Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Science of the Total Environment 693 (2019) 133425

Contents lists available at ScienceDirect

Science of the Total Environment

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

Bio-carrier and operating temperature effect on ammonia removal from


secondary wastewater effluents using moving bed biofilm
reactor (MBBR)
Amal Ashkanani a, Fares Almomani a,⁎, Majeda Khraisheh a, Rahul Bhosale a,
Muhammad Tawalbeh b, Khaled AlJaml a
a
Department of Chemical Engineering, College of Engineering, Qatar University, P.O. Box 2713, Doha, Qatar
b
Sustainable & Renewable Energy Engineering Department, College of Engineering, University of Sharjah, P.O. Box 27272, Sharjah, United Arab Emirates

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

• Ammonia a harmful constituent linked


to acute/chronic toxicities in surface
water
• Impact of biocarriers' characteristics on
ammonia removal by MMBR was evalu-
ated.
• MBBR technology is feasible choice to
remove ammonia from Wastewater.
• MBBR achieved effective removal of
micronutrients at low temperatures.
• Cold temperatures develop thick nitrify-
ing biofilm, less cells and lower activity.

a r t i c l e i n f o a b s t r a c t

Article history: This study investigates the impact of bio-carriers' surface area and shape, wastewater chemistry and operating
Received 4 February 2019 temperature on ammonia removal from real wastewater effluents using Moving bed biofilm reactors (MBBRs)
Received in revised form 15 June 2019 operated with three different AnoxKaldness bio-carriers (K3, K5, and M). The study concludes the surface area
Accepted 15 July 2019
loading rate, specific surface area, and shape of bio-carrier affect ammonia removal under real conditions.
Available online 16 July 2019
MBBR kinetics and sensitivity for temperature changes were affected by bio-carrier type. High surface area bio-
Keywords:
carriers resulted in low ammonia removal and bio-carrier clogging. Significant ammonia removals of 1.420 ±
Secondary effluents 0.06 and 1.103 ± 0.06 g − N/m2. d were achieved by K3(As = 500 m2/m3) at 35 and 20 °C, respectively. Lower
Performance removals were obtained by high surface area bio-carrier K5 (1.123 ± 0.06 and 0.920 ± 0.06 g − N/m2. d) and
Organic loading M (0.456 ± 0.05 and 0.295 ± 0.05 g − N/m2. d) at 35 and 20 °C, respectively. Theta model successfully represents
Total nitrogen ammonia removal kinetics with θ values of 1.12, 1.06 and 1.13 for bio-carrier K3, K5 and M respectively. MBBR
Biofilm growth technology is a feasible choice for treatment of real wastewater effluents containing high ammonia
Clogging concentrations.
Nitrification
© 2019 Elsevier B.V. All rights reserved.

1. Introduction
⁎ Corresponding author at: College of Engineering, Department of Chemical
Engineering, Qatar University, P.O. Box 2713, Doha, Qatar. Conventional wastewater treatment plants (CWWTPs) represent
E-mail address: falmomani@qu.edu.qa (F. Almomani). the majority of treatment processes operated worldwide. CWWTPs

https://doi.org/10.1016/j.scitotenv.2019.07.231
0048-9697/© 2019 Elsevier B.V. All rights reserved.
2 A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425

can effectively remove organic constituent from the wastewater but Furthermore, a fundamental gap in knowledge exists with respect to
they have a limited ability to nitrify ammonia (Nsenga Kumwimba the response of attached biofilm and embedded microbial communities
and Meng, 2019; Houtman, 2010; Al Momani, 2006; Al Momani et al., to effluent wastewater from real treatment processes. In Addition, the
2004). Ammonia is one of the harmful constituents released from effect of bio-carriers loading rate, shape and surface area on the system
WWTPs that is responsible for the occurrence of acute and chronic tox- performance is still lacking and not fully understood. To the best of our
icities in lakes and rivers (Shi et al., 2019; Li et al., 2019; Judd et al., 2017; knowledge, the effect of bio-carrier type and shape on the performance
Almomani and Khraisheh, 2016; Allagui et al., 2014; Allagui et al., 2012; of MBBR was investigated from organic matter perspectives (Ødegaard
Vitse et al., 2005; Marsalek et al., 1997; Oswin and Salomon, 1963). The et al., 1994), or using synthetic wastewater in case of ammonia removal
wastewater effluent regulations in different countries require a mini- (Forrest et al., 2016). There are no published works that investigate the
mum of secondary level wastewater treatment or equivalent effect of these parameters on nitrifying bacteria under real conditions.
(Suchetana et al., 2017; Almomani et al., 2014). In addition, the regula- Accordingly, this study focuses on advancing the fundamental under-
tions specify the effluent concentration limit for organic matters, solids, standing of the effect of different operational conditions on the ammo-
and ammonia (Znad et al., 2018; AlMomani and Örmeci, 2016). Sharma nia removal from real wastewaters. Key parameters such as bio-carrier
and Ahlert (1977), Welander et al. (1997), Zhu and Chen, 2002 and loading rate, size, shape and water chemistry were tested to help the
Pelissari et al. (2017) have shown that the deficiencies in nitrogen com- understanding ammonia removal mechanism under such realistic con-
pound removal in conventional treatment processes are related to the ditions. To achieve this purpose, three lab-scale MBBRs were designed
limited growth of nitrifying bacteria on suspended systems. Hence, and operated using effluents of secondary wastewater process
there is a need for advanced treatment technologies that can be used (ESWW). The impact of bio-carrier shapes and surface area on ammonia
for the removal of these constituents under real operational conditions. removals were considered. Additionally, the study presents new knowl-
Moving bed biofilm reactors (MBBRs) have shown significant am- edge regarding the effects of temperature on the biofilm attachment to
monia removal from wastewater under different working conditions the MBBR bio-carrier. The results will help in the development of funda-
(Delatolla et al., 2010; Delatolla et al., 2009; Andreottola et al., 2002; mental knowledge necessary to effectively design a low-cost solution to
Rusten et al., 1995; Ødegaard et al., 1994; Rusten et al., 1994) and upgrade current DWWTPs.
wide range of temperatures 1–21 °C (Almomani et al., 2014; Zhang
et al., 2014; Welander and Mattiasson, 2003; Welander et al., 1997). 2. Material and methods
The MBBRs are cost-effective continual flow process that benefits from
a small footprint, ease of installation, control and operation and minimal 2.1. Experimental setup
sludge production (90% lower than traditional WWT) (Young et al.,
2017a, 2017b). The installation of MBBR at the upstream of the Ammonia removal experiments were carried out in three identical
CWWTPs where the C/N ratio is low will reduce the effect of organic jacketed MBBRs designated as R1, R2, and R3 as presented in Fig. 1. The
matter on nitrifying bacteria and consequently promote higher rates total operating volume of each reactor is 2.0 L. Each MBBR has a feed
of nitrification. The MBBR systems do not require backwash and the op- pump, pH meter, dissolved oxygen electrode (DO), aeration pipe, dis-
erational capacity can be increased by increasing the percentage fill of charge lines, and air supply. The bio-carriers occupy 30% of the reactor
bio-carrier (Liu et al., 2018). Moreover, the reported resistance to toxic- volume. The Reactor temperature was controlled by adjusting the
ity spikes, shock loads and fluctuation in loading rate are also factors cooling/heating water flow rate through the jacket.
adding to the advantageous use of MBBR systems (Torresi et al., 2019; Three different types of bio-carriers (AnoxKaldness-K3,
Alizadeh et al., 2019; Oleszkiewicz and Barnard, 2006). AnoxKaldness-K5, and AnoxKaldness-M) were used in each reactor.
Typically, the performance of MBBRs depends on wastewater quality The specific surface area available for K3, K5 and M bio-carriers is 500
parameters, temperature, and pH (Bering et al., 2018; Abtahi et al., m2/m3, 800 m2/m3 and 1200 m2/m3, respectively. The feed to all the
2018). Most of the published works ignored the effect of water matrix three reactors was effluents of secondary wastewater (ESWW) of mu-
on the nitrification process and focus on evaluating ammonia removal nicipal wastewater treatment plant from Doha, Qatar. The characteris-
using synthetic wastewater (Rodgers and Zhan, 2004a, 2004b). Al- tics of the ESWW fed to the MBBRs are provided in Table 1.
though the synthetic wastewater might not contain exact concentra-
tions of ammonia as real wastewater, the effects of micro- 2.2. Experimental procedure
constituents, organic matter and water matrix interference are ignored
which affects the reliability of the results. Hoang et al. (2014) investi- Initially, the reactors were fed with SWW and operated at 20 °C for
gated ammonia removal from synthetic wastewater at 20 and 1 °C, 50 days. The SWW was prepared as per the procedure presented by
the reported average removal rates were 82.4 ± 2.5 and 14.8 ± Almomani et al. (2014). The reactor temperature was increased to 30
3.4 g N/m3.d, respectively. Experiments carried out with real wastewa- °C and then to 35 °C keeping the feed as SWW. The feed composition
ter (lagoon effluent) using the same experimental setup showed aver- was then changed by mixing SWW and ESWW in different portion
age removal rates rate of 260.0 g-N/m3·d and 110 g-N/m3·d, starting with 80:20% SWW:ESWW, and increased progressively until
confirming the discrepancy between real and synthetic wastewater reached 100% ESWW. All through the startup period, the temperature
(Almomani et al., 2014). Temperature and ammonia loading are other was fixed at 35 °C, hydraulic retention time (HRT) was maintained at
important factors to consider during the operation of the MMBRs. 5.25 h and the dissolved oxygen (DO) was kept in the range
Houda et al. (2015) confirmed that the nitrification treatment of efflu- 7.4–8.8 mg/L. Knowles et al. (1965) suggested keeping DO in the
ent of urban wastewater treatment depends on ammonia loading and MBRR as high as possible to reduce oxygen demand stress by nitrifying
temperature with higher ammonia removal observed at low ammonia bacteria. Once the feed to the MBBRs reached 100% ESWW the reactors
concentration (20.52 ± 4.89 mg/L) and high temperature (35.45 ± were operated for 160 days at different temperatures. The operation of
3.18 °C). Shore et al. (2012) observed very low ammonia removal the MBBRs can be identified in four phases related to the operational
from industrial wastewater at a temperature higher than 40 °C. temperature; (1) phase 1 operated from day 0 to 40 at 35 °C,
Delatolla et al. (2010) noticed a 73% reduction (0.60 to 0.16 kg-N/m3. (2) phase 2 operated from day 40 to 80 at 20 °C, (3) phase 3 operated
d) in ammonia removal as the temperature decreased from 14 to 3 °C. from day 80 to 120 at 10 °C, and (4) Phase 4 operated from day 120 to
The performance of MBBR under different operational conditions of 160 at 4 °C. The operational parameters (DO, pH, average nitrogen
real wastewater still not well-studied. Few works have considered the source (S0N), average carbon source (S0c), flow rate (Q) and hydraulic
effect of organic matters on the nitrification process by MBBRs (Young retention time (HRT) and other parameters of the MBBRs are summa-
et al., 2017a, 2017b; Young et al., 2016; Almomani et al., 2014). rized in Table 2. Samples were withdrawn daily in each period for
A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425 3

