You are on page 1of 14

Bioresource Technology 361 (2022) 127702

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Review

Simultaneous nitrification–denitrification in biofilm systems for wastewater


treatment: Key factors, potential routes, and engineered applications
Francesco Di Capua a, *, Francesca Iannacone b, Fabrizio Sabba c, Giovanni Esposito d
a
Department of Civil, Environmental, Land, Building Engineering and Chemistry, Polytechnic University of Bari, Bari 70125, Italy
b
Acqua & Sole s.r.l., Via Giulio Natta, Vellezzo Bellini, PV, Italy
c
Black & Veatch, KS, United States
d
Department of Civil, Architectural and Environmental Engineering, University of Naples Federico II, Via Claudio 21, Naples 80125, Italy

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Simultaneous nitrification–denitrification
(SND) is emerging for water treatment.
• Shortcut SND can cut organic and en­
ergy consumption as well as CO2 emis­
sion by >20 %.
• Nitrous oxide emission can be reduced
by adjusting operational and physical
factors.
• Moving bed and aerobic granular bio­
films are mature technologies for effi­
cient SND.
• SND biofilms can enable the combined
removal of nitrogen and phosphorus.

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

Keywords: Simultaneous nitrification–denitrification (SND) is an advantageous bioprocess that allows the complete removal
Simultaneous nitrification denitrification of ammonia nitrogen through sequential redox reactions leading to nitrogen gas production. SND can govern
Biofilm nitrogen removal in single-stage biofilm systems, such as the moving bed biofilm reactor and aerobic granular
Nitrous oxide
sludge system, as oxygen gradients allow the development of multilayered biofilms including nitrifying and
Moving bed biofilm reactor
Aerobic granular sludge
denitrifying bacteria. Environmental and operational conditions can strongly influence SND performance, bio­
film development and biochemical pathways. Recent advances have outlined the possibility to reduce the carbon
and energy consumption of the process via the “shortcut pathway”, and simultaneously remove both N and
phosphorus under specific operational conditions, opening new possibilities for wastewater treatment. This work
critically reviews the factors influencing SND and its application in biofilm systems from laboratory to full scale.
Operational strategies to enhance SND efficiency and hints to reduce nitrous oxide emission and operational costs
are provided.

* Corresponding author.
E-mail address: francesco.dicapua@poliba.it (F. Di Capua).

https://doi.org/10.1016/j.biortech.2022.127702
Received 1 July 2022; Received in revised form 20 July 2022; Accepted 22 July 2022
Available online 26 July 2022
0960-8524/© 2022 Elsevier Ltd. All rights reserved.
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

1. Introduction conditions by denitrifying bacteria (DNB) using organic carbon as en­


ergy source and producing alkalinity.
Nitrogen (N) is an essential nutrient for all living forms (Chrispim Both nitrification and denitrification produce N2O, a potent green­
et al., 2019). However, excessive discharge of N in surface waters can house gas, as a process intermediate. During nitrification, N2O may be
determine an abnormal growth of algae and, consequently, a depletion produced by AOB via two main pathways: reduction of NO–2 as terminal
of dissolved oxygen (DO) (Butcher, 2016). This phenomenon, known as electron acceptor to N2O (pathway known as nitrifier denitrification)
eutrophication, represents a serious environmental problem affecting and incomplete oxidation of hydroxylamine (NH2OH), which is first
water quality in rivers, lakes, and estuaries around the world. Phos­ chemically decomposed to NO, and then biologically reduced to N2O
phorus (P) can also contribute to eutrophication by increasing higher (hydroxylamine pathway) (Sabba et al., 2015). Similar to nitrification,
primary production in superficial water bodies. The main sources of N incomplete denitrification can also lead to the accumulation and emis­
and P in surface waters are represented by the run-off from agricultural sion of N2O as a process intermediate (Zhou et al., 2022a). Differently
land and wastewater discharge from households and industry (van from nitrification, N2O can be produced but also consumed during
Puijenbroek et al., 2019). Biological N removal (BNR) can effectively denitrification (Sabba et al., 2017b).
remove N compounds from wastewaters and is recognized as the most BNR in WWTPs is generally achieved via a pre-denitrification cycle,
economical process to reach the stringent discharge limits set by the where heterotrophic denitrification is followed by an aerobic phase for
national legislations (He et al., 2020). Similarly, P can be removed the combined oxidation of organics and NH+ 4 (Sun et al., 2010). How­
biologically or chemically, the former being more complex to implement ever, this approach is considered energy-demanding, especially due to
but also eco-sustainable, the latter being generally easier to apply but aeration, and economically unsustainable for the treatment of waste­
more expensive and less eco-friendly (Di Capua et al., 2022). water produced in small communities or rural areas due to large struc­
Simultaneous nitrification–denitrification (SND) is capable to tural footprints and energy cost, blower deterioration and maintenance,
completely remove N in a single bioreactor under specific operating and high sludge production (Yan et al., 2019a). In recent years, novel
conditions, which differentiates this process from sequential nitrifica­ biological processes, i.e., anaerobic NH+4 oxidation (anammox) and SND,
tion and denitrification typically carried out in separate bioreactors at have been increasingly investigated with the aim to reduce the WWTP
municipal wastewater treatment plants (WWTPs). Up to date, many operating costs while ensuring compliance with N discharge limits in
researchers have studied the underlying mechanisms and the effects of water bodies. Anammox bacteria (AnAOB) oxidize NH+ 4 directly to N2
various environmental factors and operating conditions on SND and its under anoxic conditions, obtaining energy for growth from the oxidation
combination with P removal, including the feed carbon-to-nitrogen (C/ of NO–2 to NO–3 (Hu et al., 2019). NO–2 can be produced by partial nitri­
N) ratio, DO level, hydraulic retention time (HRT) and aeration pattern. fication (nitritation) of NH+ 4 or through partial denitrification (deni­
Recent studies show that biofilm technologies can effectively promote tritation) of NO–3. Despite the various advantages attributed to anammox
SND as well as the combined removal of N and P under specific condi­ in terms of reduced energy consumption and sludge production (Rahimi
tions (Iannacone et al., 2021; Roots et al., 2020). However, critical et al., 2020), anammox application to mainstream wastewater treatment
evaluation and comparison of these technologies are needed to point out is limited by 1) the low maximum specific growth rate of AnAOB
their strengths and weaknesses and promote SND application at full (0.05–0.33 d− 1) (Zhang et al., 2017a), 2) excess discharge of NO–3 in the
scale. effluent, 3) high operational temperatures (i.e., 30–40 ◦ C), and 4) the
This work discusses the fundamental mechanisms behind the SND need for a strict control of temperature, pH, and organic concentration
process and provides a comprehensive and critical overview of SND for optimal AnAOB growth.
applications in biofilm reactors. Advantages and criticalities of each SND represents a suitable and promising alternative to pre-
technology are evaluated, and guidelines for a proper selection of denitrification in WWTPs for the simultaneous removal of C and N as
reactor and operating conditions to implement a successful SND process it offers several advantages: (1) the carbon demand and sludge pro­
are provided. Knowledge gaps and research needs are identified with the duction are reduced by over 30 % (Ma et al., 2017), (2) alkalinity supply
objective to unravel the potential of SND in biofilm reactors. by denitrification helps to maintain a circumneutral pH, (3) there is no
need for NO–3 recirculation, (4) lower energy for aeration (Zinatizadeh
2. Overview of nitrogen removal bioprocesses and Ghaytooli, 2015), and (5) a smaller footprint is required (Seifi and
Fazaelipoor, 2012). Potential disadvantages of the SND process include
Nitrification is a biological process carried out under aerobic con­ (1) lower N removal compared to separate denitrification and nitrifi­
ditions and typically consists of two consequential oxidation steps. In the cation, (2) significant N2O accumulation, and (3) process instability due
first step, ammonia oxidizing bacteria (AOB) oxidize ammonium (NH+ 4) to competition among the different microbial families coexisting in the
to nitrite (NO–2) by consuming 1.5 mol of O2 per mol of N. In this reac­ system. In the last 30 years, the SND process has gained significant
tion, the enzymes ammonia monooxygenase (AMO) and hydroxylamine attention from the scientific community and WWTP operators. Biofilm
oxidoreductase (HAO) are involved, along with hydroxylamine technologies have been increasingly developed and applied in full-scale
(NH2OH) as an intermediate product. During the second step, nitrite WWTPs as they offer several advantages over suspended-growth sys­
oxidizing bacteria (NOB) use the enzyme nitrite oxidoreductase (NXR) tems: (1) higher biomass concentration, (2) lower space requirements,
to convert NO–2 to NO–3 and consume 0.5 mol of O2 per mol of N (Rahimi (3) reduced HRT and sludge production and (4) more stable perfor­
et al., 2020). AOB and NOB typically use only inorganic carbon sources, mance (Zhao et al., 2019). In addition, biofilm systems allow the coex­
consuming almost 7.1 mg CaCO3 for each mg N-NH+ 4 oxidized to N-NO3

istence of a diverse and complex microbial community involved in
(Rittmann and McCarty, 2012). Besides AOB and NOB, bacteria capable nutrient removal, which makes these systems perfectly suitable for the
of complete NH+ 4 oxidation to NO3, i.e., Comammox, have been

SND process. The most utilized biofilm-based technologies applied for
discovered in 2015 (Van Kessel et al., 2015) and found to be the SND include the moving bed biofilm reactor (MBBR) (Tables 1 and 2),
dominant nitrifying microorganisms in mainstream low-oxygen nitrifi­ hybrid biofilm-membrane bioreactor (HMBR) systems, and aerobic
cation reactors (Roots et al., 2019). Denitrification involves the granular sludge (AGS) (Table 3).
sequential reduction of NO–3 to NO–2, nitric oxide (NO), nitrous oxide
(N2O), and N2. This process is typically carried out under anoxic

2
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

3. Factors influencing simultaneous nitrification–denitrification

Zinatizadeh and Ghaytooli (2015)


in biofilms

Sandip and Kalyanraman (2019)


3.1. Environmental factors

Chu and Wang (2011)


Zhang et al. (2017b)
pH and temperature can influence both the efficiency and pathway

Song et al. (2019)

Feng et al. (2012)


Liu et al. (2020a)
Liu et al. (2020a)

Gu et al. (2018)

Fu et al. (2010)
of nitrification. AOB can outcompete NOB at temperatures >25 ◦ C,
while NOB grow much faster at lower temperatures (Liu et al., 2021;
Filling ratio (%) HRT (h) DO (mg L-1) Carbon source Feed TN (mg L-1) Feed C/N COD RE (%) TN RE (%) Reference

Sabba et al., 2022). Optimal pH levels are reported in the range of


8.2–8.4 for AOB and 7.2–7.9 for NOB (Gu et al., 2012). Therefore,
relatively high temperature and pH values can help AOB to outcompete
NOB and maintain partial nitrification, being required for shortcut SND
62–76
45–72
45–78
43–62
34–42
47–59

85–95
26–46
26–50
25–54
20–55

26–51
(see Section 4). It should be noted that AOB and NOB consume alkalinity
≤30
84
75

59
due to release of H+ during nitrification (Rittmann and McCarty, 2012).
In SND systems, acidity production by nitrification is counterbalanced
by the supply of alkalinity due to denitrification, which helps to keep a
79–86

86–91

50–88
54–85

94–96
>94c
>94c

circumneutral pH.
<90
94c

90c
72c
85
81
83

81

3.7b, 8.1b
3.7b, 8.1b
4.5–13.4

3.2. Operational parameters


~11b
≤ 5b
≤ 5b
5-9b
5-9b

6.3b
3.5
3.5

7.1
5b
5b
5
5

3.2.1. Dissolved oxygen concentration


DO concentration can significantly impact microbial activity and
community composition within the biofilm. SND is based on the co-
existence of AOB, NOB and DNB in the biofilm, achieved through DO
40 ± 10
40 ± 10

25(±3)
44(±5)
44(±5)
30–70
30–70

28–33
28–33

concentration gradients generated by diffusional limitations (Liu et al.,


100
40
40

51
16

2020a) (Fig. 1). The DO concentration in the bulk liquid can be regu­

lated to create gradients determining the formation of an anoxic micro-


BOD5 Glucose
BOD5 Glucose

environment within the biofilm where denitrification can occur, while


Glucose
Glucose
Glucose
Glucose

Glucose
Glucose

Glucose
Glucose
Glucose

nitrifying bacteria can thrive in the external region where DO is more


BOD5
BOD5

available (Massoompour et al., 2020).



The suitable DO concentration reported for maintaining SND varies


over a wide range (0.8–7.0 mg L− 1) and is affected by several factors
0.6–0.8
0.6–0.8

5.0–6.5
5.0–6.5

5.0–6.5
2.0–4.0
2.0–4.0
4.0–6.0

2.0–4.0

that include carbon source availability, reactor configuration and


0.75
0.75
3.0
3.0

operation, biofilm thickness, granule size and carrier type (Tables 1–3).


