You are on page 1of 11

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/273154942

Nitrogen transforming community in a


horizontal subsurface-flow constructed
wetland

Article in Water Research · February 2015


DOI: 10.1016/j.watres.2015.02.018 · Source: PubMed

CITATIONS READS

21 362

7 authors, including:

Oksana Coban Marion Martienssen


NASA Brandenburg University of Technology Cottb…
4 PUBLICATIONS 30 CITATIONS 52 PUBLICATIONS 351 CITATIONS

SEE PROFILE SEE PROFILE

Mike S M Jetten
Radboud University
797 PUBLICATIONS 29,907 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Development and testing of a novel photocatalytic system for efficient cogeneration of clean water
and hydrogen for ecosafe agriculture View project

NITROLIMIT - Stickstofflimitation in Binnengewässern; www.nitrolimit.de View project

All content following this page was uploaded by Oksana Coban on 19 March 2015.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2

Available online at www.sciencedirect.com

ScienceDirect

journal homepage: www.elsevier.com/locate/watres

Nitrogen transforming community in a horizontal


subsurface-flow constructed wetland

Oksana Coban a,*, Peter Kuschk b, Uwe Kappelmeyer b, Oliver Spott c,


Marion Martienssen d, Mike S.M. Jetten e, Kay Knoeller a
a
Department of Catchment Hydrology, UFZ e Helmholtz Centre for Environmental Research, Theodor-Lieser-Str. 4,
06120 Halle/Saale, Germany
b
Department of Environmental Biotechnology, UFZ e Helmholtz Centre for Environmental Research, Permoserstr.
15, 04318 Leipzig, Germany
c
Department of Soil Physics, UFZ e Helmholtz Centre for Environmental Research, Theodor-Lieser-Str. 4, 06120
Halle/Saale, Germany
d
Department of Biotechnology for Water Treatment, BTU-Cottbus-Senftenberg, Cottbus, Germany
e
Department of Microbiology, Institute for Water and Wetland Research, Radboud University of Nijmegen,
Nijmegen, The Netherlands

article info abstract

Article history: Constructed wetlands are important ecosystems with respect to nitrogen cycling. Here we
Received 21 December 2014 studied the activity and abundance of nitrogen transforming bacteria as well as the spatial
Received in revised form distribution of nitrification, anaerobic ammonium oxidation (anammox), and denitrifica-
9 February 2015 tion processes in a horizontal subsurface-flow constructed wetland. The functional genes
Accepted 10 February 2015 of the nitrogen cycle were evenly distributed in a linear way along the flow path with
Available online 19 February 2015 prevalence at the superficial points. The same trend was observed for the nitrification and
denitrification turnover rates using isotope labeling techniques. It was also shown that
Keywords: only short-term incubations should be used to measure denitrification turnover rates.
Nitrification Significant nitrate consumption under aerobic conditions diminishes nitrification rates and
Anammox should therefore be taken into account when estimating nitrification turnover rates. This
Aerobic denitrification nitrate consumption was due to aerobic denitrification, the rate of which was comparable
Abundance to that for anaerobic denitrification. Consequently, denitrification should not be considered
Activity as an exclusively anaerobic process. Phylogenetic analysis of hydrazine synthase (hzsA)
Constructed wetland gene clones indicated the presence of Brocadia and Kuenenia anammox species in the
constructed wetland. Although anammox bacteria were detected by molecular methods,
anammox activity could not be measured and hence this process appears to be of low
importance in nitrogen transformations in these freshwater ecosystems.
© 2015 Elsevier Ltd. All rights reserved.

* Corresponding author. Tel.: þ49 341 235 1696; fax: þ49 341 235 45 1696.
E-mail address: oksana.voloshchenko@ufz.de (O. Coban).
http://dx.doi.org/10.1016/j.watres.2015.02.018
0043-1354/© 2015 Elsevier Ltd. All rights reserved.
204 w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2

