Recent Developments of The Mainstream Anammox Processes Challenges and Opportinities

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Environmental Chemical Engineering 9 (2021) 105583

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Recent developments of the mainstream anammox processes: Challenges


and opportunities
Hoang Phuc Trinh a, 1, Sang-Hoon Lee a, 1, Garam Jeong b, Hyeokjun Yoon b, Hee-Deung Park a, *
a
School of Civil, Environmental and Architectural Engineering, Korea University, Seoul, South Korea
b
Biological and Genetic Resources Assessment Division, National Institute of Biological Resources, Incheon, South Korea

A R T I C L E I N F O A B S T R A C T

Editor: Dr. Z. Wen Anaerobic ammonium oxidation (anammox) processes are gaining attention owing to their advantages over the
conventional biological nitrogen removal processes. Although technologies related to sidestream anammox
Keywords: processes are being adopted as the latest advanced solutions, the application of these processes to mainstream
Anaerobic ammonium oxidation (anammox) wastewater is still under investigation. Here, we have comprehensively reviewed the challenges and opportu­
Mainstream applications
nities of mainstream anammox processes. Mainstream processes face a variety of challenges that limit their stable
Autotrophic nitrogen removal
operation such as the difficulty of out-selection of nitrite-oxidizing bacteria, high organic carbon to nitrogen (C/
Deammonification
N) ratio, retention of anammox bacteria, and the influence of high concentrations of ammonia and nitrite
compounds. For widespread applications of mainstream anammox processes in full-scale reactors in wastewater
treatment plants (WWTPs), efficient strategies are necessary to manage high carbon to nitrogen ratios, improve
performance in low-intensity wastewater, and retain anammox bacteria. Key elements in the design, operation,
and maintenance of mainstream anammox processes in full-scale WWTPs are also suggested. Overall, this review
provides a critical perspective on the challenges and opportunities associated with mainstream anammox pro­
cesses, which will be of use to researchers and engineers in future.

1. Introduction substantially limited by high operational costs (mostly owing to aera­


tion), generation of secondary pollutants, and the need to supply carbon
As a result of the increasing demand of water for life and production, resources [5,6]. To address these limitations, researchers are increas­
more than 730 million tons of sewage along with other discharges are ingly focusing on anaerobic ammonium oxidation (anammox) processes
generated and released into water bodies every year, and 80% of these (Eq. (4)), mainly because of their energy- and cost-saving properties
discharges are left untreated, which is a serious challenge for sustainable associated with the reduction of aeration (0.9 k Wh kg–1 N against
development [1]. Wastewater treatment is mostly based on physical, 3.5 k Wh kg–1 N) and excess sludge treatment [5], as well as the lack of
chemical, and biological processes [2]. In general, treatments based on carbon supply (0 kgCH3OH kg− 1 N against 3 kgCH3OH kg− 1 N) [7–9]
biological processes, which rely on the metabolic capability of micro­ (Fig. 1). In addition, by operating a completely autotrophic
organisms provide a more sustainable performance and have lower nitrogen-removal over nitrite (CANON) reactor, not only the settling of
operational costs compared to physical and chemical processes [2,3]. biomass was improved, but the aeration energy for oxygen supply was
Nutrients such as nitrogen and phosphorus cause eutrophication in the reduced by 63% compared to conventional nitrification-denitrification
aquatic environment, which hinder the quality of domestic and indus­ nitrogen removal systems [9]. A DEMON system was evaluated to
trial water supplies [4]. The demand for nitrogen removal during consume 25% less energy cost than conventional nitrogen removal
wastewater treatment has increased, especially for treatment of organic systems [9] (0.75 €/kg of N removed was estimated for a partial nitri­
wastewater with high ammonia concentrations. Accordingly, biological fication anammox system [10]). Moreover, by adjusting the WWTP
nitrogen removal processes based on nitrification ((Eqs. (1, 2)), and using the anammox process the methanol supply was reduced by 64%,
denitrification ((Eq. (3)) are conventionally employed in most waste­ thus saving up to € 380, 000 every year [6]. Therefore, anammox pro­
water treatment plants (WWTPs). However, these technologies are cesses in ammonium (NH+ 4 )-rich influent wastewater treatment have

* Correspondence to: School of Civil, Environmental and Architectural Engineering, Korea University, Anam-Ro 145, Seongbuk-Gu, Seoul 02841, South Korea.
E-mail address: heedeung@korea.ac.kr (H.-D. Park).
1
Both authors contributed equally to this work.

https://doi.org/10.1016/j.jece.2021.105583
Received 22 February 2021; Received in revised form 9 April 2021; Accepted 24 April 2021
Available online 30 April 2021
2213-3437/© 2021 Elsevier Ltd. All rights reserved.
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

become more promising areas of research, and the prospects for their outcompete anammox bacteria, and nitritation (i.e., conversion of NH+ 4
wide applications are increasing [11]. into NO-2) processes are difficult to maintain through the selective
retention of ammonium-oxidizing bacteria (AOB) against
NH+ -
4 + 1.5O2 → NO2 + H2O + 2H
+
(1)
nitrite-oxidizing bacteria (NOB) [21]. Furthermore, several other chal­
NO-2 + 0.5O2 → NO-3 (2) lenges exist, including the efficient retention of anammox bacteria under
conditions of low nitrogen concentration, negative effects of
NO-3 → NO-2 → NO → N2O → N2 (3) low-temperature wastewater on microbial communities, performance of
practical applications in different areas, high effluent concentration of
NH+ + 1.32NO-2
+ 0.066HCO-3
+ 0.13H → 1.02N2 + +
0.26NO-3 +
NO-3, and organic residues in mainstream treatment [22]. Therefore, this
4
0.066CH2O0.5N0.15 + 2.03H2O (4)
research subject requires more careful consideration and intensive
Some studies have reported the removal of NH+
and generation of
4 studies.
nitrogen gas (N2 gas) via an unknown mechanism during the anaerobic In this study, we primarily focused on providing a comprehensive
remineralization of lake sediments and organic matter [12,13]. These review of the technical challenges faced during the application of
processes began to receive attention from the academic world when the mainstream anammox processes in full-scale WWTPs, including (i) dif­
first anammox bacteria were reported in 1999 [14]. Anaerobic ficulties in NOB out-selection, (ii) high organic carbon to nitrogen (C/N)
ammonium-oxidizing bacteria (AnAOB) are chemolithoautotrophs that ratios, (iii) retention of anammox bacteria, and (iv) the influence of NH+
4
can adopt NH+ -
4 and nitrite (NO2) as a primary electron donor and ter­ and NO-2-N concentrations. Furthermore, in this manuscript, we discuss
minal electron acceptor, respectively [10,15,16]. Anammox processes the potential applications of anammox processes in mainstream waste­
normally occur under two sequential pathways. First, under aerobic water treatment such as (i) management of high C/N ratio to maxi­
conditions, partial nitritation oxidizes half of the influent ammonium to mizing the recovery of C based energy, (ii) improvement in the
nitrite through the metabolism of autotrophic ammonium-oxidizing performance of low-strength wastewater, (iii) effective retention of
microorganisms with dominant AOB. Such microbes include Nitro­ anammox bacteria, and (iv) practical applications and mechanisms of
somonas eutropha, N. europaea, and Nitrosospira [9]. Second, under anammox process in full-scale WWTPs. Accordingly, this review pro­
anaerobic conditions, anammox processes convert ammonium and ni­ vides a novel aspect of the applications of the mainstream anammox
trite into nitrogen gas by the action of anammox bacteria [9]. These processes along with its operational conditions and design factors, as
anammox bacteria are gram-negative bacteria that contain high lipid well as the potential of its full-scale operation for future development.
content and are known to exhibit red color owing to their unique cy­
tochrome during the enrichment culture. To date, a total of five anam­ 2. Challenges in mainstream anammox processes
mox bacterial genera (Brocadia, Kuenenia, Anammoxoglobus, Jettenia,
and Scalindua) [17] have been reported, and a new bacterium, named Nitritation is the step that involves the microbial oxidation of NH+
4 to
Anammoximicrobium moscowii belonging to the phylum Planctomycetes, NO-2, and it is the common stage for anammox processes. In a previous
recently have been proposed [18]. study [23], AOB enrichment and NOB washout were achieved during
Although the anammox processes for the treatment of sidestream the nitritation step by applying selective pressure on two groups of
wastewater have been actively adopted worldwide, those for the treat­ bacteria. The strategies for controlling nitritation include providing
ment of mainstream wastewater have been mostly conducted at labo­ low-dissolved oxygen (DO) (0.3–1.5 mgO2 L− 1) [24], high-temperature
ratory scales, and are limited in their full-scale applications [19]. In a (35 ◦ C), short sludge retention time (SRT) (2–3 days) [25], control of
review by Cho et al. [20], high NO-2 concentration, high salinity, sul­ aeration time [26], and sludge treatment based on free ammonia (FA)
fides, toxic metal elements, and toxic organic compounds were pre­ and free nitrous acid (FNA) [27]. Nevertheless, for the practical appli­
sented as potential inhibitors that could damage anammox bacteria and cation of mainstream wastewater treatment, several critical issues
adversely affect the nitrogen removal performance. However, for should be addressed (Fig. 2).
mainstream wastewater treatment, high chemical oxygen demand to
nitrogen (COD/N) ratios may cause heterotrophic bacteria to

Fig. 1. Nitrogen cycle in autotrophic nitrogen removal technologies and conventional nitrification/denitrification technologies [55].

2
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

Fig. 2. Challenges for mainstream anammox processes application.