The concentration of DO in the bulk liquid affects its penetration depth


SND performance and operating conditions of continuous-flow MBBRs treating municipal wastewater.

in the biofilm. The set DO concentration must ensure a balance between


4–12
4–12
5–7
12
12
10
12

14
14
10

the bacterial communities involved in SND and between the production


8
8
8
8
5
5

and consumption of NOx. High DO concentration in the bulk liquid can


limit the formation of the anoxic zone or push the zone deeper, while
low DO concentrations can limit nitrification (Layer et al., 2020; Wang
20–40
20–40

10–30

20–40

et al., 2020a). Cao et al. (2017) investigated the effect of DO concen­


16.7
30
30
35
35

10
10
30

50
50

20

30

trations ranging from 1.5 to 5.5 mg L− 1 in five identical moving bed


PE–PQAS 10–Fe2O3

PE–PQAS 10–Fe2O3

sequencing batch reactors (MBBR-SBRs) (Table 2). After an acclimati­


PU–zeolite (SiO2)

zation period, the highest total nitrogen (TN) removal efficiency (RE)
was observed at a DO concentration of 2.5 mg L− 1 (84 %), while COD RE
Specific surface (m2 m-3) Materials

was >90 % in all systems. Higher DO concentrations (3.5–5.5 mg L− 1)


inhibited denitrification, while the relative abundance of AOB and NOB
PCL
PU

PU

PU

PU
PU
PE

PE
PE

PE
PE

PE

increased. Lower DO concentration (1.5 mg L− 1) limited nitrification,


resulting in the lowest relative abundance of AOB in the biofilm and in a
significant increase of N-NH+ 4 concentration in the effluent (up to 13.8
mg L− 1).
In continuous-flow MBBRs, promising SND performances have been
measured as total organic carbon (TOC) RE.

achieved under microaerobic conditions, i.e., DO levels ≤1.0 mg L− 1


560–600
560–600
560–600
560–600

0.846a
0.846a

0.346a
1200

1120

(Table 1) (Iannacone et al., 2019; Liu et al., 2020a; Liu et al., 2021; Liu
500
280

500
500

900

500

et al., 2020b). Iannacone et al. (2019) obtained TN REs up to 68 % in a


continuous-flow MBBR at stable DO concentration of 1.0 ± 0.2 mg L− 1.


Liu et al. (2020a) investigated the effects of three different DO con­
Zeolite powder-based PU sponges

centrations (2.5 ± 0.5 mg L− 1, 1.5 ± 0.5 mg L− 1 and 0.8 ± 0.3 mg L− 1)


in a continuous-flow MBBR, demonstrating that DO levels <1.0 mg L− 1
available as COD/N.
reported as m g .
Biodegradable polymer

were beneficial to SND (TN RE up to 53 %) with high N-NH+ 4 RE (85 %).


2 − 1
Cubic-shaped sponges

Polyethylene carriers
Conventional carrier

Conventional carrier

At DO levels >1.5 mg L− 1, TN RE decreased to almost 30 % due to


Sponges biocarriers

limited denitrification activity. However, under oxygen-limited condi­


Novel carrier

Novel carrier
Carrier type

tions, heterotrophic aerobic bacteria (HAB) may outcompete nitrifiers


Bio-carriers

PUF carrier
PU carrier
Kaldnes-3
PE plastic
PU foam

due to the lower growth rate and affinity for O2 of AOB and NOB
Ring R2
Table 1

compared to HAB (Jia et al., 2020). Slow adaptation of nitrifiers to low


b
a

O2 conditions could be required to avoid suppression of NH+ 4 oxidation

3
F. Di Capua et al.
Table 2
SND in MBBR-SBR systems.
Carrier type Specific Materials Filling Cycle AN O phase A phase Settling Idle Discharge Range Carbon Feed TN Feed C/ COD TN RE Reference
surface ratio Length phase (min) (min) (min) (min) (min) DO (mg source (mg L-1) Nb RE (%)
(m2 m-3) (%) (min) (min) L-1) (%)

PU foam – PU 30 – – 600 – – 120 – 1.5–5.5 (C6H10O5)n 25 ~12 92 42–85 Wang et al.


(2020a)
a
AC CBMC 2.70 AC, PE 40, 55 360 – 350 – – – 5 6.0 Acetate 55 5 82–94 68–88 Massoompour
Conventional 0.2478a PE 40, 55 360 – 350 – – – 5 6.0 Acetate 55 5 82–94 59–76 et al. (2020)
carrier
Combined – PP – – 175 90 175 30 – 5 2.0 Glucose 100 3.8 94 73 Chai et al.,
fibers (2019)
Novel carrier 650 PQAS-10 30 240 – 200 40 – – – 1.0–1.5 – 50 5 81–84 78–80 Liu et al.,
+ Fe2O3 (2018)
+ PE
PUF – PU 30. 600, – – – – – – 0.5–1.5 Acetate 89 ~4.5 85 60 Chen et al.
480 (2018a)
HX9KL, 500 PE 40, 60 240, – 90, 50 150, 90 0, 10 – 0, 3 0.8–1.1 – 35(±7) 5.5–6 78–95 50–76 Ferrentino
BIOMASTER 360 47(±6) et al. (2018)
BCN 012KLS
4

polyurethane 382 PU + rice 15.8 720 – 240, 60 120–480 – – 30 0.1–2.0 Acetate, 35(±4) 1.8–7.6 64–83 41–89 Feng et al.,
filler + with husk yeast, (2018)
rice husk sucrose
Foam carrier – PU 30 720 – 600 – – 120 – 1.5–5.5 (C6H10O5)n 22–31 ~10.2 > 90 42–84 Cao et al.
(2017)
Plastic fiber – – 12 720 50 190–270 380–460 – – – 1.5–2.2 Glucose 50 1–4 50–90 83–98 Ge et al. (2017)
Cylindrical – PP 30 – 90 210 90 30 40 – 0.2–2.5 – 35 11.4 95 94 Yin et al.
carrier (2015)
PU foam 1621 PU – 360 – 297 – 50 – – 0.1–0.9 Glucose 65–227 1.8–10.5 – 55–79 Tan et al.,
2013)
PU foam –. PU 8 – 60 60 – 90 570 – 0–7 Peptone, 48 5 65 57–100 Lim et al.,
sucrose, (2012)
acetate
Bio–carrier K3 500 PE 30 480 60 240 120 30 – – – Acetate 35 15.7 – 71.5 Lo et al., (2010)

AN = Anaerobic phase.
O = Aerobic phase.

Bioresource Technology 361 (2022) 127702


A = Anoxic phase.
a
reported as m2g− 1.
b
available as COD/N.
F. Di Capua et al.
Table 3
SND applications in AGS systems.
Cycle length (h) Cycle AN O (min) A (min) Settling Discharge Granule H/D DO (mg L–1) (O) Carbon source Feed TN Feed C/N COD RE TN RE Reference
(min) (min) (min) diameter (mm) (mg L–1) (%) (%)

3–6 AN–O 20–60 30 – 250 – 13.3–25 3.3–6.7 1.5 18.75 1.0–2.5 BOD5 31–34 ~5a 69–84 31–71 Campo et al. (2020)
6 AN–O–A 120 90 144 2 – – – – Acetate, Succinate (1:1, 20 ~10a 90 ~80 He et al. (2020)
1:3, 3:1)
5.6 AN – O 90 240 – 5 1 > 0.3 8.4 2 Acetate – propionate – 30–43 11–26a 83–93 45–77 Layer et al. (2019)
BOD5
4 AN–A–O, 0–45–55 104–220 0–30 10 – 0.2 10 8–10 Acetate, propionate 100 1.6–6.5 91–98 19–75 Pishgar et al. (2019)
AN/O, O
6 AN–O 54 270–285 – 20–5 1 1.5–0.7 10 – Acetate, Glucose, 110 ~ 6a 87–96 34–78 Rollemberg et al. (2019)
Ethanol
a
6 AN–O–A 120 90 144 2 – – 5.2 7–8 Acetate 2–55 4–20 94–95 75–95 Wang et al. (2018)
3.6 O – 150 – 20 40 0.09–0.4 – 2 BOD5 91 ~14.7a 92 87 Piotr and Cydzik-
kwiatkowska (2018)
6 AN–O–A 120 90 144 2 – – 5 5 Acetate 20 ~10a 89.2 88.5 He et al. (2017c)
6 AN–O–A 120 90 144 2 – – 5 0.8–1.2 Acetate, glucose 20 ~10a 91–81 94–81 He et al. (2017a)
6 AN–O–A 120 90 144 2 – 1.4 ± 0.6 1.8 ± 5 3.5–4.5 1.7–2.3 Acetate 20 ~10a 95 64–94 Yan et al. (2019b)
0.9 0.8–1.2
6 AN–O–A 120 90 144 2 – – 5 5 Acetate 19 10.7a 94 94 He et al., (2016b)
8 AN–O–A 180 180–140 90–148 20–2 – 1.0–0.6 6.3 5 Acetate 40–50 ~7a 80–90 88–97 He et al. (2016a)
8 AN–O–A 180 180–150 90–138 20–2 – 1.2 5.2 5 Acetate 15 ~10a 87 97 He et al. (2016c)
6 AN–O 130 160 – 60 – 1.6–0.3 4.2 0.8–1.6 Acetate 50 ~16a – 65 Lu et al. (2016)
4 O – 180 – 5 2 0.5–1.7 8 > 4.0 Acetate 250–330 ~1a 100 33–77 Yan et al. (2016)
4 AN–O 10–20 184–219 – 35–10 – 0.5 9.7 – BOD5 83 ~7a 92 61 Wagner et al. (2015)
5

6.5–3 AN–O–A – 60–300 – 30 – >0.2 – 1.8–2.5 BOD5 50 ~ 10a 87 86 Pronk et al. (2015)
3 O – 162–167 – 2–7 6 > 1.5 11.1 7–8 Acetate 77–154 8–16 > 90 > 90 Di Bella and Torregrossa
(2013)
3–4 O – 164–224 – 15 0.5 0.3–0.7 22.2 – BOD5 76–91 4.7–11.2a 78 – 92 66–97 Wagner and Helena
(2013)
3 AN–O 60 112 – 3 5 0.9–1.1 – 20–80 % of Propionate 50 8–1.2a 100 40–95 Lochmatter et al. (2013)
saturation
a
6 AN–O–A 45–100 35–160 3–40 2–40 – 0.2 – 2 BOD5 + Acetate 67 ~5 100 75–84 Coma et al. (2012)
5.4 O – 114 – 3 3 2.4 7.2 6–8 Sucrose, acetate, 40 ~ 10a >80 >75 Isanta et al. (2012)
propionate
4 A–O – 213 10 2 5 1.5–0.7 20 2 – 60 ~ 10a 80–85 68–96 Chen et al. (2011)
4 O – 130–167 – 8–45 5 0.8 5 8 BOD5 60 ~ 16.7a 80 50 Liu et al. (2010)
2.5–4 A–O 90 240 5 – – – 3.2 2–3 Acetate–BOD5 43–147 0.4–2.3 64–89 47 Wang et al. (2009)
– AN–O–A 30 90 30 – – > 1.0 – – Acetate 60 2.5 ~ 100 ~ 100 Kishida et al. (2008)
6 AN–O 90 120 – 0.5 – 1.0 – 2–5 Acetate 60 ~ 10a ~ 100 ~ 100 Kishida et al. (2006)\
3 O – 169 – 3 5 1.0 – 10–100 % or Acetate – 8.3a – 8–45 Mosquera-Corral et al.
saturation (2005)

Bioresource Technology 361 (2022) 127702


4–6 O, AN–O, 0–120 230 0–120 2 4 0.4–1.9 13.3 0.5–2.0 Ethanol 25–150 3–5a 92 20–40 Yang et al. (2003)
AN–A