process occurs naturally (Burgin et al., 2011). While anammox


1. Introduction bacteria have been detected in a range of aquatic ecosystems
(Jetten et al., 2009), almost no investigations were done in
Nitrogen (N) compounds, such as ammonium (NHþ 4 ) and ni- terrestrial ecosystems, despite molecular evidence of the
trate (NO 3 ), are some of the most widespread contaminants in presence of anammox bacteria in diverse soils (Hu et al., 2011;
ground- and wastewater (Harrington and McInnes, 2009; Humbert et al., 2010). Due to the mosaic of aerobic and
Singleton et al., 2007). For decades it was believed that the anaerobic zones as well as the usually low dissolved oxygen
only N transformations responsible for returning fixed nitro- concentration, CWs are assumed to offer favorable conditions
gen to the atmosphere were nitrification, the aerobic stepwise for anammox (Zhu et al., 2010). Although anammox organ-
oxidation of NHþ  
4 via nitrite (NO2 ) to NO3 , and denitrification, isms have been detected in natural wetland soils (Humbert
anaerobic NO 3 or NO 
2 reduction with nitrous oxide (N2O) or et al., 2012), information about their activity and contribu-
dinitrogen gas (N2) as a final product. In the meantime, how- tion to N turnover, particularly in CWs, is rather limited. Some
ever, the process of anaerobic ammonium oxidation (anam- research has been done on enrichment and enhancement of
mox) was discovered, extending the N cycle by showing that anammox processes in CWs (Paredes et al., 2007; Tao et al.,
NHþ 4 can also be oxidized under anaerobic conditions using 2011; Wang and Li, 2011; Zhu et al., 2011). Anammox's
NO 2 as the electron acceptor to produce N2 (Mulder et al., contribution to total N2 production has also been investigated,
1995). but only in the sediments of full-scale surface flow CWs
Constructed wetlands (CWs) are man-made systems treating domestic sewage (Erler et al., 2008). Such systems are
designed to include specific modified characteristics of broadly representative of CWs together with horizontal
wetland ecosystems for improved treatment capacity (Kadlec subsurface-flow (HSSF) CWs, and the HSSF CWs are charac-
and Wallace, 2008). CWs can provide an effective contaminant terized by lower dissolved oxygen concentrations (Vymazal
degradation zone due to enhancement of microbial activity in and Kro € pfelova
 , 2008), so this could potentially increase
the plant's rhizosphere (Stottmeister et al., 2003). CWs can be anammox contributions. Thus the role of anammox in N
used to treat a wide range of contaminants including N-con- removal in systems such as HSSF CWs as well as its rela-
taining compounds. To date, only few studies have focused on tionship with other N transformations needs to be investi-
the microorganisms performing the nitrification and denitri- gated (Zhu et al., 2010).
fication processes in CWs, despite their crucial role in N- The objectives of the present study were: i) to explore the
removal (Chon et al., 2011; Correa-Galeote et al., 2013; Song spatial distribution of both activity and bacterial abundance of
et al., 2012). While the overall N turnover in ecosystems can be nitrification and denitrification, ii) to quantify rates of aerobic
investigated by a wide range of methods, measurements of denitrification, and iii) to investigate the role of anammox in N
microbial activity and its correlation with the copy numbers of removal in HSSF CW.
specific functional genes will add to our understanding of the
systems (Ruiz-Rueda et al., 2009; Song et al., 2012; Wang et al.,
2013). Furthermore, there are only a few reported measure-
ments on nitrification and denitrification turnover rates in 2. Materials and methods
CWs using isotope labeling techniques (Erler et al., 2008; Scott
et al., 2008; Stepanauskas et al., 1996). 2.1. Constructed wetland design
Denitrification is commonly considered to be an anaerobic
process and even small amounts of oxygen are reported to The CW in the study site was built as a part of CoTra
inhibit the activity of the denitrifying enzymes and suppress (Compartment Transfer) project in 2007 at Leuna near Leipzig,
their synthesis in a stepwise manner (Bryan, 1981). However, a Germany. Leuna has served as a major chemical
number of laboratory studies with batch cultures revealed manufacturing site since the beginning of the 20th century,
that denitrification can also take place in the presence of ox- and accidental spills, improper handling, and damages due to
ygen, a process referred to as aerobic denitrification heavy bombing during the World War II have left the under-
(Robertson et al., 1995; Robertson and Kuenen, 1984). In CWs, lying groundwater highly contaminated with benzene,
nitrification occurs in the plant's rhizosphere, as the plant's methyl-tert-butyl ether (MTBE), and NHþ 4 (Martienssen et al.,
roots deliver oxygen, and/or in the top layer of the water body. 2006).
This would also be the zone where aerobic denitrification The CW consisted of a basin (length 5 m  width
could potentially occur. However, studies in natural and near- 1.1 m  depth 0.6 m), was operated as horizontal subsurface
natural environments such as CWs to verify and quantify flow (HSSF), and planted with common reed (Phragmites aus-
rates of aerobic denitrification are lacking (Gao et al., 2010). tralis). The system was filled with gravel (grain size 2e3.2 mm)
Anammox accounts for over 50% of N loss in marine eco- up to a height of 0.5 m and the water level was set to 0.4 m,
systems (Kuypers et al., 2003; Thamdrup and Dalsgaard, 2002). resulting in a vadose zone of 0.1 m. The CW was operated in
However, to date, the role of anammox in freshwater eco- continuous flow regime with flow rate of 7 L h1 and the
systems has not yet been explored in depth (Humbert et al., theoretical hydraulic retention time (assuming no water loss)
2010; Jasper et al., 2014; Zhu et al., 2010). So far, anammox was 6.88 days. Inflow water was pumped in from adjacent
research has focused mainly on quantifying turnover rates contaminated groundwater and had the following chemical
(Burgin et al., 2011; Erler et al., 2008) and/or detecting anam- composition: NHþ 23.4 ± 5.0 mg L1, benzene
4 eN
mox bacteria using molecular biological tools (qPCR, FISH) in 4.1 ± 4.0 mg L , MTBE 0.4 ± 0.4 mg L1, TOC 17.5 ± 7.0 mg L1,
1

various ecosystems in order to reveal if and where this COD 45.0 ± 22.0 mg L1, BOD5 21.0 ± 14.0 mg L1, and pH
w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2 205

7.2 ± 0.4. Nitrate was absent in inflow and only a trace amount (SP) 1 is located at a distance of 1 m from the inflow at a depth
(0.2 ± 0.2 mg L1) of NO2 was detected. of 0.2 m, SP2 is 1 m from the inflow at a depth of 0.3 m, and SP3
Physico-chemical parameters and N removal efficiencies is 4 m from the inflow at a depth of 0.2 m. The qPCR products
were described by Coban et al. (2014). Briefly, the average targeting the hzsA gene were purified using the MinElute PCR
seasonal air temperature was 18.0  C in summer, 12.7  C in Purification Kit (Qiagen, Chatsworth, CA), ligated, and cloned
autumn, and 16.3  C in spring. The water loss calculated from using the pGEM-T Easy Vector System according the manu-
inflow and outflow streams varied between 14% in spring and factures protocol (Promega, Madison, WI, USA). Sixty-eight
99% in summer. Redox was always negative and fluctuated colonies were picked up, amplified with M13 primers (M13-f
between 71 at the top sampling points (0.2 m) and GTTTTCCCAGTCACGAC, M13-r CAGGAAACAGCTATGAC)
119 mV at the deepest sampling points (0.4 m). The CW had and sequencing was performed at the Macrogen sequencing
the following average NHþ þ
4 eN removal: 0.63 g NH4 eN m
2 1
d facility (Macrogen Inc., Amsterdam, Netherlands). The hzsA
þ 2 1
in summer, 0.37 g NH4 eN m d in autumn, and 0.25 g gene sequences obtained in this study are available in NCBI
NHþ 4 eN m
2 1
d in spring. under the accession numbers KP943049eKP943112. Sequences
were aligned using BioEdit Sequence Alignment Editor pro-
2.2. DNA extraction and quantitative polymerase chain gram (version BioEdit v7.2.5; North Carolina State University,
reaction assay Raleigh, NC). Phylogenetic analysis was carried out using the
Mega 6.0 (Tamura et al., 2013).
Samples of gravel and roots in the HSSF CW were taken from 9
sampling points according to distance from the flow path (1 m, 2.4. Measuring anammox and denitrification rates with
2.5 m, and 4 m), and according to depth (0.2 m, 0.3 m, and the 15N isotopic tracing method
0.4 m). DNA was isolated from 0.82 ± 0.06 g of mixed sample of
gravel and roots using the Fast DNA® SPIN Kit for Soil and the To estimate potential anammox and denitrification turnover
FastPrep® Instrument (MP Biomedicals, Santa Ana, CA). The rates we used a sampling scheme similar to that used for the
copy numbers of nitrogen transformation genes were deter- molecular biology investigations. This time samples were
mined by the quantitative polymerase chain reaction (qPCR) taken from 3 points: at a distance of 1 m from the inflow at a
method. qPCR was performed using the StepOnePlus Real- depth of 0.2 m, 2.5 m from the inflow at a depth of 0.3 m, and
Time PCR System (Applied Biosystems) and published 4 m from the inflow at a depth of 0.4 m. The presence, activity,
primer sets for hzsA (Harhangi et al., 2012), nirS, nirK, amoA and and potential of anammox and denitrifying bacteria were
16S rRNA (for sequences, coding genes and annealing tem- measured as described by Risgaard-Petersen et al. (2004). Ho-
perature, see Table 1). A dilution series of cloned amoA, nirS, mogenized samples of gravel and roots of known weight and
nirK and hzsA gene fragments (pGEM-T easy; Promega, Madi- density (16.28 ± 1.27 g, 1.98 ± 0.14 g cm3) were transferred to
son, WI, USA) were used to prepare DNA standards with 22 mL glass vials together with inflow water, purged with N2,
known quantities of target DNA (insert source see Table 1). and sealed. The experiment was performed in three repli-
Reactions were carried out using KAPA™ SYBR® FAST qPCR cates. The slurries were then pre-incubated for 24 h to remove
MasterMix. PCR efficiency for the different assays ranged be- all ambient O2 and NO x in the incubation media. Subse-
tween 84 and 98%. Diluting 1:10 and 1:100 of DNA extracts was quently, 100 ml of N2-purged stock solution of each isotopic
performed in order to avoid any inhibition of PCR by inhibitors mixture, i.e. (1) 15NHþ 15
4 ( N at.%: 98), (2)
15
NO 15
3 ( N at.%: 98),
contained in the gravel material. 15 þ 15 
and (3) NH4 ( N at.%: 98) þ NO3 was added resulting in the
final concentration of about 100 mM N. Incubations in the dark
2.3. Cloning, sequencing, and phylogenetic analysis of and at 20  C were stopped at intervals 0 h, 2 h, 4 h, and 6 h by
anammox bacteria adding 200 ml of a 7 M ZnCl2 solution. The experiment was
repeated using a set NHþ 4 1 mM þ
15
NO
3 200 mM and longer
For the cloning, three locations inside the CW were chosen incubation times of 12 h, 24 h, and 48 h. Incubations with
based on the highest hzsA gene abundances: sampling point 15
NHþ 4 were used as control to ensure that anoxic conditions