2.1. The out-selection of NOB strategies to nitritation were based on controlling the DO of the bio­
reactors, which involves operating DO at set-points of 0.5–0.9 mg L–1,
The nitritation step significantly affects the successful operation of and then reducing DO levels to 0.06 mg L–1 to determine the NH+ 4 or
anammox applications [28]. Therefore, the balancing of the different NO-2 reduction slopes at the decreased DO level [37]. However, the
functions of the microbial groups is important [22]. The most chal­ reliability of equipment, such as the air blower, is critical for the stable
lenging task is considered to be the complete out-selection of NOBs, such operation of the process [38]. In general, the influence of wastewater,
as under low-strength nitrogen and low-temperature conditions, espe­ reactor configurations, and their operations on the optimum microbial
cially in the case of the frequently varied NH+ 4 -N feeding that triggers community, as well as the selection of microbial communities in the
over-aeration [22]. In the sidestream deammonification with high NH+ 4, mainstream process require further investigation.
FA and FNA were adopted as parameters for controlling the nitritation
process [29,30]; however, this strategy no longer guarantees successful
mainstream processes under low-nitrogen concentrations [19]. 2.2. High ratio of organic carbon to nitrogen
Previously, it was reported that under FA concentration of
0.1–1.0 mg L–1, the NO-2 oxidation by NOB was inhibited, while the NH+ The C/N ratio of conventional domestic influent wastewater is
4
oxidation by AOB was inhibited at 10–150 mg L–1 [31]. Moreover, the excessively high (ranging from 10 to 12 [39,40]); therefore, it is
selective inhibition of NOB via NO-2 accumulation could occur at an FA generally unsuitable for mainstream nitritation/anammox processes
concentration of 1.0–10 mg L–1 [31]. Balmelle et al. [32] reported that [22]. The optimum ratio for a partial nitritation/anammox (PN/A)
the AOB activity was reduced by 40% at an FA concentration of treatment is usually approximately 2–5 [41], or even lower than 0.5 [39,
25 mg L–1, whereas Kim et al. [33] exhibited reduced NOB activity 42]. The high concentration of organic matter in the influent wastewater
(greater than 50%) at an FA concentration of 0.7 mg L–1. An FA level of adversely affects the growth of anammox bacteria [43]; however,
10.6 mgNH3 L–1 was recommended to maintain the stable NO-2 accu­ determining the exact reasons for this phenomenon remains a significant
mulation rate of 84% [34]. Owing to the presence of Nitrous acid challenge. The suggested mechanisms for the inhibitory effect of organic
(HNO2), nitrifying organisms began to be inhibited by FNA concentra­ carbon on the anammox process are “out-competition and inhibition”
tions of 0.22–2.8 mg L–1 [31]. Wang et al. [35] reported that FNA was and ‘‘metabolic pathway conversion inhibition” [44]. First, Chamchoi
substantially more toxic to NOB than AOB under anoxic conditions. et al. [45] and Güven et al. [46] reported that under high COD con­
Although pH control is an alternative, it is not an economically centrations, the heterotrophic bacteria can grow faster than anammox
feasible approach to nitritation owing to the expensive cost of chemicals. bacteria because they compete for space and nutrients each other, sug­
In general, an effective strategy for controlling nitritation is to adjust the gesting that the activity of anammox bacteria decreases owing to spatial
operation temperature higher than 25 ◦ C, to grew AOB relatively faster and nutritional limitations. Chen et al. [47] also demonstrated that the
than NOB [36]. However, this strategy is also not feasible for practical growth of denitrifying bacteria is rapidly stimulated by the addition of
full-scale applications. Until now, most conventional engineering glucose in an anammox reactor, whereas that of anammox bacteria is
unaffected by glucose supplementation [46]. Second, several

3
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

researchers [46,48,49] observed reduced anammox activity in the ni­ efficiency (NRE), for treating mainstream and sidestream wastewater
trogen removal process, and this indicates that anammox bacteria under different C/N ratios are compared in Fig. 3. At low C/N ratios
-
metabolized organic carbons rather than NH+ 4 and NO2 as substrates. (0.0–1.0), anammox processes exhibited high NRE in both mainstream
Interestingly, when acetate was supplemented during the reactor oper­ and sidestream wastewater treatment (71–96%). However, the NRE
ation as a substrate, anammox bacteria revealed different metabolic gradually decreased in the sidestream anammox processes (58.0–85.4%)
properties for “metabolic pathway conversion inhibition” [46]. when the C/N ratio increased over 1.0, while that in the mainstream
Several recent nitritation/anammox reactors have been operated in anammox processes decreased abruptly (4–77.3%). In general, the NREs
mainstream wastewater treatment under different organic C/N ratios, in the mainstream anammox systems are not stable compared to those in
and the challenges faced during the operation period are summarized in the sidestream anammox systems. Therefore, further investigations on
Table 1. Studies have reported that a C/N ratio of 2.7–2.9 triggered a the practical applications and control strategies for mainstream anam­
reduction in both AOB and anammox activities [24]. In contrast, under mox processes are difficult.
the increased C/N ratio conditions from 1.1 to 1.5, the nitrogen removal
process improved because of the increased number of anammox bacte­ 2.3. Retention of anammox bacteria
ria, while, their grow rate is decreased under 2.0 of C/N ratio. These
results indicate that these phenomena could be attributed to the limited For the successful start-up and process stability of anammox pro­
production of NO-2 [50]. Moreover, AOB can be washed out from the cesses, an biomass retention mechanism should be efficiently controlled
reactor at a C/N ratio of 0.8 [25]. Even in the study conducted by Chen to maintain a favorable microbial community owing to the relatively
et al. [51], when the C/N ratio of influent wastewater increased from 0.5 lower growth rates of autotrophs [50]. The accurate doubling time of
to 0.75, the efficiency of nitrogen removal from the nitritation/a­ anammox bacteria remains unclear owing to the lack of a pure culture to
nammox process decreased from 79% to 52%. Based on the operational date [39]. The reported doubling time of anammox bacteria varies be­
data observed in the laboratory-scale reactors (Table 2), the perfor­ tween studies, with a typical range of 10–12 days [8,52]. However,
mance of the anammox processes, in terms of nitrogen removal depending on the species and growth conditions, or even at the
pilot/full-scale at low-temperatures, the observed doubling times can
Table 1 increase to 25–30 days [53,54]. Furthermore, because the concentration
Performance of anammox process for treating mainstream-like wastewater of NH+ 4 is negligible in municipal wastewater, the start-up and mainte­
under different organic carbon to nitrogen ratio. nance of anammox activity remains a challenge [19]. Consequently, to
Reactor C/N Inf. TN NRE (%) Problems Reference efficiently handle the retention of anammox biomass in the mainstream
type (mgCOD (mgN L–1) nitritation/anammox process, an extended SRT should be applied,
mg–1N) especially at temperatures below 20 ◦ C [22,39].
SBR 0.6–2.0 22.0±5.9 70.0–40.0 COD [24] Brocadia, Kuenenia, and Anammoxoglobus species are reported to be
increased, distributed in not only the natural environment but also wastewater
the diffusion treatment facilities currently in operation [55]. Among them, Brocadia
limitation
species are the most commonly observed in wastewater treatment fa­
impacted
AnAOB more cilities, both in the sidestream and mainstream anammox processes
than NOB, (including moving bed biofilm reactor (MBBR), as well as in simulta­
reduced on neous nitrification and anammox reactors, denitrification reactors,
both the AOB sequencing batch reactors (SBRs), up-flow anaerobic sludge blanket
and AnAOB
activities
(UASB) reactors, membrane bioreactors, etc.) [56–61].
SBR 1.1–2.0 59.3 30.8–73.5 AnAOB [50] Interestingly, although the Brocadia species was the first anammox
activity and species to be reported [62], the Kuenenia species is known to be the most
abundance extensively studied anammox species in terms of cell structure, physi­
increased
ology, biochemistry, and comparative genomics research [63]. In
2.0–2.5 59.3 73.5–77.3 AnAOB [50]
activity particular, the key proteins involved in the genome of Kuenenia stutt­
stopped gartiensis (e.g., hydrazine synthase, hydrazine dehydrogenase, hydrox­
increasing as ylamine dehydrogenase, and an S-layer protein with an encoded
increasing hypothesized anammox pathway) have been experimentally validated
amounts of
nitrogen was
in recent studies [64–66]. The most interesting feature of the anammox
removed species was the experimental verification of a key intermediate hydra­
UASB 1.0# 45.9 88.4 Lower carbon [104] zine produced from HNO2 [67]; this suggests that the Kuenenia species
resulted in might include novel metabolic pathways for nitrogen.
high removal
Furthermore, some studies have reported that Kuenenia species,
rate
NRBC 0.5–0.75 100.0–200.0 79.0–52.0 Decrease in [51] which are known to live in freshwater environments [68], are enriched
nitrogen from wastewater treatment facilities operated by the anammox pro­
removal cesses. For example, Akaboci et al. [69] reported that Kuenenia species
efficiency are predominant in the process of operating a granular one-stage partial
MBBR 3.2# 21.8±5.2 73.0–91.0 No [57]
interruption
nitritation/anammox SRT at a low-temperature (≤ 15 ◦ C), thus sug­
because gesting that these species could play a crucial role in anammox pro­
nitrogen load cesses. Similarly, Gu et al. [70] reported that during the operation of A-B
was less processes, which consist of an anaerobic fixed-bed reactor and SBR,
H–MBBR 3.2# 21.8±5.2 63.0 Less stable in [57]
Kuenenia species were enriched in the biofilm of the MBBR system.
removing
nitrogen due Interestingly, Yang et al. [71] reported that in the process of altering
to increase in partial nitritation/anammox processes from the sidestream to the
nitrate mainstream, the Kuenenia species has been predominated instead of the
produced Brocadia species. In addition to the Brocadia species, the development of
Note: Inf. TN, Influent Nitrogen Concentration; NRE, Nitrogen Removal Effi­ the Kuenenia species in WWTPs has been widely reported [14,18,48,
ciency; “#” indicates estimated number based on the given information. 72–75].

4
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

Table 2
Performance of anammox process for treating wastewater under different operational conditions.
Reactor Temperature HRT C/N (mgCOD Inf. TN (mgN NLR (mgN L–1 NRE (%) Major Anammox species (Candidatus genus) Reference
type (oC) (h) mg–1N) L–1) d–1)

H–MBBR 15.0 13.0# 3.2 21.8 109.0# 63.0 Brocadia [57]


MBBR 9.0# 73.0–91.0
MBBR 30.0 24.0# 0.7 862.0 1450.0 82.0 Kuenenia stuttgartiensis [79]
MBfR 20.8 4.0 – 50.0 310.0# 91.7–94.7 Brocadiae [127]
MBR 30.0 40.0# – – 1000.0 80.0–85.0 Brocadia sp. 40, Brocadia Caroliniensis [60]
SBR 25.0 11.3# 0.6 22.0 46.8 70.0 – [24]
2.0 40.0
SBR 30.0 6.0# 1.1 59.3# 126.3 30.8 – [50]
1.5 113.8 63.1
2.0 114.0 73.5
2.5 122.5 77.3
SBR 35.0 6.0 – 68.0# 272.0# 92.0 Kuenenia (61%), Brocadia anammoxidans [128]
(32%), Brocadia fulgida (6%)
SBR 35.0 0.8 – 200.0 800.0# 75.0# – [100]
SBR 24.0 31.2# 1.4# 291.0 223.0# 80.0 Brocadia Scalindua [56]
SBR 20.0 6.4 – 337.5# 1200.0–1500.0 60.0 Brocadia fulgida [129]
SBR 15.0 24.0# – 590.0 775.0# 36.5# Brocadia fulgida [101]
SBR 25.0 42.5# – 740.0 533.0# 95.0 Brocadia fulgida [130]
SBR 15.0 14.8# – 800.0# 1230.0–1340.0 70.5# Kuenenia [89]
SBR 30.3 28.8# – 938.0 840.0# 51.0# Brocadia fulgida [101]
SBR 35.0 24.0 – – 380.0# 85.0 Brocadia sp. 40 [61]
SBR 15.0 24.0# – – 300.0 16.7# Kuenenia stuttgartiensis [80]
UASB 28.5 4.6 1.0 45.9 400.0 88.4 – [104]
UASB 30.0 0.1 0.5 51.41# – 92.3 – [131]
16.0 0.3 78.5
UASB 11.0 – 1.3 70.0# 1800.0 82.0 Brocadia anammoxidans [58]
ABR 30.1 72# 0.6–0.8 114.8# 38.27# 89.0–96.0 – [47]
1.2–2.3 4.0
ABR 30.0 – 1.4 178.1# – 58.0 – [47]
1.3 193.7# 63.0
SBAR 10.0 7.2 – 160.0 530.0 48.0 Brocadia fulgida [21]
20.0 85.0
15.0 73.0
CSTF 20.0 – – 100.0# 2150.0 74.0 Brocadia fulgida [15]
30.0 2090.0 78.0
CSTR 25.0 12.0 – 250.0 2500.0 77.3 Kuenenia [129]
SNAD 22.6–28.6 4.0# 2.4 57.7# 346.2# 85.4 Kuenenia Brocadia [59]
CANON 35.0 7.0 0.3 246.0 843.0# 85.0 Scalindua [132]
1.3 68.1
NRBC 35.0 8.0 0.8 250.0 690.0 52.0 – [51]
13.0 0.5 200.0 690.0 79.0
SHARON 29.0 28.2# – 657.0 560.0# 92.0 – [53]

Note: Temp., Temperature; HRT, Hydraulic Retention Time; Inf. TN, Influent Nitrogen Concentration; NLR, Nitrogen Loading Rate; NRE, Nitrogen Removal Efficiency;
“#” indicates estimated number based on the given information.