AN = Anaerobic phase.
O = Aerobic phase.
A = Anoxic phase.
a
available as COD/N.
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

exposed to O2. Therefore, HAB have a competitive advantage for


consuming the organic compounds from the bulk liquid over DNB,
which occupy the inner anoxic portion of the biofilm. During diffusion
from the bulk liquid across the biofilm, a large amount of organics is
consumed by HAB, while the residual organic carbon may not be suffi­
cient to allow denitrification. Wang et al. (2018) showed that decreasing
the COD/TN ratio from 20 to 4 in an AGS-SBR system reduced the TN RE
of SND from 95 % to around 75 % due to carbon deficiency.
Possible strategies to overcome the lack of carbon in real municipal
wastewater include (1) the addition of liquid carbon sources, such as
ethanol, methanol or acetic acid and (2) the addition of a solid carbon
source, such as biodegradable polymers (BDPs) (Liu et al., 2021). BDPs
include poly-3-hydroxybutyric acid (PHB), polycaprolactone (PCL),
polylactic acid (PLA) and poly butanediol succinate (PBS) that can be
hydrolyzed by extracellular enzymes secreted by specific microorgan­
isms. Their hydrolysis products can then be utilized by HAB and DNB as
sources of organic carbon (Chu and Wang, 2011; Han et al., 2018). BDPs
are receiving increasing attention as organic supplement at WWTPs,
since they are slow-release electron donor/nutrient sources that can also
be used as biofilm carriers (Table 2). Low-cost alternatives to BDPs are
natural biopolymers such as starch or lignocellulose. Feng et al. (2018)
used a biodegradable support made of rice husk and lignocellulosic
materials mainly composed of biodegradable cellulose, branched
hemicelluloses and recalcitrant lignin for SND in a sequencing batch
biofilm reactor. TN RE > 50 % was obtained at a feed COD/N ratio of
2.8, while lower REs were obtained at a COD/N ratio of 1.8.
Although positive effects have been highlighted, the increase of feed
C/N ratio in SND systems should be controlled. Several studies pointed
out that excess organic carbon in the feed may limit nitrification effi­
ciency. Indeed, HAB overgrowth can limit DO diffusion within biofilm
Fig. 1. Schematics of a co-diffusional combined nitrifying and denitrifying and was shown to inhibit nitrifying activity (Iannacone et al., 2019).
biofilm and the parameters affecting N2O emissions.
4. Shortcut simultaneous nitrification–denitrification
(Iannacone et al., 2019). Hence, DO must be strictly controlled during
reactor start-up and continuous operation to sustain nitrification and Shortcut SND (also known as SND via NO–2 or simultaneous partial
maximize TN RE. nitrification denitrification, SPND) involves the oxidation of NH+
4 to NO2

In AGS systems, DO concentrations in the range of 1.5–3.0 mg L− 1 by AOB followed by the reduction of NO2 to N2 by DNB. Therefore, in

are generally applied to enhance SND performance (Bengtsson et al., this process nitritation is followed by denitritation. This showcases
2018; Layer et al., 2020). The long-term stability of AGS is sustained by shortcut SND as an advantageous bioprocess compared to SND for the
high shear forces typically provided by aeration, which also determines treatment of wastewaters with low C/N ratio, as it requires 40 % less
nitrification and denitrification efficiencies (He et al., 2019a; Yan et al., carbon for denitrification in addition to 25 % lower O2 consumption in
2019b). Yan et al. (2019b) studied the effect of three DO concentration the aerobic phase, up to 2-fold faster N-NO−x reduction, 20 % lower CO2
ranges (6–7, 4–5, 2–3 mg L− 1) on AGS performances and dynamic emission and 33–55 % lower biomass production (Peng and Zhu, 2006;
changes in sludge particle size. With the decrease of DO concentration Rahimi et al., 2020). NO–2 reduction rate is reported to be 1.5–2 times
and air shear force the percentage of particle size with a diameter < 0.5 higher than NO–3 reduction, which leads to faster N removal (Rahimi
mm increased by 0.9–6.4 %, indicating that operation at low DO values et al., 2020).
led to partial break-up of the granules.
Intermittent aeration (IA) can be a suitable choice for the SND pro­ 4.1. Strategies for inhibition of nitrite oxidizing bacteria
cess, as it can promote the activities of both nitrifying and denitrifying
bacteria. In addition, IA can significantly limit the operational costs for To allow the growth of AOB and simultaneously inhibit NOB activity,
wastewater treatment, as aeration accounts for >50 % of the energy it is essential to maintain specific operational conditions. According to
costs of a municipal WWTP (Moura et al., 2012). Iannacone et al. (2021; the literature, selective inhibition of NOB can be achieved at pH > 7.5,
2020) demonstrated that IA can effectively sustain both complete and DO levels in the range of 1–2 mg L− 1, SRT < 5 days, temperature >
shortcut SND in continuous-flow MBBRs treating synthetic municipal 25 ◦ C, free ammonia (FA) and free nitrous acid (FNA) concentrations of
wastewater, obtaining a nearly 90 % TIN removal at DO concentrations 0.1–5 and 0.011–0.22 mg N L− 1 in respective order, and by alternating
ranging between 0.2 and 3.0 mg L− 1. aerobic and anoxic conditions (Iannacone et al., 2021).

3.2.2. Feed carbon-to-nitrogen ratio 4.1.1. High free ammonia and free nitrous acid concentrations
Municipal wastewaters are generally characterized by low C/N ra­ The concentrations of FA and FNA in the bulk liquid depend on those
tios, which may result in low TN removal due to lack of organic electron of NH+ 4 and NO2, respectively. High FA levels can lead to proton

donor for denitrification (Campo et al., 2020; Chai et al., 2019). As a imbalance, potassium deficiency, change of intracellular pH, increase of
result, additional organics such as methanol, glucose or sodium acetate energy requirement and inhibition of enzymatic reactions (Di Capua
may be required (Chai et al., 2019; Han et al., 2018). Likewise, low SND et al., 2021). FNA has been indicated as an uncoupler affecting energy
performance in biofilm systems is often linked to the lack of organic generation and being capable of inhibiting the expression of enzymes
carbon for denitrification (Layer et al., 2020; Wang et al., 2020b). The such as NO–2 and NO reductases (Zhou et al., 2011). The inhibitory
outer portion of the biofilm is generally occupied by HAB, being more thresholds for AOB and NOB are different, NOB being more sensitive

6
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

than AOB to both FA and FNA. AOB are inhibited by FA in the range of affect the duration of the lag-phase. In the study of Iannacone et al.
10–605 mg N L− 1, while values between 0.1 and 5 mg N L− 1 are suffi­ (2021), shortcut SND was successfully maintained in a continuous-flow
cient to inhibit NOB (Liu et al., 2020b). FNA concentrations that inhibit MBBR by adopting aeration cycles with a microaerobic/anoxic phase
AOB vary between 0.42 and 1.72 mg N L− 1, while NOB inhibition was (DO = 0.2–1.0 mg L− 1) of 20–25 min, which is in agreement with the
observed already at concentrations of 0.011–0.07 mg N L− 1. FNA con­ results of Gilbert et al. (2014).
centrations of 0.026–0.22 mg N L− 1 were reported to completely inhibit
NOB activity (Jia et al., 2020; Liu et al., 2020b). Exposing NOB to 4.1.4. High salinity
inhibitory levels of FA and FNA is not an easy task to achieve since the High salt content has negative effects on nutrient removal and in­
concentration of N-NH+ 4 and pH depend on the influent wastewater. fluences N removal pathways. NOB are more sensitive to salt than AOB,
Additionally, NO–2 accumulation in the system and subsequent discharge and an increase in salinity can lead to their inhibition. Xia et al. (2019)
should be avoided due to high NO–2 toxicity to bacteria and aquatic reported that shortcut SND became the main N removal pathway under
species (Iannacone et al., 2021). salinity between 1.6 % and 2.4 %. The authors studied SND in a hybrid
sequencing batch biofilm reactor fed with synthetic domestic sewage
4.1.2. Control of temperature and solid retention time with salinity ranging from 0 % to 2.4 % (0–24 g L-1 of seawater crystals).
At temperatures <20 ◦ C, the maximum specific growth rate of NOB An increase in salinity to 1.4–2.4 % inhibited NOB activity as NO–2
(μNOB AOB
max ) is generally higher than that of AOB (μmax ) (Shourjeh et al.,
started to accumulate in the bioreactor. As a result, higher salinity
2021). Temperatures >25 C are recommended for AOB to outgrow NOB
◦ values reduced TN RE by 8.5–17.3 %, while N-NH+ 4 and COD REs were
and ensure a stable process (Zhang et al., 2009). However, due to high not significantly affected. Similarly, He et al. (2019b) observed that NOB
specific heat of water (4.184 kJ kg− 1 K− 1 at 20 ◦ C), it is not economically population decreased to nearly undetectable levels when salinity was
feasible to increase the temperature of municipal wastewater. Main­ increased from 1 % to 2 % in an AGS-SBR system performing SND. These
taining an adequately low SRT at these temperatures can help to data indicate that salinity levels >1 % favor shortcut SND over complete
washout NOB from the reactor, although successful NOB inhibition has SND, although decrease in TN removal might be observed due to inhi­
also been obtained at long SRT under DO limitation (Peng and Zhu, bition of denitritation. However, an increase of salinity above 2.5 % was
2006). Therefore, DO is a major control parameter to obtain successful observed to negatively impact also AOB activity (She et al., 2017).
inhibition of NOB and can be controlled in biofilm systems to sustain
shortcut SND in bioreactors . 5. Nitrous oxide emissions in simultaneous
nitrification–denitrification biofilms
4.1.3. Control of dissolved oxygen
DO control has been shown as the most effective strategy for main­ This section focuses merely on combined nitrifying and denitrifying
taining nitritation at temperatures <20 ◦ C (Gilbert et al., 2015). O2 is the biofilms in a co-diffusional configuration (where the electron donor and
terminal electron acceptor in the oxidation–reduction respiratory chains acceptor diffuse from the bulk liquid) (Fig. 1); however, the key pa­
of both AOB and NOB. DO half saturation constant for AOB (KAOB rameters leading to increased or decreased N2O emissions discussed
DO ) falls
in the range of 0.2–0.4 mg O2 L− 1, being lower than that for NOB (KNOB below can also apply to counter-diffusional biofilms (Kinh et al., 2017b;
DO ),
i.e., 0.7–2.0 mg O2 L− 1, which means that AOB have a higher affinity for a).
O2 than NOB (Liu et al., 2020b). This implies that NOB are often at a
disadvantage when competing with AOB for DO (Chai et al., 2019; Liu 5.1. Carbon-to-nitrogen ratio and biofilm thickness
et al., 2020b; Rahimi et al., 2020). However, it has been observed that
NOB affinity for O2 can significantly increase after long-term operation C/N ratio plays an important role for N2O emissions in combined
at low DO concentrations, making NOB better competitors for DO than nitrifying and denitrifying biofilms. In these biofilms, most of the
AOB (Liu and Wang, 2013). Peng and Zhu (2006) suggested to maintain organic carbon is typically utilized by HAB in the exterior portion of the
a DO range of 1.0–1.5 mg L− 1 to select for AOB over NOB and obtain a biofilm (Fig. 1), leaving rather limiting or no amount of organic carbon
successful nitritation. Campo et al. (2020) showed that about 56 % of TN available for denitrification (Kinh et al., 2017b). When adequate levels
could be removed by SND via NO–2 and 44 % via NO–3 in an AGS-SBR of organic carbon are not available, intermediates such as NO and N2O
treating real domestic wastewater with low COD/N ratio (2.8–3.8) by can accumulate within the biofilm (Fig. 1) (Kinh et al., 2017a). The
maintaining (1) a low DO concentration during the aerobic phase geometry of the combined nitrifying and denitrifying biofilms attributes
(1.4–1.6 mg L− 1), (2) a low DO/NH+ thickness an important role. As mentioned earlier, N2O can undergo
4 ratio and (3) a low SRT of biomass
flocs (8–10 days). mechanisms of production and consumption by AOB and DNB within the
IA has been proposed as an effective method to sustain nitritation. biofilm depth (Schreiber et al., 2012). Eldyasti et al. (2014) evaluated
One of the NOB suppression mechanisms that has been attributed to IA is N2O emissions from denitrifying fluidized bed bioreactor biofilm and
the establishment of a lag-phase in NOB activity after the transition from showed that N2O emissions decreased with the increase in biofilm
the anoxic to the aerobic period (Gilbert et al., 2014). Under anoxic thickness. The consumption of N2O likely depends on the thickness of
conditions, NOB experience the inactivation of the NXR enzyme and its the deeper layer where, if full penetration of organic carbon occurs, DNB
reactivation under the subsequent aerobic phase. However, in the aer­ can further reduce N2O to N2 (Fig. 1).
obic phase, AOB activity recovers faster than NOB activity, leading to a
competitive advantage of AOB over NOB. Gilbert et al. (2014) indicated 5.2. Dissolved oxygen and temperature
that 15 min was the minimum duration of the anoxic phase able to
generate a lag-phase in NO–2 oxidation longer than that of NH+ DO and temperature are also important factors linked to N2O emis­
4 oxida­
tion. Therefore, IA with anoxic periods of 15–20 min and aerobic periods sions from SND biofilms. The complex stratification along with varying
shorter than the NOB lag-phase are sufficient to suppress NOB activity. DO and substrate concentrations can lead to low DO values, DO tran­
The length of this lag-phase (up to 15 min) was observed to be species- sitions from high to low values and DO fluctuations within the biofilm
specific and to depend on the DO levels experienced by the biomass (Sabba et al., 2018). A decrease in DO concentration can lead to higher
during the aerobic period. The lag-phase for biomass adapted to DO rates of N2O formation by AOB via the nitrifier denitrification pathway
levels of 0.9–1.0 mg L− 1 was distinctively longer than the one for (Chen et al., 2018b). Also, the microaerophilic layer formed within a
biomass adapted to lower DO levels (≤0.4 mg L− 1). In contrast, anoxic combined nitrifying and denitrifying biofilm can trigger more NO and
periods longer than 15–20 min and temperature (10–30 ◦ C) did not N2O production by DNB due to the inhibition of the nitrous oxide
reductase (NoS) enzyme in the presence of O2 (Kinh et al., 2017a; Guo