Table 1 e Primers and standard information of performed qPCRs.


Gene Primer Sequence (50 -30 ) Annealing Reference Standard clone
temperature ( C)
16S rRNA Nad F TCCTACGGGAGGCAGCAGT 57 Nadkarni et al. (2002) Pseudomonas putida
Nad R GGACTACCAGGGTATCT
AATCCTGTT
Hydrazine synthase, hzsA hzsA 1597F WTYGGKTATCARTATGTAG 55 Harhangi et al. (2012) Kuenenia stuttgartiensis
hzsA 1857R AAABGGYGAATCATARTGGC
ɑ subunit of ammonium amoA 1F GGGGTTTCTACTGGTGGT 57 Rotthauwe et al. (1997) Nitrosomonas europaea
monooxigenase, amoA amoA 2R CCCCTCKGSAAAGCCTTCTTC
Dissimilatory nitrite nirS cd3AF GTSAACGTSAAGGARACSGG 57 Throback et al. (2004) Pseudomonas stutzeri
reductase, nirS nirS R3cd GASTTCGGRTGSGTCTTGA
Dissimilatory nitrite nirK 1F GGMATGGTKCCSTGGCA 45 Braker et al. (1998) Blastobacter denitrificans
reductase, nirK nirK 5R GCCTCGATCAGRTTRTGGTT
206 w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2

were created, incubations with 15NO 3 were used for denitri- 2.6. Calculation method and data analysis
fication turnover quantification, and incubations with
15
NHþ  þ
4 þ NO3 as well as NH4 þ
15
NO3 were used to determine In the ammonia oxidation rates experiment, the net nitrifi-
the denitrification and anammox turnover rates. cation rates in 15NHþ4 amendments were determined using the
The analyses of 15N2 abundance in the samples were con- isotope mixing equation (e.g. Spott et al. (2006)) and the 15NO
3
ducted using a GCMS-QP2010Plus (Shimadzu). The system is amendments were used to determine the gross nitrification
equipped with a ShinCarbon ST column (100/120 mesh, rate and the NO 3 consumption rates by the pool dilution
1.33 m  1 mm ID, Restek) connected downstream via an approach using Equations (1) and (2) from Wessel and Tietema
automatic 6-port-valve (Valco Instruments Co. Inc.) to a Mol- (1992):
sieve 5A column (10 m  0.53 mm ID, 0.50 mm film, Agilent) and
lnðft kÞ
helium as the carrier gas. Gas samples were applied via a f0 k W0  Wt
sample loop attached to a second automatic 6-port-valve p¼ ln Wt
 (1)
W0
t
(Valco Instruments Co. Inc.), which is assembled between the
injector outlet and ShinCarbon ST inlet. The concentration of 2 3
N2 was calculated based on a calibration function gained from lnðft kÞ
6 7 W W
6 f k 7 0 t
standard gas calibration (N2 2530 ppm, loop size: 0.01, 0.5, and c ¼ 61 þ ln0 Wt 7  (2)
4 5 t
2 mL, n ¼ 5 for each loop size). For N2 analysis a gas aliquot of W 0

500 mL was withdrawn from the headspace of a sample vial by


gas-tight syringe and flushed through a 100 mL sample loop where p e gross production rate of the enriched pool,
(attached to the 6-port-valve). Afterwards, the sample loop
was immediately switched into the GCMS carrier gas stream c e gross consumption rate of the enriched pool,
and mass 28N2, 29N2, and 30N2 were determined. The system f e 15N abundance of the enriched pool,
set-up for N2 analysis was as follows: 40  C oven temperature, k e naturally present 15N abundance,
200  C I/F and ion source temperature, total flow 14 mL min1, W e amount of 14N plus 15N in the enriched pool,
column flow 8 mL min1. and t e time.
The rates of anammox and denitrification were calculated
from the mole fraction of 29N2 and 30N2 in the sampled vial Differences between different sample treatments over
atmosphere of samples amended with 15NHþ 4 þ
14
NO 3 and
distance and depth were determined using one-way analysis
15  of variance (ANOVA F-test). The statistical analyses were
NO3 respectively using the equations of Spott and Stange
(2011). Given that the HSSF CW had maximum N removal in performed by IBM SPSS Statistics 21 software, and the differ-
summer, the incubation temperature for the calculation of ences were regarded as significant at p < 0.05. Error bars
potential turnover rates under simulated in situ conditions in represent ±1 standard deviation (SD).
this system was chosen based on summer average air tem-
perature (þ20  C) (Coban et al., 2014).
3. Results and discussion
2.5. Potential ammonia oxidation rates
3.1. Ammonia oxidizing bacterial abundance and
To determine the potential turnover rates for ammonia activity
oxidation, we used the pool enrichment/dilution method ac-
cording to Wessel and Tietema (1992). The sampling points Abundance of ammonia oxidizing genes was quantified by
were chosen as for the denitrification and anammox in- qPCR targeting a subunit of ammonium monooxygenase
cubations. Homogenized samples of gravel and roots from (amoA), which is the key enzyme in the aerobic ammonia
HSSF CW of known weight and density (178.55 ± 16.78 g, oxidation process. The genomic material of ammonia
1.98 ± 0.14 g cm3) were placed in 500 mL Schott bottles with oxidizing bacteria was detected at all points in the CW with
0.1 L of inflow water. The experiment was performed in three the highest abundance at the depth of 0.2 m, 2.2  106 copy
replicates. Samples were closed with air access and incubated number g1 sample, and the lowest at 0.4 m depth, 6.8  104
at þ20  C in the dark with shaking. Subsequently, two sets of copy number g1 sample (F(2,33) ¼ 12.968, p ¼ 0.000) (Fig. 1).
incubations were set up. To the first set 5 mL of 10% 15NHþ 4 was The abundance of the amoA gene varied with the distance
added to the final concentration of 177.38 mM NHþ 4 ; to the from the inflow along the flow path at each depth (p < 0.05).
second 1 mL of 10% 15NO 3 was added resulting in the final Also, the net nitrification rates in 15NHþ
4 amendments were
concentration of 7.08 mM NO 3. determined using the isotope mixing equation and the 15NO 3
Between 6 and 8 mL of sample was taken with a syringe amendments were used to determine the gross nitrification
immediately after substrate addition, and then again after 3, rate and the NO 3 consumption rates using Equations (1) and
12, and 24 h. Samples were filtered (filter 0.2 mm) and stored in (2) respectively. In the latter case, NHþ 4 was not additionally
þ4  C until analysis. In all incubations 15N abundance and added or primarily present in incubations and therefore it
concentration of NO-3 were measured. For measurements, the derived from the mineralization of the organic material. Thus,
SPINMAS technique was used which is a direct coupling of an in these incubations the nitrification rate is equal to miner-
SPIN (sample preparation unit for inorganic nitrogen) to a alization. The results are presented in Table 2. Like the dis-
common quadropole Mass Spectrometer (GAM 400, InProcess tribution of the ammonia oxidizing genes, the gross
Instruments GmbH, Bremen, Germany) (Stange et al., 2007). nitrification rates depended on the location of the sample, i.e.
w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2 207