Other species belonging to anammox bacteria, such as the Anam­ significant part of the autotrophic nitrogen removal process, the high
moxoglobus, Jettenia, and Scalindua species, are known to be primarily concentration of NH+ 4 adversely affects anammox bacteria [11]. A low
distributed in landfill leachate anaerobic treatment systems, freshwater, concentration of NH+ 4 related to high cell density would lead to
and marine ecosystems. As such, they are limitedly observed in sewage decreased anammox bacterial activity [14]. Moreover, the feeding of
-
treatment systems [48,55,76,77]. insufficient amounts of NH+ 4 and NO2 would stimulate the growth of
unwanted bacteria, thereby interfering with anammox activity [14].
Because NO-2 is an important substrate of anammox bacteria, its optimal
2.4. Effect of nitrogen concentration concentrations would be necessary for the anammox processes. How­
ever, a high concentration of HNO2 also inhibits anammox activity [11].
Various loading rates of nitrogen and influent nitrogen concentra­ Unlike AOB and NOB, as mentioned in Section 2.1, the inhibitory con­
tions in previous studies can significantly affect the nitrogen removal centrations of FA and FNA on the growth of anammox bacteria were
rate (NRR), dominant bacterial species, growth capability, and optimal reported to be above 50 mg L–1 and 5 μg L–1, respectively [44].
potential of long-term anammox processes [78], and they are presented A critical concentration of FA was observed at a range of
in Table 2 and Fig. 4. Generally, in the range of nitrogen loading rates 35–40 mg L–1 [80]. In contrast, at a lower FA concentration of
(NLR; 300–900 mg L–1 d–1), NRE was higher under the sidestream 13 mg L–1, the inhibition of anammox activity was reported in a previ­
anammox processes (68–94%) than that under the mainstream anam­ ous study [81]. In the results obtained by Bae et al. [82], a specific
mox processes (52–96% or even lower than 60%). Interestingly, when anammox activity (SAA) decrease of 21.2% was observed when the FA
the NLR continued to increase up to 1000–2500 mgN L–1 d–1, the increased drastically from 1.1 to 21.5 mg L–1 (total nitrogen concen­
sidestream anammox processes still maintained their high performance tration was 160 mg L–1). Waki et al. [81] reported that FA concentra­
(60–85%) in the nitrogen removal process. Fig. 4 also demonstrated that tions of 13–90 mg L–1 could adversely impact the performance of
the NRE for mainstream anammox systems was unstable, although these anammox processes. However, FA concentrations below 13–15 mg L–1
systems were operated under the same NLR conditions. were not considered relevant to the anammox processes [83]; even an
NH+4 is considered to be the major form of nitrogen source in the FA concentration of approximately 20 mg L–1 has been reported to exert
influent wastewater for most WWTPs [79], and although it is a

5
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

Fig. 3. Performance of anammox process for treating mainstream- (hollow circle) and sidestream-like (solid triangle) wastewater under different C/N ratio values.

Fig. 4. Comparison of nitrogen loading rate in the performance of anammox process for treating mainstream- (hollow circle) and sidestream-like (solid triangle)
wastewater under different nitrogen loading rates.

no effect on the stability of the process [80]. addition, the inhibitory influence of FNA was observed at a concentra­
Anammox bacteria are more sensitive to FNA concentrations than FA tion of approximately 213 g L–1 [44,80]. The most significant decrease
[84]. Under lower pH values (pH < 7.1), the toxic effect of NO-2-N was in SAA was reported by Bae et al. [82]: the FNA increased from 0.05 to
most significantly triggered by FNA [85]. Previous studies [16,62,80] 2.4 mg L–1. However, Isaka et al. [84] determined that the anammox
reported that the critical NO-2-N concentration was in the range of bacterial activity was inhibited by the higher than 280 mg L–1 of NO-2-N.
100–350 mg L–1, while Niu et al. [86] inferred that a high removal ef­ Similarly, Strous et al. [87] also determined that the anammox activity
ficiency was realized for FA < 5 mg L–1 and FNA < 0.002 mg L–1. In was completely inhibited when the concentration of NO-2-N exceeded

6
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

100 mg L–1. Generally, for the stable anammox system operation, TN, with a pore size of 20 µm to extract soluble COD (sCOD) and NH+ 4
FA, and FNA concentrations should be below 320 mg L–1 [86], concentration in an influent ranged of 44–88 mg(sCOD) L–1 and 35–50
15 mg L–1 [88], and 1.5 μg L–1 [88], respectively. mgN d–1, respectively, for the followed deammonification processes,
such as IFAS reactor or MBBR, which achieved an NRE and NRR of
3. Opportunities for the application of mainstream anammox 70 ± 4% and 55 ± 6 gN m–3 d–1, respectively, by combining the deni­
processes trification and anammox processes [92]. A high capacity for obtaining
COD with high food-to-microorganism (F/M) ratios at a relatively short
Although anammox processes encounter several challenges that HRT and low SRT has been demonstrated via HRAS processes. The
must be addressed, they provide the interesting possibility of trans­ influent sCOD, colloidal matter, and particulate in the form of sludge
forming most of the biodegradable organic matter arriving in the were concentrated by the high-rate aerobic system HRAS processes,
WWTPs into biogas (i.e., methanization), as they do not require biode­ which could then be utilized for energy recovery via anaerobic digestion
gradable COD for denitrification [39]. This application can save energy [93]. Wett et al. [37] successfully applied a full-scale HRAS to remove
based on the demand on many WWTPs, as well as reduce operating costs COD before it was adopted in mainstream aerobic reactors that were
[57]. According to the important issues already mentioned, proposed strengthened with AOB, and anammox bacteria obtained from a side­
solutions have been continuously investigated to improve and extend stream DEMON reactor. A hydrocyclone classifier was also used to retain
these processes in their wide applications. anammox granules by selecting a high-density sludge fraction, while
intermittent aeration was adopted to repress the NOB activity [37].
Moreover, by operating full-scale HRAS processes with 1 h of HRT and
3.1. Management of high C/N ratio for maximizing C energy recovery
0.3 days of SRT to obtain a high Biochemical Oxygen Demand (BOD)
removal efficiency (~90%) and treat the effluent with a single-stage
Anaerobic digestion processes are conventionally used to convert
nitritation/anammox system at 19 ◦ C over 10 months, approximately
COD to recoverable energy [89]. The maximum production of CH4 from
46% of total NRE was achieved [94].
biodegradable COD is approximately 0.35 Nm3CH4 kg–1COD [39],
The EBPR process was applied for the simultaneous removal of
which can be used to operate heat facilities, motors and provide income
organic carbon and phosphorus from domestic wastewater. A system
for the industry generating the wastewater as well as WWTPs [90].
comprising two SBRs was adopted in the treatment of practical domestic
However, this direct application is potentially inhibited by low-strength
wastewater by integrating enhanced biological phosphorus removal
and low-temperature water, and the recovery of chemical energy is a
with nitratation/anammox [95]. Under intermittent anaerobic-aerobic
serious challenge [39]. The general concept is demonstrated in the A-B
conditions, phosphorus was removed in the first SBR with a short SRT
processes (Fig. 5) based on the modification presented by Li et al. [22]
of 2–3 days to prevent nitrification; then, the nitratation/anammox SBR
and Xu et al. [19], which was designed to maximize the capturing ca­
was responsible for treating the remaining NH+ 4 with a highest NRR of
pacity of carbon resources from wastewater before they pass through
0.12 kgN m–3 d–1 [95]. The CEPT process removes particulate and
autotrophic metabolisms (e.g., nitritation/anammox). The “A” process
colloidal matter via coagulation and flocculation by adding chemicals
involves biodegradation of COD and methanization (e.g., organic mat­
(e.g., polymers and metal salts) to primary clarifiers [22,96]. However,
ters and biogas) through several techniques such as the anaerobic pro­
the CEPT might be ineffective in removing sCOD, and this could affect
cess, high rate activated sludge (HRAS), integrated fixed-film activated
the nitritation/anammox performance. Therefore, certain factors (e.g.,
sludge (IFAS), enhanced biological phosphorus removal (EBPR), chem­
the raw wastewater characteristics and sCOD from the primary clarifers)
ically enhanced primary treatment (CEBT), and bioelectrochemical
should be thoroughly evaluated before designing CEPT as a pretreat­
system. Because anammox processes do not require organics for deni­
ment. The bioelectrochemical system behaves as a pretreatment alter­
trification, organic mass (COD) are transformed into methane for energy
native to directly generate electricity or enhance nitrogen removal
production, which could increase methane production by 47%, and
through current generation [97]. More than 90% of the suspended solids
cause energy reduction for air supply by 50% in comparison with con­
and 75% of the total COD were treated by a microbial fuel cell system.
ventional biological nitrogen removal techniques [91]. The “B” process
Then, the remaining nitrogen at the anode effluent was treated by
is specifically contributed to treat the remaining nutrients (nitrogen)
nitritation–anammox in the cathode [98,99].
passing through the autotrophic metabolic pathway.
For the anaerobic process, a UASB reactor was operated at 20 ◦ C to
pretreat the primary wastewater, and then filtered using a textile filter

Fig. 5. Flow scheme of the “A-B processes” for COD capture prior to mainstream nitritation/anammox applications.