7
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

et al., 2017). Qi et al. (2022) found that N2O-reducing DNB can recover biofilms where there is a higher N2O mass transfer towards the bulk
from exposure to DO and that this recovery is species-dependent. Zhou rather than the interior denitrifying portion of the biofilm, where N2O
et al. (2022b) found that temperature can influence the abundance and can be reduced (Sabba et al., 2018). Iannacone et al. (2020) operated a
composition of the N2O-reducing DNB. continuous-flow MBBR alternating microaerobic and aerobic conditions
to achieve carbon, N and P removal through simultaneous nitrification
5.3. pH and nitrite concentration and denitrification coupled to P removal. The authors found that IA
helped enriching for a denitrifying community and therefore allowing to
The chemical equilibrium between NO–2 and FNA is determined by shift the N2O mass transfer toward the anoxic zone where N2O was
NO–2 concentration, pH and temperature (Di Capua et al., 2021). High consumed.
nitrification rates in a combined nitrifying and denitrifying biofilm
could lead to high NO–2 production with a decrease of pH; this can drive 6. Biofilm reactors performing simultaneous
the formation of FNA that in turn inhibits N2O consumption (Zhou et al., nitrification–denitrification
2008). This mechanism has been recently shown to occur in denitrifying
mixed cultures (Zhou et al., 2022a). Zhou et al. (2022a) suggest that this SND has been investigated under various operational conditions and
inhibition can lead to the redistribution of electrons between different reactor configurations. In this section, the configurations, and key fea­
denitrification enzymes such as nitrite reductase (NiR) and NoS, with the tures of the most used biofilm reactors for SND, i.e., MBBR, AGS, and
latter being the most impacted and causing an overall reduction in the HMBR (Fig. 2), are critically described with the scope to provide
N2O consumption activity of DNB. However, FNA and pH have signifi­ guidelines for successful and cost-effective operation.
cant impacts on AOB populations as well (Sabba et al., 2022), raising the
importance of community composition and how shifts in community/­ 6.1. Moving bed biofilm reactors
predominant species might be a strategy leading to lower N2O produc­
tion (Suenaga et al., 2018). MBBRs have been widely applied for nutrient removal from waste­
water (Gu et al., 2018; Wang et al., 2018). The presence of mobile
5.4. Reactor operation carriers in the reactor enables the formation of a stratified biofilm with a
bacterial community profile influenced by wastewater composition and
Reactor operation can influence N2O emissions and the microbial operating conditions (Khanongnuch et al., 2019). Substrates are trans­
community overall behavior. Yu et al. (2021) investigated the effects of ported from the bulk into a co-diffusional biofilm through diffusion
HRT on N2O production rates during nitrification in lab-scale biological mechanisms allowing the formation of anaerobic, anoxic and aerobic
aerated filters. The authors tested three different HRTs (4, 6 and 8 h) and layers in the MBBR biofilm (Suarez et al., 2019). Compared to fixed-bed
found that the maximum and minimum N2O production rate occurred at biofilm systems, the MBBR does not suffer from clogging or channeling
HRT of 4 and 6 h, respectively. The proximity of the nitrifying portion of issues and no periodical backwashing is needed (Chu and Wang, 2011).
the biofilm to the bulk liquid allows for the N2O produced to be released In recent years, the effect of various factors, i.e., carrier type, C/N ratio,
in the bulk and a lower HRT leads to higher emissions. This is especially DO concentration and filling ratio, on the SND process has been inves­
true if the bulk is aerated in a combined nitrifying and denitrifying tigated in MBBR systems performing SND. Table 1 lists the performance

Fig. 2. Biofilm development and stratification in bioreactors applied for SND.

8
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

and operating conditions of continuous-flow MBBRs applied for the SND carriers accelerated biofilm formation during start-up period by
process. reducing repulsion and determined higher REs compared to traditional
PE carriers. The authors reported a material cost of the surface-modified
6.1.1. Carrier type carrier similar to that of conventional carriers (around 46 USD ton− 1),
Carrier material, surface area and roughness play a key role in although the step of dispersing the materials (i.e., clinoptilolite) into PE
determining the performance of a MBBR, since these characteristics granules would require additional energy.
affect both microbial adhesion and growth (Massoompour et al., 2020;
Zhao et al., 2019). Additionally, the development, thickness, and ge­ 6.1.2. Filling ratio
ometry of the MBBR biofilm can significantly influence the diffusion and MBBRs are usually operated with filling ratios between 20 % and 70
modulate gradients of different parameters, including DO (Fig. 1). Up to % (McQuarrie and Boltz, 2011; Ødegaard, 2006). Values < 70 % are
date, various biofilm carriers have been tested to support SND in MBBRs, recommended in order to assure adequate mixing, while ratios > 20 %
including K-series AnoxKaldnesTM made in polyethylene (PE), poly­ guarantee sufficient concentration of attached-growth biomass in the
urethane (PU) sponge, high density PE (HDPE) carriers, ceramic carrier system (di Biase et al., 2019). The choice of the filling ratio affects both
or zeolite powders combined with PU sponge (Z-PU) (Table 1). the bioreactor performance and operational costs, i.e., energy con­
Of all carrier types used in MBBR applications, sponges feature the sumption and purchase of carriers. Feng et al. (2012) evaluated the ef­
largest specific surface area (SSA) and a high porosity (around 97 %), fect of three polyurethane foam (PUF) carrier filling ratios (20 %, 30 %
being useful for microbial growth (Al-Amshawee et al., 2020). The high and 40 %) on SND in three parallel MBBRs with an aeration zone and a
SSA of sponges promotes biofilm development, while porosity allows to settling zone. TN RE in the MBBR filled with 40 % of PUF carriers was 2
reduce clogging (Nguyen et al., 2010). Sandip and Kalyanraman (2019) times higher than with 20 % filling ratio (Table 1). The filling ratio also
compared the SND performance of a MBBR filled with PU foam carriers affected biofilm structure and nitrification activity. Biofilm in the MBBR
(280 m2 m-3) to that of a MBBR containing conventional PE carriers with a 20 % filling ratio was thick and dense, while in the other two
similar in shape to the AnoxKaldnes K1 (500 m2 m-3) (Table 1). A MBBRs the carriers looked hollow with a thin biofilm. Microprofiles
maximum TN RE of 59 % was achieved in the MBBR with PU foam analysis revealed that dense biofilm could limit DO diffusion into the
carriers, 17 % higher than the average TN RE observed in the MBBR with inner layer and inhibit nitrifier growth, being the relative abundance of
PE carriers. This difference was mainly due to the higher nitrifying ac­ the nitrifying bacteria Nitrosomonas and Nitrospira in MBBRs with filling
tivity observed with PU foam carriers, being NH+ 4 RE 11 % higher than ratios of 30 % and 40 % higher than with 20 % filling ratio. Similarly,
the average observed in the MBBR with PE carriers, which was attrib­ Ferrentino et al. (2018) observed that the increase of filling ratio from
uted to the higher biofilm-liquid contact area of PU foam compared to 40 % to 60 % in a SBR-MBBR performing SND determined an increase of
PE carriers. The effect of sponge size on SND was investigated by Lim nitrification activity and, consequently, an increase of TN RE from an
et al. (2011) in four MBBR-SBRs operated in parallel with foam cubes of average value of 50 % to 66 %.
8, 27, 64 and 125 mL, which showed TN REs of 37 %, 31 %, 24 % and 19 Zhang et al. (2017b) studied the SND process in three MBBRs filled
%, respectively. Larger PU foam cubes were not fully covered by with PUF at filling ratios of 10 %, 20 % and 30 %, in which average TN
biomass, which limited TN RE, while a robust biofilm developed on REs of 77 %, 86 % and 87 % were observed, respectively (Table 1). The
smaller cubes, establishing a DO gradient along the PU foam inward highest NH+ 4 oxidation and denitrification rates (2.2 mg N-NH4 g bio­
+
− 1 − 1 − 1 − 1
depth favorable to the SND process. For comparison, one bioreactor was mass h and 5.1 mg N-NO3 g biomass h , respectively) were

operated without carrier addition, showing a TN RE of only 15 % due to achieved in the MBBR filled at 30 %. No significant difference in
poor denitrification as a result of high DO and low COD concentrations maximum TN RE (94–95 %) and SND efficiency (98–100 %) was
in the bulk. Batch tests revealed the occurrence of NO−x reduction in observed between the MBBRs filled at 20 % and 30 %. Similar to Feng
MBBR biofilms even without the addition of a carbon source, which et al. (2012), the concentration of attached biomass per gram of carrier
could occur due to storage of carbon in the deep biofilm layers. was higher at low filling ratios (10 % and 20 %) than at the highest
Carriers made with recycled waste material have attracted the in­ tested value of 30 %. On the other hand, suspended biomass was more
terest of researchers in recent years, being a cost-effective and envi­ abundant at 30 % filling ratio, due to a more frequent carrier collision
ronmentally friendly solution for MBBR operation (Sabba et al., 2017a). leading to biomass detachment, which resulted in an overall higher
Massoompour et al. (2020) proposed a new carrier made from waste biomass concentration at 30 % filling ratio. In contrast, Lim et al. (2012)
activated carbon (AC) and PE carriers mixed through a chemical and observed that increasing the filling ratio from 20 % to 30 % and 40 % in
thermal process. The authors compared the performance of two parallel three IA MBBR-SBRs packed with 8-mL PUF cubes reduced the con­
MBBRs, one filled with conventional PE carriers, and the other with centration of suspended biomass, which was mainly responsible for
modified PE with AC carriers (Table 2). The SSA of the modified carriers nitrification, while the MBBR biofilm was mainly responsible for deni­
was 10.9 times higher than that of conventional carriers, which trification. As a result, increasing the filling ratio led to a decrease of
increased the attached biomass by up to 20 %. The speed of microbial NH+ 4 removal. However, the denitrification efficiency increased due to a
adhesion and growth of biofilm on modified PE carriers was 26 % higher larger anoxic zone and carbon storage in the biofilm, leading to TN REs
than with conventional carriers, being enhanced by the hydrophobic of 86 %, 100 % and 96 % at filling ratios of 20 %, 30 % and 40 %,
nature and high porosity of AC (Massoompour et al., 2020). TN RE was respectively. As a result, 30 % was considered the optimal filling ratio in
improved by almost 17 % for MBBR equipped with modified carriers due the studied IA system, as it could maximize the overall N removal by
to the development of a larger anoxic area favoring denitrification, SND. Despite the different biomass dynamics ongoing in the described
being nitrification efficiencies similar in both MBBRs. Again, improve­ MBBR systems, filling ratios ≥ 30 % can be recommended for reactor
ment of denitrification was the main mechanism for achieving higher TN operation to enhance N removal.
REs by using novel carriers, which allowed to modulate DO gradient and
establish larger anoxic conditions within the biofilm. 6.1.3. Hydraulic retention time
Pure PP, PE and HDPE carriers are characterized by a negative Changes of HRT can lead to short- and long-term effects on SND
charge surface similar to the surface charge of biofilm, which can performance. Prolonging the HRT has been observed to exert positive
hamper biofilm adhesion on carriers due to repulsions between microbes effects on nitrification and improve TN RE in MBBRs. Feng et al. (2012)
and carriers. Liu et al. (2020b; a) proposed a surface-modified carrier showed that increasing the HRT from 5 h to 7 h improved the TN RE (i.e.,
made of PE granules with polyquaternium-10 (PQAS-10), Fe2O3 and 2 24.8–34.1 %) by 18–29 % in three MBBR with filling ratios of 20 %, 30 %
wt% clinoptilolite, characterized by a strong NH+ 4 adsorption capacity, and 40 %. The increase of HRT in the three reactors did not affect the
which benefited the enrichment of nitrifying bacteria. Surface-modified COD RE, which remained stable at 81 %, but enhanced the NH+ 4 RE. In