Fig. 1. Schematic of experimental set up for continues flow MBBR.

analysis. The HRTs of the reactors were controlled by adjusting the at 105 °C for 5 h, and ignited it at 550 °C for 15 min. The difference in
speed of the feed pump. Table 2 shows that the DO for all the MBBRs weight was reported as TAS and VAS, respectively. The statistical signif-
ranged from 7.4 to 8.8 mg/L, the pH variation is very small in the icance of the measured TAS and VAS with respect to average value was
range 7.7 to 8.2, average nitrogen source range from 65.2 ± 4.1 to evaluated using the Student's paired t-test. The test was conducted to
67.5 ± 4.2 gN/m3 and average carbon source range from 73.2 ± 5.5 to investigate whether the differences between the individual and average
83.1 ± 8.1 g CODdiss/m3. The feed flow rate (Q) and HRT were fixed at measurements is significant. The analysis was performed by one-way
6.40 L/day and 5.25 h. analysis of variance (ANOVA) using Prism GraphPad statistics software
(Version 7.04, USA). All statistical tests were carried out at a significance
2.3. Analytical methods level of 5%. DO was measured using Orion Star™ DO-meter (Model-
A223).
Samples withdrawn from the MBBRs on daily base were tested for The biofilm Live/Dead test was carried out using LIVE/DEAD kit
− −
ammonia (NH+ 4 − N), pH, nitrite (NO2 ), nitrate NO3 , total phosphorus (FilmTracer™) and following the procedure explained in our previous
and dissolved chemical oxygen demand (CODdiss). The total attached work (Almomani and Khraisheh, 2016). Biofilm morphology (surface
solid (TAS) and volatile attached solids (VAS) to bio-carrier were mea- coverage and thickness) was determined using a variable pressure elec-
sured on days 20, 40, 60, 80, 100, 140 and 160 of operation. The chem- tron scanning microscope (VPSEM). Measurements were performed on
ical oxygen demand (COD) was measured using HACH COD reagents 20, 40, 80, and 160 days of operation.
following the procedure outlined in the Standard Methods, test #
5220D (APHA, 1985; Eaton et al., 2005). Ammonia was tested according 3. Results and discussion
to the Standard Methods, method # 4500–NH3 B and C using Nessler re-
agent and HACH spectrophotometer (DR 2000 HACH, CO, USA) at 3.1. Feed properties
425 nm. Total phosphorous (Ptot) was measured following test #
4500-P from the standard methods (APHA, 1985). NO− 2 was tested at Fig. 2a, b and c show the total nitrogen (TN), ammonia (NH+ 4 − N),
505 nm using HACH diazotization powder pillows (Nitrite Method nitrite (NO− −
2 ), nitrate (NO3 ), COD, Ptot of the influent to MMBRs (R1, R2
8507) and HACH spectrophotometer. NO− 3 was measured at 500 nm and R3) as well as the operating temperature during 160 days of exper-
using a HACH cadmium reduction method (Nitrate Method 10020) imental phases. The water quality parameters fluctuate within the nor-
and HACH spectrophotometer. mal range as results of the change in secondary WWTP performance.
Total attached solids (TAS) and volatile attached solid (VAS) were The concentrations of TN and NH4 in the MBBRs feed ranged from
determined by scrubbing the biofilm attached to one bio-carrier, dry it 63.8 to 68.4 and 59.9 to 65 mg-N/L, respectively. The concentrations of
these two parameters although fluctuate up and down are remained
relatively constant through the operational period apart from high con-
Table 1
Water quality parameters of the effluents of secondary wastewater (ESWW) fed to the
centration value on days 40 and 60 of operation. The change in the con-
MBBRs. centrations of the feed is normal practice for treatment process dealing
with real wastewater. In addition, the fluctuating in influent water qual-
Parameter Lower value Higher value Average
ity parameters to the reactors gives a chance to test the reactor perfor-
NH3 (mgN/L) 58. 65.78 62.00 mance under varied feed concentration. Nitrate (NO− 3 ) and nitrite
NO2 (mgN/L) 0.010 0.098 0.042
(NO− 2 ) ranged from 2.99 to 3.23 and 0.0105 to 0.071 mg-N/L, respec-
NO3 (mgN/L) 2.80 3.84 3.305
COD (mgN/L) 63.27 71.04 66.96 tively; with high variation in the concentration of NO− 3 in the reactors
PO4 (mgN/L) 11.72 12.32 12.40 feed. The COD and Ptot concentrations were in the range at 62.5 to
pH 7.23 8.75 7.76 72.5 mg COD/L and 11.1 to 13.2 mg P/L. COD showed high fluctuation
C:N ratio (gCODs/gN) 1.02 1.04 1.02 during days 20 to 70 of operation. The SALR which depends on the
C:N:P ratio (gCOD:gN:gP) 4.4:5.3 4.4:5.3 4.4:5.3
bio-carries surface area is considered high for R1, medium for R2 and
4 A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425

Table 2
Operational parameters of the MBBRs.