7
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

3.2. Performance improvement of low-strength wastewater mainstream wastewater to achieve high biomass retention have been
reported, which involves the use of membrane bioreactors, biofilm
Both the anammox and partial nitritation processes are significantly technology, SBR, MBBR, hydrocyclone and screen, screening or sieving,
influenced by the conditions of the low-strength wastewater treatment. and rotating drum screen. Fernández et al. [39] reported that the most
In fact, the nitritation step is considered as one of the primary processes, widely studied and implemented mechanisms for obtaining high SRT of
because this step selects and grows the AOB effectively, outcompetes the anammox biomass rely on the formation of biofilms, whose high density
-
NOB, and obtains the oxidation of NH+ 4 to NO2 at approximately 50%. provides a significantly high retention of biomass in the reactors.
However, the treatment of these types of wastewater is very difficult, Although membrane bioreactors have been adopted for research at
especially in low-temperature regions [39]. Recently, most of the laboratory scales to obtain the total retention of biomass [80], they are
autotrophic nitrogen removal systems reported in the literature were rarely employed in full-scale practical applications [39]. Biofilm tech­
operated at temperature conditions of over 25 ◦ C and influent nitrogen nology is being adopted in some of the recent applications of municipal
concentrations of more than 100 mgN L–1 [79]. Although the concen­ wastewater treatment by the PN/A system [80]. In addition, the SBR has
tration ratio of AOB to NOB in the sludge is relatively constant, the been employed for anammox reactors and can be used in combination
reduction of temperature could be more beneficial to the highly with biofilm systems as it allows very high SRTs close to the total
competitive NOB, because of their higher activity than AOB and their retention of biomass [39,50,108]. Recent studies have focused on
higher production of NO-3 [100]. A significantly reduced SAA has been municipal water treatment via the SBR system [21,50,108] with the
observed when the operational temperature was decreased below 25 ◦ C expectation of practically applying the anammox technology in the
[101]. Moreover, the virtually negligible FA could inhibit anammox future owing to its high biomass retention efficiency and the operational
processes in the case of low NH+ 4 concentration, usually approximately flexibility of bioreactors. In addition, anammox immobilized cells have
or less than 50 mg L–1 [36]. Therefore, the analysis of low-temperature been adopted in previous studies for synthetic wastewater owing to their
and low-nitrogen concentrations in wastewater application opens up high biomass content, prevention of washout biomass, and lower
new avenues for applying anammox processes in designing toxicity to microorganisms [20]. By operating a hybrid PN/A MBBR over
energy-saving WWTPs. a broad simulated SRT ranged from 6.8–24.5 days, the direct competi­
Currently, there increase an interest in anammox research under tion for NO-2-N between NOB and AnAOB (retained in the biofilm and
low-temperature and low-nitrogen conditions [100]. The Brocadia spe­ behaving as a “NO-2-N-sink”) triggered a difference in the actual growth
cies (e.g., Brocadia sinica and Brocadia fulgida) is usually involved in rates of AOB and NOB (i.e., μNOB < μAOB in flocs) and allowed for the
processes operated at a temperature range of 5–6 ◦ C. Kuenenia stuttgar­ selective NOB washout while retaining AOB [107]. However, studies on
tiensis was also determined in a continuous bioreactor system that their practical application in wastewater and improvement of polymeric
operated at temperatures lower than 20 ◦ C [100,102]. Lotti et al. [103] gel disposal to the environment need further consideration.
reported that an adapted anammox sludge exhibits a higher specific Other approaches for securing effective retention for anammox are
anammox rate at a lower temperature. At 15 ◦ C, the PN/A process was hydrocyclone and screen. Waste sludge in the sidestream deammonifi­
stably operated on pretreated municipal wastewater with a nitrogen cation was returned to the mainstream deammonification processes for
influent concentration of 21 ± 5 mgNH+ 4 -N L , and it gained an NRR of
–1
the separation of anammox aggregates by hydrocyclones [37]. By
30 mgN L–1 d–1 in the range of conventional nutrient removal systems returning anammox aggregates from cyclone separation, the SRT of
[104]. By controlling DO up to 2.5 mgO2 L–1 for the repression of the anammox bacteria in the system can be considerably prolonged [19].
NOB activity, this could lead the efficient suppression of the nitritation Larger aggregates from the underflow stream of a hydrocyclone
at both 20 and 15 ◦ C, thereby also leading the sufficient NRRs of 0.44 demonstrated that they exhibited significantly higher anammox activity
and 0.40 gNTotal L–1 d–1 with optimal NREs of 86% and 73%, respec­ than the smaller ones [109]. In general, anammox aggregates are larger
tively. Moreover, under the same conditions with 15 ◦ C, an estimated but less compressible than AOB and NOB [19]. Moreover, based on these
growth rate of 0.017 d− 1 was observed, fulfilling a decisive prerequisite findings, screening or sieving has been developed for the selective sep­
for the application of the processes in mainstream WWTPs [105]. In a aration and retention of anammox aggregates [110]. This approach re­
mainstream IFAS deammonification process, 10% of the primary lies on the essential concept that NOB prefers to grow in flocs, while
effluent wastewater before the biological phosphorus and COD removal anammox bacteria exist in granules [111]. By using a vibrating screen
stages was directly diverted to the deammonification reactor to raise the with more moderate normal forces (by water sprinklers) and additional
influent C/N ratio from 2.3 to 3.1 g(sCOD) g–1NH+ 4 -N. In this system, tangential forces (by applying mechanical vibration) to allow for a
72% of TN removal was maintained during the decline in temperature gentle separation of granules entrapped in flocs and shearing of the
from 20 ◦ C to 12 ◦ C, and the anammox metabolism routed about 53% surface of granules, a better balancing retention efficiency of anammox
of nitrogen removal, and it exhibited no negative impact on anammox bacteria granules (41% of the anammox activity) and washout of NOB
activity owing to the increased carbon in the influent wastewater [106]. (92% activity washout) were recorded, which resulted in an enhanced
In a hybrid PN/A MBBR system with an NH+ 4 concentration of 23 mg L ,
–1
NOB out-selection (AOB/NOB ratio of 2.3) and a total NRE of 70%
the AOB and NOB were enriched preferentially in the flocs, while the (effluent total nitrogen concentration < 10 mgN L–1) with the COD/N
anammox bacteria were enriched in the biofilm. The actual growth rates ratio of influent wastewater at 1.4 [111]. For the mainstream deam­
of both AOB and NOB decreased, and the washout of NOB from the flocs monification full-scale reactor in the Strass WWTP (Austria), superior
was observed, whereas the minor NOB fraction remained in the biofilm anammox bacteria retention efficiencies, achieved by the rotating drum
inhibited by the decreased DO concentration (from 1.2 to 0.17 mg L–1) screen (η = 72%) over the cyclone (η = 42%), triggered a prospected
[107]. To date, it was suggested that the anammox population has been increase in capacity by 80–90% and enhanced the success of mainstream
activated under conditions of low nitrogen concentration and deammonification by decreasing its dependency on nitrite residuals
low-temperature, and this has become a promising scenario for appli­ [112]. Research on the operation and choice of physical selectors is
cation in mainstream WWTPs [43,103]. required to directly determine the operational strategy and footprint
necessary to obtain higher anammox bacteria retention and NOB
3.3. Effective retention of anammox bacteria out-selection in mainstream deammonification, as well as to reduce the
demand for tight aerobic SRT control [112].
The growth of anammox bacteria is significantly slow, and the op­
timum anammox biomass retention is essential for its application in 3.4. Commercialized processes for mainstream treatment
different wastewaters, especially under low-temperature conditions [43,
80]. Efforts on research and applications in lab-scale and full-scale Owing to the lack of pure cultures, stringency of operational

8
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

conditions and low growth rate of anammox bacteria, the practical mainstream WWTP conditions, and new granules were enriched and
application of anammox processes is limited by its research progress efficiently retained in the system [94]. These results indicate that the
[113]. However, the application of anammox to nitrogen removal pro­ proposed reactor with granular sludge has a high propensity to be suc­
cesses would replace the conventional denitrification step, thereby cessfully applied for the autotrophic nitrogen removal from the main­
saving half of the nitrification aeration costs, and thus facilitating in the stream of WWTPs [115]. Successful full-scale anammox application on
reduction of operational costs [114]. Owing to continuous and suc­ high nitrogen concentration wastewater aided this accomplishment of
cessful studies, the first full-scale granular anammox reactor with mainstream wastewater treatment with several efficient techniques for
loading rate of up to 750 kgN d–1 was introduced for sludge treatment in anammox enrichment and full-scale application in practical treatment
the municipal wastewater treatment plant of Rotterdam, the (Fig. 6).
Netherlands in 2002, and it symbolically represented the beginning of A full-scale one-stage PN/A MBBR process in Klagshamn WWTP,
the commercial application of anammox processes [8]. Moreover, the Malmö, southern Sweden, which was designed to treat up to 200 kgN
one-stage partial nitritation/anammox reactor (4 m3), which was used d–1, was started up without inoculum within 111 days [116]. Heated
for the supplementation of the actual effluent to the A-stage of the above dilution water has been reported as a prerequisite for the rapid start-up,
WWTP, was operated separately at temperature conditions of summer and it provided significant results, such that the NRR reached 1.8 gN m3
(23.2 ± 1.3 ◦ C) and winter (13.4 ± 1.1 ◦ C), and it exhibited total NRR L–1 and the nitrogen removal was greater than 80% at 35 ◦ C [116].
values of 0.223 ± 0.029 and 0.097 ± 0.016kgN m3 d–1, respectively [94, Another typical full-scale enhanced nitrogen removal via the in situ
115]. Some studies have reported that heterotrophic biomass was effi­ enrichment of anammox bacteria for mainstream wastewater was suc­
ciently washed out of the system, while anammox bacteria grew under cessfully applied in the Xi’an municipal WWTP located in Shanxi

Fig. 6. Efficient approaches for anammox enrichment and full-scale application in practical treatment. (a) Hydrocyclones in combination with the mainstream and
sidestream system; (b) hydrocyclones for enriched biomass in SBR; (c) chemical supply for the improvement of processes; (d) dilution of water and heat exchange for
the improvement of growth rate; (e) moving carriers for the enhancement of efficiency; and (f) robust separation for retaining biomass.