9
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

another study, Zinatizadeh and Ghaytooli (2015) observed increased carrier-attached biofilm. Similarly, Han et al. (2018) reported a slower
growth and activity of AOB and NOB following increase of HRT and DO increase of the TMP in a HMBR seeded with biocarriers (sponge coupled
concentration from 4 to 8 h and from 2 to 3 mg L− 1, respectively. Bac­ with PBS granules) compared to a CMBR, corresponding to 1.5-fold
terial growth led to a higher biofilm thickness, which promoted more longer operational time of the HMBR compared to the CMBR. The
favorable anoxic conditions for DNB and higher TN REs. However, it reduction of cake layer was attributed to physical mechanisms, e.g., the
should be noted that increasing the HRT can significantly prolong the frictional force carried out by biocarriers on the submerged membrane,
SRT of both suspended and attached-growth biomass and promote the back-transport effect from the membrane surface to the bulk solution
growth of NOB in the system, which is deleterious if shortcut SND is the due to turbulence of suspended carriers or impact of biocarriers on the
desired bioprocess (see Section 4.1.2). To the best of our knowledge, membrane with consequent shaking of this, as well as to the increased
literature lacks information regarding the effect of HRT on the evolution EPS hydrophobicity, which could considerably improve the flocculation
of the microbial community in MBBRs performing SND, which should be capacity of the sludge and decrease membrane fouling (Han et al.,
investigated in future studies. 2018). However, negative effects of biocarrier addition on the MBR
performance have also been reported. These include the breaking up of
6.1.4. Combined nitrogen and phosphorus removal in moving bed biofilm sludge flocs increasing the number of small particles as well as TOC
reactors levels and, consequently, accelerating membrane fouling. Huang et al.
Recent studies have outlined that the combined removal of N-NH+ 4 (2008) showed that maintaining a low carrier-to-suspended biomass
and P-PO3−4 is achievable under microaerobic/IA conditions in MBBR. ratio reduces the breaking up of sludge flocs, resulting in an overall
Iannacone et al. (2019) observed a P-PO3−4 RE up to 72 % combined to a positive effect of carrier addition on membrane fouling.
TIN RE up to 68 % in a MBBR operated under stable microaerobic Although interesting performances have been highlighted, the
conditions (DO = 1.0 ± 0.2 mg L-1) at a feed C/N ratio of 4.2. In another operation of HMBRs would be rather expensive due to the high cost of
study, MBBR operation under IA conditions (DO alternating between 0.2 membranes and carriers and, therefore, a cost-benefit evaluation for full
and 3 mg L-1) could further improve P-PO3− 4 removal combined to SND, scale application would be needed (Meng et al., 2009).
reaching an average efficiency of 75 % (Iannacone et al., 2020). Pre-
treatment of inoculum by cultivation at pH 8.2 (±0.2), 26–28 ◦ C and 6.3. Aerobic granular sludge systems
SRT of 4 day effectively inhibited NOB growth, allowing shortcut SND to
occur in the MBBR under continuous-flow conditions. Decrease of car­ AGS combines the characteristics of activated sludge and biofilm
bon consumption due to shortcut SND could be likely used by phos­ systems due to presence of suspended microbial aggregates without any
phorus accumulating organisms (PAO)-like bacteria, leading to a PO3− 4 supporting carrier and a biofilm-like structure (Wang et al., 2009).
RE of 81–88 % (Iannacone et al., 2021). Microbial community analyses Aerobic granules are spherical bacterial aggregates that grow under
via Illumina sequencing revealed that the genus Hydrogenophaga was aerobic conditions and are usually denser and heavier than regular CAS
dominant in the MBBRs under both microaerobic/IA conditions. Several flocks (Sarma and Tay, 2018). Granule formation is generally promoted
studies have reported Hydrogenophaga as putative PAO being able to by providing a feast and famine regime consisting in alternating
accumulate P under aerobic conditions (Iannacone et al., 2021). Another anaerobic and aerobic conditions to select for slow-growing microor­
potential mechanism enabling P removal in MBBR is the occurrence of ganisms against fast-growing aerobic heterotrophs. The establishment of
transient anaerobic zones in deeper layers of the microbial biofilm under aerobic, anoxic and anaerobic zones within the granule allows the
non-aerated conditions, leading to the establishment of the aerobic- coexistence of nitrifiers, DNB and PAO, providing the conditions for
aerobic/anoxic alternation triggering typical PAO metabolism. Howev­ combined SND and P removal (Yuan et al., 2019). Up to date, AGS has
er, no typical PAO was detected in the microbial community of the been recognized as a promising and emerging technology for waste­
MBBR biofilm, suggesting that Hydrogenophaga likely had a major role in water treatment as alternative to CAS thanks to several advantages, such
P removal under (micro)aerobic conditions. as excellent settleability, small footprint, high biomass retention and the
ability to remove simultaneously C, N and P (Chyi et al., 2020).
6.2. Hybrid biofilm-membrane bioreactor systems Table 3 lists the operational parameters and SND performance of
AGS reactors. The most common setup to enrich for aerobic granules
Biofilm reactors can be combined with MBR to improve the system performing SND and to maximize C, N and P removal is to cultivate AGS
performance and reduce the effect of suspended solids on membrane in a SBR (Bengtsson et al., 2018; de Sousa Rollemberg et al., 2018). SND
fouling (Wang et al., 2020b). Yang et al. (2009) compared the SND in AGS-SBRs has been successfully established in lab-scale SBRs with a
performance of a HMBR filled with nonwovens carriers at 30 % filling height/diameter ratio (H/D) between 5 and 10. Despite the excellent TN
ratio to that of a conventional MBR (CMBR). Both reactors showed COD REs obtained through the SND process, the poor long-term stability of
REs higher than 95 %, while TN REs reached values up to 89 % for the aerobic granules due to granule desegregation into filamentous fractions
HMBR and 70 % for the CMBR. The HMBR showed a more pronounced represents a serious issue (Wang et al., 2018). AGS stability is affected
improvement of the AOB, NOB and HAB activity compared to the CMBR, by: (1) feed C/N ratio; (2) aeration intensity and (3) organic loading rate
and a more stable response to variations of the feed COD/TN ratio. (He et al., 2019a). A possible solution to ensure long stability is to enrich
Membrane biofouling is the most critical issue for MBR operation, as the bacterial community with slow growing microorganisms (metabolic
it reduces membrane permeability and increases the transmembrane selection), such as PAO and glycogen-accumulating organisms (GAO),
pressure (TMP), leading to high operational costs due to frequent characterized by high granulation capacity (Campo et al., 2020). The
chemical cleaning or replacement of membranes (Asik et al., 2021; Ucar proliferation of these microorganisms also leads to suppression of HAB
et al., 2021; 2020). It has been demonstrated that biocarrier addition to due to the lack of carbon source under aerobic conditions, stored by
MBR systems can alleviate fouling due to scouring of the membrane GAO and PAO under the previous anaerobic phase (de Sousa Rollemberg
surface and/or to a physical–chemical effect on the bulk sludge char­ et al., 2018).
acterization (Chen et al., 2016). Wang et al. (2020b) showed that adding Studies on SND in AGS systems have been mainly conducted at lab-
static biocarriers improved nitrification and denitrification while scale using VFAs as sources of organic carbon. If the growth of PAO and
reducing the abundance of bacteria responsible for significant EPS GAO is crucial for the formation of stable granules, granulation may be
secretion and consequent membrane fouling. Meanwhile, the authors hampered during the treatment of real municipal wastewater containing
also observed a nearly 8 % improvement of the TN RE in the HMBR low VFA concentrations (Layer et al., 2020; 2019). Indeed, Sarma and
compared to CMBR due to better denitrification, which was attributed to Tay (2018) reported low VFA concentrations in real municipal waste­
the formation of anoxic and aerobic microenvironments within the water (between 22 and 92 mg COD L− 1), with non-diffusible organic

10
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

carbon being nearly 50 % of the total influent COD. synthetic wastewater containing 100 % VFA as organic carbon, which
led to the lowest accumulation of NO–3. In contrast, the presence of non-
6.3.1. Granule diameter diffusible organic carbon reduced TN RE due to poorer denitrification,
The diameter of AGS granules is a factor influencing N removal via leading to higher NO–3 accumulation (>10 mg L− 1).
SND (De Kreuk et al., 2005). On average, aerobic granules feature di­
ameters between 0.1 and 3.0 mm (Dahalan et al., 2015). Derlon et al. 6.3.3. Aeration strategy
(2016) pointed out that typical granule size developed with real Besides the lack of sufficient electron donor for denitrification,
municipal wastewater varied between 0.2 and 1.3 mm (Table 3). These another aspect that may limit SND in AGS systems is the formation of
values are smaller than that observed in granules cultivated with syn­ limited anoxic zones within the granules. If O2 penetration is too fast, it
thetic wastewater, usually reported to be higher than 2.0 mm. The size may reduce the time span of the anoxic phase and limit denitrification.
of the granules together with the DO penetration depth determines the To improve TN REs, several strategies could be considered, such as pre-
extension of the anoxic zone (Liou et al., 2021). Limited anoxic zones in or post-denitrification or optimizing the aeration strategy. IA could be
smaller granules negatively affect SND efficiency (Layer et al., 2020). implemented during the aerobic phase to enhance the persistence of the
We et al. (2020) reported that a granule diameter of 0.9 mm could anoxic zones and favor denitrification. Layer et al. (2020) tested two
significantly improve TN RE. In contrast, Derlon et al. (2016) did not different strategies, i.e., 2-step aeration and IA, to improve SND and TN
observe the occurrence of SND due to the small size of granules (0.25 < removal in AGS systems treating real municipal wastewater. 2-step
d < 0.63 mm), which limited denitrification due to limited extension aeration consisted in maintaining the DO at 2.0 mg L− 1 for 15 min
and/or carbon availability of the anoxic zone (Liu et al., 2010; Wagner and at 0.5 mg L− 1 for the residual 270 min of the aerobic phase, while IA
et al., 2015; Wang et al., 2009). Chen et al. (2011) reported an increase was maintained by turning on aeration every 15 min (DO range = 0–2.0
of TN RE from 68 % to 72 % with the increase of the granule size from mg L− 1). Denitrification efficiency increased from 14 to 37 % to 65–79 %
0.7 to 1.5 mm. In order to achieve an appropriate development of the when applying these strategies, resulting in an increase in TN RE from
anoxic and aerobic zones, the optimal average diameter of aerobic 13 % to > 65 %. In another study with real wastewater, Campo et al.
granules should be around 2.0 mm (Campo et al., 2020). (2020) applied an aeration phase with two sub-phases, the first being DO
controlled (1.4–1.6 mg L− 1), the second without control to allow DO
6.3.2. Source of organic carbon depletion and denitrification of residual NO−x . The applied DO control
Carbon source influences the microbial community composition and strategy promoted high TN RE (71 %) and suppressed NOB activity,
density of the granules (Layer et al., 2019), and thus, can affect N reducing carbon requirement for N removal.
removal in AGS systems. de Sousa Rollemberg et al. (2019) studied the
effects of three different carbon sources (acetate, ethanol and glucose) 6.3.4. Shear force
on SND efficiency in three AGS-SBR systems. Acetate feeding resulted in One of the main factors to control granule stability is the hydraulic
the formation of large granules with an average diameter of 1.5 mm, shear force, which is mainly affected by aeration intensity, mixing speed
favored microbial diversity, and led to TN RE of 72 % and total phos­ and reactor shape (He et al., 2019a; Yan et al., 2019b). Shear stress in
phorus (TP) RE of 42 %, being respectively 20–30 % and 35–100 % AGS systems arises from liquid/gas flows and attrition between particles
higher than the REs achieved with the other substrates. However, partial (Bengtsson et al., 2018). When exposed to high shear stress (>2 cm s− 1),
disintegration of the granules was observed and attributed to the for­ aerobic granules become more compact and stable, while low superficial
mation of a thick biofilm limiting O2 diffusion and leading to the for­ gas velocity (SGV) generally results in larger granules with a less
mation of anaerobic gaseous catabolites within the granules. Ethanol- compact structure and low settling velocity (de Sousa Rollemberg et al.,
fed reactors led to stable granules and good TN RE (53 %), but TP RE 2018). This is likely due to different bacterial stratification and pro­
was low (31 %). Glucose showed the lowest TN and TP REs (44 % and 21 duced polymers during the granulation process. However, granule sta­
%, respectively) as well as the lowest microbial diversity. N removal in bility may also be achieved at low SGV depending on the influent
AGS systems fed with acetate and ethanol mainly occurred via SND strength. Devlin et al. (2017) showed that stable granulation was ob­
during the aerobic period. NO–3 and NO–2 accumulation with acetate was tained at SGV as low as 0.41 cm s− 1 when low-strength wastewater (340
higher than with ethanol, indicating that ethanol feeding promoted mg COD L− 1) was fed to the system, resulting in a TN RE of 62 % and
denitrification. NO–2 accumulated more than NO–3, meaning that part of COD and TP REs > 90 %. In contrast, utilization of medium-strength
NO–2 was directly reduced to N2 through the shortcut SND pathway. (630 mg COD L− 1) and high-strength (1300 mg COD L− 1) wastewater
Layer et al. (2019) used four different influent compositions char­ resulted in poor granulation and filamentous growth, which could be
acterized by increasing concentrations of non-diffusible organic carbon attributed to incomplete carbon uptake during the anaerobic phase with
to study how carbon diffusivity influences SND and the granulation subsequent proliferation of filamentous HAB. Moreover, the granules
process under anaerobic/aerobic conditions. The diffusible organic obtained with medium-strength wastewater were incapable of nitrifi­
fraction was represented by VFA and soluble fermentable substrates cation, resulting in a TN RE of only 30 %, mostly due to N assimilation
(glucose and amino acids), while the non-diffusible fraction was repre­ into new biomass. In another study, He et al. (2017b) studied the effect
sented by particulate substrates such as peptone and starch. Results of three low SGVs (0.17, 0.11 and 0.04 cm s− 1) for the treatment of low-
showed that the presence of diffusible organic carbon in wastewater strength wastewater (feed C/N ratio of 3.3) on an AGS system operated
promoted granulation. The higher the concentration of diffusible carbon on an anaerobic/oxic/anoxic mode. The granule size increased from 1.5
in the influent, the higher was the growth in the deep layers of the mm to 1.54, 1.69 and 1.77 mm at decreasing SGV, while the settling
granules, leading to the formation of dense and large granules (d = 1–3 velocity decreased from 58 to 51 and 40 m h− 1. Nevertheless, the system
mm). In contrast, the presence of non-diffusible organic substrate remained stable, and no sludge bulking nor granule disintegration
hampered the granulation process and resulted in slower reactor start- occurred, indicating a great adaptability of the granules to decreasing
up, smaller granules (0.25–0.63 mm) and presence of 20–40 % (% of SGV. The decrease in SGV improved SND performance, as TN RE
total suspended solids) of sludge flocs in the reactors. Particulate organic increased from 64 % to 94 % when SGV was decreased from 0.17 to 0.04
carbon cannot diffuse in granules, so the high availability of organics in cm s− 1, although more time (+30 min) was required to achieve complete
the aerobic phase supports the growth of HAB. This can lead to the nitrification. Increase in N removal was attributed to better anoxic
formation of flocs and filamentous granules with poor settling properties conditions in the granules at lower SGVs, which may result from the
and nutrient removal capability. The share of diffusible organic carbon increased particle size and EPS concentration of the granule. This
in the feed had a significant impact on SND performance of the AGS improved SND during the aerobic phase and denitrification during the
systems. The highest TN RE of 77 % was achieved in the reactor fed with anoxic phase. Applying low SGVs also results in a reduced energy

