Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Environmental Management 253 (2020) 109655

Contents lists available at ScienceDirect

Journal of Environmental Management


journal homepage: http://www.elsevier.com/locate/jenvman

Research article

Ozone combined with ceramic membranes for water treatment: Impact on


HO� radical formation and mitigation of bromate
Khaled Ibn Abdul Hamid a, b, Peter J. Scales c, Sebastien Allard d, Jean-Philippe Croue d,
Shobha Muthukumaran a, b, Mikel Duke a, *
a
Institute for Sustainable Industries & Liveable Cities, Victoria University, PO Box 14428, Melbourne, VIC, 8001, Australia
b
College of Engineering and Science, Victoria University, PO Box 14428, Melbourne, VIC, 8001, Australia
c
Department of Chemical Engineering, The University of Melbourne, Parkville, VIC, 3010, Australia
d
Curtin Water Quality Research Centre, Curtin University, GPO Box U1987, Perth, WA, 6845, Australia

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

Keywords: The beneficial effect of combining ozone with ceramic membrane filtration (CMF) to enhance membrane flux
Ceramic membrane performances during water treatment (e.g., wastewater and drinking water) could be related to the formation of
Ozone hydroxyl (HO�) radicals from the interaction of ozone with ceramic membrane. To explore this effect, para-
pCBA
chlorobenzoic acid was used to probe HO� radical activity during a combined ozone/CMF process using a
Hydroxyl radical
Bromate formation
0.1 μm pore size membrane supplied by Metawater, Japan. Tests were then extended to explore the impact on
bromate formation downstream CMF, a well-known undesirable by-product from ozone use in water treatment.
Ozone reduction by the membrane and its module appeared to be more associated with physical degassing, but a
noticeable formation of HO� radicals was observed during the interaction of ozone with the ceramic membrane.
CMF treatment of ozonated potable water containing bromide showed a reduced bromate formation of 50%
when the water was recirculated to the filtration module containing the ceramic membrane, compared to the
experiment performed with an empty module. Single pass experiments showed bromate mitigation of around
10%. The mitigation of bromate formation was attributed to reduced overall ozone exposure by deagassing ef­
fect, but also potentially from suppression of the oxidation of Br and HOBr/BrO to BrO3 due to the catalytic
degradation of ozone via a HO� radical pathway.

1. Introduction preceded by more aggressive oxidative environments such as ozone


(Karnik et al., 2005). Ozone oxidizes electron-rich compounds con­
Most applications of membrane filtration utilise polymeric mem­ taining carbon-carbon double bonds, activated aromatic systems and
branes. They are very reliable and cost effective. However, some ap­ non-protonated amines (von Gunten, 2003a). This in turn has a powerful
plications require a more chemically resilient, longer life and high effect to remove colour, taste and odor in drinking water production.
physical integrity membrane and ceramic membranes are useful to this The widespread application of ozone to secondary effluent (SE) is more
need (Hausmann et al., 2013). The material robustness of ceramic limited (Oneby et al., 2010).
membranes permits operation under high pressure, temperature and The chemical action of ozone on water is well known to be compli­
operating flux whilst still retaining polymeric style rejection character­ mentary to physical separation on membranes whereby the combined
istics (Lehman and Liu, 2009). However, like any membrane process, process of ozonation and ceramic membranes has been shown to offer
ceramic membrane performances are improved with a several advantages. Schlichter et al. (2004) found that a minimum
suitable pre-treatment in order to manage fouling and maintain long concentration of 0.05 mg O3/L led to stable high permeate fluxes during
term operation. For water treatment applications, ceramic membranes ceramic microfiltration (MF) and ultrafiltration (UF) of river water
may be used in conjunction with conventional pre-treatments such as without back-flushing (Schlichter et al., 2004). For wastewater recy­
coagulation, but the uniquely inert material structure allows them to be cling, ozone and coagulation were also observed to lead to higher