9
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

province, China, with a capacity of 250, 000 m3 d–1. In addition, this was sieves) that retains anammox biomass in the reactor for mainstream
an anaerobic-anoxic-oxic (A2O) (HRT 2.0 h, 2.6 h and 6.4 h respec­ applications [125]. Preliminary results indicate that this application
tively) process and it was upgraded by adding moving carriers seems to be efficient in suppressing the NOB activity while gaining
(approximately 25% filling) to the anoxic zone, and equipping sub­ N-removal rate up to 0.2–0.3 kgN m3 d–1 at 16–18 ◦ C, which is higher
mersible mixers and guide walls to ensure complete interaction between than what is commonly achieved in conventional biological nitrogen
wastewater and the carriers, as well as installing intercept screens in the removal activated sludge systems [125].
outlet to prevent loss of the suspended carriers [117]. In approximately Up to date, Aqualia (AquELAN® with granular biomass), Paques
2 years of continuous operation, when the influent COD/TN ratio varied (ANNAMOX® with granular biomass), Veolia (ANITA® Mox with MBBR
from 1.2 to 7.9, nitrogen was efficiently removed with an average and IFAS), DC water (DEMON® with granular biomass) and DEMON
effluent total inorganic nitrogen (TIN) concentration of 7.6 mg L–1. In GmbH are leading industrial corporations which have made efforts to
addition, by batch tests, 15N-stable isotope tracing tests and mass bal­ develop marketable products for the practical application of anammox
ance analysis determined that NO-2 could be produced via denitration, processes [126]. Applications of this eco-friendly and economic pro­
and its adoption by anammox bacteria along with 15.9% of nitrogen loss cesses are expected to become a commercial technique for mainstream
was likely attributed to the anammox processes being carried out by the wastewater treatment in the near future.
anoxic–carrier biofilm [117,118]. Moreover, metagenome sequencing
analysis showed that the relative abundance of anammox bacteria in the 4. Future prospects of mainstream anammox application
anoxic-carrier biofilms was significantly higher than that of the reported
level in the flocculent sludge of conventional WWTPs [118]. Over the past decade, new discoveries and experiments have
By augmenting AOB and anammox biomass from sidestream SBR, significantly improved our understanding of mainstream anammox
preserve anammox bacteria with hydrocyclone, and inhibiting NOB by processes. However, the full-scale application of these processes requires
controlling aeration, as indicated by the higher concentrations of nitrite technical solutions to address immediate challenges. First, such tech­
rather than nitrate in the effluent and quantified by aerobic activity tests nologies must solve the problem of applying anammox processes to low-
(approximately 75% inhibition), the Strass WWTP (Austria) achieved concentration nitrogen wastewater. The partial nitrification/anammox
stable mainstream anammox technologies [37,119,120]. Similar to the process can be considered to operate under stable conditions in
Strass WWTP, the demonstration projects at Glarnerland WWTP, where municipal wastewater with low-concentration nitrogen inflows. Second,
the anammox activity in the mainstream was enhanced by transferring a the removal of NOB from the treatment system is also a significant
major part of the produced waste sludge in the sidestream, realized the operational challenge. Alternating aeration is a useful strategy applied
feasibility of the deammonification concept when achieving an to mainstream anammox processes for NOB inhibition as it is difficult to
ammonia-removal efficiency greater 90% (maximum load of 200 kg restore activity in a limited time while increasing anammox activity
NH+ 4 -N) [121]. However, investigations were conducted to understand if during the transition from anoxic to aerobic, even in low temperate
anammox can complement the conventional nitrification/deni­ conditions. In addition, FA and FNA have recently been proposed for
trification processes in a temperate zone at the Marselisborg WWTP NOB selection in mainstream processes due to a more potent biocidal
(Aarhus, Denmark) by feeding excess sludge from a sidestream tank to effect on NOB compared to AOB and AnAOB. Third, organic carbon
mainstream tanks. It was found that after more than 3 years of opera­ should be minimized in the anammox reactor, as organic matter nega­
tion, approximately 1% of NRE was contributed by anammox with the tively affects the anammox reaction. The A-B process was developed to
conversion rates being 800–900 times lower than that achieved by maximize the capture capacity of carbon sources from influent waste­
deammonification through 15N-labeling experiments at different sets of water to manage high C/N ratios before the anammox reaction takes
temperature (10, 20, and 30 ◦ C) [122]. This observation could be place. The anaerobic reactor of the A-B process not only helps to reduce
explained by the lack of adaptation to lower temperatures as well as the organic carbon content but also contributes to the production of biogas
small abundance of anammox bacteria in the reactors, due to which the by anaerobic digesters. Methane in biogas can be utilized to produce
anammox processes were excluded by the dominant nitrification [122]. heat and/or electric energy, which is generally used for the operation of
The measurements of 15N-labeling rate also proved the obvious depen­ wastewater treatment plants. The supply of NH4Cl can improve low-
dence of anammox activities on temperature (7.7 ± 0.6 and 72.9 ± 7.2 intensity influents as well as reduce the C/N ratio. However, this
μmol gVSS h–1 at 10 ◦ C and 30 ◦ C, respectively) [122]. Another project chemical method must be carefully evaluated based on the operational
applied by two-stage mainstream deammonification processes without expenses for gaining stability. Finally, the biggest challenge is main­
bioaugmentation for nitrogen polishing was conducted at Chesapeake taining a very slow growth of AnAOB in mainstream anammox pro­
Elizabeth WWTP (VA, USA) [123]. Partially nitrified effluent waste­ cesses. The effective maintenance of AnAOB can be ensured by applying
water was operated as the first-stage activated sludge process by con­ several methods that have been proven useful in recent years. According
-
trolling the ratio of NH+ 4 to NO2 and nitrate (NOx-N), followed by to laboratory-based studies, full-scale reactors, membrane bioreactors,
anammox MBBR as the second stage [123]. The results indicated that MBBR, and SBR systems were successful in holding high concentrations
the system was partially successful in NOB outcompeting, which adopts of AnAOB, and they provided operational flexibility. Considering the
transient aeration with a high DO set-point of 1.0 mg/L and short SRT of economic way to keep AnAOB in the mainstream anammox processes,
approximately 0.1 d at 20 ◦ C. In addition, the performance of anammox physical devices such as hydro-cyclones, sieving equipment, and me­
MBBR polishing was stable, although the nitrogen removal was limited chanical vibrating screens can be selected. Based on the suggested ap­
by the influent COD/NH4 ratio of 8:11 [123]. In a full-scale anammox proaches, companies can develop marketable products and secure
mainstream technology, Changi Water Reclamation Plant (200,000 m3 sludges enriched with AnAOB for practical applications to mainstream
d–1), Singapore, intermittent aeration as a spatial fashion (controlling anammox processes.
DO at 1.5–2.0 mgO2 L–1 by ceramic) was applied to oxidize approxi­
-
mately 72% NH+ 4 to NO2 and then removed by anammox processes in 5. Conclusions
the anoxic zone (the DO dropped to below 0.1 mgO2 L–1 within 1 min
when aeration was stopped) [124]. In the ANITA™ Mox processes, a In this review, we have provided the challenges and opportunities of
single-stage deammonification process utilizing MBBR technology was mainstream anammox processes by analyzing their experimental results
successfully applied to treat mainstream effluent with low sCOD at and associated performances as reported in recent publications. Four
65 mg L–1 content using the IFAS configuration (controlled DO at major challenges have been identified for the application of the anam­
0.2–0.8 mg L–1) after the HRAS process [125]. A unique feature of IFAS mox process to mainstream wastewater: (1) low concentration of
ANITA™ Mox is its easy and robust separation system (i.e., media ammonia in the influent, (2) the difficulty of NOB out-selection, (3) high