11
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

requirement and operational costs. However, despite better TN removal Chu, L., Wang, J., 2011. Comparison of polyurethane foam and biodegradable polymer
as carriers in moving bed biofilm reactor for treating wastewater with a low C/N
and energy saving, the poor granule settleability obtained at the lowest
ratio. Chemosphere 83 (1), 63–68.
SGV tested suggests that higher SGVs should be applied for stable Chyi, A., We, E., Aris, A., Azimah, N., Zain, M., 2020. A review of the treatment of
operation. low–medium strength domestic wastewater using aerobic granulation technology.
Environ. Sci. Water Res. Technol. 6, 464–490.
Coma, M., Verawaty, M., Pijuan, M., Yuan, Z., Bond, P.L., 2012. Enhancing aerobic
7. Conclusions granulation for biological nutrient removal from domestic wastewater. Bioresour.
Technol. 103 (1), 101–108.
Dahalan, F.A., Abdullah, N., Yuzir, A., Olsson, G., Salmiati, Hamdzah, M., Din, M.F.M.,
SND is a promising bioprocess that can be applied to replace con­ Ahmad, S.A., Khalil, K.A., Anuar, A.N., Noor, Z.Z., Ujang, Z., 2015. A proposed
ventional N removal at WWTPs, and more convenient if performed via aerobic granules size development scheme for aerobic granulation process.
NO–2-shortcut. However, applying SND increases BNR complexity as Bioresour. Technol. 181, 291–296.
de Kreuk, M.K., Heijnen, J.J., van Loosdrecht, M.C.M., 2005. Simultaneous COD,
environmental and biological factors must be attentively controlled to
nitrogen, and phosphate removal by aerobic granular sludge. Biotechnol. Bioeng. 90
maintain high performances and low N2O emissions. At lab-scale, SND (6), 761–769.
biofilms demonstrated high N RE (>90 %) for all technologies: AGS and de Sousa Rollemberg, S.L., Mendes Barros, A.R., Milen Firmino, P.I., Bezerra dos
Santos, A., 2018. Aerobic granular sludge: Cultivation parameters and removal
MBBR systems also allowed P RE > 80 %. Operational factors are
mechanisms. Bioresour. Technol. 270, 678–688.
technology-specific and must be optimized accordingly. Further de Sousa Rollemberg, S.L., de Oliveira, L.Q., Barros, A.R.M., Melo, V.M.M., Firmino, P.I.
research at pilot and full scale is required to assess biofilm stability M., dos Santos, A.B., 2019. Effects of carbon source on the formation, stability,
under fluctuating feed conditions to determine impacts on SND bioactivity and biodiversity of the aerobic granule sludge. Bioresour. Technol. 278,
195–204.
efficiency. Derlon, N., Wagner, J., da Costa, R.H.R., Morgenroth, E., 2016. Formation of aerobic
granules for the treatment of real and low-strength municipal wastewater using a
sequencing batch reactor operated at constant volume. Water Res. 105, 341–350.
CRediT authorship contribution statement
Devlin, T.R., di Biase, A., Kowalski, M., Oleszkiewicz, J.A., 2017. Granulation of
activated sludge under low hydrodynamic shear and different wastewater
Francesco Di Capua: Conceptualization, Investigation, Supervision, characteristics. Bioresour. Technol. 224, 229–235.
Di Bella, G., Torregrossa, M., 2013. Simultaneous nitrogen and organic carbon removal
Writing – original draft, Writing – review & editing. Francesca Ianna­
in aerobic granular sludge reactors operated with high dissolved oxygen
cone: Visualization, Investigation, Writing – original draft. Fabrizio concentration. Bioresour. Technol. 142, 706–713.
Sabba: Visualization, Writing – original draft, Writing – review & di Biase, A., Kowalski, M.S., Devlin, T.R., Oleszkiewicz, J.A., 2019. Moving bed biofilm
editing. Giovanni Esposito: Supervision, Writing – review & editing. reactor technology in municipal wastewater treatment: a review. J. Environ.
Manage. 247, 849–866.
Di Capua, F., Adani, F., Pirozzi, F., Esposito, G., Giordano, A., 2021. Air side-stream
ammonia stripping in a thin film evaporator coupled to high-solid anaerobic
Declaration of Competing Interest digestion of sewage sludge: process performance and interactions. J. Environ.
Manage. 295, 113075.
Di Capua, F., de Sario, S., Ferraro, A., Petrella, A., Race, M., Pirozzi, F., Fratino, U.,
The authors declare that they have no known competing financial Spasiano, D., 2022. Phosphorous removal and recovery from urban wastewater:
interests or personal relationships that could have appeared to influence Current practices and new directions. Sci. Total Environ. 823, 153750.
the work reported in this paper. Eldyasti, A., Nakhla, G., Zhu, J., 2014. Influence of biofilm thickness on nitrous oxide
(N2O) emissions from denitrifying fluidized bed bioreactors (DFBBRs). J. Biotechnol.
192, 281–290.
Data availability Feng, L., Jia, R., Zeng, Z., Yang, G., Xu, X., 2018. Simultaneous nitrification –
denitrification and microbial community profile in an oxygen-limiting intermittent
No data was used for the research described in the article. aeration SBBR with biodegradable carriers. Biodegradation 29 (5), 473–486.
Feng, Q., Wang, Y., Wang, T., Zheng, H., Chu, L., Zhang, C., Chen, H., Kong, X., Xing, X.
H., 2012. Effects of packing rates of cubic-shaped polyurethane foam carriers on the
References microbial community and the removal of organics and nitrogen in moving bed
biofilm reactors. Bioresour. Technol. 117, 201–207.
Ferrentino, R., Ferraro, A., Mattei, M.R., Esposito, G., Andreottola, G., 2018. Process
Al-Amshawee, S., Yunus, M.Y.B.M., Vo, D.-V., Tran, N.H., 2020. Biocarriers for biofilm
performance optimization and mathematical modelling of a SBR-MBBR treatment at
immobilization in wastewater treatments: a review. Environ. Chem. Lett. 18 (6),
low oxygen concentration. Process Biochem. 75, 230–239.
1925–1945.
Fu, B.o., Liao, X., Ding, L., Ren, H., 2010. Characterization of microbial community in an
Asik, G., Yilmaz, T., Di Capua, F., Ucar, D., Esposito, G., Sahinkaya, E., 2021. Sequential
aerobic moving bed biofilm reactor applied for simultaneous nitrification and
sulfur-based denitrification/denitritation and nanofiltration processes for drinking
denitrification. World J. Microbiol. Biotechnol. 26 (11), 1981–1990.
water treatment. J. Environ. Manage. 295, 113083.
Ge, G., Zhao, J., Li, X., Ding, X., Chen, A., Chen, Y., Hu, B., Wang, S., 2017. Effects of
Bengtsson, S., de Blois, M., Wilén, B.-M., Gustavsson, D., 2018. Treatment of municipal
influent COD/N ratios on nitrous oxide emission in a sequencing biofilm batch
wastewater with aerobic granular sludge. Crit. Rev. Environ. Sci. Technol. 48 (2),
reactor for simultaneous nitrogen and phosphorus removal. Sci. Rep. 7.
119–166.
Gilbert, E.M., Agrawal, S., Brunner, F., Schwartz, T., Horn, H., Lackner, S., 2014.
Butcher, R.W., 2016. Studies in the ecology of rivers : VII. The algae of organically
Response of different Nitrospira species to anoxic periods depends on operational
enriched waters. J. Ecol. 35, 186–191.
DO. Environ. Sci. Technol. 48, 2934–2941.
Campo, R., Sguanci, S., Caffaz, S., Mazzoli, L., Ramazzotti, M., Lubello, C., Lotti, T.,
Gilbert, E.M., Agrawal, S., Schwartz, T., Horn, H., Lackner, S., 2015. Comparing different
2020. Efficient carbon, nitrogen and phosphorus removal from low C/N real
reactor configurations for Partial Nitritation/Anammox at low temperatures. Water
domestic wastewater with aerobic granular sludge. Bioresour. Technol. 305, 122961.
Res. 81, 92–100.
Cao, Y., Zhang, C., Rong, H., Zheng, G., Zhao, L., 2017. The effect of dissolved oxygen
Gu, Y.Q., Li, T.T., Li, H.Q., 2018. Biofilm formation monitored by confocal laser scanning
concentration (DO) on oxygen diffusion and bacterial community structure in
microscopy during startup of MBBR operated under different intermittent aeration
moving bed sequencing batch reactor (MBSBR). Water Res. 108, 86–94.
modes. Process Biochem. 74, 132–140.
Chai, H., Xiang, Y., Chen, R., Shao, Z., Gu, L., Li, L., He, Q., 2019. Enhanced simultaneous
Gu, S., Wang, S., Yang, Q., Yang, P., Peng, Y., 2012. Start up partial nitrification at low
nitrification and denitrification in treating low carbon-to-nitrogen ratio wastewater:
temperature with a real-time control strategy based on blower frequency and pH.
treatment performance and nitrogen removal pathway. Bioresour. Technol. 280,
Bioresour. Technol. 112, 34–41.
51–58.
Guo, G., Wang, Y., Hao, T., Wu, D., Chen, G.H., 2017. Enzymatic nitrous oxide emissions
Chen, F., Bi, X., Ng, H.Y., 2016. Effects of bio-carriers on membrane fouling mitigation in
from wastewater treatment. Front. Environ. Sci. Eng. 2018 121 12, 1–12.
moving bed membrane bioreactor. J. Memb. Sci. 499, 134–142.
Han, F., Ye, W., Wei, D., Xu, W., Du, B., Wei, Q., 2018. Simultaneous nitrification-
Chen, H., Li, A., Wang, Q., Cui, D.i., Cui, C., Ma, F., 2018a. Nitrogen removal
denitrification and membrane fouling alleviation in a submerged biofilm membrane
performance and microbial community of an enhanced multistage A/O biofilm
bioreactor with coupling of sponge and biodegradable PBS carrier. Bioresour.
reactor treating low- strength domestic wastewater. Biodegradation 29 (3), 285–299.
Technol. 270, 156–165.
Chen, F.-Y., Liu, Y.-Q., Tay, J.-H., Ning, P., 2011. Operational strategies for nitrogen
He, Q., Wang, H., Yang, X., Zhou, J., Ye, Y., Chen, D., 2016a. Culture of denitrifying
removal in granular sequencing batch reactor. J. Hazard. Mater. 189 (1-2), 342–348.
phosphorus removal granules with different influent wastewater. Desalin. Water
Chen, X., Yuan, Z., Ni, B.J., 2018b. Nitrite accumulation inside sludge flocs significantly
Treat. 57, 17247–17254.
influencing nitrous oxide production by ammonium-oxidizing bacteria. Water Res.
He, Q., Zhang, S., Zou, Z., Zheng, L., Wang, H., 2016b. Unraveling characteristics of
143, 99–108.
simultaneous nitrification, denitrification and phosphorus removal (SNDPR) in an
Chrispim, M.C., Scholz, M., Nolasco, M.A., 2019. Phosphorus recovery from municipal
aerobic granular sequencing batch reactor. Bioresour. Technol. 220, 651–655.
wastewater treatment: critical review of challenges and opportunities for developing
countries. J. Environ. Manage. 248, 109268.