Fig. 2 e Comparison of measured 29N2 data with modelled


29
N2 data when assuming an increasing contribution of
Fig. 1 e The spatial variations of ammonia oxidizing either only denitrification N2 or only anammox N2 to N2
bacterial abundance targeting the amoA gene and the atmosphere with natural 15N abundance in the HSSF CW
nitrification rates calculated using the mixing equation incubations with NHþ 4 þ
15
NO¡
3 (98 at%).

(net) and using Equation (1) (gross) in samples of gravel and


roots taken from the HSSF CW at three distances along the
flow path and at three depths for each distance.
high NO3 consumption can also take place and this should be
accounted for when calculating the nitrification rates.
Therefore, the approach of Wessel and Tietema (1992) should
depth and distance along the flow path (F(2,23) ¼ 5.428, be used to measure the nitrification turnover rate.
p ¼ 0.012) (Fig. 1). Furthermore, the highest nitrification rates
and ammonia oxidizing bacterial abundance were detected at 3.2. Potential anammox activity
the 0.2 and 0.3 m depths, which can be explained by the
enhanced microbial activity in zones with the highest root To determine potential activity for anammox and denitrifi-
density, i.e. the top sampling points. This could imply that the cation, the isotope labeling experiments were performed with
plant's rhizosphere plays an important role in oxygen delivery three sets of incubations (15NHþ 4,
15
NO3 , and
15
NHþ 
4 þ NO3 ) and
þ 15 
as no oxygen is permitted into the system via the surface in a subsequent set NH4 þ NO3 . The isotope incubations
the HSSF CW (Vymazal and Kro € pfelova
 , 2008). amended with only 15NHþ 4 had no accumulation of
29
N2 or
30 
The gross nitrification rates were significantly higher than N2, indicating that all ambient O2 and NOx had been
the net nitrification rates (F(1,49) ¼ 15.599, p ¼ 0.000). This can consumed during the 24 h preincubations. When both 15NHþ 4
be explained by the fact that the applied isotope mixing and NO-3 were added, no accumulation of 29N2 or 30N2 could be
equation does not consider a possible NO 3 consumption. In observed (i.e. no 15N enrichment in sampled N2 could be
contrast, the nitrification estimation method from Wessel and detected) in any incubation, which implies that 15NHþ 4 was not
Tietema (1992) produces the result as ‘gross nitrification’ and utilized for N2 production within the incubation period, and
thus accounts for a possible NO 3 consumption. At depths of hence hinted at an absence of an active anammox process
0.2 m and 0.3 m, where high NO 3 consumption were detected, (Fig. 2). In fact, when using the equations of Spott and Stange
the nitrification rates calculated on the basis of those two (2011) to calculate the fraction of anammox N2 (i.e. hybrid N2)
methods were quite different. Likewise, at 0.4 m depth, where in treatments with NHþ 4 and
15
NO 3 , the resulting contributions
NO 3 consumption was not significant, there was also no dif- were 0.08% to the total N2 mixture, which is just below the
ference between net nitrification and gross nitrification rates. detection limit of 0.1%. Moreover, the observed accumulation
However, at this depth the nitrification rates were also much of 29N2 within sampled N2 was extremely weak (0.78 %e0.84
lower. It can be concluded that under a high nitrification rate %), which also indicated the absence of an active fraction of

Table 2 e Calculated turnover rates for N-turnover processes at different points in HSSF CW.
Distance Depth Net nitrification Gross nitrification Nitrate consumption Anaerobic Anaerobic
from [m] (mixing equation) nmol NO 3 h
1 1
g nmol NO 3 h
1 1
g denitrification denitrification
inflow [m] nmol NO 3 h
1 1
g sample sample (short-term) nmol (long-term) nmol
sample N2 h1 g1 sample N2 h1 g1 sample
1 0.2 1.6 ± 0.9 7.1 ± 0.3 109.3 ± 89.0 3.8 ± 2.7 1.1 ± 0.7
2.5 0.3 2.1 ± 0.2 6.8 ± 3.7 164.2 ± 201.0 3.4 ± 0.7 0.7 ± 0.4
4 0.4 1.3 ± 0.1 1.3 ± 0.3 7.3 ± 12.2 0.6 ± 0.2 0.7 ± 0.3
208 w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2

anammox bacteria in the HSSF CW (Fig. 2). The possible in-


hibition of the anammox bacteria could be the use of NO 3 as
substrate instead of NO 2 . However, given that anammox
bacteria perform dissimilatory NO -
3 reduction to NO2 (Kartal
 