10
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

organic carbon to nitrogen ratio, and (4) retention of slow-growing [17] X. Wen, Y. Hong, J. Wu, Y. Wang, Optimization of a method for diversity analysis
of anammox bacteria using high-throughput sequencing of 16S rRNA gene
AnAOB. Possible and efficient strategies have been proposed to over­
amplicon, J. Microbiol. Methods 178 (2020), 106066.
come these four challenges. Partial nitritation and subsequent anam­ [18] S.V. Khramenkov, M.N. Kozlov, M.V. Kevbrina, A.G. Dorofeev, E.A. Kazakova, V.
mox, aeration control, removal of organic carbon before anammox A. Grachev, B.B. Kuznetsov, D.Y. Polyakov, Y.A. Nikolaev, A novel bacterium
reaction, and physical separation of AnAOB have been suggested to carrying out anaerobic ammonium oxidation in a reactor for biological treatment
of the filtrate of wastewater fermented sludge, Microbiology 82 (2013) 628–636.
overcome the main challenges. Discussion of the main challenges and [19] G. Xu, Y. Zhou, Q. Yang, Z.M.-P. Lee, J. Gu, W. Lay, Y. Cao, Y. Liu, The challenges
strategies to solve these challenges would be useful in establishing future of mainstream deammonification process for municipal used water treatment,
research directions for the development of mainstream anammox pro­ Appl. Microbiol. Biotechnol. 99 (2015) 2485–2490.
[20] S. Cho, C. Kambey, V.K. Nguyen, Performance of anammox processes for
cesses. Currently, there are limited cases in which mainstream anammox wastewater treatment: a critical review on effects of operational conditions and
processes have been applied at full scale. However, as analyzed in this environmental stresses, Water 12 (2020) 20.
review, if new processes are developed considering the solutions to the [21] T. Lotti, R. Kleerebezem, Z. Hu, B. Kartal, M.S.M. Jetten, M.C.M. Van Loosdrecht,
Simultaneous partial nitritation and anammox at low temperature with granular
factors that hinder the application of mainstream anammox, it is ex­ sludge, Water Res. 66 (2014) 111–121.
pected that the mainstream anammox processes will be popular in the [22] X. Li, S. Klaus, C. Bott, Z. He, Status, challenges, and perspectives of mainstream
near future. nitritation–anammox for wastewater treatment: Li et al. Water Environ. Res. 90
(2018) 634–649.
[23] R. Du, Y. Peng, J. Ji, L. Shi, R. Gao, X. Li, Partial denitrification providing nitrite:
Declaration of Competing Interest opportunities of extending application for anammox, Environ. Int. 131 (2019),
105001.
[24] M. Han, H. De Clippeleir, A. Al-Omari, B. Wett, S.E. Vlaeminck, C. Bott,
The authors declare that they have no known competing financial S. Murthy, Impact of carbon to nitrogen ratio and aeration regime on mainstream
interests or personal relationships that could have appeared to influence deammonification, Water Sci. Technol. 74 (2016) 375–384.
the work reported in this paper. [25] A. Mosquera-Corral, F. Gonzalez, J.L. Campos, R. Méndez, Partial nitrification in
a SHARON reactor in the presence of salts and organic carbon compounds,
Process Biochem. 40 (2005) 3109–3118.
Acknowledgement [26] M. Jia, K. Solon, D. Vandeplassche, H. Venugopal, E.I.P. Volcke, Model-based
evaluation of an integrated high-rate activated sludge and mainstream anammox
system, Chem. Eng. J. 382 (2020), 122878.
This work was supported by a grant from the National Institute of [27] S. Jenni, S.E. Vlaeminck, E. Morgenroth, K.M. Udert, Successful application of
Biological Resources (NIBR), funded by the Ministry of Environment nitritation/anammox to wastewater with elevated organic carbon to ammonia
(MOE) of South Korea (NIBR202019102) and the National Research ratios, Water Res. 49 (2014) 316–326.
[28] E.M. Gilbert, S. Agrawal, S.M. Karst, H. Horn, P.H. Nielsen, S. Lackner, Low
Foundation of Korea (NRF) (NRF-2020R1A6A1A03045059). temperature partial nitritation/anammox in a moving bed biofilm reactor treating
low strength wastewater, Environ. Sci. Technol. 48 (2014) 8784–8792.
References [29] B. Ma, S. Wang, S. Cao, Y. Miao, F. Jia, R. Du, Y. Peng, Biological nitrogen
removal from sewage via anammox: recent advances, Bioresour. Technol. 200
(2016) 981–990.
[1] A. Boretti, L. Rosa, Reassessing the projections of the world water development
[30] S. Park, W. Bae, Modeling kinetics of ammonium oxidation and nitrite oxidation
report, NPJ Clean Water 2 (2019) 1–6.
under simultaneous inhibition by free ammonia and free nitrous acid, Process
[2] H. Yi, M. Li, X. Huo, G. Zeng, C. Lai, D. Huang, Z. An, L. Qin, X. Liu, B. Li, Recent
Biochem. 44 (2009) 631–640.
development of advanced biotechnology for wastewater treatment, Crit. Rev.
[31] A.C. Anthonisen, R.C. Loehr, T.B.S. Prakasam, E.G. Srinath, Inhibition of
Biotechnol. 40 (2020) 99–118.
nitrification by ammonia and nitrous acid, J. Water Pollut. Control Fed. 48 (1976)
[3] C. Yang, H. Qian, X. Li, Y. Cheng, H. He, G. Zeng, J. Xi, Simultaneous removal of
835–852.
multicomponent VOCs in biofilters, Trends Biotechnol. 36 (2018) 673–685.
[32] B. Balmelle, K.M. Nguyen, Bl Capdeville, J.C. Cornier, A. Deguin, Study of factors
[4] R.O. Carey, K.W. Migliaccio, Contribution of wastewater treatment plant
controlling nitrite build-up in biological processes for water nitrification, Water
effluents to nutrient dynamics in aquatic systems: a review, Environ. Manag. 44
Sci. Technol. 26 (1992) 1017–1025.
(2009) 205–217.
[33] D.-J. Kim, D.-I. Lee, J. Keller, Effect of temperature and free ammonia on
[5] B. Wang, Y. Guo, M. Zhao, B. Li, Y. Peng, Achieving energy-efficient nitrogen
nitrification and nitrite accumulation in landfill leachate and analysis of its
removal and excess sludge reutilization by partial nitritation and simultaneous
nitrifying bacterial community by FISH, Bioresour. Technol. 97 (2006) 459–468.
anammox denitrification and sludge fermentation process, Chemosphere 218
[34] Q. Sui, C. Liu, J. Zhang, H. Dong, Z. Zhu, Y. Wang, Response of nitrite
(2019) 705–714.
accumulation and microbial community to free ammonia and dissolved oxygen
[6] M. Azari, U. Walter, V. Rekers, J.-D. Gu, M. Denecke, More than a decade of
treatment of high ammonium wastewater, Appl. Microbiol. Biotechnol. 100
experience of landfill leachate treatment with a full-scale anammox plant
(2016) 4177–4187.
combining activated sludge and activated carbon biofilm, Chemosphere 174
[35] Q. Wang, L. Ye, G. Jiang, S. Hu, Z. Yuan, Side-stream sludge treatment using free
(2017) 117–126.
nitrous acid selectively eliminates nitrite oxidizing bacteria and achieves the
[7] J.G. Kuenen, Anammox bacteria: from discovery to application, Nat. Rev.
nitrite pathway, Water Res. 55 (2014) 245–255.
Microbiol. 6 (2008) 320–326.
[36] M. Negulescu, Municipal Waste Water Treatment, Elsevier, 2011.
[8] W.R.L. Van der Star, W.R. Abma, D. Blommers, J.-W. Mulder, T. Tokutomi,
[37] B. Wett, A. Omari, S.M. Podmirseg, M. Han, O. Akintayo, M. Gómez Brandón,
M. Strous, C. Picioreanu, M.C.M. van Loosdrecht, Startup of reactors for anoxic
S. Murthy, C. Bott, M. Hell, I. Takács, Going for mainstream deammonification
ammonium oxidation: experiences from the first full-scale anammox reactor in
from bench to full scale for maximized resource efficiency, Water Sci. Technol. 68
Rotterdam, Water Res. 41 (2007) 4149–4163.
(2013) 283–289.
[9] A. Gonzalez-Martinez, B. Muñoz-Palazon, A. Rodriguez-Sanchez, J. Gonzalez-
[38] R. Wen, Y. Jin, W. Zhang, Application of the anammox in China—a review, Int. J.
Lopez, New concepts in anammox processes for wastewater nitrogen removal:
Environ. Res. Public Health 17 (2020) 1090.
recent advances and future prospects, FEMS Microbiol. Lett. 365 (2018) fny031.
[39] I. Fernández, J. Dosta, J. Mata-Álvarez, A critical review of the future trends and
[10] H.J.M. Op den Camp, B. Kartal, D. Guven, L.A.M.P. Van Niftrik, S.C.M. Haaijer,
perspectives for the implementation of Anammox in the main line of municipal
W.R.L. Van Der Star, K.T. Van De Pas-Schoonen, A. Cabezas, Z. Ying, M.
WWTPs, Desalin. Water Treat. 57 (2016) 27890–27898.
C. Schmid, Global Impact and Application of the Anaerobic Ammonium-oxidizing
[40] G. Tchobanoglus, F. Burton, H.D. Stensel, Wastewater engineering: treatment and
(Anammox) Bacteria, Portland Press Ltd, 2006.
reuse, Am. Water Works Assoc. J. 95 (2003) 201.
[11] F.X. Tan, M. Huang, D.J. Wu, Z.L. Zhu, Research progress on the anammox
[41] S. Lackner, A. Terada, B.F. Smets, Heterotrophic activity compromises
technology. Advanced Materials Research, Trans Tech Publications, 2012,
autotrophic nitrogen removal in membrane-aerated biofilms: results of a
pp. 2015–2020.
modeling study, Water Res. 42 (2008) 1102–1112.
[12] R.J. Allgeier, W.H. Peterson, C. Juday, E.A. Birge, The anaerobic fermentation of
[42] G.T. Daigger, Oxygen and carbon requirements for biological nitrogen removal
lake deposits, Int. Rev. Gesamten Hydrobiol. Hydrogr. 26 (1932) 444–461.
processes accomplishing nitrification, nitritation, and anammox, Water Environ.
[13] F.A. Richards, Anoxic Basins and Fjords, Academic Press, 1965.
Res. 86 (2014) 204–209.
[14] M. Strous, J.A. Fuerst, E.H.M. Kramer, S. Logemann, G. Muyzer, K.T. Van de Pas-
[43] J. Frijns, J. Hofman, M. Nederlof, The potential of (waste) water as energy carrier,
Schoonen, R.I. Webb, J.G. Kuenen, M.S.M. Jetten, Missing lithotroph identified as
Energy Convers. Manag. 65 (2013) 357–363.
new planctomycete, Nature 400 (1999) 446–449.
[44] R.-C. Jin, G.-F. Yang, J.-J. Yu, P. Zheng, The inhibition of the Anammox process: a
[15] J.A.S. Guillén, P.R.C. Guardado, C.M.L. Vazquez, L.M. de Oliveira Cruz,
review, Chem. Eng. J. 197 (2012) 67–79.
D. Brdjanovic, J.B. Van Lier, Anammox cultivation in a closed sponge-bed
[45] N. Chamchoi, S. Nitisoravut, J.E. Schmidt, Inactivation of ANAMMOX
trickling filter, Bioresour. Technol. 186 (2015) 252–260.
communities under concurrent operation of anaerobic ammonium oxidation
[16] K. Egli, U. Fanger, P.J.J. Alvarez, H. Siegrist, J.R. van der Meer, A.J.B. Zehnder,
(ANAMMOX) and denitrification, Bioresour. Technol. 99 (2008) 3331–3336.
Enrichment and characterization of an anammox bacterium from a rotating
[46] D. Güven, A. Dapena, B. Kartal, M.C. Schmid, B. Maas, K. van de Pas-Schoonen,
biological contactor treating ammonium-rich leachate, Arch. Microbiol. 175
S. Sozen, R. Mendez, H.J.M.O. den Camp, M.S.M. Jetten, Propionate oxidation by
(2001) 198–207.