12
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

He, Q., Zhou, J., Wang, H., Zhang, J., Wei, L., 2016c. Microbial population dynamics Liu, T., He, X., Jia, G., Xu, J., Quan, X., You, S., 2020a. Simultaneous nitrification and
during sludge granulation in an A/O/A sequencing batch reactor. Bioresour. denitrification process using novel surface-modified suspended carriers for the
Technol. 214, 1–8. treatment of real domestic wastewater. Chemosphere 247, 125831.
He, Q., Song, Q., Zhang, S., Zhang, W., Wang, H., 2017a. Simultaneous nitrification, Liu, T., Jia, G., Xu, J., He, X., Quan, X., 2021. Simultaneous nitrification and
denitrification and phosphorus removal in an aerobic granular sequencing batch denitrification in continuous flow MBBR with novel surface-modified carriers.
reactor with mixed carbon sources: reactor performance, extracellular polymeric Environ. Technol. 42, 3607–3617.
substances and microbial successions. Chem. Eng. J. 331, 841–849. Liu, X., Kim, M., Nakhla, G., Andalib, M., Fang, Y., 2020b. Partial nitrification-reactor
He, Q., Zhang, W., Zhang, S., Wang, H., 2017b. Enanced nitrogen removal in an aerobic configurations, and operational conditions: Performance analysis. J. Environ. Chem.
granular sequencing batch reactor performing simultaneous nitrification, Eng. 8, 103984.
endogenous denitrification and phosphorus removal with low superficial gas Liu, Y.Q., Moy, B., Kong, Y.H., Tay, J.H., 2010. Formation, physical characteristics and
velocity. Chem. Eng. J. 326, 1223–1231. microbial community structure of aerobic granules in a pilot-scale sequencing batch
He, Q., Zhou, J., Song, Q., Zhang, W., Wang, H., Liu, L., 2017c. Elucidation of microbial reactor for real wastewater treatment. Enzyme Microb. Technol. 46, 520–525.
characterization of aerobic granules in a sequencing batch reactor performing Liu, G., Wang, J., 2013. Long-term low DO enriches and shifts nitrifier community in
simultaneous nitrification, denitrification and phosphorus removal at varying carbon activated sludge. Environ. Sci. Technol. 47, 5109–5117.
to phosphorus ratios. Bioresour. Technol. 241, 127–133. Lo, I.W., Lo, K.V., Mavinic, D.S., Shiskowski, D., Ramey, W., 2010. Contributions of
He, Q., Chen, L., Zhang, S., Chen, R., Wang, H., 2019a. Hydrodynamic shear force shaped biofilm and suspended sludge to nitrogen transformation and nitrous oxide emission
the microbial community and function in the aerobic granular sequencing batch in hybrid sequencing batch system. J. Environ. Sci. 22, 953–960.
reactors for low carbon to nitrogen (C/N) municipal wastewater treatment. Lochmatter, S., Gonzalez-gil, G., Holliger, C., 2013. Optimized aeration strategies for
Bioresour. Technol. 271, 48–58. nitrogen and phosphorus removal with aerobic granular sludge. Water Res. 47,
He, Q., Wang, H., Chen, L., Gao, S., Zhang, W., Song, J., Yu, J., 2019b. Elevated salinity 6187–6197.
deteriorated enhanced biological phosphorus removal in an aerobic granular sludge Lu, Y., Wang, H., Kotsopoulos, T.A., Zeng, R.J., 2016. Advanced phosphorus recovery
sequencing batch reactor performing simultaneous nitrification, denitrification and using a novel SBR system with granular sludge in simultaneous nitrification,
phosphorus removal. J. Hazard. Mater. 390, 121782. denitrification and phosphorus removal process. Appl. Microbiol. Biotechnol. 100,
He, Q., Song, J., Zhang, W., Gao, S., Wang, H., Yu, J., 2020. Enhanced simultaneous 4367–4374.
nitrification, denitrification and phosphorus removal through mixed carbon source Ma, W., Han, Y., Ma, W., Han, H., Zhu, H., Xu, C., Li, K., Wang, D., 2017. Enhanced
by aerobic granular sludge. J. Hazard. Mater. 382, 121043. nitrogen removal from coal gasification wastewater by simultaneous nitrification
Hu, Z., Wessels, H.J.C.T., Alen, T., Jetten, M.S.M., Kartal, B., 2019. Nitric oxide- and denitrification (SND) in an oxygen-limited aeration sequencing batch biofilm
dependent anaerobic ammonium oxidation oxidation. Nat. Commun. 10, 1244. reactor. Bioresour. Technol. 244, 84–91.
Huang, X., Wei, C.H., Yu, K.C., 2008. Mechanism of membrane fouling control by Massoompour, A.R., Borghei, S.M., Raie, M., 2020. Enhancement of biological nitrogen
suspended carriers in a submerged membrane bioreactor. J. Memb. Sci. 309, 7–16. removal performance using novel carriers based on the recycling of waste materials.
Iannacone, F., Di Capua, F., Granata, F., Gargano, R., Pirozzi, F., Esposito, G., 2019. Water Res. 170, 115340.
Effect of carbon-to-nitrogen ratio on simultaneous nitrification denitrification and McQuarrie, J.P., Boltz, J.P., 2011. Moving bed biofilm reactor technology: process
phosphorus removal in a microaerobic moving bed biofilm reactor. J. Environ. applications, design, and performance. Water Environ. Res. 83, 560–575.
Manage. 250, 109518. Meng, F., Chae, S., Drews, A., Kraume, M., Shin, H., 2009. Recent advances in membrane
Iannacone, F., Di Capua, F., Granata, F., Gargano, R., Esposito, G., 2020. Simultaneous bioreactors (MBRs): membrane fouling and membrane material. Water Res. 43,
nitrification, denitrification and phosphorus removal in a continuous-flow moving 1489–1512.
bed biofilm reactor alternating microaerobic and aerobic conditions. Bioresour. Mosquera-Corral, A., Kreuk, M.K.D., Heijnen, J.J., Loosdrecht, M.C.M.V., 2005. Effects of
Technol. 310, 123453. oxygen concentration on N-removal in an aerobic granular sludge reactor. Water
Iannacone, F., Di Capua, F., Granata, F., Gargano, R., Esposito, G., 2021. Shortcut Res. 39, 2676–2686.
nitrification-denitrification and biological phosphorus removal in acetate- and Moura, R.B., Damianovic, M.H.R.Z., Foresti, E., 2012. Nitrogen and carbon removal from
ethanol-fed moving bed biofilm reactors under microaerobic/aerobic conditions. synthetic wastewater in a vertical structured-bed reactor under intermittent
Bioresour. Technol. 330, 124958. aeration. J. Environ. Manage. 98, 163–167.
Isanta, E., Suárez-ojeda, M.E., Val, Á., Morales, N., Pérez, J., Carrera, J., 2012. Long term Nguyen, T.T., Ngo, H.H., Guo, W., Johnston, A., Listowski, A., 2010. Effects of sponge
operation of a granular sequencing batch reactor at pilot scale treating a low- size and type on the performance of an up-flow sponge bioreactor in primary treated
strength wastewater. Chem. Eng. J. 198–199, 163–170. sewage effluent treatment. Bioresour. Technol. 101, 1416–1420.
Jia, Y., Zhou, M., Chen, Y., Hu, Y., Luo, J., 2020. Insight into short-cut of simultaneous Ødegaard, H., 2006. Innovations in wastewater treatment: The moving bed biofilm
nitrification and denitrification process in moving bed biofilm reactor: effects of process. Water Sci. Technol. 53, 17–33.
carbon to nitrogen ratio. Chem. Eng. J. 400, 125905. Peng, Y., Zhu, G., 2006. Biological nitrogen removal with nitrification and denitrification
Khanongnuch, R., Di Capua, F., Lakaniemi, A.M., Rene, E.R., Lens, P.N.L., 2019. Long- via nitrite pathway. Appl. Microbiol. Biotechnol. 73, 15–26.
term performance evaluation of an anoxic sulfur oxidizing moving bed biofilm Piotr, Ś., Cydzik-kwiatkowska, A., 2018. Performance and microbial characteristics of
reactor under nitrate limited conditions. Environ. Sci. Water Res. Technol. 5, biomass in a full-scale aerobic granular sludge wastewater treatment plant. Environ.
1072–1081. Sci. Pollut. Res. 25, 1655–1669.
Kinh, C.T., Riya, S., Hosomi, M., Terada, A., 2017a. Identification of hotspots for NO and Pishgar, R., Dominic, J.A., Sheng, Z., Tay, J.H., 2019. Influence of operation mode and
N2O production and consumption in counter- and co-diffusion biofilms for wastewater strength on aerobic granulation at pilot scale: startup period, granular
simultaneous nitrification and denitrification. Bioresour. Technol. 245, 318–324. sludge characteristics, and effluent quality. Water Res. 160, 81–96.
Kinh, C.T., Suenaga, T., Hori, T., Riya, S., Hosomi, M., Smets, B.F., Terada, A., 2017b. Pronk, M., de Kreuk, M.K., de Bruin, B., Kamminga, P., Kleerebezem, R., van
Counter-diffusion biofilms have lower N2O emissions than co-diffusion biofilms Loosdrecht, M.C.M., 2015. Full scale performance of the aerobic granular sludge
during simultaneous nitrification and denitrification: Insights from depth-profile process for sewage treatment. Water Res. 84, 207–217.
analysis. Water Res. 124, 363–371. Qi, C., Zhou, Y., Suenaga, T., Oba, K., Lu, J., Wang, G., Zhang, L., Yoon, S., Terada, A.,
Kishida, N., Kim, J., Tsuneda, S., Sudo, R., 2006. Anaerobic/oxic/anoxic granular sludge 2022. Organic carbon determines nitrous oxide consumption activity of clade I and II
process as an effective nutrient removal process utilizing denitrifying polyphosphate- nosZ bacteria: genomic and biokinetic insights. Water Res. 209, 117910.
accumulating organisms. Water Res. 40, 2303–2310. Rahimi, S., Modin, O., Mijakovic, I., 2020. Technologies for biological removal and
Kishida, N., Tsuneda, S., Sakakibara, Y., Kim, J.H., Sudo, R., 2008. Real-time control recovery of nitrogen from wastewater. Biotechnol. Adv. 43, 107570.
strategy for simultaneous nitrogen and phosphorus removal using aerobic granular Rittmann, B.E., McCarty, P.L., 2012. Environmental Biotechnology: Principles and
sludge. Water Sci. Technol. 58, 445–450. Applications. McGraw Hill Education.
Layer, M., Adler, A., Reynaert, E., Hernandez, A., Pagni, M., Morgenroth, E., 2019. Roots, P., Wang, Y., Rosenthal, A.F., Griffin, J.S., Sabba, F., Petrovich, M., Yang, F.,
Organic substrate diffusibility governs microbial community composition, nutrient Kozak, J.A., Zhang, H., Wells, G.F., 2019. Comammox Nitrospira are the dominant
removal performance and kinetics of granulation of aerobic granular sludge. Water ammonia oxidizers in a mainstream low dissolved oxygen nitrification reactor.
Res. X 4, 100033. Water Res. 157, 396–405.
Layer, M., Garcia, M., Hernandez, A., Reynaert, E., Morgenroth, E., Derlon, N., 2020. Roots, P., Sabba, F., Rosenthal, A.F., Wang, Y., Yuan, Q., Rieger, L., Yang, F., Kozak, J.A.,
Limited simultaneous nitrification-denitrification (SND) in aerobic granular sludge Zhang, H., Wells, G.F., 2020. Integrated shortcut nitrogen and biological phosphorus
systems treating municipal wastewater: Mechanisms and practical implications. removal from mainstream wastewater: process operation and modeling. Environ.
Water Res. X 7, 100048. Sci. Water Res. Technol. 6, 566–580.
Lim, J., Seng, C., Lim, P., Ng, S., Sujari, A.A., 2011. Nitrogen removal in moving bed Sabba, F., Picioreanu, C., Pérez, J., Nerenberg, R., 2015. Hydroxylamine diffusion can
sequencing batch reactor using polyurethane foam cubes of various sizes as carrier enhance N2O emissions in nitrifying biofilms: a modeling study. Environ. Sci.
materials. Bioresour. Technol. 102, 9876–9883. Technol. 49, 1486–1494.
Lim, J., Lim, P., Seng, C., 2012. Enhancement of nitrogen removal in moving bed Sabba, F., Calhoun, J., Johnson, B.R., Daigger, G.T., Kovács, R., Takács, I., Boltz, J.,
sequencing batch reactor with intermittent aeration during REACT period. Chem. 2017a. Applications of mobile carrier biofilm modelling for wastewater treatment
Eng. J. 197, 199–203. processes. Lect. Notes Civ. Eng. 4, 508–512.
Liou, H.C., Sabba, F., Wang, Z., Wells, G., Balogun, O., 2021. Layered viscoelastic Sabba, F., McNamara, P., Redmond, E., Ruff, C., Young, M., Downing, L., 2022. Lab-scale
properties of granular biofilms. Water Res. 202, 117394. data and microbial community structure suggest shortcut nitrogen removal as the
Liu, T., Jia, G., Quan, X., 2018. Accelerated start-up and microbial community structures predominant nitrogen removal mechanism in post-aerobic digestion (PAD). Water
of simultaneous nitrification and denitrification by using novel suspended carriers. Environ. Res. 94, e10762.
Chem. Technol. Biotechnol. 93, 577–584. Sabba, F., Picioreanu, C., Boltz, J.P., Nerenberg, R., 2017b. Predicting N2O emissions
from nitrifying and denitrifying biofilms: A modeling study. Water Sci. Technol. 75,
530–538.