et al., 2007), not only NO2 but also NO3 should be utilizable
in combination with NHþ 4 for anammox. Instead of anammox,
however, in all 15NO 3 treatments high denitrifying activity
was detected.
Analysis of operating parameters in our HSSF CW and
sufficient operating time of six years to establish the microbial
community indicated the presence of all suitable conditions
for the anammox process (Zhu et al., 2010). It was shown by
Musat et al. (2010) that anammox process can be coupled to
nitrate-dependent hydrocarbon attenuation, thus the pres-
Fig. 3 e The spatial variations of anammox bacterial
ence of benzene and MTBE should not have a negative impact
abundance targeting diagnostic hzsA gene in the samples
on the anammox rates. However, given the high organic car-
of gravel and roots from the HSSF CW at three distances
bon of this inflow of contaminated groundwater and the car-
along the flow path and at three depths for each distance.
bon from the rhizodeposition products, NO 2 produced by
ammonia oxidizing bacteria may be predominantly used for
denitrification instead of anammox. As it was shown by
Reiche et al. (2010), low concentrations of contaminants and a The hzsA primers have a preference over traditional 16S
low C/N ratio of below 2 stimulates anammox, and in our rRNA anammox primers as anammox primers may target not
study a C/N ratio of about 2 was relatively high, which would only the anammox group but other Planctomycetes as well
create conditions that favor growth of denitrifying bacteria. (Humbert et al., 2010). Using traditional 16S rRNA anammox
Therefore, the anammox process might play a role in CWs, primers for the same samples gave us two orders of magni-
albeit primarily in systems where they will not be out- tude of higher copy numbers of anammox bacteria (data not
competed by denitrifying bacteria, i.e. under low C/N ratio (He shown), and these results are similar to those of Bale et al.
et al., 2012). (2014). However, even when using a highly specific target for
One advantage of treating contaminated waters with anammox, the abundance of the bacteria does not correlate
anammox in CWs would be the potential mitigation of N2O with its activity (Wang et al., 2012). While Bale et al. (2014)
emissions because, unlike denitrification, anammox does not detected anammox activity already at 4.2  105e1.4  106
produce N2O as a byproduct (Van de Graaf et al., 1997). So far, copies of the hzsA gene g1 in a marine environment, the ac-
only two studies in CWs detected active anammox process tivity of anammox bacteria in our samples could not be
without stimulation and/or enhancement (Erler et al., 2008; measured, even though the anammox bacterial abundance
Jasper et al., 2014). However, in the study of Erler et al. (2008) was similar.
it was shown that even when anammox contributed up to
30% to N removal, high N2O emissions were detected. Alter- 3.4. Biodiversity of anammox bacteria
natively, CWs were considered to be a minor source of N2O
emissions and it was suggested that if all global domestic The presence of anammox bacteria and diversity of the
wastewater were treated by wetlands the share of the trace anammox community were studied by analysis of the 68 hzsA
gas emission budget would still be less than 1% (Teiter and gene sequences obtained from the gravel and root samples
Mander, 2005). This means that absence of anammox pro- from the HSSF CW. Phylogenetic analysis of the hzsA genes
cess in CWs for waste- and groundwater treatment should not (Fig. 4) showed that most of the sequences in both SP1 and SP3
have a negative impact on global warming. were related to “Ca. Brocadia fulgida”. One hzsA clone sequence
from the SP1 and three hzsA clone sequences from the SP3
3.3. Anammox bacterial abundance were closely related to the sequence of “Ca. Brocadia anam-
moxidans”, while two other SP1 clones with two SP2 clones
The abundance of anammox genes in the biofilms attached to clustered together with “Ca. Kuenenia stuttgartiensis”. While
the gravel and roots from the HSSF CW was quantified with Brocadia could be found at locations SP1 and SP3 (i.e. 1 and 4 m
qPCR using hzsA primers, the most diagnostic phylogenetic distance to inflow), Kuenenia was detected at SP1 and SP2 (i.e.
marker for anammox bacteria (Harhangi et al., 2012). The re- 1 m distance to inflow). This reflects the presence of the
sults are shown in Fig. 3. The genomic material of anammox various microniches in CWs suitable for different anammox
bacteria was detected at all points in the CW and the resulting species (Zhu et al., 2010). Other studies also showed high
copy number was similar to the copy number of amoA gene, biodiversity of anammox in terrestrial ecosystems (Hu et al.,
between 3.7  104 and 1.4  106 copy number g1 sample. 2011; Humbert et al., 2010; Zhu et al., 2013).
Likewise the ammonia oxidizing gene copy number, hzsA gene
copies showed that bacteria were most abundant at 0.2 m and 3.5. Denitrifying bacterial abundance and activity
least abundant at 0.4 m depth (F(2,24) ¼ 7.157, p ¼ 0.004). The
abundance of the hzsA gene varied with the distance from the Abundance of denitrifying genes was quantified with qPCR
inflow along the flow path at each depth (p < 0.05). targeting nirS and nirK, which encode for key enzymes in
w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2 209

Given that the highest denitrification rates were observed


at 0.2 m and 0.3 m depths, i.e. in zones with higher root den-
sity, we can assume that root exudates play the main role in
organic carbon delivery for the denitrification process.
Another possible conclusion is that denitrification is not
inhibited by the presence of oxygen. This will be discussed
further.

3.6. Aerobic denitrification potential

While measuring nitrification rates in 15NO 3 amendments,


significant NO 3 consumption under aerobic conditions was
observed (Table 2). There are several possible routes through
which these NO 
3 losses can occur. Dissimilatory NO3 reduc-
tion (DNRA) is a microbial process that transforms NO3 to NHþ

4
via formation of NO 2 in anaerobic or low O2 environments.
Fig. 4 e Phylogenetic analysis of hzsA gene sequences from DNRA was shown to be relevant under reduced conditions
samples of gravel and roots from HSSF CW (68 in total). The (Eh ¼ 200 mV) (Buresh and Patrick, 1981). However, today
evolutionary history was inferred using the neighbor- there is growing evidence that DNRA can also take place in the
joining method. The optimal tree with the sum of branch presence of O2, especially under high C:N ratios of about 10
length ¼ 1.055 is shown. Bootstrap values of ≥50 (500 and low NO 3 concentrations (Fazzolari et al., 1998), and it is
replicates) are shown at the branches. The tree is drawn to stimulated by presence of macrophytes (Nijburg and
scale, with branch lengths in the same units as those of the Laanbroek, 1997). Still, the rates of DNRA in CWs related to
evolutionary distances used to infer the phylogenetic tree. the total NO 3 consumption reported in the literature are in the
The evolutionary distances were computed using the 2e9 % range (Fazzolari et al., 1998) and therefore could not be
JukeseCantor method and are in the units of the number of of high importance in our incubations considering the sig-
base substitutions per site. The bar represents 5% nificant NO 3 consumption. Immobilization (microbial uptake)
sequence divergence. All positions containing gaps and could be another NO þ
3 consumption route. However, NH4 is a
missing data were eliminated. There were a total of 258 
preferred N source for microorganisms and NO3 uptake by
positions in final dataset. microorganisms is inhibited when both NHþ 
4 and NO3 are
present (Recous et al., 1990). Taking into account the occur-
rence of nitrification in these incubations, NHþ 4 was present
and therefore, NO 3 uptake should be negligible. Denitrifica-
denitrification that are responsible for NO 2 reduction. The tion was assumed to be an exclusively anaerobic process
results showed high copy numbers of denitrifying genes, be- which occur only with absence of oxygen (Bryan, 1981).
tween 1.4  108 and 7.7  108 g1 sample (Fig. 5). Furthermore, However, in the laboratory studies evidence was found for
denitrifying bacteria were most abundant at the depth of denitrification to occur under aerobic conditions as well
0.2 m and least abundant at the depth of 0.4 m (Martienssen and Schops, 1999; Robertson et al., 1995) and to
(F(2,32) ¼ 12.278, p ¼ 0.000). The nirS and nirK genes did not vary
with distance from the inflow along the flow path at each
depth (p > 0.05).
In the next step, the potential denitrification rate was
calculated with the isotope labeling technique using the
approach of Spott and Stange (2011). The results are presented
in Table 2. Interestingly, there was a significant difference
between the rates calculated from the short-term and long-
term incubations (F(1,50) ¼ 17.157, p ¼ 0.000) (Fig. 5). With
short-term incubation, the maximum denitrification rates
were detected at the 0.2 m depth and the minimum at the
0.4 m depth (F(2,22) ¼ 6.714, p ¼ 0.005), showing a similar trend
to the distribution of the denitrifying genes. However, in the
long-term experimental set-up there was no difference be-
tween depths (F(2,24) ¼ 1.181, p ¼ 0.324) and the rates were
lower. This might be attributable to a decrease of turnover
rates over time in a batch experiment with a decrease of Fig. 5 e The spatial variations of denitrification bacterial
substrate concentration, according to MichaeliseMenten ki- abundance targeting the nirS and nirK gene and the
netics. This would also smooth out the difference between denitrification rates for the short-term and long-term
different sampling points. Therefore, only short-term in- experimental set-ups in the samples of gravel and roots
cubations (up to 6 h) should be used to measure denitrification from the HSSF CW at three distances along the flow path
rates in future experiments. and at three depths for each distance.
210 w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2