* Corresponding author.
E-mail addresses: Khaled.ibnabdulhamid@live.vu.edu.au (K. Ibn Abdul Hamid), peterjs@unimelb.edu.au (P.J. Scales), S.Allard@curtin.edu.au (S. Allard), Jean-
Philippe.Croue@curtin.edu.au (J.-P. Croue), Shobha.muthukumaran@vu.edu.au (S. Muthukumaran), Mikel.duke@vu.edu.au (M. Duke).

https://doi.org/10.1016/j.jenvman.2019.109655
Received 2 September 2018; Received in revised form 18 September 2019; Accepted 28 September 2019
Available online 22 October 2019
0301-4797/© 2019 Published by Elsevier Ltd.
K. Ibn Abdul Hamid et al. Journal of Environmental Management 253 (2020) 109655

sustainable ceramic MF fluxes when considered as a pre-treatment to 100 μg/L (von Gunten, 2003b). The formation of bromate is a result of
RO, where further benefits such as reduced RO biofouling were observed the oxidation of bromide through a combination of ozone and HO�
(Myat et al., 2018). Kim et al. (2008) showed that their hybrid ozona­ radical reactions (Elovitz and von Gunten, 1999).
tion–ceramic membrane system significantly reduced membrane fouling The goal of this study was to acquire a better understanding of ozone
and fouling behaviour was strongly dependent upon ozone concentra­ interaction with a ceramic membrane. The project also aimed to inves­
tion and hydrodynamic conditions (Kim et al., 2008). Park et al. (2012) tigate the mechanism of catalytic ozone decomposition into HO� radicals
investigated the characteristics of natural organic matter (NOM) treated in combined ozone and ceramic membrane treatment process. Subse­
by a hybrid ozone/ceramic membrane process. Changes in permeate flux quent experiments with bromide-containing water were also investi­
against filtration time were observed for a cross flow membrane filtra­ gated to elucidate the impact of ozone/ceramic MF (CMF) on the
tion system at a flux of 44 L/m2⋅h. A noticeable flux increase (25%) was formation of bromate.
observed after ozone pre-treatment. The reduction in fouling was
attributed to the reaction of molecular ozone and/or HO� radicals 2. Theory and reaction mechanism
generated by ozone decomposition with NOM (Park et al., 2012). Karnik
et al. (2005) investigated samples taken from Lake Lansing (Haslett, It is well-known that the material used to make ceramic membranes
Michigan). The permeate flux through a titania coated ceramic mem­ can catalyse the breakdown of ozone to HO� radicals (Gracia et al.,
brane was found to be significantly affected by both ozonation and the 2000a; Gracia et al., 2000b; Karnik et al., 2005; Legube and Karpel Vel
pH of the feed water in the system. The effect of ozone was attributed to Leitner, 1999; Munter, 2001). However, tracking HO� radical formation
the formation of HO� or other radicals at the membrane surface that during continuous CMF would provide new insights into the ozone/CMF
reduced the extent of membrane fouling. Lehman and Liu (2009) studied chemistry. HO� radical formation and kinetics can be extracted using
the effect of ozonation and ceramic membrane on the SE of a pilot plant pseudo-first order kinetics. Due to difficulties in directly measure HO�
located in Chino, California. They found that ozone treatment was radicals, the transient concentration of HO� radicals during ozonation
effective in degrading colloidal NOM which was likely responsible for can be calculated using an ozone resistant probe, para-chlorobenzoic
the majority of membrane fouling (Lehman and Liu, 2009). acid (pCBA). pCBA reacts primarily with HO� radicals (kHO�,
9 1 1
Recently, Wei et al. (2016) investigated the hybrid process of in-situ pCBA ¼ 5 � 10 M s ) (Neta and Dorfman, 1968), and insignificantly
ozonation with ceramic UF membrane for the treatment of algal-rich with O3 (kO3,pCBA < 0.15 M 1 s 1) (David Yao and Haag, 1991) with the
water (Wei et al., 2016). It was observed that in-situ ozonation can concentration of HO� radicals extracted by measuring the pCBA con­
effectively mitigate membrane fouling by forming a more porous and centration (Elovitz and von Gunten, 1999). Tert-butanol can be used to
thinner cake layer. Moreover, in-situ ozonation can alleviate membrane study the reactivity of ozone alone since it is an effective scavenger of
pore blocking by inducing accelerated organics degradation within the HO� radicals. The reaction rate constants between tert-butanol and
membrane pores. Song et al. (2017) compared the pre-ozonation and either HO� or O3 have been reported as kHO� ¼ 5.9 � 108 M 1 s 1 and
in-situ ozonation on mitigation of ceramic UF membrane fouling caused kO3 ¼ 3.0 � 10 3 M 1 s 1, respectively (de Vera et al., 2015; Hoign�
e and
by alginate (Song et al., 2017). It was concluded that, in-situ ozonation Bader, 1983; Ma and Graham, 2000; Staehelin and Hoigne, 1985; von
with ceramic membrane catalyzed ozone decomposition with increased Gunten and Hoigne, 1994).
generation of HO� radicals resulting in alleviated reversible and irre­ Elovitz and von Gunten (Elovitz and von Gunten, 1999) proposed the
versible fouling. These findings separately conclude that, the low fouling Rct concept, which was defined as the ratio of HO� radical exposure vs.
and high flux effect of ozone on ceramic membranes was either due to a ozone exposure shown in Equation (1).
catalytic effect by the membrane or altered fouling layer chemistries on
the membrane surface (Dow et al., 2013). Some are reporting that (1)
ceramic materials are responsible for formation of hydroxyl radicals via
catalytic decomposition (Zhu et al., 2011). However, it has been re­ where the time-integrated ozone concentration, ʃ[O3]dt is the ozone
ported in some other studies (Ikhlaq et al., 2015; Nawrocki and Fijołek, exposure. The ozone exposure can be calculated from the integral of the
2013) that, aluminium oxide (the major component of ceramic mem­ ‘ozone concentration vs. time’ data. By following the ozone exposure and
brane) is not a catalyst of ozone decomposition in water. Despite these the decay of pCBA, the Rct value can be obtained using Equation (2).
observations and viable explanations of the unique and positive per­
formance effects when ozone and ceramic membranes are combined, the (2)
actual mechanism of ozone on the membrane has not been confirmed in When the pCBA concentration is plotted with respect to O3 exposure
a dedicated study (Nawrocki and Kasprzyk-Hordern, 2010). Therefore, (product of ozone concentration and time) the slope is then Rct � kHO�,
investigating the catalytic effect of ozone on a well-known commercial 9 1 1
pCBA. Since kHO ,pCBA is reported as 5 � 10 M s (Elovitz and von

