ArticlePDF AvailableLiterature Review

Anaerobic ammonium oxidation (anammox) in different natural ecosystems

Authors:

Abstract and Figures

Anammox (anaerobic ammonium oxidation), which is a reaction that oxidizes ammonium to dinitrogen gas using nitrite as the electron acceptor under anoxic conditions, was an important discovery in the nitrogen cycle. The reaction is mediated by a specialized group of planctomycete-like bacteria that were first discovered in man-made ecosystems. Subsequently, many studies have reported on the ubiquitous distribution of anammox bacteria in various natural habitats, including anoxic marine sediments and water columns, freshwater sediments and water columns, terrestrial ecosystems and some special ecosystems, such as petroleum reservoirs. Previous studies have estimated that the anammox process is responsible for 50% of the marine nitrogen loss. Recently, the anammox process was reported to account for 9-40% and 4-37% of the nitrogen loss in inland lakes and agricultural soils respectively. These findings indicate the great potential for the anammox process to occur in freshwater and terrestrial ecosystems. The distribution of different anammox bacteria and their contribution to nitrogen loss have been described in different natural habitats, demonstrating that the anammox process is strongly influenced by the local environmental conditions. The present mini-review summarizes the current knowledge of the ecological distribution of anammox bacteria, their contribution to nitrogen loss in various natural ecosystems and the effects of major influential factors on the anammox process.
Content may be subject to copyright.
ICoN2and the NCycle16 1811
Anaerobic ammonium oxidation (anammox) in
different natural ecosystems
Bao-lan Hu, Li-dong Shen, Xiang-yang Xu and Ping Zheng1
Department of Environmental Engineering, Zhejiang University, Hangzhou 310058, China
Abstract
Anammox (anaerobic ammonium oxidation), which is a reaction that oxidizes ammonium to dinitrogen
gas using nitrite as the electron acceptor under anoxic conditions, was an important discovery in the
nitrogen cycle. The reaction is mediated by a specialized group of planctomycete-like bacteria that were
first discovered in man-made ecosystems. Subsequently, many studies have reported on the ubiquitous
distribution of anammox bacteria in various natural habitats, including anoxic marine sediments and water
columns, freshwater sediments and water columns, terrestrial ecosystems and some special ecosystems,
such as petroleum reservoirs. Previous studies have estimated that the anammox process is responsible
for 50% of the marine nitrogen loss. Recently, the anammox process was reported to account for 9–
40% and 4–37% of the nitrogen loss in inland lakes and agricultural soils respectively. These findings
indicate the great potential for the anammox process to occur in freshwater and terrestrial ecosystems.
The distribution of different anammox bacteria and their contribution to nitrogen loss have been described
in different natural habitats, demonstrating that the anammox process is strongly influenced by the local
environmental conditions. The present mini-review summarizes the current knowledge of the ecological
distribution of anammox bacteria, their contribution to nitrogen loss in various natural ecosystems and the
effects of major influential factors on the anammox process.
Introduction
The process of anammox (anaerobic ammonium oxidation),
which refers to the oxidation of ammonium coupled with
the reduction of nitrite under anoxic conditions, has been
predicted to be a more thermodynamically favourable
process than aerobic ammonium oxidation [1], yet the
anammox process was not discovered until nearly 20 years
later in a wastewater-treatment plant in The Netherlands
[2]. The process is mediated by anammox bacteria, a
deep-branching monophyletic group of bacteria within
the phylum Planctomycetes. To date, anammox bacteria
have not been cultured in the laboratory; however, at
least five genera and 13 species have been identified
using culture-independent molecular techniques. These taxa
include the following: Candidatus ‘Brocadia’ (Ca. ‘Brocadia
anammoxidans’ [3], Ca. ‘Brocadia fulgida’ [4] and Ca.
‘Brocadia sinica’ [5]); Candidatus ‘Kuenenia stuttgartiensis’
[6]; Candidatus ‘Scalindua’ (Ca. ‘Scalindua brodae’ [7], Ca.
‘Scalindua wagneri’ [7], Ca. ‘Scalindua sorokinii’ [8], Ca.
‘Scalindua arabica’ [9], Ca. ‘Scalindua sinooilfield’ [10] and
Ca. ‘Scalindua zhenghei’ [11]); Candidatus ‘Anammoxo-
globus’ (Ca. ‘Anammoxoglobus propionicus’ [12] and Ca.
Key words: anaerobic ammonium oxidation (anammox), ecological distribution, environmental
conditions, natural ecosystem, nitrogen loss.
Abbreviations used: anammox, anaerobic ammonium oxidation; DNRA, dissimilatory nitrite
reduction to ammonium; OMZ, oxygen minimum zone.
1To whom correspondence should be addressed (email pzheng@zju.edu.cn).
‘Anammoxoglobus sulfate’ [13]) and Candidatus ‘Jettenia
asiatica’ [14]. Henceforth, these genera will be referred to
simply as Brocadia,Kuenenia, Scalindua,Anammoxoglobus
and Jettenia respectively.
Anammox bacteria have been detected in various natural
habitats, such as anoxic marine sediments [15–17] and
water columns [18–20], freshwater sediments [21] and water
columns [22], terrestrial ecosystems [23,24] and some special
ecosystems, such as petroleum reservoirs [10]. All of the
available evidence indicates that the anammox process is
critically important in the marine nitrogen cycle, and the
relative contribution of the anammox process to the total
production of dinitrogen gas (N2) has been estimated to be
50% in the ocean [25]. In addition to marine environments,
anammox activity has been detected in natural freshwater
and terrestrial environments [22,24], indicating that the
anammox process may play an even more significant role
in the global nitrogen cycle than previously thought.
The ecological distribution of anammox bacteria and their
contribution to the nitrogen loss in natural ecosystems are
influenced by local environmental conditions: the organic
content [26,27], NOxconcentration [28], environmental
stability [29], salinity [16,17] and temperature [22] have been
described as key influencing factors. The present mini-review
summarizes the recent findings concerning the distribution
of anammox bacteria, their contribution to N2production in
various natural habitats and the major factors influencing the
anammox process.
Biochem. Soc. Trans. (2011) 39, 1811–1816; doi:10.1042/BST20110711 C
The Authors Journal compilation C
2011 Biochemical Society
1812 Biochemical Society Transactions (2011) Volume 39, part 6
Table 1 Ecological distribution of anammox bacteria and their contribution to N2production in various natural habitats
ND, no data.