be persistent at high O2 levels (Lloyd et al., 1987). Aerobic


denitrification or co-respiration implies the simultaneous use Acknowledgments
of both O2 and NO 3 as oxidizing agents and can be performed
by various genera of microorganisms (Robertson and Kuenen, This research was completed within the framework of the
1984). Gao et al. (2010) concluded that O2 dynamics did not Marie Curie Initial Training Network ADVOCATE e Advancing
strongly affect N consumption by denitrification in the pres- sustainable in situ remediation for contaminated land and
ence of abundant NO groundwater, funded by the European Commission, Marie
x , but rather denitrification coexisted
with O2 respiration. Therefore, the NO Curie Actions Project No. 265063, SAFIRA project and the
3 consumption rates
observed under aerobic conditions in our incubations could be Helmholtz Interdisciplinary Graduate School for Environ-
attributed to aerobic denitrification. mental Research (HIGRADE). MSMJ was supported by ERC AG
Studies on aerobic denitrification rates in natural envi- 232937 and 339880. We are also grateful to Ines Ma€ usezahl of
ronments are largely missing, except for one study in marine the molecular biology laboratory for her help with the mo-
sediments (Gao et al., 2010). Ours is the first report of aerobic lecular biological analysis of the samples. Thank also goes to
denitrification rates in freshwater ecosystems. At each depth Anne Carney for her help with English proofreading of this
examined in the incubations, the potential denitrification rate manuscript.
under aerobic conditions was much higher than that
measured under anaerobic conditions. However, the rates of
aerobic denitrification might be overestimated given that: 1) references
the starting NO 3 concentrations for the aerobic incubations
were an order of magnitude higher than for the anaerobic
ones; 2) the aerobic denitrification (NO 3 consumption) is
Bale, N.J., Villanueva, L., Fan, H.X., Stal, L.J., Hopmans, E.C.,
expressed in nmol NO 1
g1 sample and the anaerobic Schouten, S., Damste, J.S.S., 2014. Occurrence and activity of
3 h
anammox bacteria in surface sediments of the southern North
denitrification in nmol N2 h1 g1 sample; and 3) 2 mol of NO 3
Sea. FEMS Microbiol. Ecol. 89 (1), 99e110.
transform to 1 mol of N2. The previous findings in the sea Braker, G., Fesefeldt, A., Witzel, K.P., 1998. Development of PCR
sediments reported quite similar rates of denitrification under primer systems for amplification of nitrite reductase genes
aerobic and anaerobic conditions (Gao et al., 2010). The high- (nirK and nirS) to detect denitrifying bacteria in environmental
est denitrification rate was not observed in the deepest depth samples. Appl. Environ. Microbiol. 64 (10), 3769e3775.
interval with the lowest redox potential, but rather in the Bryan, B.A., 1981. Physiology and Biochemistry of Denitrification.
surface 0.2 m depth. It is well related to the nitrification rates Buresh, R.J., Patrick, W.H., 1981. Nitrate reduction to ammonium
and organic nitrogen in an estuarine sediment. Soil Biol.
and implies simultaneous occurrence of nitrification and
Biochem. 13 (4), 279e283.
denitrification, as has been demonstrated by the stable Burgin, A.J., Yang, W.H., Hamilton, S.K., Silver, W.L., 2011. Beyond
isotope approach in Coban et al. (2014). Our CW was fed with carbon and nitrogen: how the microbial energy economy
NHþ 4 rich contaminated groundwater and therefore it would couples elemental cycles in diverse ecosystems. Front. Ecol.
be consistent to assume that denitrifying bacteria should be Environ. 9 (1), 44e52.
located in close proximity to nitrifying bacteria in order to Chon, K., Chang, J.S., Lee, E., Lee, J., Ryu, J., Cho, J., 2011.
Abundance of denitrifying genes coding for nitrate (narG),
obtain NO 3 from them. Biofilm formation is advantageous for
nitrite (nirS), and nitrous oxide (nosZ) reductases in estuarine
this as it creates an oxiceanoxic interface and therefore pro-
versus wastewater effluent-fed constructed wetlands. Ecol.
vides process stratification (Fukada et al., 2004). Eng. 37 (1), 64e69.
Coban, O., Kuschk, P., Wells, N., Strauch, G., Knoeller, K., 2014.
Microbial nitrogen transformation in constructed wetlands
4. Conclusions treating contaminated groundwater. Environ. Sci. Pollut. Res.
1e11.
Correa-Galeote, D., Marco, D.E., Tortosa, G., Bru, D.,
In this study, the contribution and diversity of various nitro-
Philippot, L., Bedmar, E.J., 2013. Spatial distribution of N-
gen cycle processes in a horizontal subsurface-flow con- cycling microbial communities showed complex patterns
structed wetland were determined with molecular markers in constructed wetland sediments. FEMS Microbiol. Ecol. 83
and stable isotope tests. The functional genes of the nitrogen (2), 340e351.
cycle were abundant along the flow path with prevalence at Erler, D.V., Eyre, B.D., Davison, L., 2008. The contribution of
the superficial points. The same trend was observed for the anammox and denitrification to sediment N2 production in a
nitrification and denitrification turnover rates. Furthermore, surface flow constructed wetland. Environ. Sci. Technol. 42
(24), 9144e9150.
significant nitrate consumption under aerobic conditions was
Fazzolari, E., Nicolardot, B., Germon, J.C., 1998. Simultaneous
detected and this should be taken into account when esti- effects of increasing levels of glucose and oxygen partial
mating nitrification rates. This nitrate consumption was pressures on denitrification and dissimilatory nitrate
apparently due to aerobic denitrification. Therefore, in con- reduction to ammonium in repacked soil cores. Eur. J. Soil Biol.
structed wetlands denitrification should not be considered as 34 (1), 47e52.
an exclusively anaerobic process. Although anammox bacte- Fukada, T., Hiscock, K.M., Dennis, P.F., 2004. A dual-isotope
approach to the nitrogen hydrochemistry of an urban aquifer.
ria were detected by molecular-biological techniques, anam-
Appl. Geochem. 19 (5), 709e719.
mox activity was absent and therefore this process appeared
Gao, H., Schreiber, F., Collins, G., Jensen, M.M., Kostka, J.E.,
to be of low importance in nitrogen transformations in the Lavik, G., de Beer, D., Zhou, H.Y., Kuypers, M.M.M., 2010.
studied ecosystem of constructed wetland due to presence of Aerobic denitrification in permeable Wadden Sea sediments.
oxygen and high concentrations of organic carbon. ISME J. 4 (3), 417e426.
w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2 211