water treatment ceramic membrane from Metawater would be highly Gunten, 1999), the Rct value can also be obtained from the slope. As the
significant. ozone concentration can be easily measured, the Rct value allows the
Moreover, when ozone is applied to bromide-containing waters, calculation of the HO� concentration in the ozonation process. The
bromate is generated (Han et al., 2014; Moslemi et al., 2011; Ozekin concept of tracking Rct has become beneficial to model the degradation
et al., 1998; Pinkernell and von Gunten, 2001; von Gunten and Hoigne, of micro-pollutants during the treatment of natural water or wastewater
1994; von Gunten and Oliveras, 1998; Zhao et al., 2013). Bromate is by ozonation (Elovitz and von Gunten, 1999). Since Rct is a useful tool to
declared as a potential human carcinogen as it was found to be a gen­ indirectly quantify the relative contribution of HO� radicals in ozonation
otoxic carcinogen inducing compound (e.g., renal cell tumors in rats) system, the catalytic effect of combined ozonation and CMF to generate
(Kurokawa et al., 1990; Moslemi et al., 2012a). The guideline value of HO� radicals during the treatment of SE can be explored using the same
bromate issued by the World Health Organization (WHO) is 25 μg/L. A concept.
maximum contaminant level of 10 μg/L bromate is established by both
the European Union (EU) and the United States Environmental Protec­ 3. Materials and methods
tion Agency (USEPA) (USEPA, 2002; WHO, 1996). Therefore, an un­
derstanding of bromate formation in the unique ozone-ceramic 3.1. Membrane and materials
membrane process is highly significant. In natural waters, bromide
levels are highly variable with a range of 10–1000 μg/L (von Gunten, A multi-channel ceramic membrane was provided by Metawater Co.,
2003b). Bromide itself is not toxic (WHO, 2009), but the formation of Ltd. The 100 nm membrane is a scaled down version of the Metawater
bromate can become a serious issue for bromide concentrations above