Reactor R1 R2 R3

Parameter T = 35 °C T = 20 °C T = 10 °C T = 4 °C T = 35 T = 20 °C T = 10 °C T = 4 °C T = 35 °C T = 20 T = 10 T = 4 °C
°C °C °C

Time of operation (day) 40 40 40 40 40 40 40 40 40 40 40 40


DO (mg/L) 7.4 8.3 8.6 8.8 7.4 8.3 8.6 8.8 7.4 8.3 8.6 8.8
pH 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2 7.7–8.2
Average nitrogen source (S0N,avg) 67.5 ± 67.5 ± 67.7 ± 65.2 ± 67.5 ± 67.5 ± 67.7 ± 65.2 ± 67.5 ± 67.5 ± 67.7 ± 65.2 ±
(gN/m3) 4.5 4.5 5.1 4.2 4.5 4.5 5.1 4.2 4.5 4.5 5.1 4.2
Average carbon source (S0C) (g 73.2 ± 73.2 ± 79.9 ± 83.1 ± 73.2 ± 73.2 ± 79.9 ± 83.1 ± 73.2 ± 73.2 ± 79.9 ± 83.1 ±
CODdiss/m3) 5.5 5.5 8.1 8.1 5.5 5.5 8.1 8.1 5.5 5.5 8.1 8.1
Surface area loading rate (SALR) 0.612 ± 0.611 ± 0.604 ± 0.60 ± 0.60 ± 0.382 ± 0.372 ± 0.368 ± 0.368 ± 0.25 ± 0.25 ± 0.252 ±
(g/m2.d) 0.05 0.05 0.07 0.01 0.01 0.05 0.05 0.05 0.05 0.03 0.03 0.03
Nitrogen loading rate (NLR) 305 ± 10 305 ± 10 303 ± 15 300 ± 6 300 ± 6 305 ± 10 305 ± 10 299 ± 12 299 ± 12 307 ± 309 ± 303 ± 12
(kgN/m3.day) 10 10
Organic loading rate (OLR) 313 ± 8 313 ± 8 306 ± 10 302 ± 6 302 ± 6 313 ± 8 311 ± 10 303 ± 12 303 ± 12 311 ± 8 308 ± 307 ± 12
(gCOD/m3.day) 10
Total phosphorous (g/m3) 13.1 ± 13.1 ± 13.1 ± 13.1 ± 13.1 ± 13.1 ± 13.1 ± 13.1 ± 13.1 ± 13.1 ± 13.1 ± 13.1 ±
2.6 2.6 2.6 2.6 2.6 2.6 2.6 2.6 2.6 2.6 2.6 2.6
Q (L/Day) 6.40 6.40 6.40 6.40 6.40 6.40 6.40 6.40 6.40 6.40 6.40 6.40
HRT (hr) 5.25 5.25 5.25 5.25 5.25 5.25 5.25 5.25 5.25 5.25 5.25 5.25
C:N ratio (gCODs/gN) 1.2:1 1.2:1 1.3:1 1.4:1 1.2:1 1.2:1 1.3:1 1.4:1 1.2:1 1.2:1 1.3:1 1.4:1
C:N:P ratio (gCOD:gN:gP) 6.1:5.6:1 6.1:5.6:1 5.0:4.2:1 6.4:5.1:1 6.1:5.6:1 6.1:5.6:1 5.0:4.2:1 6.4:5.1:1 6.1:5.6:1 6.1:5.6:1 5.0:4.2:1 6.4:5.1:1

low for R3. Table 2 shows that SALR for R1 ranged from 0.600 to 0.612 that the pH of the wastewater determine the nitrification efficiency and
gN/m2.day, 0.368 to 0.382 gN/m2.day for R2 and 0.250 to 0.252 gN/m2. suggested high ammonia removals for the wastewater with carbon to
day for R3. The OLR and NLR for all reactors are in the same order of nitrogen ratio (C/N) ratio around 1. Almomani and Khraisheh (2016)
magnitude ranging from 302 to 313 gCOD/m3.day and 229 to 309 showed that a C/N ratio in the range of 1.12 to 2.24 kgCODdiss/m3.day re-
kgN/m3.day, respectively. The C:N ratio ranged from 1.2:1 to 1.4:1 sulted in significant ammonia removals. Higher C/N ratio might en-
while the C:N:P ranged from 5.0:4.2:1 to 6.4:5.0:1. hance the growth of heterotrophic bacteria and hinder the rate of
nitrification. On other hand, wastewater with low C/N ratio will have
3.2. Effect of bio-carrier on ammonia removal limited heterotrophic population and high nitrification. Lazarova et al.
(1999) also suggested a C/N ratio around one for the highest ammonia
Fig. 3 shows the surface area ammonia removal rate (SARR, g removal rate.
− N/m2. d) as a function of time for the three bio-carriers at different The average SARR for K3, K5, and M at 35 °C were 1.420 ± 0.06, 1.123
operational temperature. The SARR values were calculated at steady ± 0.06 and 0.456 ± 0.05 g − N/m2. d. and at 20 °C were 1.103 ± 0.06,
state conditions using Eq. (1): 0.920 ± 0.06 and 0.295 ± 0.05 g − N/m2. d, respectively. Previous
works showed that MBBRs operated at temperatures ≥20 °C with DO ~
 
ðm3   9 mg/L and pH ~ 8.0 resulted in high ammonia removal (Wijffels et al.,
Q CNHþ ;in −CNHþ ;out  g  1995b; Figueroa and Silverstein, 1992; Knowles et al., 1965). In the
Day 4 4

SARR ¼   m3 ð1Þ present study, oxygen was supplied to the MBBR via ceramic diffuser
2 and the DO was kept around maximum solubility limit in the range
Vr ðm Þ  As
3  %F
m3 7.5 to 8.8 mg/L and the pH was in the range 7.7 to 8.2 (Table 2), suggest-
ing that oxygen is not the limiting factor since it was at the maximum
where CNH+4 , in is ammonia concentration in the influent to MBBR, CNH+4 , solubility limit the conditions that enhance ammonia oxidation. The
3
outis effluent ammonia concentration from the MBBR (g-N/m ), Q is the MBBRs operated with bio-carrier K3 and K5 at T = 35 °C resulted in a sig-
flow rate of feed (m3/day);Vr is the total volume of the reactor (m3), nificant ammonia reduction from ~62.0 mgN/L to 8.2 and 13.4 mg-N/L,
“As” bio-carrier specific surface area (m2/m3) and %F is the percent fill respectively. However, the MBBR operated with bio-carrier M at the
of the reactor with bio-carrier (30% in this study). The three bio- same temperature showed an effluent ammonia concentration of
carriers have demonstrated stable SARRs within the same temperature. 36 mg-N/L suggesting low ammonia removal. The obtained results con-
The SARRs stabilized during the first 4 days of temperature change firm that the type of bio-carrier has a direct effect on ammonia removals
confirming the ability of the MBBR to respond to irregular changes in from wastewater. It was expected that increasing the bio-carrier specific
operational conditions. Under all studied conditions, the SARR values surface area (As) will improve the biofilm availability and thus increase
are different for each bio-carriers with the general trends showing ammonia removal. However, contradictory results were observed ex-
that the SARR for K3 N K5 N M. Previous studies showed a longer stabili- perimentally. The MBBRs operated with bio-carrier K3 with As of 500
zation times for nitrifiers especially at low temperatures (Andersson m2/m3 resulted in higher SARR compared with bio-carrier K5 with As
et al., 2001; Wijffels et al., 1995a). Delatolla et al. (2009) and Hoang of 800 m2/m3 and M with As of 1200 m2/m3. The obtained trends
et al. (2014) studied ammonia removals from synthetic wastewater at were related to low mass transfer of oxygen from water to biofilm (i.e
4 and 1 °C and showed that the system required one to two months to diffusion of oxygen) with respect to biofilm density. Bio-carrier with
reach stable removal rate. Almomani et al. (2014) showed that ammo- high surface area showed a high biofilm density that required more ox-
nia removal rates stabilized in the first 10 days after the temperature ygen for ammonia oxidation. Although DO was kept at maximum in the
change. The obtained results suggest differences in the mechanism of reactor, the mass transfer of O2 from gas to liquid for high-density bio-
ammonia removal between synthetic and real wastewater with higher film was low due to a partial or total clogging of bio-carrier cells and
removal rate and short adaptation time observed in the case of real thus ammonia oxidation decrease. Low surface area bio-carrier showed
wastewater. The significant ammonia removal obtained for real waste- low-density biofilm, high mass transfer of O2 from gas to liquid and sig-
water can be related to the effect of micro-nutrient, organic matter and nificant ammonia removals. Further work is required to confirm these
buffering capacity of this wastewater. Whang et al. (2009) have shown observations. Forrest et al. (2016) reported the same clogging problem
A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425 5