Location 16S rRNA affiliation Contribution (%) Reference(s)
Marine sediments
Skagerak (North Sea) ND 24–67 [26]
Thames Estuary (U.K.) ND 1–8 [27]
Randers Fjord (Denmark) Scalindua 5–24 [28]
Greenland Sea (Greenland) Scalindua >19 [15]
Disko Bay (Greenland) Scalindua ND [15]
Cascadia Basin (U.S.A) ND 40–42 [32]
Gullmarsfjorden (Sweden) ND 23–47 [14,50]
Barents Sea (Greenland) Scalindua ND [15]
Golfo Dulce (Costa Rica) Scalindua ND [15]
Young Sound (Greenland) Scalindua ND [15]
North Sea (North of the Friesian Front) Scalindua ND [15]
Yodo Estuary (Japan) Scalindua,Brocadia,Kuenenia 1–2 [30]
Chesapeake Bay (U.S.A.) Scalindua 0–22 [16]
Cape Fear River Estuary (U.S.A.) Scalindua,Brocadia,Kuenenia,Jettenia 4–17 [17]
North Atlantic ND 33–65 [33]
Jiaozhou Bay (China) Scalindua ND [51]
Mai Po Marshes (Hong Kong) Scalindua,Kuenenia ND [31]
South China Sea (China) Scalindua ND [11]
Equatorial Pacific Scalindua ND [52]
Haiphong (Vietnam) Scalindua,Brocadia,Kuenenia <2 [53]
Marine water columns
Golfo Dulce (Costa Rica) Scalindua 19–35 [35]
Black Sea Scalindua 10–15 [8,38]
Namibian waters Scalindua Approximately 100 [18,39]
Peruvian waters Scalindua Approximately 100 [20]
Northern Chile Scalindua Approximately 100 [19]
Arabian Sea Scalindua <13 [36,37]
Freshwater ecosystems
Lake Tanganyika (Kigoma) Scalindua 9–13 [22]
Wintergreen Lake (U.K.) Scalindua ND [42]
Xinyi River (China) Brocadia ND [21]
Groundwater (Canada) Scalindua,Brocadia,Kuenenia,Jettenia 18–36 [41]
Lake Kitaura (Japan) Brocadia,Kuenenia,Anammoxoglobus <40 [40]
Terrestrial ecosystems
Various terrestrial habitats (Switzerland) Scalindua,Brocadia,Kuenenia,Jettenia ND [23]
Peat soil (The Netherlands) Brocadia,Kuenenia,Jettenia ND [43]
Paddy soil (Southern China) Brocadia,Kuenenia,Jettenia,Anammoxoglobus 4–37 [24]
Special ecosystems
Hot springs (U.S.A.) Brocadia,Kuenenia ND [46]
Hydrothermal vents (Mid-Atlantic Ridge) Scalindua,Kuenenia ND [47]
Marine sponge (Norway) Scalindua 0–1 [48]
Marine sponge (U.S.A.) Brocadia ND [49]
Petroleum reservoirs (China) Scalindua,Brocadia,Kuenenia,Jettenia ND [10]
Anammox in marine ecosystems
Anoxic sediments
Dalsgaard and Thamdrup [26] first detected anammox
activities in the sediments of two continental shelf sites
of the Skagerrak in the Danish Belt seaway. Subsequently,
anammox activities were detected in various marine sediments
(Table 1). Although organisms belonging to the Brocadia and
Kuenenia genera were found in some coastal and estuarine
sediments [17,30,31], the majority of anammox bacteria in
marine sediments were affiliated with the Scalindua genus
and showed surprisingly low diversity [8,9,15,18,20]. The
C
The Authors Journal compilation C
2011 Biochemical Society
ICoN2and the NCycle16 1813
contribution of the anammox process to regional nitrogen
loss is highly variable in marine sediments (Table 1); indeed,
the reported amounts of N2production by the anammox
process have ranged from 20 to 80% in shelf and deep
sediments [26,32,33]. However, it has also been reported
that the anammox process contributed to less than 20%
of the N2production in shallow coastal and estuarine
sediments [16,17,27]. Therefore the relative contribution of
anammox to nitrogen loss is positively correlated with water
depth. Furthermore, the water depth and organic content are
negatively correlated because a larger fraction of the organic
matter is mineralized during transport to the sediment within
a deep water column. Thus the organic content of deep
sediment is low [29]. The higher electron donor (organic
matter) availability in organic-rich shallow sediments leads
to increased competition for NO2between denitrifiers and
anammox bacteria. Yet anammox bacteria are slow-growing
organisms [25] and are less competitive for NO2than the
denitrifiers in organic-rich shallow sediments. In electron-
donor-limited deep sediments, nitrate-reducing organisms
produce more NO2for the anammox process owing to
the shortage of sufficient organic matter. In addition, the
relative importance of the anammox process is directly
related to the availability of NO3[16,28]: a higher NO3
concentration leads to a higher nitrate reduction rate and a
greater release of NO2for anammox. The stability of the
environment may also be important for regulating the relative
importance of the anammox process [29], an effect that is
due to the low growth rate of the anammox bacteria. The
anammox process is only significant in stable environments
that allow a prolonged period for the bacterial population to
develop [29]. Recent studies have shown that the distribution
of anammox bacteria and their contribution to nitrogen
loss is correlated with the salinity [17,31]. Scalindua is
the most abundant genus in environments with higher salt
concentrations and the most halotolerant of the anammox
genera [34]. In addition, previous studies showed that higher
contributions of the anammox process to nitrogen loss were
found in environments with higher salinity [17] and that the
abundances of Brocadia and Kuenenia species were negatively
correlated with the salinity [17,31].
Anoxic water columns
In 2003, two parallel studies demonstrated that anammox
bacteria were responsible for a substantial portion of the
nitrogen loss in the anoxic water columns of the Black Sea
and Golfo Dulce, in which 10–35% of the total nitrogen
loss was attributable to the anammox process [8,35]. Recent
studies indicated that the anammox process was responsible
for a greater percentage (more than 50%) of the nitrogen
loss in marine water columns, especially in OMZs (oxygen
minimum zones) [18–20], and anammox bacteria were
reported to be the dominant N2producer in the OMZs of
Namibia, Chile and Peru [18–20]. Denitrification was initially
reported as the dominant driver of nitrogen loss in the OMZ
of the Arabian Sea [36]; however, a direct link between the
DNRA (dissimilatory nitrite reduction to ammonium) and
the anammox process through the 15NO2signal-mediated
production of 15N15 N was easily mistaken as a signature for
denitrification [37]. This finding indicates that anammox–
DNRA coupling, rather than denitrification, was responsible
for the massive nitrogen loss in the OMZ of the Arabian Sea.
As observed in marine sediments, the dominant anammox
species in marine water columns is closely related to the
Scalindua genus (Table 1). However, the factors influencing
the anammox process in marine water columns are not
well known. Recent findings have indicated that aerobic
ammonium oxidizers provided NO2and created anoxic
microenvironments for the anammox bacteria through the
consumption of oxygen in the OMZs of the Black Sea and
Namibian Sea [38,39]. Therefore anammox bacteria may be
dependent on the activity of aerobic ammonium oxidizers
under oxygen-limited conditions. The NH4+in the water
column is released through the mineralization of organic
matter by both denitrifiers and DNRA organisms [20,37].
The availability of organic matter also acts as an important
factor controlling the release of NO2from NO3for the
anammox process. Lam et al. [20] showed that, in the OMZ
of Peru, anammox bacteria obtained at least 67% of their
NO2from nitrate reduction using organic matter as the
electron donors, whereas less than 33% of the NO2was
derived from aerobic ammonium oxidation. Therefore the
availability of organic matter and its mineralization rate are
important factors influencing the anammox process in marine
water columns.
Anammox in freshwater ecosystems
Although there is growing evidence for the widespread occur-
rence of anammox bacteria in marine environments, evidence
for the existence of anammox bacteria in natural freshwater
habitats is limited (Table 1). The first direct evidence of
anammox bacteria in freshwater ecosystems was provided
by Schubert et al. [22], who have reported that the anammox
process contributed to 9–13% of the N2production and was
responsible for 0.2 Tg of the fixed inorganic nitrogen loss
per year in Lake Tanganyika, the second largest freshwater
body in the world. Temperature has been identified as
an important factor influencing anammox activity in this
freshwater ecosystem [22]. A higher relative contribution
of the anammox process to nitrogen loss was reported in
a eutrophic freshwater lake, Lake Kitaura, in which up to
40% of the N2production was correlated to anammox
activity [40]. A positive correlation between the NO3
concentration and the relative importance of the anammox
process was found in this lake, suggesting that the NO3
concentration was the key factor for the development
of anammox bacterial populations and their activities in
this freshwater habitat [40]. Anammox activities were also
recently detected in ammonium-contaminated groundwater
sites in Canada, where anammox activity was responsible for
18–36% of the nitrogen loss [41]. Because the highest relative
contribution of the anammox process to nitrogen loss was
C
The Authors Journal compilation C
2011 Biochemical Society
1814 Biochemical Society Transactions (2011) Volume 39, part 6
found in sites with high concentrations of dissolved organic
matter, NH4+and NO3, researchers hypothesized that
these factors influenced anammox activity [41]. In addition,
anammox bacteria have also been detected in the sediments
of the Xinyi River [21] and Lake Wintergreen [42]. Unlike
marine environments, in which the anammox communities
were exclusively dominated by the Scalindua genus, different
dominant anammox species have been observed in freshwater
ecosystems. In Lake Tanganyika and Lake Wintergreen,
the Scalindua genus was the dominant anammox species
[22,42], whereas in the sediments of the Xinyi River, Lake
Kitaura and Canada’s groundwater environments, Brocadia
organisms were the most common representatives [21,40,41].
