammonia oxidizers in a laboratory- scale model system. Environ Microbiol 14:3146–3158
  12. 1 2 3 Kartal B, Maalcke WJ, de Almeida NM, Cirpus I, Gloerich J, Geerts W, Op den Camp HJ, Harhangi HR, Janssen-Megens EM, Francoijs KJ, Stunnenberg HG, Keltjens JT, Jetten MS, Strous M. (2011). "Molecular mechanism of anaerobic ammonium oxidation". Nature. 479 (7371): 127–130. doi:10.1038/nature10453. PMID 21964329.
  13. Jaeschke et al. 2009. 16s rRNA gene and lipid biomarker evidence for anaerobic ammonium-oxidizing bacteria (anammox) in California and Nevada hot springs. FEMS Microbiol. Ecol. 343-350
  14. Byrne, N., Strous, M., Crepeau, V, et al. 2008. Presence and activity of anaerobic ammonium-oxidizing bacteria at deep-sea hydrothermal vents. The ISME Journal.
  15. Allgeier, R. J. et al. The anaerobic fermentation of lake deposits. International Review of Hydrobiology 26(5-6), 444-461 (1932)
  16. F. A. Richards. Anoxic basins and fjordsin. Chemical Oceanography, J.P. Ripley and G. Skirrow, Eds., pp 611-645, Academic Press, London, UK, 1965
  17. Arrigo, K. R. Marine microorganisms and global nutrient cycles. Nature 437(7057), 349-355 (2005)
  18. Broda, E. Two kinds of lithotrophs missing in nature. Zeitschrift fur Allgemeine Mikrobiologie 17(6), 491-493 (1977)
  19. Kuenen, J. G. Anammox bacteria: from discovery to application. Nature Reviews Microbiology 6(4), 320-326 (2008)
  20. A. A. van de Graaf, A. Mulder, H. Slijkhuis, L. A. Robertson, and J. G. Kuenen, “Anoxic ammonium oxidation,” in Proceedings of the 5th European Congress on Biotechnology, C. Christiansen, L. Munck, and J. Villadsen, Eds., pp. 338–391, Copenhagen, Denmark, 1990
  21. A. Mulder, A. A. Van De Graaf, L. A. Robertson, and J. G. Kuenen, “Anaerobic ammonium oxidation discovered in a denitrifying fluidized bed reactor,” FEMS Microbiology Ecology, vol. 16, no. 3, pp. 177–184, 1995
  22. A. A. Van de Graaf, A. Mulder, P. De Bruijn, M. S. M. Jetten, L. A. Robertson, and J. G. Kuenen, “Anaerobic oxidation of ammonium is a biologically mediated process,” Applied and Environmental Microbiology, vol. 61, no. 4, pp. 1246–1251, 1995
  23. M. Strous, J. J. Heijnen, J. G. Kuenen, and M. S. M. Jetten, “The sequencing batch reactor as a powerful tool for the study of slowly growing anaerobic ammonium-oxidizing microorganisms,” Applied Microbiology and Biotechnology, vol. 50, no. 5, pp. 589–596, 1998
  24. 1 2 M. Strous, J. A. Fuerst, E. H. M. Kramer et al., “Missing lithotroph identified as new planctomycete,” Nature, vol. 400, no. 6743, pp. 446–449, 1999
  25. J. G. Kuenen and M. S. M. Jetten, “Extraordinary anaerobic ammonium oxidising bacteria,” ASM News, vol. 67, pp. 456–463, 2001
  26. C. A. Francis, J. M. Beman, and M. M. M. Kuypers, “New processes and players in the nitrogen cycle: the microbial ecology of anaerobic and archaeal ammonia oxidation,” ISME Journal, vol. 1, no. 1, pp. 19–27, 2007
  27. B. Thamdrup and T. Dalsgaard, “Production of N2 through anaerobic ammonium oxidation coupled to nitrate reduction in marine sediments,” Applied and Environmental Microbiology, vol. 68, no. 3, pp. 1312–1318, 2002
  28. Van De Graaf, A. A. et al. Metabolic pathway of anaerobic ammonium oxidation on the basis of 15N studies in a fluidized bed reactor. Microbiology 143(7), 2415-2412 (1997)