Harhangi, H.R., Le Roy, M., van Alen, T., Hu, B.L., Groen, J., Nijburg, J.W., Laanbroek, H.J., 1997. The fate of N-15-nitrate in
Kartal, B., Tringe, S.G., Quan, Z.X., Jetten, M.S.M., Op den healthy and declining Phragmites australis stands. Microb. Ecol.
Camp, H.J.M., 2012. Hydrazine synthase, a unique 34 (3), 254e262.
phylomarker with which to study the presence and Paredes, D., Kuschk, P., Stange, F., Muller, R.A., Koser, H., 2007.
biodiversity of anammox bacteria. Appl. Environ. Microbiol. 78 Model experiments on improving nitrogen removal in
(3), 752e758. laboratory scale subsurface constructed wetlands by
Harrington, R., McInnes, R., 2009. Integrated Constructed enhancing the anaerobic ammonia oxidation. Water Sci.
Wetlands (ICW) for livestock wastewater management. Technol. 56 (3), 145e150.
Bioresour. Technol. 100 (22), 5498e5505. Recous, S., Mary, B., Faurie, G., 1990. Microbial immobilization of
He, Y.L., Tao, W.D., Wang, Z.Y., Shayya, W., 2012. Effects of pH ammonium and nitrate in cultivated soils. Soil Biol. Biochem.
and seasonal temperature variation on simultaneous partial 22 (7), 913e922.
nitrification and anammox in free-water surface wetlands. J. Reiche, N., Lorenz, W., Borsdorf, H., 2010. Development and
Environ. Manag. 110, 103e109. application of dynamic air chambers for measurement of
Hu, B.L., Rush, D., van der Biezen, E., Zheng, P., van Mullekom, M., volatilization fluxes of benzene and MTBE from constructed
Schouten, S., Damste, J.S.S., Smolders, A.J.P., Jetten, M.S.M., wetlands planted with common reed. Chemosphere 79 (2),
Kartal, B., 2011. New anaerobic, ammonium-oxidizing 162e168.
community enriched from peat soil. Appl. Environ. Microbiol. Risgaard-Petersen, N., Meyer, R.L., Schmid, M., Jetten, M.S.M.,
77 (3), 966e971. Enrich-Prast, A., Rysgaard, S., Revsbech, N.P., 2004. Anaerobic
Humbert, S., Tarnawski, S., Fromin, N., Mallet, M.P., Aragno, M., ammonium oxidation in an estuarine sediment. Aquat.
Zopfi, J., 2010. Molecular detection of anammox bacteria in Microb. Ecol. 36 (3), 293e304.
terrestrial ecosystems: distribution and diversity. ISME J. 4 (3), Robertson, L.A., Dalsgaard, T., Revsbech, N.P., Kuenen, J.G., 1995.
450e454. Confirmation of aerobic denitrication in batch cultures, using
Humbert, S., Zopfi, J., Tarnawski, S.E., 2012. Abundance of gas-chromatography and 15N mass-spectrometry. FEMS
anammox bacteria in different wetland soils. Environ. Microbiol. Ecol. 18 (2), 113e119.
Microbiol. Rep. 4 (5), 484e490. Robertson, L.A., Kuenen, J.G., 1984. Aerobic denitrification e a
Jasper, J.T., Jones, Z.L., Sharp, J.O., Sedlak, D.L., 2014. Nitrate controversy revived. Arch. Microbiol. 139 (4), 351e354.
removal in shallow, open-water treatment wetlands. Environ. Rotthauwe, J.H., Witzel, K.P., Liesack, W., 1997. The ammonia
Sci. Technol. 48 (19), 11512e11520. monooxygenase structural gene amoA as a functional marker:
Jetten, M.S.M., van Niftrik, L., Strous, M., Kartal, B., Keltjens, J.T., molecular fine-scale analysis of natural ammonia-oxidizing
Op den Camp, H.J.M., 2009. Biochemistry and molecular populations. Appl. Environ. Microbiol. 63 (12), 4704e4712.
biology of anammox bacteria. Crit. Rev. Biochem. Mol. Biol. 44 Ruiz-Rueda, O., Hallin, S., Baneras, L., 2009. Structure and
(2e3), 65e84. function of denitrifying and nitrifying bacterial communities
Kadlec, R.H., Wallace, S.D., 2008. Treatment Wetlands, second ed. in relation to the plant species in a constructed wetland. FEMS
CRC Press. Microbiol. Ecol. 67 (2), 308e319.
Kartal, B., Kuypers, M.M.M., Lavik, G., Schalk, J., den Camp, H., Scott, J.T., McCarthy, M.J., Gardner, W.S., Doyle, R.D., 2008.
Jetten, M.S.M., Strous, M., 2007. Anammox bacteria disguised Denitrification, dissimilatory nitrate reduction to ammonium,
as denitrifiers: nitrate reduction to dinitrogen gas via nitrite and nitrogen fixation along a nitrate concentration gradient in
and ammonium. Environ. Microbiol. 9 (3), 635e642. a created freshwater wetland. Biogeochemistry 87 (1), 99e111.
Kuypers, M.M.M., Sliekers, A.O., Lavik, G., Schmid, M., Singleton, M.J., Esser, B.K., Moran, J.E., Hudson, G.B.,
Jorgensen, B.B., Kuenen, J.G., Damste, J.S.S., Strous, M., McNab, W.W., Harter, T., 2007. Saturated zone denitrification:
Jetten, M.S.M., 2003. Anaerobic ammonium oxidation by potential for natural attenuation of nitrate contamination in
anammox bacteria in the Black Sea. Nature 422 (6932), shallow groundwater under dairy operations. Environ. Sci.
608e611. Technol. 41 (3), 759e765.
Lloyd, D., Boddy, L., Davies, K.J., 1987. Persistence of bacterial Song, K.Y., Kang, H., Zhang, L., Mitsch, W.J., 2012. Seasonal and
denitrification capacity under aerobic conditions: the rule spatial variations of denitrification and denitrifying bacterial
rather than the exception. FEMS Microbiol. Lett. 45 (3), community structure in created riverine wetlands. Ecol. Eng.
185e190. 38 (1), 130e134.
Martienssen, M., Fabritius, H., Kukla, S., Balcke, G.U., Spott, O., Russow, R., Apelt, B., Stange, C.F., 2006. A 15N-aided
Hasselwander, E., Schirmer, M., 2006. Determination of artificial atmosphere gas flow technique for online
naturally occurring MTBE biodegradation by analysing determination of soil N2 release using the zeolite Ko € strolith
metabolites and biodegradation by-products. J. Contam. SX6®. Rapid Commun. Mass Spectrom. 20 (22), 3267e3274.
Hydrol. 87 (1e2), 37e53. Spott, O., Stange, C.F., 2011. Formation of hybrid N2O in a
Martienssen, M., Schops, R., 1999. Population dynamics of suspended soil due to co-denitrification of NH2OH. J. Plant
denitrifying bacteria in a model biocommunity. Water Res. 33 Nutr. Soil Sci. 174 (4), 554e567.
(3), 639e646. Stange, C.F., Spott, O., Apelt, B., Russow, R.W.B., 2007. Automated
Mulder, A., van de Graaf, A.A., Robertson, L.A., Kuenen, J.G., and rapid online determination of 15N abundance and
1995. Anaerobic ammonium oxidation discovered in a concentration of ammonium, nitrite, or nitrate in aqueous
denitrifying fluidized bed reactor. FEMS Microbiol. Ecol. 16 samples by the SPINMAS technique. Isot. Environ. Health
(3), 177e184. Stud. 43 (3), 227e236.
Musat, F., Wilkes, H., Behrends, A., Woebken, D., Widdel, F., 2010. Stepanauskas, R., Davidsson, E.T., Leonardson, L., 1996. Nitrogen
Microbial nitrate-dependent cyclohexane degradation coupled transformations in wetland soil cores measured by N-15
with anaerobic ammonium oxidation. ISME J. 4 (10), isotope pairing and dilution at four infiltration rates. Appl.
1290e1301. Environ. Microbiol. 62 (7), 2345e2351.
Nadkarni, M.A., Martin, F.E., Jacques, N.A., Hunter, N., 2002. Stottmeister, U., Wiessner, A., Kuschk, P., Kappelmeyer, U.,
Determination of bacterial load by real-time PCR using a Kastner, M., Bederski, O., Muller, R.A., Moormann, H., 2003.
broad-range (universal) probe and primers set. Microbiol.-SGM Effects of plants and microorganisms in constructed wetlands
148, 257e266. for wastewater treatment. Biotechnol. Adv. 22 (1e2), 93e117.
212 w a t e r r e s e a r c h 7 4 ( 2 0 1 5 ) 2 0 3 e2 1 2