These findings suggest that freshwater environments are more
favourable for the growth of different groups of anammox
bacteria than marine environments.
Anammox in terrestrial ecosystems
Because anammox depends on the concomitant presence of
oxidized and reduced inorganic nitrogen compounds under
anoxic conditions, the oxic/anoxic interfaces in terrestrial
areas may provide suitable habitats for anammox bacteria
[23]. However, little is known about the presence of anammox
bacteria in terrestrial ecosystems (Table 1). Humbert et al. [23]
first reported the distribution of diverse anammox bacteria in
agricultural and permafrost soils, which contained Brocadia,
Kuenenia,Scalindua and Jettenia. Four different anammox
genera were also simultaneously detected in fertilized paddy
soil in Southern China [24], and three distinct anammox
genera were found together in nitrogen-loaded peat soil [43].
These results suggest a higher diversity of anammox bacteria
in terrestrial ecosystems than usually observed in aquatic
habitats and may be mediated by the larger variety of suitable
niches for anammox bacteria in terrestrial habitats [23].
The dominant anammox species in the reported terrestrial
environments were affiliated with the Brocadia and Kuenenia
genera [23,24,44], demonstrating that Brocadia and Kuenenia
organisms exhibited a better adaptation capacity in these
ecosystems [23]. Brocadia and Kuenenia organisms possess
a more mixotrophic metabolism than previously thought
[4,12,44]. These microbes use ferrous iron and a variety of
organic compounds, such as formate, acetate, propionate
and methylamines, as electron donors [4,12,44] and employ
ferric iron and manganese oxides as electron acceptors
during their metabolic activities [44]. Therefore the versatile
metabolism of Brocadia and Kuenenia organisms may be
the main reason for their better adaptation in heterogeneous
terrestrial environments. Anammox activity was reported
to account for 4–37% of the soil N2production in a
fertilized paddy soil, and the substantial contribution of
anammox to nitrogen loss in the paddy field was related
to the high concentrations of NH4+that were introduced
by the fertilization [24]. Moreover, Hu et al. [43] obtained
an enriched anammox culture from a nitrogen-loaded peat
soil and showed a significant amount of anammox activity.
In controlled environments, the presence of slowly released
organic matter (e.g. humic acids) was identified as an
important factor that influenced the distribution of anammox
bacteria (Brocadia and Jettenia) in peat soil [43].
Anammox in special ecosystems
Anammox bacteria are active at 6–43C, with an optimal
temperature of 35C in laboratory bioreactors [45]. However,
it was recently observed that this process also occurred at
52C in hot springs [46], 72C in petroleum reservoirs [10]
and even at 85C in hydrothermal vents [47]. The dominant
anammox species in these high-temperature habitats were
Brocadia or Kuenenia, but not Scalindua (Table 1). This result
is likely to be due to the higher optimal growth temperatures
of Brocadia and Kuenenia organisms (35C in bioreactors)
compared with that of Scalindua organisms (12–15Cin
marine ecosystems) [46]. The production of long-chain
fatty acids in anammox bacteria at elevated temperatures
mediates their adaption to high-temperature environments
[46]. The versatile metabolism of some anammox species
(such as Brocadia and Kuenenia) is critical for survival in
high-temperature habitats [10]. Byrne et al. [47] detected
the activity of anammox bacteria in a hydrothermal vent,
indicating that anammox may play an important role in the
nitrogen cycle in thermophilic and mesophilic environments.
Anammox bacteria were also detected in some marine
sponges, which are ancient animals [48,49]. The majority of
anammox sequences found in marine sponges were affiliated
with the Scalindua or Brocadia genus. Hoffmann et al.
[48] detected the activity of anammox bacteria in marine
sponges and showed that anammox activity contributed to
1.25% of the N2production; in addition, the nitrification
and denitrification processes were also found in the marine
sponges, and the related and complex interactions of these
nitrogen-cycling processes were mainly controlled by the
metabolic waste products (e.g. NH4+and organic matter) of
the sponge host. Therefore the anammox process in marine
sponges was largely influenced by the amount of NH4+and
organic matter produced by the sponge host.
Conclusions and outlook
With the current knowledge of the anammox process in
natural ecosystems, it is clear that anammox bacteria have
a widespread distribution in various natural habitats. Among
the anammox bacteria that have been described to date,
Scalindua organisms appear to be the most widespread
anammox species in natural ecosystems. However, Brocadia
and Kuenenia species appear to be more likely to live in
freshwater and terrestrial ecosystems. The published data
have indicated that anammox activity was responsible for
50% of the marine nitrogen loss and may also play an
important role in the nitrogen cycle of freshwater and
terrestrial ecosystems. In contrast with marine ecosystems,
the detection of anammox bacteria, the measurement of
anammox activity and elucidation of environmental factors
influencing the anammox process are still lacking for
C
The Authors Journal compilation C
2011 Biochemical Society
ICoN2and the NCycle16 1815
freshwater and terrestrial environments. Therefore more
studies on the anammox process in a broader range of
freshwater and terrestrial habitats are required to obtain
a better comprehension of the ecological distribution of
anammox bacteria and their potential importance in marine
ecosystems as natural ecosystems. In addition, further study
will enhance our understanding of the influential factors that
control anammox activity in natural environments.
Funding
Our research was supported by the National Hi-Tech Research
and Development Program of China (863) [grant number
2006AA06Z332] and the Fundamental Research Funds for the Central
Universities [grant number 2010QNA6017].
References
1 Broda, E. (1977) Two kinds of lithotrophs missing in nature. Z. Allg.
Microbiol. 17, 491–493
2 Mulder, A., van de Graaf, A.A., Robertson, L.A. and Kuenen, J.G. (1995)
Anaerobic ammonium oxidation discovered in a denitrifying fluidized
bed reactor. FEMS Microbiol. Ecol. 16, 177–184
3 Strous, M., Fuerst, J.A., Kramer, E.H.M., Logemann, S., Muyzer, G., van de
Pas-Schoonen, K.T., Webb, R.I., Kuenen, J.G. and Jetten, M.S.M. (1999)
Missing lithotroph identified as new planctomycete. Nature 400,
446–449
4 Kartal, B., van Niftrik, L.A., Rattray, J., van de Vossenberg, J., Schmid,
M.C., Damst ´e, J.S.S., Jetten, M.S.M. and Strous, M. (2008) Candidatus
‘Brocadia fulgida’: an autofluorescent anaerobic ammonium oxidizing
bacterium. FEMS Microbiol. Ecol. 63, 46–55
5 Hu, B.L., Zheng, P., Tang, C.J., Chen, J.W., van der Biezen, E., Zhang, L., Ni,
B.J., Jetten, M.S.M., Yan, J., Yu, H.Q. and Kartal, B. (2010) Identification
and quantification of anammox bacteria in eight nitrogen removal
reactors. Water Res. 44, 5014–5020
6 Schmid, M.C., Twachtmann, U., Klein, M., Strous, M., Juretschko, S., Jetten,
M.S.M., Metzger, J.W., Schleifer, K.H. and Wagner, M. (2000) Molecular
evidence for genus level diversity of bacteria capable of catalyzing
anaerobic ammonium oxidation. Syst. Appl. Microbiol. 23, 93–106
7 Schmid, M.C., Walsh, K., Webb, R.I., Rijpstra, W.I., van de Pas-Schoonen,
K., Verbruggen, M.J., Hill, T., Moffett, B., Fuerst, J., Schouten, S. et al.