2
K. Ibn Abdul Hamid et al. Journal of Environmental Management 253 (2020) 109655

commercial unit that is commonly operated for full scale water treat­ The samples taken from the beaker was used as background data for
ment applications. The 10-cm long membrane had 55 channels, each comparison. To ensure that the recovered sample volume was repre­
with 2.5 mm channel diameter and is mostly made of aluminium oxide. sentative of the solution in contact with the ceramic membrane, sample
The total filtration surface area of the membrane was 0.04 m2. The was discarded for the first 40 s before collecting it into the reagent tube.
characteristics of the membrane are given in Table 1 pCBA was pur­ This 40 s delay was obtained by calculating the breakthrough of indigo
chased from Supelco. Tert-butanol and potassium indigo-trisulfonate solution from the ceramic membrane. The breakthrough of indigo was
were purchased from Sigma Aldrich. Bromide and bromate standards calculated by pumping diluted indigo solution into the module until a
were prepared using potassium bromide (KBr, MERCK Pty Limited) and complete fill of the membrane and module was ensured. Ultra-pure
sodium bromate (NaBrO3, Merck Pty Limited). water was then pumped through the module. The displacement of the
indigo solution contained within the membrane module was observed
by measuring the UV absorbance of the permeate at 600 nm. Samples
3.2. Experimental setup
were taken from the feed beaker and the membrane permeates for
bromate analysis. After each set of experiments, the ceramic membrane
The schematic diagram of the experimental setup is shown in Fig. 1.
was rinsed thoroughly with ultra-pure water.
The membrane was inserted in a custom made stainless steel module
(also provided by Metawater), and was held in a vertical position with
3.4. Analytical methods
clamps at its two ends. The feed was pumped using a low speed piston
pump (Fluid Metering, Inc, QG 150) at a flow rate of 48 mL/min.
3.4.1. Ozone concentration
Stainless steel fittings (Swagelok) and high pressure nylon tubes were
The concentration of dissolved ozone in solution was determined by
used for connecting the membrane rig. An ozone generator (American
the Indigo Method (Bader and Hoign�e, 1981). The method is based on
Ozone Systems Inc.) was used to generate a stock ozone solution. Stock
decolourization of the indigo reagent by ozone, where the loss of colour
solutions were prepared by continuously bubbling the generated ozone
is directly proportional to the ozone concentration. High purity indigo
into ice-cooled ultra-pure water through a Dreschel bottle as described
trisulfonate (>80%) with a molar absorptivity of about 20,
by Bader and Hoign�e (1981) (Bader and Hoign�e, 1981). The concen­
000 M 1 cm 1 at 600 nm was used. To measure the residual ozone, the
tration of the stock solutions was approximately 50 mg-O3/L, which
absorbance of indigo trisulfonate after reaction with the sample was
were standardised by measurement of the UV absorbance (ε258
1 subtracted from that of an ozone free blank. The absorbance at 600 nm
nm ¼ 3000 M cm 1). Glass feed beakers and syringes for injecting stock
was measured using a UV mini-1240 UV–vis spectrophotometer
solution were used to avoid ozone decomposition.
(Shimadzu).