11
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

and methanol inhibition of anaerobic ammonium-oxidizing bacteria, Appl. Application, eco-physiology and biodiversity of anaerobic ammonium-oxidizing
Environ. Microbiol. 71 (2005) 1066–1071. bacteria, Rev. Environ. Sci. Bio/Technol. 3 (2004) 255–264.
[47] C. Chen, F. Sun, H. Zhang, J. Wang, Y. Shen, X. Liang, Evaluation of COD effect on [73] C.R. Penton, A.H. Devol, J.M. Tiedje, Molecular evidence for the broad
anammox process and microbial communities in the anaerobic baffled reactor distribution of anaerobic ammonium-oxidizing bacteria in freshwater and marine
(ABR), Bioresour. Technol. 216 (2016) 571–578. sediments, Appl. Environ. Microbiol. 72 (2006) 6829–6832.
[48] B. Kartal, J. Rattray, L.A. van Niftrik, J. van de Vossenberg, M.C. Schmid, R. [74] I. Tsushima, Y. Ogasawara, T. Kindaichi, H. Satoh, S. Okabe, Development of
I. Webb, S. Schouten, J.A. Fuerst, J.S. Damsté, M.S.M. Jetten, Candidatus high-rate anaerobic ammonium-oxidizing (anammox) biofilm reactors, Water
“Anammoxoglobus propionicus” a new propionate oxidizing species of anaerobic Res. 41 (2007) 1623–1634.
ammonium oxidizing bacteria, Syst. Appl. Microbiol. 30 (2007) 39–49. [75] J. van de Vossenberg, D. Woebken, W.J. Maalcke, H.J.C.T. Wessels, B.E. Dutilh,
[49] B. Kartal, L.A. Van Niftrik, J. Rattray, J.L.C.M. Van De Vossenberg, M.C. Schmid, B. Kartal, E.M. Janssen-Megens, G. Roeselers, J. Yan, D. Speth, The metagenome
J. Sinninghe Damsté, M.S.M. Jetten, M. Strous, Candidatu s ‘Brocadia fulgida’: an of the marine anammox bacterium ‘Candidatus Scalindua profunda’illustrates the
autofluorescent anaerobic ammonium oxidizing bacterium, FEMS Microbiol. versatility of this globally important nitrogen cycle bacterium, Environ.
Ecol. 63 (2008) 46–55. Microbiol. 15 (2013) 1275–1289.
[50] Y. Miao, Y. Peng, L. Zhang, B. Li, X. Li, L. Wu, S. Wang, Partial nitrification- [76] M.M.M. Kuypers, A.O. Sliekers, G. Lavik, M.C. Schmid, B.B. Jørgensen, J.
anammox (PNA) treating sewage with intermittent aeration mode: effect of G. Kuenen, J.S.S. Damsté, M. Strous, M.S.M. Jetten, Anaerobic ammonium
influent C/N ratios, Chem. Eng. J. 334 (2018) 664–672. oxidation by anammox bacteria in the Black Sea, Nature 422 (2003) 608–611.
[51] H. Chen, S. Liu, F. Yang, Y. Xue, T. Wang, The development of simultaneous [77] M. Ali, M. Oshiki, T. Awata, K. Isobe, Z. Kimura, H. Yoshikawa, D. Hira,
partial nitrification, ANAMMOX and denitrification (SNAD) process in a single T. Kindaichi, H. Satoh, T. Fujii, Physiological characterization of anaerobic
reactor for nitrogen removal, Bioresour. Technol. 100 (2009) 1548–1554. ammonium oxidizing bacterium ‘Candidatus Jettenia caeni’, Environ. Microbiol.
[52] M. Strous, J.J. Heijnen, J.G. Kuenen, M.S.M. Jetten, The sequencing batch reactor 17 (2015) 2172–2189.
as a powerful tool for the study of slowly growing anaerobic ammonium- [78] H. Ma, Y. Zhang, Y. Xue, Y.-Y. Li, A new process for simultaneous nitrogen
oxidizing microorganisms, Appl. Microbiol. Biotechnol. 50 (1998) 589–596. removal and phosphorus recovery using an anammox expanded bed reactor,
[53] C. Fux, M. Boehler, P. Huber, I. Brunner, H. Siegrist, Biological treatment of Bioresour. Technol. 267 (2018) 201–208.
ammonium-rich wastewater by partial nitritation and subsequent anaerobic [79] S.W.H. Van Hulle, H.J.P. Vandeweyer, B.D. Meesschaert, P.A. Vanrolleghem,
ammonium oxidation (anammox) in a pilot plant, J. Biotechnol. 99 (2002) P. Dejans, A. Dumoulin, Engineering aspects and practical application of
295–306. autotrophic nitrogen removal from nitrogen rich streams, Chem. Eng. J. 162
[54] T.L.G. Hendrickx, Y. Wang, C. Kampman, G. Zeeman, H. Temmink, C.J. (2010) 1–20.
N. Buisman, Autotrophic nitrogen removal from low strength waste water at low [80] J. Dosta, I. Fernández, J.R. Vázquez-Padín, A. Mosquera-Corral, J.L. Campos,
temperature, Water Res. 46 (2012) 2187–2193. J. Mata-Alvarez, R. Méndez, Short-and long-term effects of temperature on the
[55] A. Gonzalez-Martinez, B. Muñoz-Palazon, A. Rodriguez-Sanchez, J. Gonzalez- Anammox process, J. Hazard. Mater. 154 (2008) 688–693.
Lopez, New concepts in anammox processes for wastewater nitrogen removal: [81] M. Waki, T. Tokutomi, H. Yokoyama, Y. Tanaka, Nitrogen removal from animal
recent advances and future prospects, FEMS Microbiol. Lett. 365 (2018) fny031. waste treatment water by anammox enrichment, Bioresour. Technol. 98 (2007)
[56] A.V. del Rio, A. Pichel, N. Fernandez-Gonzalez, A. Pedrouso, A. Fra-Vázquez, 2775–2780.
N. Morales, R. Mendez, J.L. Campos, A. Mosquera-Corral, Performance and [82] H. Bae, T. Paul, D. Kim, J.-Y. Jung, Specific ANAMMOX activity (SAA) in a
microbial features of the partial nitritation-anammox process treating fish sequencing batch reactor: optimization test with statistical comparison, Environ.
canning wastewater with variable salt concentrations, J. Environ. Manag. 208 Earth Sci. 75 (2016) 1452.
(2018) 112–121. [83] I. Fernández, J. Dosta, C. Fajardo, J.L. Campos, A. Mosquera-Corral, R. Méndez,
[57] M. Laureni, P. Falås, O. Robin, A. Wick, D.G. Weissbrodt, J.L. Nielsen, T. Short-and long-term effects of ammonium and nitrite on the Anammox process,
A. Ternes, E. Morgenroth, A. Joss, Mainstream partial nitritation and anammox: J. Environ. Manag. 95 (2012) S170–S174.
long-term process stability and effluent quality at low temperatures, Water Res. [84] K. Isaka, T. Sumino, S. Tsuneda, High nitrogen removal performance at
101 (2016) 628–639. moderately low temperature utilizing anaerobic ammonium oxidation reactions,
[58] C. Reino, M.E. Suárez-Ojeda, J. Pérez, J. Carrera, Stable long-term operation of an J. Biosci. Bioeng. 103 (2007) 486–490.
upflow anammox sludge bed reactor at mainstream conditions, Water Res. 128 [85] D. Puyol, J.M. Carvajal-Arroyo, R. Sierra-Alvarez, J.A. Field, Nitrite (not free
(2018) 331–340. nitrous acid) is the main inhibitor of the anammox process at common pH
[59] Y. Miao, J. Zhang, Y. Peng, S. Wang, An improved start-up strategy for conditions, Biotechnol. Lett. 36 (2014) 547–551.
mainstream anammox process through inoculating ordinary nitrification sludge [86] Q. Niu, S. He, Y. Zhang, H. Ma, Y. Liu, Y.-Y. Li, Process stability and the recovery
and a small amount of anammox sludge, J. Hazard. Mater. 384 (2020), 121325. control associated with inhibition factors in a UASB-anammox reactor with a
[60] T. Lotti, R. Kleerebezem, C. Lubello, M.C.M. Van Loosdrecht, Physiological and long-term operation, Bioresour. Technol. 203 (2016) 132–141.
kinetic characterization of a suspended cell anammox culture, Water Res. 60 [87] M. Strous, E. Van Gerven, J.G. Kuenen, M. Jetten, Effects of aerobic and
(2014) 1–14. microaerobic conditions on anaerobic ammonium-oxidizing (anammox) sludge,
[61] A. Rosenthal, K. Ramalingam, H. Park, A. Deur, K. Beckmann, K. Chandran, Appl. Microbiol. Biotechnol. 63 (1997) 2446–2448.
J. Fillos, Anammox studies using New York city centrate to correlate [88] S. He, Y. Zhang, Q. Niu, H. Ma, Y.-Y. Li, Operation stability and recovery
performance, population dynamics and impact of toxins, Proc. Water Environ. performance in an Anammox EGSB reactor after pH shock, Ecol. Eng. 90 (2016)
Fed. 2009 (2009) 120–128. 50–56.
[62] M. Strous, J.G. Kuenen, M.S.M. Jetten, Key physiology of anaerobic ammonium [89] B. Zhang, J. Zhao, J. Zuo, X. Shi, J. Gong, H. Ren, Realizing stable operation of
oxidation, Appl. Microbiol. Biotechnol. 65 (1999) 3248–3250. anaerobic ammonia oxidation at low temperatures treating low strength synthetic
[63] S.H. Peeters, L. van Niftrik, Trending topics and open questions in anaerobic wastewater, J. Environ. Sci. 75 (2019) 193–200.
ammonium oxidation, Curr. Opin. Chem. Biol. 49 (2019) 45–52. [90] R.P. Taylor, C.L.W. Jones, R.K. Laubscher, Recovery of methane and adding value
[64] A. Dietl, C. Ferousi, W.J. Maalcke, A. Menzel, S. de Vries, J.T. Keltjens, M.S. to the digestate of biomass produced by high rate algal ponds or waste activated
M. Jetten, B. Kartal, T.R.M. Barends, The inner workings of the hydrazine sludge, used to treat brewery effluent, J. Water Process Eng. 40 (2021), 101797.
synthase multiprotein complex, Nature 527 (2015) 394–397. [91] P.L. McCarty, What is the best biological process for nitrogen removal: when and
[65] W.J. Maalcke, J. Reimann, S. de Vries, J.N. Butt, A. Dietl, N. Kip, U. Mersdorf, T. why? Environ. Sci. Technol. 52 (2018) 3835–3841.
R.M. Barends, M.S.M. Jetten, J.T. Keltjens, Characterization of anammox [92] A. Malovanyy, J. Yang, J. Trela, E. Plaza, Combination of upflow anaerobic
hydrazine dehydrogenase, a key N2-producing enzyme in the global nitrogen sludge blanket (UASB) reactor and partial nitritation/anammox moving bed
cycle, J. Biol. Chem. 291 (2016) 17077–17092. biofilm reactor (MBBR) for municipal wastewater treatment, Bioresour. Technol.
[66] M.C.F. van Teeseling, N.M. de Almeida, A. Klingl, D.R. Speth, H.J.M.O. den Camp, 180 (2015) 144–153.
R. Rachel, M.S.M. Jetten, L. van Niftrik, A new addition to the cell plan of [93] J. Jimenez, M. Miller, C. Bott, S. Murthy, H. De Clippeleir, B. Wett, High-rate
anammox bacteria:“Candidatus Kuenenia stuttgartiensis” has a protein surface activated sludge system for carbon management–evaluation of crucial process
layer as the outermost layer of the cell, J. Bacteriol. 196 (2014) 80–89. mechanisms and design parameters, Water Res. 87 (2015) 476–482.
[67] B. Kartal, W.J. Maalcke, N.M. de Almeida, I. Cirpus, J. Gloerich, W. Geerts, H.J.M. [94] T. Lotti, R. Kleerebezem, Z. Hu, B. Kartal, M.K. De Kreuk, C.S. van Erp Taalman
O. den Camp, H.R. Harhangi, E.M. Janssen-Megens, K.-J. Francoijs, Molecular Kip, J. Kruit, T.L.G. Hendrickx, M.C.M. Van Loosdrecht, Pilot-scale evaluation of
mechanism of anaerobic ammonium oxidation, Nature 479 (2011) 127–130. anammox-based mainstream nitrogen removal from municipal wastewater,
[68] P. Sonthiphand, M.W. Hall, J.D. Neufeld, Biogeography of anaerobic ammonia- Environ. Technol. 36 (2015) 1167–1177.
oxidizing (anammox) bacteria, Front. Microbiol. 5 (2014) 399. [95] Y. Yang, L. Zhang, H. Shao, S. Zhang, P. Gu, Y. Peng, Enhanced nutrients removal
[69] T.R.V. Akaboci, F. Gich, M. Ruscalleda, M.D. Balaguer, J. Colprim, Assessment of from municipal wastewater through biological phosphorus removal followed by
operational conditions towards mainstream partial nitritation-anammox stability partial nitritation/anammox, Front. Environ. Sci. Eng. 11 (2017) 8.
at moderate to low temperature: reactor performance and bacterial community, [96] V. Diamantis, W. Verstraete, A. Eftaxias, B. Bundervoet, V. Siegfried, P. Melidis,
Chem. Eng. J. 350 (2018) 192–200. A. Aivasidis, Sewage pre-concentration for maximum recovery and reuse at
[70] J. Gu, Q. Yang, Y. Liu, Mainstream anammox in a novel A-2B process for energy- decentralized level, Water Sci. Technol. 67 (2013) 1188–1193.
efficient municipal wastewater treatment with minimized sludge production, [97] W.-W. Li, H.-Q. Yu, Z. He, Towards sustainable wastewater treatment by using
Water Res. 138 (2018) 1–6. microbial fuel cells-centered technologies, Energy Environ. Sci. 7 (2014)
[71] Y. Yang, L. Zhang, J. Cheng, S. Zhang, X. Li, Y. Peng, Microbial community 911–924.
evolution in partial nitritation/anammox process: from sidestream to [98] Z. Ge, Z. He, Long-term performance of a 200 liter modularized microbial fuel cell
mainstream, Bioresour. Technol. 251 (2018) 327–333. system treating municipal wastewater: treatment, energy, and cost, Environ. Sci.
[72] B. Kartal, L. Van Niftrik, O. Sliekers, M.C. Schmid, I. Schmidt, K. Van De Pas- Water Res. Technol. 2 (2016) 274–281.
Schoonen, I. Cirpus, W. Van der Star, M. Van Loosdrecht, W.R. Abma,