− −
Fig. 2. TN, NH+ 4 − N, NO2 , NO3 , CODs and Ptot of the influent to MMBRs (R1, R2 and R3)
during 160 days of experiment phases. Conditions: phase I, T = 35 °C, phase II, T = 20 Fig. 3. Surface area ammonia removal rate (SARR, g − N/m2. d) and Arrhenius-type Theta
°C, phase III, T = 10 °C and phase IV, T = 4 °C. model regression as function of time for the three bio-carriers at different operational
temperature.

in the study carried out with the same bio-carrier using synthetic waste-
water. It was expected the NLR and OLR are the main reason for the conserved and the removal of ammonia was achieved by biological con-
clogging problem. Thus, another set of experiments was conducted version. The maximum losses in total nitrogen did not exceed 4–8% for
with NLR of 163(kg-N/m3.day) and OLR of 150 gCOD/m3.day. Reducing all MBBRs.
the NLR and OLR did not resolve the clogging problem but it was de- The effluent concentrations of nitrite ranged between 0.25 and
layed; while the clogging was observed after 45 days for the bio- 0.35 mg-N/L, 0.22–0.86 mg-N/L and 0.25 to 0.90 mg-N/L for R1, R2,
carrier operated with high OLR and NLR, it was delayed till day 90 of op- and R3, respectively, indicating that MBBRs are capable of nitrifying am-
eration for wastewater with low loading. Still, bio-carrier with low As monia in ESWW and maintained full nitrification to NO− 3 -N at each
showed higher ammonia removal rate. temperature.
Complimentary to Fig. 3, Table 3 includes the SARR values for each The percent ammonia removal from the three reactors was also cal-
reactor at each operational temperature. Table 3 also shows surface culated and showed trends similar to the SARR. The reactor operated
area carbon removal rate (SACR) (kgCOD/m2.d), the average concentra- with bio-carrier K3 has higher % NH+ 4 − N removal compared with reac-
− −
tions of NH+4 -N, NO2 -N, CODdiss, Ptot, and NO3 -N in the effluent under tors operated with K5 and M. The percentage ammonia removals at 20
steady state conditions. The average SACR was also observed to be a °C were 87.3%, 71.8% and 47.2% for the reactors operated with K3, K5
function of bio-carrier surface area and operating temperature with and M, respectively. The values decreased to 59.3%, 44.6% and 12.5%
the highest values observed in R1 (K3) and at 35 °C. The nitrogen when the temperature reduced to 4 °C. The obtained percentage ammo-
mass balance (ammonia, nitrite, and nitrate) calculated during experi- nia removals are considered lower than values reported by Forrest et al.
mental phases confirmed that nitrogen content in the reactors was (2016) using same bio-carriers. The differences in % ammonia removal
6 A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425

Table 3
Summary of ammonia, organic and phosphorous removal rates for the MBBRs during experimental phases I-IV.

Reactor R1 R2 R3

T = 35 T = 20 T = 10 T = 4 °C T = 35 T = 20 T = 10 T = 4C T = 35 T = 20 T = 10 T = 4 °C
°C °C °C °C °C °C °C °C °C

Surface area nitrogen removal rate 1.420 ± 1.103 ± 0.95 ± 0.6424 ± 1.123 ± 0.920 ± 0.653 ± 0.423 ± 0.456 ± 0.295 ± 0.142 ± 0.05 ±
(SARR) (gN/m2.d) 0.06 0.06 0.04 0.06 0.06 0.06 0.08 0.08 0.05 0.05 0.02 0.01
RNH4 (gN/m3.d) 0.509 ± 0.494 ± 0.468 ± 0.34 ± 0.39 ± 0.408 ± 0.385 ± 0.245 ± 0.13 ± 0.10 ± 0.04 ±
0.07 0.07 0.3 0.06 0.06 0.2 0.2 0.6 0.04 0.04 0.01
Surface area carbon removal rate 0.275 ± 0.253 ± 0.1660 0.24 ± 0.130 ± 0.115 ± 0.092 ± 0.062 ± 0.074 0.062 ± 0.082 ± 0.04 ±
(SACR) (kgCOD/m2.d) 0.1 0.1 ± 0.2 0.10 0.5 0.5 0.1 0.1 0.04 0.04 0.04
Effluent NH4 (mgN/L) 8.2 ± 2.1 8.2 ± 2.1 9.4 ± 2.1 25.5.7 ± 13.4 ± 17.7 ± 19.9 ± 33.7 ± 36.0 ± 34 ± 2.1 39.3 ± 51.9 ±
5.1 1.5 2.1 4.1 5.1 2.5 3.1 4.3
Effluent NO2 (mgN/L) 0.25 ± 0.25 ± 0.35 ± 0.46 ± 0.22 ± 0.25 ± 0.55 ± 0.86 ± 0.25 ± 0.56 ± 0.87 ± 0.90 ±
0.05 0.05 0.05 0.05 0.05 0.05 0.05 0.15 0.05 0.15 0.15 0.15
Effluent NO3 (mgN/L) 60 ± 6.0 60 ± 6.0 56.0 ± 6 42 ± 5.0 43.0 ± 48.0 ± 46.0 ± 36.0 ± 27.0 ± 32.0 ± 28.0 ± 18.0 ±
3.3 6.3 10.0 8.0 8.0 8.0 8.0 6.0
Effluent COD (mg/L) 41 ± 4 41 ± 3 48 ± 4 42 ± 3 42 ± 5 46 ± 4 52 ± 4 55 ± 5 55 ± 5 29 ± 6 31 ± 3 25 ± 4