(2003) Candidatus ‘Scalindua brodae’, sp. nov., Candidatus ‘Scalindua
wagneri’, sp. nov., two new species of anaerobic ammonium oxidizing
bacteria. Syst. Appl. Microbiol. 26, 529–538
8 Kuypers, M.M.M., Sliekers, A.O., Lavik, G., Schmid, M.C., Jørgensen, B.B.,
Kuenen, J.G., Damst ´e, J.S.S., Strous, M. and Jetten, M.S.M. (2003)
Anaerobic ammonium oxidation by anammox bacteria in the Black Sea.
Nature 422, 608–611
9 Woebken, D., Lam, P., Kuypers, M.M.M., Naqvi, S.W., Kartal, B., Strous,
M., Jetten, M.S.M., Fuchs, B.M. and Amann, R. (2008) A microdiversity
study of anammox bacteria reveals a novel Candidatus Scalindua
phylotype in marine oxygen minimum zones. Environ. Microbiol. 10,
3106–3119
10 Li, H., Chen, S., Mu, B.Z. and Gu, J.D. (2010) Molecular detection of
anaerobic ammonium-oxidizing (anammox) bacteria in
high-temperature petroleum reservoirs. Microb. Ecol. 60, 771–783
11 Hong, Y.G., Li, M., Cao, H.L. and Gu, J.D. (2011) 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
12 Kartal, B., Rattray, J., van Niftrik, L.A., van de Vossenberg, J., Schmid,
M.C., Webb, R.I., Schouten, S., Fuerst, J.A., Damst ´e, J.S.S., Jetten, M.S.M.
and Strous, M. (2007) Candidatus ‘Anammoxoglobus propionicus’ a new
propionate oxidizing species of anaerobic ammonium oxidizing bacteria.
Syst. Appl. Microbiol. 30, 39–49
13 Liu, S.T., Yang, F.L., Gong, Z, Meng, F.F., Chen, H.H., Quan, Z.X. and
Furukawa, K. (2008) Application of anaerobic ammonium-oxidizing
consortium to achieve completely autotrophic ammonium and sulfate
removal. Biores. Technol. 99, 6817–6825
14 Quan, Z.X., Rhee, S.K., Zuo, J.E., Yang, Y., Bae, J.W., Park, J.R., Lee, S.T. and
Park, Y.H. (2008) Diversity of ammonium-oxidizing bacteria in a granular
sludge anaerobic ammonium-oxidizing (anammox) reactor. Environ.
Microbiol. 10, 3130–3139
15 Schmid, M.C., Risgaard-Petersen, N., van de Vossenberg, J., Kuypers,
M.M.M., Lavik, G., Petersen, J., Hulth, S., Thamdrup, B., Canfield, D.,
Dalsgaard, T. et al. (2007) Anaerobic ammonium-oxidizing bacteria in
marine environments: widespread occurrence but low diversity. Environ.
Microbiol. 9, 1476–1484
16 Rich, J.J., Dale, O.R., Song, B. and Ward, B.B. (2008) Anaerobic
ammonium oxidation (Anammox) in Chesapeake Bay sediments.
Microb. Ecol. 55, 311–320
17 Dale, O.R., Tobias, C.R. and Song, B. (2009) Biogeographical distribution
of diverse anaerobic ammonium oxidizing (anammox) bacteria in Cape
Fear River Estuary. Environ. Microbiol. 11, 1194–1207
18 Kuypers, M.M.M., Lavik, G., W ¨obken, D., Schmid, M.C., Fuchs, B.M. and
Amann, R. (2005) Massive nitrogen loss from the Benguela upwelling
system through anaerobic ammonium oxidation. Proc. Natl. Acad. Sci.
U.S.A. 102, 6478–6483
19 Thamdrup, B., Dalsgaard, T., Jensen, M.M., Ulloa, O., Farias, L. and
Escribano, R. (2006) Anaerobic ammonium oxidation in the
oxygen-deficient waters off northern Chile. Limnol. Oceanogr. 51,
2145–2156
20 Lam, P., Lavik, G., Jensen, M.M., van de Vossenberg, J., Schmid, M.C.,
Woebken, D., Guti ´errez, D., Amann, R., Jetten, M.S.M. and Kuypers,
M.M.M. (2009) Revising the nitrogen cycle in the Peruvian oxygen
minimum zone. Proc. Natl. Acad. Sci. U.S.A. 106, 4752–4757
21 Zhang, Y., Ruan, X.H., Op den Camp, H.J.M., Smits, T.J.M., Jetten, M.S.M.
and Schmid, M.C. (2007) Diversity and abundance of aerobic and
anaerobic ammonium-oxidizing bacteria in freshwater sediments of the
Xinyi River (China). Environ. Microbiol. 9, 2375–2382
22 Schubert, C.J., Durisch-Kaiser, E., Wehrli, B., Thamdrup, B., Lam, P. and
Kuypers, M.M.M. (2006) Anaerobic ammonium oxidation in a tropical
freshwater system (Lake Tanganyika). Environ. Microbiol. 8, 1857–1863
23 Humbert, S., Tarnawski, S., Fromin, N., Mallet, M.P., Aragno, M. and
Zopfi, J. (2010) Molecular detection of anammox bacteria in terrestrial
ecosystems: distribution and diversity. ISME J. 4, 450–454
24 Zhu, G.B, Wang, S., Wang, Y., Wang, C., Risgaard-Petersen, N., Jetten,
M.S.M. and Yin, C.Q. (2011) Anaerobic ammonia oxidation in a fertilized
paddy soil. ISME J., doi:10.1038/ismej.2011.63
25 Jetten, M.S.M., van Niftrik, L.A., Strous, M., Kartal, B., Keltjens, J.T. and Op
den Camp, H.J.M. (2009) Biochemistry and molecular biology of
anammox bacteria. Crit. Rev. Biochem. Mol. Biol. 44, 65–84
26 Dalsgaard, T. and Thamdrup, B. (2002) Production of N2through
anaerobic ammonium oxidation coupled to nitrate reduction in marine
sediments. Appl. Environ. Microbiol. 68, 1312–1318
27 Trimmer, M., Nicholls, J.C. and Deflandre, B. (2003) Anaerobic
ammonium oxidation measured in sediments along the Thames Estuary,
United Kingdom. Appl. Environ. Microbiol. 69, 6447–6454
28 Risgaard-Petersen, N., Meyer, R.L., Schmid, M.C., Jetten, M.S.M.,
Enrich-Prast, A., Rysgaard, S. and Revsbech, N.P. (2004) Anaerobic
ammonia oxidation in an estuarine sediment. Aquat. Microb. Ecol. 36,
293–304
29 Dalsgaard, T., Thamdrup, B. and Canfield, D.E. (2005) Anaerobic
ammonium oxidation (anammox) in the marine environment. Res.
Microbiol. 156, 457–464
30 Amano, T., Yoshinaga, I., Okada, K., Yamagishi, T., Ueda, S., Obuchi, A.,
Sako, Y. and Suwa, Y. (2007) Detection of anammox activity and
diversity of anammox bacteria-related 16S rRNA genes in coastal marine
sediment in Japan. Microbes Environ. 22, 232–242
31 Li, M., Cao, H.L., Hong, Y.G. and Gu, J.D. (2011) Seasonal dynamics of
anammox bacteria in estuarial sediment of the Mai Po Nature Reserve
revealed by analyzing the 16S rRNA and hydrazine oxidoreductase (hzo)
genes. Microbes Environ. 26, 15–22
32 Engstr ¨om, P., Penton, C.R. and Devol, A.H. (2009) Anaerobic ammonium
oxidation in deep-sea sediments off the Washington margin. Limnol.