  29. 1 2 Ni, S-Q. and Zhang, J. Anaerobic Ammonium Oxidation: From Laboratory to Full-Scale Application. BioMed Research International 2013, 1-10 (2013)
  30. Jetten, M. S. M. et al. The anaerobic oxidation of ammonium. FEMS Microbiology Reviews 22(5), 421-437 (1998)
  31. Schalk, H. et al. The anaerobic oxidation of hydrazine: a novel reaction in microbial nitrogen metabolism. FEMS Microbiology 158(1), 61-67 (1998)
  32. Dilworth M. J. and Eady R. R. Hydrazine is a product of dinitrogen reduction by the vanadium-nitrogenase from Azotobacter chroococcum. Biochemical Journal 277(2), 465-468 (1991)
  33. Kartal B, de Almeida NM, Maalcke WJ, Op den Camp HJ, Jetten MS, Keltjens JT (2013). "How to make a living from anaerobic ammonium oxidation.". FEMS Microbiol Rev. 37 (3): 428–461. doi:10.1111/1574-6976.12014. PMID 23210799.
  34. Dietl A, Ferousi C, Maalcke WJ, Menzel A, de Vries S, Keltjens JT, Jetten MS, Kartal B, Barends TR. "The inner workings of the hydrazine synthase multiprotein complex". Nature. doi:10.1038/nature15517. PMID 26479033.
  35. Hooper, A. B. et al. Enzymology of the oxidation of ammonia to nitrite by bacteria. Antonie van Leeuwenhoek 71(1-2), 59-67 (1997)
  36. 1 2 3 Strous, M. et al. Deciphering the evolution and metabolism of an anammox bacterium from a community genome. Nature 440(7085), 790-794 (2006)
  37. Kartal, B. et al. Candidatus 'Brocadia fulgida': an autofluorescent anaerobic ammonium oxidizing bacterium. FEMS Microbiol. Ecol. 63, 46-55 (2008)
  38. Oshiki, M. et al. Physiological characteristics of the anaerobic ammonium-oxidizing bacterium Candidatus 'Brocadia sinica'. Microbiology 157, 1706-1713 (2011)
  39. Kartal, B. et al. Candidatus "Anammoxoglobus propionicus" a new propionate oxidizing species of anaerobic ammonium oxidizing bacteria. Syst Appl Micrbiol 30, 39-49 (2007)
  40. Quan, Z. X. et al. Diversity of ammonium-oxidizing bacteria in a granular sludge anaerobic ammonium-oxidizing (anammox) reactor. Environ Microbiol 10, 3130-3139 (2008)
  41. Hu, B. L. et al. New anaerobic, ammonium-oxidizing community enriched from peat soil. Appl Environ Microbiol 77: 966–971 (2011)
  42. 1 2 Schmid, M. et al. Candidatus “Scalindua brodae”, sp. nov., Candidatus “Scalindua wagneri”, sp. nov., two new species of anaerobic ammonium oxidizing bacteria. Syst Appl Microbiol 26: 529–538. (2003)
  43. 1 2 Woebken, D. et al. A microdiversity study of anammox bacteria reveals a novel Candidatus Scalindua phylotype in marine oxygen minimum zones. Environ Microbiol 10: 3106–3119 (2008)
  44. Van de Vossenberg, J. et al. The metagenome of the marine anammox bacterium ‘Candidatus Scalindua profunda’ illustrates the versatility of this globally important nitrogen cycle bacterium. Environ Microbiol. doi:10.1111/j. 1462-2920, 2012.02774.x. [Epub ahead of print](2012)
  45. Schubert, C. J. et al. Anaerobic ammonium oxidation in a tropical freshwater system (Lake Tanganyika). Environ Microbiol 8: 1857–1863 (2006)
  46. Hamersley, M. R. et al. Water column anammox and denitrification in a temperate permanently stratified lake (Lake Rassnitzer, Germany). Syst Appl Microbiol 32: 571–582 (2009)