are likely related to the effect of the feed quality parameters on re- 3.3. Effect of operational temperature on ammonia removal
movals; while the present study deal with a real ESWW with a lot of
constituents that might interfere with nitrogen removal bacteria, previ- The effect of changes in reactor temperature on ammonia removal
ous studies worked with synthetic wastewater which consider the ideal was highly observed in the three MBBRs. There was a significant de-
case for the ammonia removing bacteria. Luostarinen et al. (2006) re- crease in SARR values when the temperature decrease from 35 to 4 °C
ported a similar percentage of ammonia removal (~60%) from real for all MBBRs. The SARR decreased by 54.7% (from 1.42 ± 0.06 to
wastewaters with nitrogen loading rate in the range of 0.012 to 0.642 ± 0.06 gN/m2.d), 62.3% (1.123 ± 0.06 to 0.423 ± 0.08 gN/m2.d
0.016 kg.N/m3.d using sequencing batch mode. Whang et al. (2009) and 89.0% (0.456 ± 0.05 to 0.053 ± 0.01) for the reactors R1, R2 and
studied the nitrification of ammonia in a high strength petrochemical R3, respectively as the temperature reduced from 20 to 4C°. The de-
wastewater (COD = 310 mg/L and ammonia 325 mg-N/L) with C/N crease in ammonia removal at lower temperatures can be related to
ratio around 1. The average ammonia removal in the pilot-scale MBBR the decrease in the relative abundance of ammonia-oxidizing bacterium
was in the range 24–51%. The study suggests that reagent used in con- (Nitrosospira) as observed by Hoang et al. (2014). The obtained SARR as
trolling the pH determine the nitrification efficiency. Switching the PH a function of temperature agree with values in literature; Almomani and
control reagent from NaOH to Na2CO3 resulted in an increase in ammo- Khraisheh (2016) reported a decrease of 40% and 24% in SARR for
nia removal from 24 to 51%. MBBRs used to treat septic tank effluents at high and low HRTs (5.7
Ammonia oxidation by-products (nitrate and nitrite) were followed and 13.3 h). The study reports average ammonia removals of
on a daily base (see Table 3). The concentrations of nitrite were low in 0.540 kg-N/m3 and 0.279 kg-N/m3 using high and low HRTs (5.7 and
R1& R2 ranging from 0.26 to 0.46 mg-N/L. However, in R2 the NO− 2 -N 13.3 h) at 25 °C. Ammonia removal rates at low temperatures 4–7 °C
concentration under all operational conditions was high ranging from were in the range 0.15–0.28 kgN/m3.day as reported by Delatolla et al.
0.56 to 0.9. The high concentration of nitrite in the reactor may indicate (2010). Forrest et al. (2016) studied ammonia removal from synthetic
slow oxidation of ammonia and can inhibit nitrifying bacteria (Abzazou wastewater at 25 °C, the study shows SARR values for MBBRs operated
et al., 2016; Anthonisen et al., 1976). The concentrations of NO− 2 -N in with bio carrier K3 and M, under high loading conditions, in the range

the reactors is related to the oxidation of NH+ 4 − N to NO2 -N, and 1.85 to 1.86 kg-N/m2 and 1.56 to 1.71 kg-N/m2, respectively. Normal
then to NO− −
3 -N. The observed low concentration of NO2 -N in R1 & R2 loading conditions showed lower SARR in the range of 0.02 to 0.92 kg-
suggests a transfer of all oxidized ammonia to nitrate–nitrogen. On an- N/m2 and 1.56 to 1.71 kg-N/m2, respectively. Houda et al. (2015) used
other hand, R3 with high nitrite concentration implies low oxidation fluidize nitrifying process to remove ammonia from the effluent of the
and low ammonia removal. The low nitrite concentrations in R1 reac- urban wastewater treatment process. The reported ammonia removal,
tors were achieved by maintaining the DO in the reactor N7.5 mg O2/L during 170 days of operation, was in the range 90 to 94.4%. The study
to guarantee complete oxidation of ammonia as recommended by pre- confirmed the nitrification process depends on NH4 + -N loading and
vious works (Persson et al., 2002; Hao et al., 2002). The high concentra- temperature with higher ammonia removal observed at low ammonia
tion of NO− 2 -N inside the R2 & R3 can be related to low O2 mass transfer concentration (20.52 ± 4.89 mg/L) and high temperature (35.45 ±
rate because of bio-carrier clogging as discussed before. 3.18 °C). The study showed a drop in the removal rate from approxi-
NO− 3 -N concentration, on other hands, increased during the course mately 0.60 kg-N/m3.d at 14 °C to approximately 0.16 kg-N/m3.d after
of experiments as results of ammonia oxidation. The concentration of 40 days of operation at 3 °C. Shore et al. (2012) studied ammonia re-
NO− 3 -N ranged from 42 ± 5.0 to 60 ± 6.0 mgN/L for R1, 48.0 ± 6.3 to moval from high-temperature industrial wastewater using bench scale
36.0 ± 8.0 for R2 and 18.0 ± 6.0 to 32.0 ± 8 mgN/L for R3. The concen- MBRR operated at 35 and 45 °C. MBBRs were able to remove up to
tration of NO− 3 -N in the reactor were related to ammonia removals with 90% of influent ammonia from industrial wastewater at temperature
general trends indicating high NO− 3 -N for the reactor with high ammo- range 35 and 40 °C. However, no bio-treatment was observed above
nia removal and vice versa. It was also observed that decreasing the re- 45 °C. Other studies confirmed the MBBR inability to effectively nitrify
actor temperature showed lower concentrations of NO− 3 -N in effluents at 45 °C (Grunditz and Dalhammar, 2001; Jones and Hood, 1980).
reflecting low ammonia removal kinetics. The kinetics of ammonia oxidation was fitted to Arrhenius-type
The Surface Area Carbon Removal Rate (SACR) at 35 °C were 0.275, temperature Theta model (ATH-Model) using Eq. (2) in combination
0.130 and 0.062 kgCOD/m2.d in R1, R2 and R3. It is seen that R1 operated with Eq. (3) as published by Almomani et al. (2014)
with K3 with As of 500m2/m3 has the highest SACR compared with R2
and R3 operated with bio-carriers K5 and M, respectively. The SACR de- RNH4;T2 ¼ RNH4;1 :θðT 2 −T 1 Þ ð2Þ
creased by 3%, 11.5%. 16.2% when operating temperature decreased
from 35 to 20 °C. Reducing the temperature from 35 to 4 °C decreased
the SACR by 12.7%, 52.3% and 45.9% for R1, R2, and R3, respectively. θ ¼ 3:83x10−2 ln ðt Þ þ 9:83x10−1 ð3Þ
A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425 7

where RNH4,T1 and RNH4,T2 are the removal rates at T1 and T2 (°C),
respectively.
Ammonia removal rates for R1, R2 and R3 fitted to the ATH-Model
are presented in Fig. 3. ATH-Model fits very well the MBBR kinetics
with a coefficient of determinations (R2) N0.90 for all MBBRs. The kinet-
ics of the MBBR operated with bio-carrier K3 fits the ATH-Model with
two θ values (1.13 & 1.05), the first values applied for the temperature
range 35 to 10 °C, while the second one applies for the temperature
range 10 to 4 °C. The kinetics of ammonia removal rates for the MBBR
operated with bio-carrier K5 was fitted to ATH with three θ values
1.08, 1.03 and 1.01 for the temperature ranges 35 °C to 20 °C, 20 °C to
10 °C, and 10 °C to 4 °C respectively. The MBBR operated with bio-
carrier M fits the ATH with two values of θ of 1.02 and 0.99 for the tem-
perature ranges 35 °C to 20 °C and 20 °C to 4 °C, respectively The re-
ported θ values agree with values reported by Almomani and
Khraisheh (2016), Almomani et al. (2014), Dellatola et al. (2012), and
Salvetti et al. (2006). The variation in the value of θ for the same
MBBR under different temperatures provides information regarding
the reactor sensitivity to change in operational temperature. It is
known that MBBR with one θ value suggests that the reactor is stable
under all operating temperatures. However, change in θ values as the
reactor temperature change implies a change in ammonia removal ki-
netic as a function of temperature. The results of this study showed
two different values of θ for MBBRs operated with bio-carriers K3 and
M. The range of temperature where each kinetics dominant can be
used to determine the best operating condition for each bio-carrier.
While R1 is stable in the range 35 °C to 10 °C, reactor R3 is stable for
the temperature range 20 °C to 4 °C suggesting that the first reactor
can be operated at high temperature and the second reactor at low tem-
perature. It was also concluded that R2 is very sensitive to the tempera-
ture and it is not recommended to be used if a frequent change in
operational temperature is expected.
Fig. 4. (a) TAS and VAS of biofilm and (b) the BVRR on day 30, 60 and 150 of operation for
the MBBRs (R1, R2 and R3).
3.4. Characteristics of attached biofilm