Oceanogr. 54, 1643–1652
C
The Authors Journal compilation C
2011 Biochemical Society
1816 Biochemical Society Transactions (2011) Volume 39, part 6
33 Trimmer, M. and Nicholls, J.C. (2009) Production of nitrogen gas via
anammox and denitrification in intact sediment cores along a
continental shelf to slope transect in the North Atlantic. Limnol.
Oceanogr. 54, 577–589
34 Terada, A., Zhou, S. and Hosomi, M. (2011) Presence and detection of
anaerobic ammonium-oxidizing (anammox) bacteria and appraisal of
anammox process for high-strength nitrogenous wastewater treatment.
Clean Technol. Environ. Policy, doi:10.1007/s10098-011-0355-3
35 Dalsgaard, T., Canfield, D.E., Petersen, J., Thamdrup, B. and
Acu ˜na-Gonz ´alez, J. (2003) N2production by the anammox reaction in the
anoxic water column of Golfo Dulce, Costa Rica. Nature 422, 606–608
36 Ward, B.B., Devol, A.H., Rich, J.J., Chang, B.X., Bulow, S.E., Naik, H.,
Pratihary, A. and Jayakumar, A. (2009) Denitrification as the dominant
nitrogen loss process in the Arabian Sea. Nature 461, 78–81
37 Jensen, M.M., Lam, P., Revsbech, N.P., Nagel, B., Gaye, B., Jetten, M.S.M.
and Kuypers, M.M.M. (2011) Intensive nitrogen loss over the Omani
Shelf due to anammox coupled with dissimilatory nitrite reduction to
ammonium. ISME J. 5, 1660–1670
38 Lam, P., Jensen, M.M., Lavik, G., McGinnis, D.F., M ¨uller, B., Schubert, C.J.,
Amann, R., Thamdrup, B. and Kuypers, M.M.M. (2007) Linking
crenarchaeal and bacterial nitrification to anammox in the Black Sea.
Proc. Natl. Acad. Sci. U.S.A. 104, 7104–7109
39 Woebken, D., Fuchs, B.M., Kuypers, M.M.M. and Amann, R. (2007)
Potential interactions of particle-associated anammox bacteria with
bacterial and archaeal partners in the Namibian upwelling system. Appl.
Environ. Microbiol. 73, 4648–4657
40 Yoshinaga, I., Amano, T., Yamagishi, T., Okada, K., Ueda, S., Sako, Y. and
Suwa, Y. (2011) Distribution and diversity of anaerobic ammonium
oxidation (anammox) bacteria in the sediment of a eutrophic freshwater
lake, Lake Kitaura, Japan. Microbes Environ. 26, 189–197
41 Moore, T.A., Xing, Y.P., Lazenby, B., Lynch, M.D.J., Schiff, S., Robertson,
W.D., Timlin, R., Lanza, S., Ryan, M.C., Aravena, R. et al. (2011)
Prevalence of anaerobic ammonium-oxidizing bacteria in contaminated
groundwater. Environ. Sci. Technol. 45, 7217–7225
42 Penton, C.R., Devol, A.H. and Tiedje, J.M. (2006) Molecular evidence for
the broad distribution of anaerobic ammonium-oxidizing bacteria in
freshwater and marine sediments. Appl. Environ. Microbiol. 72,
6829–6832
43 Hu, B.L., Rush, D., van der Biezen, E., Zheng, P., van Mullekom, M.,
Schouten, S., Damst ´e, J.S.S., Smolders, A.J.P., Jetten, M.S.M. and Kartal, B.
(2011) New anaerobic, ammonium-oxidizing community enriched from
peat soil. Appl. Environ. Microbiol. 77, 966–971
44 Strous, M., Pelletier, E., Mangenot, S., Rattei, T., Lehner, A., Taylor, M.W.,
Horn, M., Daims, H., Bartol-Mavel, D., Wincker, P. et al. (2006)
Deciphering the evolution and metabolism of an anammox bacterium
from a community genome. Nature 440, 790–794
45 Kuenen, J.G. and Jetten, M.S.M. (2001) Extraordinary anaerobic
ammonium oxidizing bacteria. ASM News 67, 456
46 Jaeschke, A., Op den Camp, H.J.M., Harhangi, H.R., Klimiuk, A., Hopmans,
E.C., Jetten, M.S.M., Schouten, S. and Damst ´e, J.S.S. (2009) 16S rRNA
gene and lipid biomarker evidence for anaerobic ammonium-oxidizing
bacteria (anammox) in California and Nevada hot springs. FEMS
Microbiol. Ecol. 67, 343–350
47 Byrne, N., Strous, M., Cr ´epeau, V., Kartal, B., Birrien, J.L., Schmid, M.C.,
Lesongeur, F. and Godfroy, A. (2009) Presence and activity of anaerobic
ammonium-oxidizing bacteria at deep-sea hydrothermal vents. ISME J.
3, 117–123
48 Hoffmann, F., Radax, R., Woebken, D., Holtappels, M., Lavik, G., Rapp,
H.T., Schl ¨appy, M.L., Schleper, C. and Kuypers, M.M.M. (2009) Complex
nitrogen cycling in the sponge Geodia barrette.Environ.Microbiol.11,
2228–2243
49 Mohamed, N.M., Saito, K., Tal, Y. and Hill, R.T. (2010) Diversity of aerobic
and anaerobic ammonia-oxidizing bacteria in marine sponges. ISME J. 4,
38–48
50 Brandsma, J., van de Vossenberg, J., Risgaard-Petersen, N., Schmid, M.C.,
Engstr ¨om, P., Eurenius, K., Hulth, S., Jaeschke, A., Abbas, B., Hopmans,
E.C. et al. (2011) A multi-proxy study of anaerobic ammonium oxidation
in marine sediments of the Gullmar Fjord, Sweden. Environ. Microbiol.
Rep. 3, 360–366
51 Dang, H.Y., Chen, R.P., Wang, L., Guo, L.Z., Chen, P.P., Tang, Z.W., Tian, F.,
Li, S.Z. and Klotz, M.G. (2010) Environmental factors shape sediment
anammox bacterial communities in hypernutrified Jiaozhou Bay, China.
Appl. Environ. Microbiol. 76, 7036–7047
52 Hong, Y.G., Yin, B. and Zheng, T.L. (2011) Diversity and abundance of
anammox bacterial community in the deep-ocean surface sediment
from equatorial Pacific. Appl. Microbiol. Biotechnol. 89, 1233–1241
53 Amano, T., Yoshinaga, I., Yamagishi, T., Thuoc, C.V., Thu, P.T., Ueda, S.,
Kato, K., Sako, Y. and Suwa, Y. (2011) Contribution of anammox bacteria
to benthic nitrogen cycling in a mangrove forest and shrimp ponds,
Haiphong, Vietnam. Microbes Environ. 26,16
Received 17 August 2011
doi:10.1042/BST20110711
C
The Authors Journal compilation C
2011 Biochemical Society
... Nitrification is a vital process that converts ammonium into nitrite and nitrate, an essential nutrient for plant growth. We evaluated changes in potential nitrification by monitoring soil microbial activity after fertilization with NEO and other fertilizers (Hu, Shen, Xu, & Zheng, 2011;Robertson, 2015). ...