13
F. Di Capua et al. Bioresource Technology 361 (2022) 127702

Sabba, F., Terada, A., Wells, G., Smets, B.F., Nerenberg, R., 2018. Nitrous oxide emissions as revealed by microbial community structures. Bioprocess Biosyst. Eng. 43,
from biofilm processes for wastewater treatment. Appl. Microbiol. Biotechnol. 102, 1833–1846.
9815–9829. Wang, H., Song, Q., Wang, J., Zhang, H., He, Q., Zhang, W., Song, J., Zhou, J., Li, H.,
Sandip, M., Kalyanraman, V., 2019. Enhanced simultaneous nitri-denitrification in 2018. Simultaneous nitrification, denitrification and phosphorus removal in an
aerobic moving bed biofilm reactor containing polyurethane foam-based carrier aerobic granular sludge sequencing batch reactor with high dissolved oxygen: effects
media. Water Sci. Technol. 79, 510–517. of carbon to nitrogen ratios. Sci. Total Environ. 642, 1145–1152.
Sarma, S.J., Tay, J., 2018. Carbon, nitrogen and phosphorus removal mechanisms of Xia, Z., Wang, Q., She, Z., Gao, M., Zhao, Y., Guo, L., Jin, C., 2019. Nitrogen removal
aerobic granules. Crit. Rev. Biotechnol. 38, 1077–1088. pathway and dynamics of microbial community with the increase of salinity in
Schreiber, F., Wunderlin, P., Udert, K.M., Wells, G.F., 2012. Nitric oxide and nitrous simultaneous nitrification and denitrification process. Sci. Total Environ. 697,
oxide turnover in natural and engineered microbial communities: biological 134047.
pathways, chemical reactions, and novel technologies. Front. Microbiol. 3, 372. Yan, L., Zhang, S., Hao, G., Zhang, X., Ren, Y., Wen, Y., Guo, Y., 2016. Simultaneous
Seifi, M., Fazaelipoor, M.H., 2012. Modeling simultaneous nitrification and nitrification and denitrification by EPSs in aerobic granular sludge enhanced
denitrification (SND) in a fluidized bed biofilm reactor. Appl. Math. Model. 36, nitrogen removal of ammonium-nitrogen-rich wastewater. Bioresour. Technol. 202,
5603–5613. 101–106.
She, Z., Wu, L., Wang, Q., Gao, M., Jin, C., Zhao, Y., Zhao, L., Guo, L., 2017. Salinity Yan, L., Liu, S., Liu, Q., Zhang, M., Liu, Y., Wen, Y., Chen, Z., 2019a. Improved
effect on simultaneous nitrification and denitrification, microbial characteristics in a performance of simultaneous nitrification and denitrification via nitrite in an
hybrid sequencing batch biofilm reactor. Bioprocess Biosyst. Eng. 41, 67–75. oxygen-limited SBR by alternating the DO. Bioresour. Technol. 275, 153–162.
Shourjeh, M.S., Kowal, P., Lu, X., Xie, L., Drewnowski, J., 2021. Development of Yan, L., Zhang, M., Liu, Y., Liu, C., Zhang, Y., Liu, S., Yu, L., 2019b. Enhanced nitrogen
strategies for AOB and NOB competition supported by mathematical modeling in removal in an aerobic granular sequencing batch reactor under low DO
terms of successful deammonification implementation for energy-efficient WWTPs. concentration: role of extracellular polymeric substances and microbial community
Processes 9, 562. structure. Bioresour. Technol. 289, 121651.
Song, Z., Zhang, X., Hao, H., Guo, W., Song, P., Zhang, Y., 2019. Zeolite powder based Yang, S., Tay, J., Liu, Y., 2003. A novel granular sludge sequencing batch reactor for
polyurethane sponges as biocarriers in moving bed biofilm reactor for improving removal of organic and nitrogen from wastewater. J. Biotechnol. 106, 77–86.
nitrogen removal of municipal wastewater. Sci. Total Environ. 651, 1078–1086. Yang, S., Yang, F., Fu, Z., Lei, R., 2009. Comparison between a moving bed membrane
Suarez, C., Piculell, M., Modin, O., Langenheder, S., Persson, F., Hermansson, M., 2019. bioreactor and a conventional membrane bioreactor on organic carbon and nitrogen
Thickness determines microbial community structure and function in nitrifying removal. Bioresour. Technol. 100, 2369–2374.
biofilms via deterministic assembly. Sci. Rep. 9, 5110. Yin, J., Zhang, P., Li, F., Li, G., Hai, B., 2015. Simultaneous biological nitrogen and
Suenaga, T., Riya, S., Hosomi, M., Terada, A., 2018. Biokinetic characterization and phosphorus removal with a sequencing batch reactor-biofilm system. Int.
activities of N2O-reducing bacteria in response to various oxygen levels. Front. Biodeterior. Biodegradation 1–6.
Microbiol. 9, 697. Yu, C., Tu, Q., Huangfu, X., Zhu, Y., Fan, L., Ai, H., Chen, M., He, Q., 2021. Effects of
Sun, S.-P., Pellicer i Nàcher, C., Merkey, B., Zhou, Q., Xia, S.-Q., Sun, J.-H., Smets, B.F., hydraulic retention time on nitrous oxide production rates during nitrification in a
2010. Effective biological nitrogen removal treatment processes for domestic laboratory-scale biological aerated filter reactor. Environ. Technol. Innov. 21,
wastewaters with Low C/N ratios: a review. Environ. Eng. Sci. 27, 111–126. 101342.
Tan, C., Ma, F., Qiu, S., 2013. Impact of carbon to nitrogen ratio on nitrogen removal at a Yuan, Q., Gong, H., Xi, H., Xu, H., Jin, Z., Ali, N., Wang, K., 2019. Strategies to improve
low oxygen concentration in a sequencing batch biofilm reactor. Water Sci. Technol. aerobic granular sludge stability and nitrogen removal based on feeding mode and
67, 612–618. substrate. J. Environ. Sci. 84, 144–154.
Ucar, D., Yilmaz, T., Di Capua, F., Esposito, G., Sahinkaya, E., 2020. Comparison of Zhang, L., Wei, C., Zhang, K., Zhang, C., Fang, Q., Li, S., 2009. Effects of temperature on
biogenic and chemical sulfur as electron donors for autotrophic denitrification in simultaneous nitrification and denitrification via nitrite in a sequencing batch
sulfur-fed membrane bioreactor (SMBR). Bioresour. Technol. 299, 122574. biofilm reactor. Bioprocess Biosyst. Eng. 32, 175–182.
Ucar, D., Di Capua, F., Yücel, A., Nacar, T., Sahinkaya, E., 2021. Effect of nitrogen Zhang, L., Narita, Y., Gao, L., Ali, M., Oshiki, M., Okabe, S., 2017a. Maximum specific
loading on denitrification, denitritation and filtration performances of membrane growth rate of anammox bacteria revisited. Water Res. 116, 296–303.
bioreactors fed with biogenic and chemical elemental sulfur. Chem. Eng. J. 419, Zhang, X., Song, Z., Guo, W., Lu, Y., Qi, L., Wen, H., Hao, H., 2017b. Behavior of nitrogen
129514. removal in an aerobic sponge based moving bed biofilm reactor. Bioresour. Technol.
Van Kessel, M.A.H.J., Speth, D.R., Albertsen, M., Nielsen, P.H., Op Den Camp, H.J.M., 245, 1282–1285.
Kartal, B., Jetten, M.S.M., Lücker, S., 2015. Complete nitrification by a single Zhao, Y., Liu, D., Huang, W., Yang, Y., Ji, M., Nghiem, L.D., Trinh, Q.T., Tran, N.H.,
microorganism. Nature 528, 555–559. 2019. Insights into biofilm carriers for biological wastewater treatment processes:
van Puijenbroek, P.J.T.M., Beusen, A.H.W., Bouwman, A.F., 2019. Global nitrogen and current state-of-the-art, challenges, and opportunities. Bioresour. Technol. 288,
phosphorus in urban waste water based on the Shared Socio-economic pathways. 121619.
J. Environ. Manage. 231, 446–456. Zhou, Y., Pijuan, M., Zeng, R.J., Yuan, Z., 2008. Free nitrous acid inhibition on nitrous
Wagner, J., Guimarães, L.B., Akaboci, T.R.V., Costa, R.H.R., 2015. Aerobic granular oxide reduction by a denitrifying-enhanced biological phosphorus removal sludge.
sludge technology and nitrogen removal for domestic wastewater treatment. Water Environ. Sci. Technol. 42, 8260–8265.
Sci. Technol. 71, 1040–1047. Zhou, Y., Oehmen, A., Lim, M., Vadivelu, V., Ng, W.J., 2011. The role of nitrite and free
Wagner, J., Helena, R., 2013. Aerobic granulation in a sequencing batch reactor using nitrous acid (FNA) in wastewater treatment plants. Water Res. 45, 4672–4682.
real domestic wastewater. J. Environ. Eng. 139, 1391–1396. Zhou, Y., Suenaga, T., Qi, C., Riya, S., Hosomi, M., Terada, A., 2022b. Temperature and
Wang, K.M., Jiang, S.F., Zhang, Z.H., Ye, Q.Q., Zhang, Y.C., Zhou, J.H., Hong, Q.K., Yu, J. oxygen level determine N2O respiration activities of heterotrophic N2O-reducing
M., Wang, H.Y., 2020b. Impact of static biocarriers on the microbial community, bacteria: Biokinetic study. Biotechn. Bioeng. 118 (3), 1330–1341.
nitrogen removal and membrane fouling in submerged membrane bioreactor at Zhou, Y., Zhao, S., Suenaga, T., Kuroiwa, M., Riya, S., Terada, A., 2022a. Nitrous oxide-
different COD: N ratios. Bioresour. Technol. 301, 122798. sink capability of denitrifying bacteria impacted by nitrite and pH. Chem. Eng. J.
Wang, F., Lu, S., Wei, Y., Ji, M., 2009. Characteristics of aerobic granule and nitrogen 428, 132402.
and phosphorus removal in a SBR. J. Hazard. Mater. 164, 1223–1227. Zinatizadeh, A.A.L., Ghaytooli, E., 2015. Simultaneous nitrogen and carbon removal
Wang, J., Rong, H., Cao, Y., Zhang, C., 2020a. Factors affecting simultaneous nitrification from wastewater at different operating conditions in a moving bed biofilm reactor
and denitrification (SND) in a moving bed sequencing batch reactor (MBSBR) system (MBBR): Process modeling and optimization. J. Taiwan Inst. Chem. Eng. 53, 98–111.

14

You might also like