Fig. 4a illustrates TS and VAS of the biofilm attached to the bio- biofilm thicknesses were found to be in the range 69.08 to 456.11 μm
carriers on day 0, 5, 20, 40, 60, 80, 100, 140 and 160 of operation. The in R1, 108.44–348.69 μm in R2 and 155.09–366.19 μm in R3. Measure-
time of analysis was selected to cover different operating temperatures ments after 160 days showed a small decrease in the biofilm thickness
(35 °C, 20 °C, 10 °C and 4 °C). There was an increased in TAS and VAS of by 4.5%, 6.6% and 8.9% for R1, R2, and R3, respectively. The very small de-
the biofilm for all the MBBRs after 5 days following the change from crease in biofilm thickness during the 160 days of operation indicates
SWW to ESWW. After that, the values slightly decrease to stabilize stable biofilm volume and density. Hao et al. (2002) showed that at
around an average value. The small increase and fast re-stabilization the optimal DO, the biofilm thickness ranged from 0.2 to 0.7 μm, and
in biofilm (TAS and VAS) after the change in feed confirms the ability this thickness was enough to contain all required organisms for high
of MBBRs to respond to system upsets and/or irregular changes in ammonia removal. Clogging problem in R2 and R3 can be seen in the
feed conditions without a major change in biofilm characteristics. In ad- VPSEM images R2A and R3A.
dition, it was confirmed that the biofilm was stable over the 160 days of Ammonia removal with respect to biofilm volume was followed by
operation even with changes in temperature. The percentage loss from calculating the biofilm volume ammonia removal rate (BVAR) for the
the biofilm was very low and did not exceed 4–7% under all conditions. MBBRs (R1, R2, and R3) as in Eq. (4)
The high values of TAS and VAS during the first 5 days of operation is
mainly due to change in the feed characteristics, not the temperature. SARR
BVAR ¼ ð4Þ
The average values of TAS (standard deviation) as well as the VAS/TAS T bio
ratio were 70.2 mg/carrier (0.007) and 0.73(0.02) in R1, 77.4(0.03)
mg/carrier and 0.75.6(0.002) in R2 and 60.2 mg/carrier (0.005) and where Tbio is biofilm thickness. Fig. 4b displays the variation in BVAR for
0.74(0.02) in R3. The Student's paired t-test at a significance level of R1, R2 and R3 on day 30 (T = 35 °C), 60 (T = 20 °C) and 150 (T = 4 °C).
5% showed that biofilm thickness measured from different bio-carrier BVAR at day 30 of operation is significantly higher than at days 60 and
is not significantly different from the reported average value. These 150. At high temperature, the SARR was high and the biofilm thickness
values are slightly different at different operational temperatures. was low which resulted in an increase in the BVAR values. When the op-
Based on the experimental results, the reported TAS were enough to erational temperature decrease, both SARR and biofilm thickness de-
achieve significant organic matter and ammonia removals. Hoang creased resulting in lower BVAR. The obtained results give and
et al. (2014) reported significant ammonia removal rate under the information on how the SARR and biofilm changes to achieve the ex-
same biomass concentration in the treatment of SWW at 1 °C and 20 °C. pected ammonia removal. MBBRs operated at low temperature showed
The biofilms thickness was measured using VPSEM (Fig. 5). The thicker bio-film but lower the SARR and BVRR.
VPSEM images showed that biofilm on the bio-carrier is not uniform Table 4 presents the Live/Dead (L/D) ratios at different depths of bio-
and covers only the inner side of the bio-barrier. Hao et al. (2002) film and EPS analysis during the operation of the MBBRs at different
showed that the attached growth nitrifying bacteria have non-uniform temperatures. L/D analysis revealed that MBBR R1 operated with bio-
properties during the treatment of wastewater treatment. The average carriers k3 have a high L/D ratio compared with the other MBBRs.
8 A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425

R1-A R1-B

R2-A R2-B

R3-A R3-B

Fig. 5. VPSEM images of biofilms in R1, R2 and R3 after 80 day of operation.


A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425 9

Table 4
Summary of live/dead test and biofilm coverage by extracellular polymeric substance for R1, R2 and R3 at different operation time.

R1 R2 R3

Time (day) Scanning depth (μm) Live (L) Dead (D) L/D % EPS (St. Dev) Live (L) Dead (D) L/D % EPS (St. Dev) Live (L) Dead (D) L/D % EPS (St. Dev)

20 0.07 0.06 1.17 0.07 69 (3.2) 0.19 0.18 1.06 71 (4.4) 0.18 0.16 1.13 73 (4.8)
0.27 0.22 1.23 0.27 0.17 0.14 1.21 0.16 0.13 1.23
1.15 0.26 4.42 1.15 5.04 2.09 2.41 4.79 1.59 3.01
0.87 0.15 5.80 0.87 0.14 0.08 1.75 0.13 0.07 1.86
0.62 0.16 3.88 0.62 0.3 0.15 2.00 0.28 0.13 2.15
40 0.07 0.063 1.11 0.07 70 (7.8) 0.19 0.18 1.06 73 (3.0) 0.18 0.16 1.13 74.9 (5.6)
0.27 0.22 1.23 0.27 0.17 0.14 1.21 0.16 0.13 1.23
1.16 0.26 4.46 1.16 5.09 1.3 3.92 4.83 1.67 2.89
0.88 0.1475 5.97 0.88 0.14 0.08 1.75 0.13 0.07 1.86
0.63 0.16 3.94 0.63 0.3 0.15 2.00 0.29 0.13 2.23
80 0.07 0.07 1.00 0.07 69 (4.1) 0.2 0.19 1.05 73 (3.8) 0.188 0.17 1.11 75 (4.3)
0.28 0.23 1.22 0.28 0.18 0.15 1.20 0.171 0.134 1.28
1.21 0.27 4.48 1.21 5.3 1.33 3.98 5.038 1.69 2.98
0.92 0.155 5.94 0.92 0.15 0.08 1.88 0.14 0.071 1.97
0.65 0.17 3.82 0.65 0.32 0.16 2.00 0.298 0.137 2.18
160 0.56 0.14 4.00 0.56 73 (3.9) 0.66 0.18 3.67 77 (4.6) 0.621 0.22 2.82 78 (4.3)
0.41 0.07 5.86 0.41 0.55 0.14 3.93 0.523 0.44 1.19
0.26 0.06 4.33 0.26 0.66 0.17 3.88 0.627 0.23 2.73
0.49 0.17 2.88 0.49 0.4 0.19 2.11 0.38 0.23 1.65
0.55 0.1 5.50 0.55 0.53 0.135 3.93 0.493 0.29 1.70