Thesis
Full-text available
In an era of population growth, increasing food prices, scarcity of arable land, and environmental degradation of farmlands, demand for novel solutions has emerged. This entails introducing innovative fertilizer products designed to mitigate their environmental footprint. Synergized with complementary strategies, these innovations can bolster food production while safeguarding the food production system's sustainability. Nitrogen Enriched Organic fertilizer (NEO) is produced using a new method, where dinitrogen (N2) is captured from the air and through a plasma process mixed with bio-based fertilizers as nitrate (NO3 -) and nitrite (NO2-). However, a thorough product assessment is necessary to unveil potential adverse effects before introducing it to the global markets. In this context, our research has centered on examining the fertilizer's impact on soil fauna activity (measured through substrate breakdown), the abundance of key soil species (springtails and earthworms), as well as critical processes like nitrification and nutrient uptake (yield). Different fertilization regimes were employed, including mineral fertilizer, NEO, untreated biobased fertilizers, and no fertilizer across various experimental setups. Regarding soil fauna feeding activity, there were discrepancies between our two studies. However, we observed a tendency for higher feeding activity in unfertilized soil or under lower fertilization amounts, irrespective of fertilizer type. However, the initial perturbative effect of fertilization on soil fauna feeding activity subsided within a few weeks after application. Likewise, NEO and other fertilizers demonstrated no detrimental effects on the abundance and weight of earthworms or the abundance of springtails. The study also investigated the impact of NEO on soil nitrification potential and observed that although NEO initially stimulated nitrification rates in controlled settings, this effect did not persist ≈ six months after fertilization in the field. Concerning crop yields, while yielding slightly less grass than mineral fertilizers under controlled conditions with equivalent N-min input, NEO exhibited a grain yield approximately 20% lower than mineral fertilizer in the field. Albeit, NEO unveiled an advantage, yielding 20–30% more than the original cattle slurry supply. This signifies a noteworthy enhancement in crop productivity, achieved solely through using electricity and cattle slurry as inputs. In brief, the explorations did not detect any harmful effects of NEO on soil functions and key species, while improved crop yields than the feedstock from which it was derived. Thus, our research findings demonstrate that NEO constitutes a meaningful contribution despite its incremental role in transitioning global food production systems toward sustainability. Keywords: Sustainable agriculture; nitrogen; fertilization; soil health; crop productivity
... The denitrifying bacteria contributing to nitrate removal which are functioning in nitrite accumulation and or complete denitrification. The different requirements of nitrifiers and denitrifies have led to a number of reactor combinations for the removal of nitrogen from wastewater (Hu et al., 2011). There is a huge risk, and these traditional methods can therefore not be used due to several issues impending large-scale application of bio-denitrification (Wang and Wang, 2013). ...
Article
Full-text available
Anammox (an abbreviation for anaerobic ammonium oxidation) is a reaction where specialized microorganisms with anamoxasomes carry out coupled oxidation reduction process where ammonium is oxidized, and nitrite is reduced to form dinitrogen. Here, nitrite is used as the electron acceptor under anoxic conditions. Anamoxasomes (phylum Planctomycetes) are strict anaerobes hence oxygen is not required for this process for treating nitrogen-rich wastewater. The Anammox process was discovered by Mulder in 1995 to avoid the need to add additional COD (chemical oxygen demand) to the system, it functions as an effective and affordable biological nutrient removal procedure in wastewater treatment. It is mainly done to protect the quality of the water body that it is discharged into. Algal bloom which is caused by fixed nitrogen such as ammonium and nitrate is avoided. Anammox is used as a better alternative approach for the removal of nitrogen from wastewater treatment as the bacteria requires less energy, reduces CO2 emissions, produces less excess sludge. It is also known as a low-energy consuming and ecofriendly technology. We can also observe increasing importance of anammox process. Therefore, this study reviews and discusses the current developments in anammox combined processes and its impact on wastewater treatment techniques
... of anammox in terrestrial ecosystems (e.g., croplands, wetlands, forests, and grasslands) is also noteworthy (Hu et al., 2011;Wang et al., 2019). For instance, the anammox rate in croplands was reported to be 2.50 Tg N year −1 in Southern China, accounting for 10% of N loss, which may be greater than the N loss as nitrous oxide (emission factor ranging from 0.1% to 1.1%) (Symonds, 2021;Yang et al., 2015). ...
Article
Soil anammox is an environmentally-friendly way to eliminate reactive nitrogen (N) without generating nitrous oxide. Nevertheless, the current earth system models have not incorporated the anammox due to the lack of parameters in anammox rates on a global scale, limiting the accurate projection for N cycling. A global synthesis with 1212 observations from 89 peer-reviewed papers showed that the average anammox rate was 1.60 ± 0.17 nmol N g-1 h-1 in terrestrial ecosystems, with significant variations across different ecosystems. Wetlands exhibited the highest rate (2.17 ± 0.31 nmol N g-1 h-1 ), followed by croplands at 1.02 ± 0.09 nmol N g-1 h-1 . The lowest anammox rates were observed in forests and grasslands. The anammox rates were positively correlated with the mean annual temperature, mean annual precipitation, soil moisture, organic carbon (C), total N, as well as nitrite and ammonium concentrations, but negatively with the soil C:N ratio. Structural equation models revealed that the geographical variations in anammox rates were primarily influenced by the N contents (such as nitrite and ammonium) and abundance of anammox bacteria, which collectively accounted for 42% of the observed variance. Furthermore, the abundance of anammox bacteria was well simulated by the mean annual precipitation, soil moisture, and ammonium concentrations, and 51% variance of the anammox bacteria was accounted for. The key controlling factors for soil anammox rates differed from ecosystem type, e.g., organic C, total N, and ammonium contents in croplands, versus soil C:N ratio and nitrite concentrations in wetlands. The controlling factors in soil anammox rate identified by this study are useful to construct an accurate anammox module for N cycling in earth system models.
... In other studies, anammox bacteria were detected in anoxic or suboxic zones of marine and freshwater ecosystems. Anammox bacteria have also been found in extreme saline-related environments [34,36,38,39]. The factors affecting anammox bacterial diversity and distribution have been recently investigated in habitat-specific studies-for example, organic carbon influences anammox diversity in freshwater habitats, estuary marine sediments, and soil. ...
Article
Full-text available
Anaerobic ammonium oxidation (anammox) has shown success in past years for the treatment of municipal and industrial wastewater containing inorganic nutrients (i.e., nitrogen). However, the increase in polycyclic aromatic hydrocarbon (PAH)-contaminated matrices calls for new strategies for efficient and environmentally sustainable remediation. Therefore, the present review examined the literature on the connection between the anammox process and PAHs using VOSviewer to shed light on the mechanisms involved during PAH biodegradation and the key factors affecting anammox bacteria. The scientific literature thoroughly discussed here shows that PAHs can be involved in nitrogen removal by acting as electron donors, and their presence does not adversely affect the anammox bacteria. Anammox activity can be improved by regulating the operating parameters (e.g., organic load, dissolved oxygen, carbon-to-nitrogen ratio) and external supplementation (i.e., calcium nitrate) that promote changes in the microbial community (e.g., Candidatus Jettenia), favoring PAH degradation. The onset of a synergistic dissimilatory nitrate reduction to ammonium and partial denitrification can be beneficial for PAH and nitrogen removal.
Article
Full-text available
Complete nitrogen removal depends on the requirement of oxygen as an electron acceptor and organics as electron donors. In the natural environment, these electron acceptors and donors regulate the nitrogen cycle in association with several nature-based processes, such as iron dependent (feammox), sulphur dependent (sul-fammox), and nitrite dependent (annamox). These processes represent the latest discoveries in the nitrification part of the nitrogen cycle and provide the electron acceptors for anaerobic ammonium oxidation. Research findings on these processes suggest that a previously undescribed pathway is available for anaerobic ammonium oxidation, which can regulate the nitrogen cycle in natural environments. It has been shown that some redox active materials can serve simultaneously as electron acceptors and donors in the nitrogen cycle. This review provides comprehensive information of the known oxidative processes in the nitrogen cycle and highlights the emergence of a new pathway for anaerobic NH 4 + oxidation, which is dependent on insoluble redox active materials as electron donors and acceptors. This new pathway is not limited to anaerobic NH 4 + oxidation, but it also includes reductive pathways to nitrogen gas, and thus can lead to complete nitrogen removal from nitrogen environments. Recent developments and knowledge advancement in the nitrogen cycle open a new path for nature-based sustainable solutions for nitrogen management. This new knowledge can provide innovative and novel approaches for conventional NH 4 + removal processes.