12
H.P. Trinh et al. Journal of Environmental Chemical Engineering 9 (2021) 105583

[99] Y. Yang, X. Li, X. Yang, Z. He, Enhanced nitrogen removal by membrane-aerated [116] I. Dimitrova, A. Dabrowska, S. Ekström, Start-up of a full-scale partial nitritation-
nitritation-anammox in a bioelectrochemical system, Bioresour. Technol. 238 anammox MBBR without inoculum at Klagshamn WWTP, Water Sci. Technol. 81
(2017) 22–29. (2020) 2033–2042.
[100] R.-C. Jin, G.-F. Yang, Q.-Q. Zhang, C. Ma, J.-J. Yu, B.-S. Xing, The effect of sulfide [117] J. Li, Y. Peng, L. Zhang, J. Liu, X. Wang, R. Gao, L. Pang, Y. Zhou, Quantify the
inhibition on the ANAMMOX process, Water Res. 47 (2013) 1459–1469. contribution of anammox for enhanced nitrogen removal through metagenomic
[101] N. Morales, A. Val del Rio, J.R. Vázquez-Padín, R. Gutiérrez, R. Fernández- analysis and mass balance in an anoxic moving bed biofilm reactor, Water Res.
González, P. Icaran, F. Rogalla, J.L. Campos, R. Méndez, A. Mosquera-Corral, 160 (2019) 178–187.
Influence of dissolved oxygen concentration on the start-up of the anammox- [118] B. Ma, X. Xu, Y. Wei, C. Ge, Y. Peng, Recent advances in controlling denitritation
based process: ELAN®, Water Sci. Technol. 72 (2015) 520–527. for achieving denitratation/anammox in mainstream wastewater treatment
[102] Z. Hu, T. Lotti, M. de Kreuk, R. Kleerebezem, M.C.M. van Loosdrecht, J. Kruit, M. plants, Bioresour. Technol. 299 (2020), 122697.
S.M. Jetten, B. Kartal, Nitrogen removal by a nitritation-anammox bioreactor at [119] B. Wett, S.M. Podmirseg, M. Gómez-Brandón, M. Hell, G. Nyhuis, C. Bott,
low temperature, Appl. Environ. Microbiol. 79 (2013) 2807–2812. S. Murthy, Expanding DEMON sidestream deammonification technology towards
[103] T. Lotti, R. Kleerebezem, M.C.M. Van Loosdrecht, Effect of temperature change on mainstream application, Water Environ. Res. 87 (2015) 2084–2089.
anammox activity, Biotechnol. Bioeng. 112 (2015) 98–103. [120] H. De Clippeleir, N. Weissenbacher, T. Schaubroeck, M. Hell, P. Boeckx, N. Boon,
[104] B. Ma, S. Zhang, L. Zhang, P. Yi, J. Wang, S. Wang, Y. Peng, The feasibility of B. Wett, Mainstream partial nitritation/anammox: balancing overall
using a two-stage autotrophic nitrogen removal process to treat sewage, sustainability with energy savings, in: WEFTEC, New Orleans/Los Angeles, 2012.
Bioresour. Technol. 102 (2011) 8331–8334. [121] B. Wett, G. Nyhuis, I. Takács, S. Murthy, Development of enhanced
[105] H. Siegrist, D. Salzgeber, J. Eugster, A. Joss, Anammox brings WWTP closer to deammonification selector, in: Water Environment Federation, WEFTEC, 2010,
energy autarky due to increased biogas production and reduced aeration energy pp. 2–6.
for N-removal, Water Sci. Technol. 57 (2008) 383–388. [122] A. Kamp, L.D.M. Ottosen, N.B. Thøgersen, N.P. Revsbech, B. Thamdrup, M.
[106] P. Roots, A.F. Rosenthal, Q. Yuan, Y. Wang, F. Yang, J.A. Kozak, H. Zhang, G. H. Andersen, Anammox and partial nitritation in the mainstream of a wastewater
F. Wells, Optimization of the carbon to nitrogen ratio for mainstream treatment plant in a temperate region (Denmark), Water Sci. Technol. 79 (2019)
deammonification and the resulting shift in nitrification from biofilm to 1397–1405.
suspension, Environ. Sci. Water Res. Technol. 6 (2020) 3415–3427. [123] M. O’Shaughnessy, Mainstream deammonification, Water Environment Research
[107] M. Laureni, D.G. Weissbrodt, K. Villez, O. Robin, N. De Jonge, A. Rosenthal, Foundation Alexandria, VA, 2015.
G. Wells, J.L. Nielsen, E. Morgenroth, A. Joss, Biomass segregation between [124] C. Yeshi, K.B. Hong, M.C.M. van Loosdrecht, G.T. Daigger, P.H. Yi, Y.L. Wah, C.
biofilm and flocs improves the control of nitrite-oxidizing bacteria in mainstream S. Chye, Y.A. Ghani, Mainstream partial nitritation and anammox in a 200,000
partial nitritation and anammox processes, Water Res. 154 (2019) 104–116. m3/day activated sludge process in Singapore: scale-down by using laboratory
[108] K. Isaka, Y. Suwa, Y. Kimura, T. Yamagishi, T. Sumino, S. Tsuneda, Anaerobic fed-batch reactor, Water Sci. Technol. 74 (2016) 48–56.
ammonium oxidation (anammox) irreversibly inhibited by methanol, Appl. [125] R. Lemaire, H. Zhao, C. Thomson, M. Christensson, S. Piveteau, S. Hemmingsen,
Microbiol. Biotechnol. 81 (2008) 379–385. F. Veuillet, P. Zozor, J. Ochoa, Mainstream deammonification with ANITATM
[109] S.E. Vlaeminck, A. Terada, B.F. Smets, H. De Clippeleir, T. Schaubroeck, S. Bolca, Mox process, Proc. Water Environ. Fed. 2014 (2014) 2183–2197.
L. Demeestere, J. Mast, N. Boon, M. Carballa, W. Verstraete, Aggregate size and [126] T. Liu, S. Hu, J. Guo, Enhancing mainstream nitrogen removal by employing
architecture determine microbial activity balance for one-stage partial nitritation nitrate/nitrite-dependent anaerobic methane oxidation processes, Crit. Rev.
and anammox, Appl. Environ. Microbiol. 76 (2010) 900–909. Biotechnol. 39 (2019) 732–745.
[110] H. De Clippeleir, R. Jimenez, E. Giraldo, B. Wett, N. Dockett, R. Riffat, A.A.O.S. [127] G.-J. Xie, T. Liu, C. Cai, S. Hu, Z. Yuan, Achieving high-level nitrogen removal in
Murthy, Screens as a method for selective anammox retention in single stage mainstream by coupling anammox with denitrifying anaerobic methane oxidation
deammonification processes, WEF/IWA Nutrient Removal and Recovery, in: in a membrane biofilm reactor, Water Res. 131 (2018) 196–204.
Presented at the WEF/IWA Nutrient Removal and Recovery, Canada, 2013. [128] E. Isanta, T. Bezerra, I. Fernández, M.E. Suárez-Ojeda, J. Pérez, J. Carrera,
[111] M. Han, S.E. Vlaeminck, A. Al-Omari, B. Wett, C. Bott, S. Murthy, H. De Clippeleir, Microbial community shifts on an anammox reactor after a temperature shock
Uncoupling the solids retention times of flocs and granules in mainstream using 454-pyrosequencing analysis, Bioresour. Technol. 181 (2015) 207–213.
deammonification: a screen as effective out-selection tool for nitrite oxidizing [129] U. Manonmani, K. Joseph, Research advances and challenges in anammox
bacteria, Bioresour. Technol. 221 (2016) 195–204. immobilization for autotrophic nitrogen removal, J. Chem. Technol. Biotechnol.
[112] T. Van Winckel, S.E. Vlaeminck, A. Al-Omari, B. Bachmann, B. Sturm, B. Wett, 93 (2018) 2486–2497.
I. Takács, C. Bott, S.N. Murthy, H. De Clippeleir, Screen versus cyclone for [130] I. Zekker, A. Kivirüüt, E. Rikmann, A. Mandel, M. Jaagura, T. Tenno,
improved capacity and robustness for sidestream and mainstream O. Artemchuk, Sd.C. Rubin, T. Tenno, Enhanced efficiency of nitritating-
deammonification, Environ. Sci. Water Res. Technol. 5 (2019) 1769–1781. anammox sequencing batch reactor achieved at low decrease rates of
[113] S.-Q. Ni, J. Zhang, Anaerobic ammonium oxidation: from laboratory to full-scale oxidation–reduction potential, Environ. Eng. Sci. 36 (2019) 350–360.
application, BioMed Res. Int. 2013 (2013) 1–10. [131] B. Ma, Y. Peng, S. Zhang, J. Wang, Y. Gan, J. Chang, S. Wang, S. Wang, G. Zhu,
[114] M.S.M. Jetten, M. Wagner, J. Fuerst, M.C.M. van Loosdrecht, G. Kuenen, Performance of anammox UASB reactor treating low strength wastewater under
M. Strous, Microbiology and application of the anaerobic ammonium oxidation moderate and low temperatures, Bioresour. Technol. 129 (2013) 606–611.
(‘anammox’) process, Curr. Opin. Biotechnol. 12 (2001) 283–288. [132] M.J. García-Ruiz, P. Maza-Márquez, J. González-López, F. Osorio, Nitrogen
[115] M. Hoekstra, S.P. Geilvoet, T.L.G. Hendrickx, C.S. van Erp Taalman Kip, removal capacity and bacterial community dynamics of a Canon biofilter system
R. Kleerebezem, M.C.M. van Loosdrecht, Towards mainstream anammox: lessons at different organic matter concentrations, Chemosphere 193 (2018) 591–601.
learned from pilot-scale research at WWTP Dokhaven, Environ. Technol. 40
(2019) 1721–1733.

13

You might also like