Under all scanning depths, the L/D ratio for MBBR R1 is N1.2, and the biofilm reactors (MBBRs): contribution of the biofilm and suspended biomass. Sci.
Total Environ. 643, 1464–1480.
ratio ranged from 1.2 to 6. The MBBRs R2 and R3 operated with bio- Abzazou, T., Araujo, R.M., Auset, M., Salvadó, H., 2016. Tracking and quantification of
carrier K5 and M showed L/D ratio in the range of 1.05 to 3.98 and nitrifying bacteria in biofilm and mixed liquor of a partial nitrification MBBR
1.11 to 3.01, respectively. The high L/D ratio in MBBRs R1 confirms stable pilot plant using fluorescence in situ hybridization. Sci. Total Environ. 541,
1115–1123.
and significant biological activities and healthy microorganisms. The Al Momani, F., 2006. Impact of photo-oxidation technology on the aqueous solutions of
MBBRs R2 and R3 showed very low L/D ratio suggesting low biological nitrobenzene: degradation efficiency and biodegradability enhancement.
activities in the biofilm, which was observed before to have low ammo- J. Photochem. Photobiol. A Chem. 179 (1–2), 184–192.
Al Momani, F., Gonzalez, O., Sans, C., Esplugas, S., 2004. Combining photo-Fenton process
nia removal rates. The EPS analysis for the three reactors revealed that
with biological sequencing batch reactor for 2,4-dichlorophenol degradation. Water
the biofilm exists only on the internal surface of bio-carrier and covers Sci. Technol. 49 (4), 293–298.
EPS at least 69% of it (Table 2). The effect of temperature on cells L/D Alizadeh, S., Ghoshal, S., Comeau, Y., 2019. Fate and inhibitory effect of silver nano-
particles in high rate moving bed biofilm reactors. Sci. Total Environ. 647,
ratio is clear in Table 2. The L/D ratios were observed to be lower at
1199–1210.
low temperatures. Allagui, A., Sarfraz, S., Middleton, B., Almomani, F., Baranova, E.A., 2012. Ammonia
electrooxidation on NiPd nanoparticles in alkaline media: effect of pH and concentra-
tion. ECS Trans. 1897–1906.
4. Conclusion Allagui, A., Sarfraz, S., Ntais, S., Al momani, F., Baranova, E.A., 2014. Electrochemical behav-
ior of ammonia on Ni98Pd2 nano-structured catalyst. Int. J. Hydrog. Energy 39 (1),
The present study focused on the feasibility of using moving bed bio- 41–48.
Almomani, F., Khraisheh, M., 2016. Treatment of septic tank effluent using moving-bed bi-
film reactor (MBBR) system for the treatment of ammonia-containing ological reactor: kinetic and biofilm morphology. Int. J. Environ. Sci. Technol. 13 (8),
effluents under real operation conditions and steady state. The novelty 1917–1932.
of the work lies in the fact that real wastewater (effluent of secondary AlMomani, F.A., Örmeci, B., 2016. Performance of Chlorella vulgaris, Neochloris
oleoabundans, and mixed indigenous microalgae for treatment of primary effluent,
treatment tanks) was used instead of synthesized wastewater input ef-
secondary effluent and centrate. Ecol. Eng. 95, 280–289.
fluent. The results showed that the surface area loading rate and bio- Almomani, F.A., Delatolla, R., Örmeci, B., 2014. Field study of moving bed biofilm reactor
carrier specific surface should be considered when designing of MBBR. technology for post-treatment of wastewater lagoon effluent at 1°C. Environ.
Technol. 35 (13), 1596–1604.
The research has shown that bio-carriers with the larger surface area
Andersson, A., Laurent, P., Kihn, A., Prévost, M., Servais, P., 2001. Impact of temperature on
have a greater propensity to become clogged under low and high load- nitrification in biological activated carbon (BAC) filters used for drinking water treat-
ing rate and this should be considered during the design of nitrifying ment. Water Res. 35 (12), 2923–2934.
MBBR systems. There was significant different in ammonia removal ef- Andreottola, G., Foladori, P., Ragazzi, M., Villa, R., 2002. Dairy wastewater treatment in a
moving bed biofilm reactor. Water Sci. Technol. 45 (12), 321–328.
ficiencies between bio-carriers with different surface areas as a result of Anthonisen, A., Loehr, R., Prakasam, T., Srinath, E., 1976. Inhibition of nitrification by am-
oxygen mass transfer limitation for high specific surface area bio- monia and nitrous acid. J. Water Pollut. Control Fed. 835–852.
carrier. Bio-carrier type has a direct effect on the kinetics of ammonia re- APHA, A., WPCF, 1985. Standard Methods for the Examination of Water and Wastewater.
16 ed. American Public Health Association.
moval and affects the MBBR sensitivity for changes in operational tem- Bering, S., Mazur, J., Tarnowski, K., Janus, M., Mozia, S., Morawski, A.W., 2018. The appli-
peratures. The MBBR was responsive to shock temperature in a short cation of moving bed bio-reactor (MBBR) in commercial laundry wastewater treat-
time and maintained a stable and active embedded nitrifying biomass ment. Sci. Total Environ. 627, 1638–1643.
Delatolla, R., Tufenkji, N., Comeau, Y., Gadbois, A., Lamarre, D., Berk, D., 2009. Kinetic anal-
with minimum biofilm loss or washout. ysis of attached growth nitrification in cold climates. Water Sci. Technol. 60 (5).
Delatolla, R., Tufenkji, N., Comeau, Y., Gadbois, A., Lamarre, D., Berk, D., 2010. Investigation
Acknowledgment of laboratory-scale and pilot-scale attached growth ammonia removal kinetics at cold
temperature and low influent carbon. Water Quality Research Journal 45 (4),
427–436.
The authors gratefully acknowledge the financial support provided Delatolla, R., Tufenkji, N., Comeau, Y., Gadbois, A., Lamarre, D., Berk, D., 2012. Effects of
by the Qatar University Internal Grant (QUUG-CENG-CHE-14\15-11). long exposure to low temperatures on nitrifying biofilm and biomass in wastewater
treatment. Water Environment Research 84, 328–338.
Eaton, A.D., Clesceri, L.S., Greenberg, A.E., Franson, M.A.H., 2005. Standard methods for the
References examination of water and wastewater. American public health association 1015,
49–51.
Abtahi, S.M., Petermann, M., Juppeau Flambard, A., Beaufort, S., Terrisse, F., Trotouin, T., Figueroa, L.A., Silverstein, J., 1992. The effect of particulate organic matter on biofilm nitri-
Joannis Cassan, C., Albasi, C., 2018. Micropollutants removal in tertiary moving bed fication. Water Environ. Res. 64 (5), 728–733.
10 A. Ashkanani et al. / Science of the Total Environment 693 (2019) 133425