Article
Full-text available
Amidst population growth, escalating food costs, limited arable land, and farmland degradation, the adoption of innovative technologies—like organic waste recycling and nutrient recovery—is crucial for enhancing the resilience of global agri-food systems. Nitrogen-Enriched Organic fertilizer (NEO) is produced using a new method, where dinitrogen (N2) is captured from the air through a plasma process and mixed with slurries or digestates as nitrate (NO3-) and nitrite (NO2-). This process leads to solid slurry acidification and a high NO2- content, potentially yielding toxic inorganic or organic N compounds. This study investigated the impact of NEO derived from cattle slurry and biogas digestate on soil nitrification—conversion of NH4+ to NO2- and NO3- by aerobic autotrophic bacteria and archaea. We investigated and compared the potential nitrification rates in soil samples from two agricultural trials (cereal and grass) treated with NEO and other fertilizers after two consecutive fertilization years. Additionally, we examined the immediate nitrification response to NEO through 73-hour soil incubations. Our results revealed that NEO significantly stimulated nitrification rates in agitated soil slurries, regardless of the feedstock used, surpassing rates observed in ammonium controls. Similarly, this pattern was also observed in loosely placed soil samples, with high nitrification rates occurring with NEO and ammonium chloride. Interestingly, the differences in nitrification rates between field-fertilized soil samples were minimal and inconsequential, suggesting that while NEO exhibits a rapid boost in nitrification rates shortly after application, this effect is not sustained ≈ six months after fertilization under field conditions. Consequently, NEO indicates its potential as an environmentally benign fertilizer without adversely affecting soil nitrification. Keywords: environment, innovative fertilizers, nitrogen, manure, sustainability, resilience, waste management
Article
Low carbon to nitrogen ratio (C/N) wastewater contains low organic carbon sources, leading to incomplete removal of nitrogen (N) and phosphorus (P) in the processes of denitrification and anaerobic P release. Due to N and P pollutants in water bodies, there is a serious threat to sustainable development, human health, and water ecosystems. Hence, it is of great significance to develop the energy-saving, efficient, and sustainable N and P removal technology. As functional microorganisms of simultaneous nitrification, denitrification, and phosphorus removal (SNDPR) process, glycogen accumulating organisms (GAOs) and phosphorus accumulation organisms (PAOs) can fully store the carbon source of raw wastewater as the intracellular carbon source while performing N and P removal. Additionally, with the novel autotrophic biological N removal technologies such as anaerobic ammonium oxidation (Anammox), ferric ammonium oxidation (Feammox), and nitrate-dependent ferrous oxidation (NDFO), the inorganic carbon sources (CO2, HCO3⁻, CO3²⁻) can be used as carbon sources, thus achieving N removal. These novel biological nitrogen removal (BNR) processes effectively solve insufficient carbon sources in wastewater with low C/N. In this paper, the recent findings and potential applications of the novel technologies such as SNDPR, Anammox, Feammox, and NDFO are reviewed, and the effectiveness and development trends of the novel technologies in wastewater treatment are discussed. Besides, a new process model is developed for the deep treatment of wastewater in practical engineering. Finally, the development opportunities and challenges of the novel BNR process in future practical engineering applications are summarized and forecasted.
Technical Report
Full-text available
https://www.ecotoxcentre.ch/projects/soil-ecotoxicology/monitoring-concept-for-plant-protection-products-in-soils
Preprint
Full-text available
In the face of population growth, rising food production costs, limited arable land availability, and farmland environmental degradation, novel technologies are crucial to bolster the resilience of global agri-food systems. Nitrogen-Enriched Organic fertilizer (NEO) is produced using a new method, where dinitrogen (N2) is captured from the air through a plasma process and mixed with bio-based fertilizers as nitrate (NO3-) and nitrite (NO2-). This process leads to solid slurry acidification and a high NO2- content, potentially yielding toxic inorganic or organic N compounds. In this study, we investigated the impact of NEO, derived from cattle slurry and biogas digestate, on soil nitrification, which involves the conversion of NH4+ to NO2- and NO3- by aerobic autotrophic bacteria and archaea. We investigated and compared the potential nitrification rates in soil samples from two agricultural trials (cereal and grass) treated with NEO and other fertilizers after two consecutive fertilization years. Additionally, we examined the immediate nitrification response to NEO through 72-hour bottle incubations. Our results revealed that NEO significantly stimulated nitrification rates in agitated soil slurries, regardless of the feedstock used, surpassing rates observed in ammonium controls. Similarly, this pattern was also observed in loosely placed soil samples, with high nitrification rates occurring with NEO and ammonium chloride. Surprisingly, the differences in nitrification rates between field-fertilized soil samples were minimal and inconsequential, suggesting that while NEO exhibits a rapid boost in nitrification rates shortly after application, this effect is not sustained ≈ six months after fertilization under field conditions. Consequently, NEO indicates its potential as an environmentally benign fertilizer without adversely affecting soil nitrifier communities.
Article
Full-text available
Primary production in over half of the world’s oceans is limited by fixed nitrogen availability. The main loss term from the fixed nitrogen inventory is the production of dinitrogen gas (N2) by heterotrophic denitrification or the more recently discovered autotrophic process, anaerobic ammonia oxidation (anammox). Oceanic oxygen minimum zones (OMZ) are responsible for about 35% of oceanic N2 production and up to half of that occurs in the Arabian Sea1. Although denitrification was long thought to be the only loss term, it has recently been argued that anammox alone is responsible for fixed nitrogen loss in the OMZs2–4. Here we measure denitrification and anammox rates and quantify the abundance of denitrifying and anammox bacteria in the OMZ regions of the Eastern Tropical South Pacific and the Arabian Sea. We find that denitrification rather than anammox dominates the N2 loss term in the Arabian Sea, the largest and most intense OMZ in the world ocean. In seven of eight experiments in the Arabian Sea denitrification is responsible for 87–99% of the total N2 production. The dominance of denitrification is reproducible using two independent isotope incubation methods. In contrast, anammox is dominant in the Eastern Tropical South Pacific OMZ, as detected using one of the isotope incubation methods, as previously reported3,5. The abundance of denitrifying bacteria always exceeded that of anammox bacteria by up to 7- and 19-fold in the Eastern Tropical South Pacific and Arabian Sea, respectively. Geographic and temporal variability in carbon supply may be responsible for the different contributions of denitrification and anammox in these two OMZs. The large contribution of denitrification to N2 loss in the Arabian Sea indicates the global significance of denitrification to the oceanic nitrogen budget.
Article
Full-text available
The occurrence and significance of the anammox (anaerobic ammonium oxidation) process relative to denitrification was studied in photosynthetically active sediment from 2 shallow-water estuaries: Randers Fjord and Norsminde Fjord, Denmark. Anammox accounted for 5 to 24 % of N-2 production in Randers Fjord sediment, whereas no indication was seen of the process in sediment from Norsminde Fjord, It is suggested that the presence of anammox in Randers Fjord and its absence from Norsminde Fjord is associated with differences in the availability of NO3- + NO2- (NOx-) in the suboxic zone of the sediment. In Randers Fjord, NOx- is present in the water column throughout the year and NOx- porewater profiles showed that NOx- penetrates into the suboxic zone of the sediment. In Norsminde Fjord, NOx- is absent from the water column during the summer months and, via assimilation, benthic microalgae may prevent penetration of NOx- into the suboxic zone of the sediment. Volume-specific anammox rates in Randers Fjord were comparable with rates measured previously in Skagerrak sediment by other investigators, but denitrification rates were 10 to 15 times higher. Thus, anammox contributes less to N-2, production in Randers Fjord than in Skagerrak sediment, We propose that the lower contribution of anammox in Randers Fjord is linked to the higher availability of easily accessible carbon, which supports a higher population of denitrifying bacteria. Amplification of DNA extracted from the sediment samples from Randers Fjord using planctomycete-specific primers yielded 16S rRNA gene sequences closely related to candidatus Scalindua sorokinii found in the Black Sea by other investigators. The present study thus confirms the link between the presence of bacteria affiliated with candidatus S. sorokinii and the anammox reaction in marine environments. Anammox rates in sediment with intact chemical gradients were estimated using both N-15 and microsensor techniques. Anammox rates estimated with microsensors were less than 22 % of the rates measured with isotopes. It is suggested that this discrepancy was due to the presence of fauna, because the applied N-15 technique captures total N-2 production while the microsensor technique only captures diffusion-controlled N-2 production at the sediment surface. This hypothesis was verified by consistent agreement between the methods when applied to defaunated sediments.