  47. Ligi, T. et al. The genetic potential of N2 emission via denitrification and ANAMMOX from the soils and sediments of a created riverine treatment wetland complex. Ecol. Eng. In Press (2015)
  48. Schmid, M. C. et al. Anaerobic ammonium-oxidizing bacteria in marine environments: widespread occurrence but low diversity (2007)
  49. Dang, H. et al. Environmental factors shape sediment anammox bacterial communities in hypernutrified Jiaozhou Bay, China. Appl Environ Microbiol 76: 7036–7047 (2010)
  50. Hong, Y. G. et al. Residence of habitat-specific anammox bacteria in the deep-sea subsurface sediments of the South China Sea: analyses of marker gene abundance with physical chemical parameters. Microb Ecol 62: 36–47 (2011a)
  51. Hong, Y. G. et al. Diversity and abundance of anammox bacterial community in the deep-ocean surface sediment from equatorial Pacific. Appl Microbiol Biotechnol 89: 1233–1241 (2011b)
  52. Li, M. et al. Spatial distribution and abundances of ammonia-oxidizing archaea (AOA) and ammonia-oxidizing bacteria (AOB) in mangrove sediments. Appl Microbiol Biotechnol 89: 1243–1254 (2011)
  53. Fuerst J. A. & Sagulenko E. Beyond the bacterium: planctomycetes challenge our concepts of microbial structure and function. Nat Rev Microbiol 9: 403–413 (2011)
  54. Wagner M & Horn M (2006) The Planctomycetes, Verrucomicrobia, Chlamydiae and sister phyla comprise a superphylum with biotechnological and medical relevance. Curr Opin Biotechnol 17: 241–249
  55. Jetten MSM, Op den Camp HJM, Kuenen JG & Strous M (2010) Description of the order Brocadiales. Bergey’s Manual of Systematic Bacteriology, Vol 4 (Krieg NR, Ludwig W, Whitman WB, Hedlund BP, Paster BJ, Staley JT, Ward N, Brown D & Parte A, eds), pp. 596–603. Springer, Heidelberg
  56. Kartal B, Kuenen JG, van Loosdrecht MC (2010). "Sewage treatment with anammox". Science. 328 (5979): 702–703. doi:10.1126/science.1185941. PMID 20448175.
  57. Knight, Helen (May 7, 2010). "Bugs will give us free power while cleaning our sewage". New Scientist. Retrieved May 2010. Check date values in: |access-date= (help)
  58. van der Star WRL, Abma WR, Blommers D, Mulder J-W, Tokutomi T, Strous M, Picioreanu C, Van Loosdrecht MCM (2007) Startup of reactors for anoxic ammonium oxidation: experiences from the first full-scale anammox reactor in Rotterdam. Water Res 41:4149–4163
  59. 1 2 3 Hu Z, Lotti T, Lotti T, de Kreuk M, Kleerebezem R, van Loosdrecht M, Kruit J, Jetten MSM, Kartal B (2013) Nitrogen removal by a nitritation-anammox bioreactor at low temperature. Appl Environ Microbiol. doi:10.1128/ AEM.03987-12
  60. van Loosdrecht MCM (2008) Innovative nitrogen removal. In: Henze M, van Loosdrecht MCM, Ekama GA, Brdjanovic D (eds) Biological wastewater treatment: principles, modelling and design. IWA Publishing, London, pp 139–155
  61. Siegrist H, Salzgeber D, Eugster J, Joss A (2008) Anammox brings WWTP closer to energy autarky due to increased biogas production and reduced aeration energy for N-removal. Water Sci Technol 57:383–388
  62. van Dongen U, Jetten MSM, van Loosdrecht MCM (2001) The SHARON((R))-Anammox((R)) process for treatment of ammonium rich wastewater. Water Sci Technol 44: 153–160
This article is issued from Wikipedia - version of the 11/10/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.