Tamura, K., Stecher, G., Peterson, D., Filipski, A., Kumar, S., 2013. ditch system of a constructed wetland. J. Soils Sediments 13
MEGA6: molecular evolutionary genetics analysis version 6.0. (9), 1626e1635.
Mol. Biol. Evol. 30 (12), 2725e2729. Wang, L., Li, T., 2011. Anaerobic ammonium oxidation in
Tao, W.D., Wen, J.F., Huchzermeier, M., 2011. Batch operation of constructed wetlands with bio-contact oxidation as
biofilter e free-water surface wetland series for enhancing pretreatment. Ecol. Eng. 37 (8), 1225e1230.
nitritation and anammox. Water Environ. Res. 83 (6), 541e548. Wang, S.Y., Zhu, G.B., Peng, Y.Z., Jetten, M.S.M., Yin, C.Q., 2012.
Teiter, S., Mander, U., 2005. Emission of N2O, N2, CH4, and CO2 Anammox bacterial abundance, activity, and contribution in
from constructed wetlands for wastewater treatment and riparian sediments of the Pearl river estuary. Environ. Sci.
from riparian buffer zones. Ecol. Eng. 25 (5), 528e541. Technol. 46 (16), 8834e8842.
Thamdrup, B., Dalsgaard, T., 2002. Production of N2 through Wessel, W.W., Tietema, A., 1992. Calculating gross N
anaerobic ammonium oxidation coupled to nitrate reduction transformation rates of 15N pool dilution experiments with
in marine sediments. Appl. Environ. Microbiol. 68 (3), acid forest litter: analytical and numerical approaches. Soil
1312e1318. Biol. Biochem. 24 (10), 931e942.
Throback, I.N., Enwall, K., Jarvis, A., Hallin, S., 2004. Reassessing Zhu, G., Wang, S., Wang, W., Wang, Y., Zhou, L., Jiang, B., Op den
PCR primers targeting nirS, nirK and nosZ genes for community Camp, H.J.M., Risgaard-Petersen, N., Schwark, L., Peng, Y.,
surveys of denitrifying bacteria with DGGE. FEMS Microbiol. Hefting, M.M., Jetten, M.S.M., Yin, C., 2013. Hotspots of
Ecol. 49 (3), 401e417. anaerobic ammonium oxidation at land-freshwater
Van de Graaf, A.A., de Bruijn, P., Robertson, L.A., Jetten, M.S.M., interfaces. Nat. Geosci. 6 (2), 103e107.
Kuenen, J.G., 1997. Metabolic pathway of anaerobic Zhu, G.B., Jetten, M.S.M., Kuschk, P., Ettwig, K.F., Yin, C.Q., 2010.
ammonium oxidation on the basis of 15N studies in a fluidized Potential roles of anaerobic ammonium and methane
bed reactor. Microbiol.-UK 143, 2415e2421. oxidation in the nitrogen cycle of wetland ecosystems. Appl.
Vymazal, J., Kro€ pfelova
 , L., 2008. In: Vymazal, J. (Ed.), Wastewater Microbiol. Biotechnol. 86 (4), 1043e1055.
Treatment, Plant Dynamics and Management in Constructed Zhu, G.B., Wang, S.Y., Feng, X.J., Fan, G.N., Jetten, M.S.M.,
and Natural Wetlands. Springer, Netherlands, pp. 311e317. Yin, C.Q., 2011. Anammox bacterial abundance, biodiversity
Wang, C.X., Zhu, G.B., Wang, W.D., Yin, C.Q., 2013. Preliminary and activity in a constructed wetland. Environ. Sci. Technol.
study on the distribution of ammonia oxidizers and their 45 (23), 9951e9958.
contribution to potential ammonia oxidation in the plant-bed/

View publication stats

You might also like