Article
Full-text available
Anaerobic ammonium oxidation (anammox) has become a main focus in oceanography and wastewater treatment1, 2. It is also the nitrogen cycle's major remaining biochemical enigma. Among its features, the occurrence of hydrazine as a free intermediate of catabolism3, 4, the biosynthesis of ladderane lipids5, 6 and the role of cytoplasm differentiation7 are unique in biology. Here we use environmental genomics8, 9—the reconstruction of genomic data directly from the environment—to assemble the genome of the uncultured anammox bacterium Kuenenia stuttgartiensis10 from a complex bioreactor community. The genome data illuminate the evolutionary history of the Planctomycetes and allow us to expose the genetic blueprint of the organism's special properties. Most significantly, we identified candidate genes responsible for ladderane biosynthesis and biological hydrazine metabolism, and discovered unexpected metabolic versatility.
Article
Active expression of putative ammonia monooxygenase gene subunit A (amoA) of marine group I Crenarchaeota has been detected in the Black Sea water column. It reached its maximum, as quantified by reverse-transcription quantitative PCR, exactly at the nitrate maximum or the nitrification zone modeled in the lower oxic zone. Crenarchaeal amoA expression could explain 74.5% of the nitrite variations in the lower oxic zone. In comparison, amoA expression by ?-proteobacterial ammonia-oxidizing bacteria (AOB) showed two distinct maxima, one in the modeled nitrification zone and one in the suboxic zone. Neither the amoA expression by crenarchaea nor that by ?-proteobacterial AOB was significantly elevated in this latter zone. Nitrification in the suboxic zone, most likely microaerobic in nature, was verified by 15NO2 ? and 15N15N production in 15NH4 + incubations with no measurable oxygen. It provided a direct local source of nitrite for anammox in the suboxic zone. Both ammonia-oxidizing crenarchaea and ?-proteobacterial AOB were important nitrifiers in the Black Sea and were likely coupled to anammox in indirect and direct manners respectively. Each process supplied about half of the nitrite required by anammox, based on 15N-incubation experiments and modeled calculations. Because anammox is a major nitrogen loss in marine suboxic waters, such nitrification–anammox coupling potentially occurring also in oceanic oxygen minimum zones would act as a short circuit connecting regenerated ammonium to direct nitrogen loss, thus reducing the presumed direct contribution from deep-sea nitrate.
Article
We measured the production of N2 gas from anammox and denitrification simultaneously in intact sediment cores at six sites along a transect of the continental shelf (50 m) and deeper slope (2000 m) in the North Atlantic. Maximum rates of total N2 production were measured on the shelf and were largely due to denitrification, with anammox contributing, on average, 33% of this production. On the continental slope, the production of N2 gas decreased but the proportion due to anammox reached a maximum of 65%. This change in both amount and dominant pathway of N2 production could be explained largely by the concentration of organic carbon at each site. With increasing carbon the total production of N2 increased rapidly while the response of anammox was not significant. On the continental slope, total N2 production fell below 2 μmol N m22 h21 and anammox was strongly related (r = 0.95) to denitrification but the relative magnitude of anammox to denitrification (1.65 : 1) suggested that anammox could not be fuelled by NO-2 from denitrification alone. On the shelf, however, where total N2 production was predominantly greater than 2 μmol N m-2 h-1, no relationship between anammox and denitrification was found and anammox remained constant at 1.4 mmol N m22 h21. Despite the constancy and greater availability of NO-3 and lower temperatures on the continental slope, the significance of anammox to the total production of N2 appears primarily controlled by the overall rate of N2 production. © 2009, by the American Society of Limnology and Oceanography, Inc.
Article
A first assessment of anammox activity, which is a unique N-2 emission process, was conducted in samples of coastal marine sediment from Japan with a N-15 tracer. The occurrence and diversity of bacteria possibly responsible for the anammox process were also evaluated by selective PCR-amplification of the 16S rRNA gene for known anammox bacteria. Anammox activity, detected by measuring (NN)-N-14-N-15 gas production, was only found in samples collected at the intertidal sand bank located at the Yodo River estuary. In the Yodo River samples, 16S rRNA gene fragments affiliated with the known anammox bacterial lineage were also recovered, and the two major phylotypes were both "Candidatus Scalindua wagneri" relatives with 95% and 98% sequence similarity. Even from the other samples in which no recognizable anammox activity was detected, 16S rRNA gene fragments related to known anammox bacteria, but not to "Ca. S. wagneri", were detected. This is the first report of anammox-mediated N-2 emission in coastal marine environments in Japan. Notably, the PCR-based analysis allowed us to discover unexpected phylogenetic diversity of anammox bacteria-related 16S rRNA gene sequences. The selective PCR primer set developed in this study could be a powerful tool to unveil the ecology of anammox bacteria in natural environments.
Article
The significance of anaerobic ammonium oxidation (anammox) for nitrogen removal in deep continental margin sediments was studied with 15N amendments to suboxic sediments collected from 2800–3100-m water depth at eight sites in the Cascadia Basin (eastern North Pacific Ocean). Consistent with earlier data from deep continental margin sediments, pore-water distributions of inorganic N indicated NH4 removal from suboxic zone sediments, likely due to reaction with nitrate. Anammox rates estimated from suboxic sediment incubations with 15N-labled substrates ranged between 0.065 and 1.7 nmol N mL21 h21 (wet sediment), which suggested that anammox was responsible for the observed NH4 removal. Anammox and denitrification rates derived from NH4 and NO3 pore-water profiles were 32–82 mmol N m22 d21 and 50–110 mmol N m22 d21, respectively. The average contribution of anammox to total N2 production was 40% (15N-amended sediment incubations) to 42% (flux from pore-water inorganic N), which does not support earlier reports that suggested that the relative importance of anammox increased with water depth and thereby should dominate over denitrification at depths greater than 1000 m.
Article
Anaerobic ammonium oxidation (anammox) is an important process for nitrogen removal in marine pelagic and benthic environments and represents a major sink in the global nitrogen cycle. We applied a suite of complementary methods for the detection and enumeration of anammox activity and anammox bacteria in marine sediments of the Gullmar Fjord, and compared the results obtained with each technique. 15N labelling experiments showed that nitrogen removal through N2 production was essentially limited to the upper 2 cm of the sediment, where anammox contributed 23–47% of the total production. The presence of marine anammox bacteria belonging to the genus ‘Candidatus Scalindua’ was shown by 16S rRNA gene sequence comparison. FISH counts of anammox bacteria correlated well with anammox activity, while quantitative PCR may have underestimated the number of anammox bacterial 16S rRNA gene copies at this site. Potential nitrogen conversion by anammox ranged from 0.6 to 4.8 fmol N cell−1 day−1, in agreement with previous measurements in the marine environment and in bioreactors. Finally, intact ladderane glycerophospholipid concentrations better reflected anammox activity and abundance than ladderane core lipid concentrations, most likely because the core lipid fraction contained a substantial fossil component, especially deeper in the sediment.