Forrest, D., Delatolla, R., Kennedy, K., 2016. Carrier effects on tertiary nitrifying moving Rodgers, M., Zhan, X.-M., 2004b. Biological nitrogen removal using a vertically moving
bed biofilm reactor: an examination of performance, biofilm and biologically pro- biofilm system. Bioresour. Technol. 93 (3), 313–319.
duced solids. Environ. Technol. 37 (6), 662–671. Rusten, B., Siljudalen, J.G., Nordeidet, B., 1994. Upgrading to nitrogen removal with the
Grunditz, C., Dalhammar, G., 2001. Development of nitrification inhibition assays using KMT moving bed biofilm process. Water Sci. Technol. 29 (12), 185–195.
pure cultures of Nitrosomonas and Nitrobacter. Water Res. 35 (2), 433–440. Rusten, B., Hem, L.J., Ødegaard, H., 1995. Nitrification of municipal wastewater in moving-
Hao, X., Heijnen, J.J., MCM, V.L., 2002. Model-based evaluation of temperature and inflow bed biofilm reactors. Water Environ. Res. 67 (1), 75–86.
variations on a partial nitrification—ANAMMOX biofilm process. Water Res. 36 (19), Salvetti, R., Azzellino, A., Canziani, R., Bonomo, L., 2006. Effects of temperature on tertiary
4839–4849. nitrification in moving-bed biofilm reactors. Water Res. 40 (15), 2981–2993.
Hoang, V., Delatolla, R., Abujamel, T., Mottawea, W., Gadbois, A., Laflamme, E., Stintzi, A., Sharma, B., Ahlert, R., 1977. Nitrification and nitrogen removal. Water Res. 11 (10),
2014. Nitrifying moving bed biofilm reactor (MBBR) biofilm and biomass response 897–925.
to long term exposure to 1 C. Water Res. 49, 215–224. Shi, Q., Wang, W., Chen, M., Zhang, H., Xu, S., 2019. Ammonia induces Treg/Th1 imbalance
Houda, N., Abdelwaheb, C., Asma, B.R., Ines, M., Ahmed, L., Abdennaceur, H., 2015. Tertiary with triggered NF-κB pathway leading to chicken respiratory inflammation response.
nitrification using moving-bed biofilm reactor: a case study in Tunisia. Curr. Sci. Total Environ. 659, 354–362.
Microbiol. 70 (4), 602–609. Shore, J.L., M'Coy, W.S., Gunsch, C.K., Deshusses, M.A., 2012. Application of a moving bed
Houtman, C.J., 2010. Emerging contaminants in surface waters and their relevance for the biofilm reactor for tertiary ammonia treatment in high temperature industrial waste-
production of drinking water in Europe. J. Integr. Environ. Sci. 7 (4), 271–295. water. Bioresour. Technol. 112, 51–60.
Jones, R.D., Hood, M.A., 1980. Effects of temperature, pH, salinity, and inorganic nitrogen Suchetana, B., Rajagopalan, B., Silverstein, J., 2017. Assessment of wastewater treatment
on the rate of ammonium oxidation by nitrifiers isolated from wetland environ- facility compliance with decreasing ammonia discharge limits using a regression
ments. Microb. Ecol. 6 (4), 339–347. tree model. Sci. Total Environ. 598, 249–257.
Judd, S., AlMomani, F.A.O., Znad, H., Al Ketife, A.M.D., 2017. The cost benefit of algal tech- Torresi, E., Tang, K., Deng, J., Sund, C., Smets, B.F., Christensson, M., Andersen, H.R., 2019.
nology for combined CO2 mitigation and nutrient abatement. Renew. Sust. Energ. Removal of micropollutants during biological phosphorus removal: impact of redox
Rev. 71, 379–387. conditions in MBBR. Sci. Total Environ. 663, 496–506.
Knowles, G., Downing, A., Barret, M., 1965. Determination of kinetic nitrifying biomass in Vitse, F., Cooper, M., Botte, G.G., 2005. On the use of ammonia electrolysis for hydrogen
biological filters used in drinking water production. J. Ind. Microbiol. Biotechnol. 38, production (vol 142, pg 18, 2005). J. Power Sources 152 (1), 311–312.
263–278. Welander, U., Henrysson, T., Welander, T., 1997. Nitrification of landfill leachate using
Lazarova, V., Bellahcen, D., Manem, J., Stahl, D.A., Rittmann, B.E., 1999. Influence of oper- suspended-carrier biofilm technology. Water Research 31 (9), 2351–2355.
ating conditions on population dynamics in nitrifying biofilms. Water Sci. Technol. 39 Welander, U., Mattiasson, B., 2003. Denitrification at low temperatures using a suspended
(7), 5–11. carrier biofilm process. Water Res. 37 (10), 2394–2398.
Li, X., Yuan, Y., Huang, Y., Bi, Z., 2019. Simultaneous removal of ammonia and nitrate by Welander, U., Henrysson, T., Welander, T., 1997. Nitrification of landfill leachate using
coupled S0-driven autotrophic denitrification and Anammox process in fluorine- suspended-carrier biofilm technology. Water Res. 31 (9), 2351–2355.
containing semiconductor wastewater. Sci. Total Environ. 661, 235–242. Whang, L.M., Yang, K.H., Yang, Y.F., Han, Y.L., Chen, Y.J., Cheng, S.S., 2009. Microbial ecol-
Liu, X., Wang, L., Pang, L., 2018. Application of a novel strain Corynebacterium pollutisoli ogy and performance of ammonia oxidizing bacteria (AOB) in biological processes
SPH6 to improve nitrogen removal in an anaerobic/aerobic-moving bed biofilm reac- treating petrochemical wastewater with high strength of ammonia: effect of
tor (A/O-MBBR). Bioresour. Technol. 269, 113–120. Na2CO3 addition. Water Sci. Technol. 59 (2), 223–231.
Luostarinen, S., Luste, S., Valentín, L., Rintala, J., 2006. Nitrogen removal from on-site Wijffels, R., Hunik, J., Leenen, E., Günther, A., de Castro, J.O., Tramper, J., Englund, G.,
treated anaerobic effluents using intermittently aerated moving bed biofilm reactors Bakketun, Å., 1995a. Effects of diffusion limitation on immobilized nitrifying microor-
at low temperatures. Water Res. 40 (8), 1607–1615. ganisms at low temperatures. Biotechnol. Bioeng. 45 (1), 1–9.
Marsalek, J., Lawrence, J., Servos, M., Busnarda, J., Munger, K., Adare, K., Jefferson, C., Kent, Wijffels, R., De Gooijer, C., Schepers, A., Beuling, E., Mallée, L., Tramper, J., 1995b. Dynamic
R., Wong, M., 1997. The impacts of municipal wastewater effluents on Canadian wa- modeling of immobilized Nitrosomonas europaea: implementation of diffusion limi-
ters: a review (Water Quality Research Journal of Canada 1997, volume 32, no. 4, pg. tation over expanding microcolonies. Enzym. Microb. Technol. 17 (5), 462–471.
659-713) PA Chambers, M. Allard 2, SL Walker 3. Water Qual. Res. J. Can. 32 (4), Young, B., Delatolla, R., Ren, B., Kennedy, K., Laflamme, E., Stintzi, A., 2016. Pilot-scale ter-
659–713. tiary MBBR nitrification at 1°C: characterization of ammonia removal rate, solids
Nsenga Kumwimba, M., Meng, F., 2019. Roles of ammonia-oxidizing bacteria in improv- settleability and biofilm characteristics. Environ. Technol. 37, 2124–2132.
ing metabolism and cometabolism of trace organic chemicals in biological wastewa- Young, B., Delatolla, R., Kennedy, K., LaFlamme, E., Stintzi, A., 2017a. Post carbon removal
ter treatment processes: a review. Sci. Total Environ. 659, 419–441. nitrifying MBBR operation at high loading and exposure to starvation conditions.
Ødegaard, H., Rusten, B., Westrum, T., 1994. A new moving bed biofilm reactor- Bioresour. Technol. 239, 318–325.
applications and results. Water Sci. Technol. 29 (10−11), 157–165. Young, B., Delatolla, R., Kennedy, K., Laflamme, E., Stintzi, A., 2017b. Low temperature
Oleszkiewicz, J.A., Barnard, J.L., 2006. Nutrient removal technology in North America and MBBR nitrification: microbiome analysis. Water Res. 111, 224–233.
the European Union: a review. Water Quality Research Journal 41 (4), 449–462. Zhang, S., Wang, Y., He, W., Wu, M., Xing, M., Yang, J., Gao, N., Pan, M., 2014. Impacts of
Oswin, H., Salomon, M., 1963. The anodic oxidation of ammonia at platinum black elec- temperature and nitrifying community on nitrification kinetics in a moving-bed bio-
trodes in aqueous KOH electrolyte. Can. J. Chem. 41 (7), 1686–1694. film reactor treating polluted raw water. Chem. Eng. J. 236, 242–250.
Pelissari, C., Ávila, C., Trein, C.M., García, J., de Armas, R.D., Sezerino, P.H., 2017. Nitrogen Zhu, Z., Chen, D., 2002. Nitrogen fertilizer use in China–contributions to food production,
transforming bacteria within a full-scale partially saturated vertical subsurface flow impacts on the environment and best management strategies. Nutr. Cycl.
constructed wetland treating urban wastewater. Sci. Total Environ. 574, 390–399. Agroecosyst. 63 (2–3), 117–127.
Persson, F., Wik, T., Sörensson, F., Hermansson, M., 2002. Distribution and activity of am- Znad, H., Al Ketife, A.M., Judd, S., AlMomani, F., Vuthaluru, H.B., 2018. Bioremediation and
monia oxidizing bacteria in a large full-scale trickling filter. Water Res. 36 (6), nutrient removal from wastewater by Chlorella vulgaris. Ecol. Eng. 110, 1–7.
1439–1448.
Rodgers, M., Zhan, X.-M., 2004a. Biological nitrogen removal using a vertically moving
biofilm system. Bioresour. Technol. 93 (3), 313–319.

You might also like