3.3. Experimental procedure 3.4.2. pCBA concentration


pCBA was quantified by high performance liquid chromatography
A fixed volume (200 mL and 1000 mL for the recirculating sample (HPLC) (Agilent Technologies 1200 Series) with UV detection at
and single pass test, respectively) of tap water (Dissolved Organic Car­ 240 nm. The flow rate and eluent were 0.7 mL/min of 70% methanol
bon ¼ 0.29 mg/L; Conductivity ¼ 220 μS/cm; Turbidity ¼ 0.34 NTU) and 30% 10 mM phosphoric acid, respectively. The column was a
was added to the glass beaker (1200 mL). pCBA and bromide was spiked 4.6 mm � 250 mm Kinetex XB-C18 (Phenomenex) 5 μm particle size,
at 0.5 μM and 6.25 μM, respectively. The measured pH value was always with a pre-column attachment. A low detection limit of about 0.025 μM
7.5 (�0.1). The beaker was covered with Parafilm and placed on a (4 mg/L) can be achieved by HPLC.
magnetic stirrer. First, the solution was stirred vigorously for 10 s to
ensure good homogeneity. Then, the stirring was lowered and an aliquot 3.4.3. Bromate concentration
of the stock ozone solution was spiked into the beaker using a glass Bromate was measured via ion chromatography using a Dionex
syringe so that the effective ozone concentration in the beaker reached ICS3000 (AG9HC/AS9HC) followed by a post-column reaction, ac­
approximately 10 mg O3/L. The ozone decomposition in the beaker was cording to the standard operating procedure from an existing method
used as the background data. For the tests, when ozone was spiked into previously reported by Salhi and von Gunten (1999) (Salhi and von
the beaker, the pump was turned on immediately. After each specified Gunten, 1999). All samples were measured in duplicate and blank an­
reaction time, recirculated samples (3 mL) were collected into a tube alyses were performed. The limits of detection (LODs) were calculated
containing buffered indigo trisulphonate to quench the ozone reaction. for every analytical batch. The average LOD for bromate was 4 nM
After 15 min, a sample (3.5 mL) was collected for bromate analysis. (0.5 μgL 1).
In order to investigate ozone stability in the membrane module, the
filtration unit was filled up with fresh ozonized water before the pump 4. Results and discussion
was stopped and the ozone decay was measured after a certain period of
time. This was done for four different contact times (1 min, 3 min, 5 min 4.1. Ozone decay and HO� radical formation
and 7 min) over the course of a single ozonized water feed run. In a
similar experiment, the solution containing pCBA was replaced by a Fig. 2 shows the residual ozone concentration in tap water solution
solution containing 50 mM tert-butanol. After each specified reaction recirculated through the ceramic membrane (fitted inside the module),
time, samples (11 mL) from the feed beaker and the permeate outlet the empty module, and in the stirred feed beaker. The residual ozone
were collected into a test tube containing buffered indigo trisulphonate. concentration was similar in all compartments for the first 2 min of
testing. However, after 3 min, the residual ozone concentration in the
Table 1 empty module and the module containing the membrane declined at a
Properties of the multi-channel ceramic membrane. faster rate. The decrease in ozone concentration started after 2 min and
Properties of membrane 10 s which corresponded to the time needed to receive recirculated
permeate to the feed beaker. After 13 min (Fig. 2), the residual ozone
Manufacturer Metawater Co., Ltd.
Length (cm) 10 concentration for the membrane and empty module were much lower
Outer diameter (mm) 30 than for the stirred feed beaker. During the experiments, the sample was
Channel diameter (mm) 2.5 recirculating through the membrane and the module. The water was
Number of channels 55 passed through the pump and piping to the membrane and or module
Membrane surface area (m2) 0.04
and then discharged from the permeate side of the module. While

3
K. Ibn Abdul Hamid et al. Journal of Environmental Management 253 (2020) 109655

Fig. 1. Generation of ozone stock solution (a) and schematic of the experimental setup for investigating HO� radical formation across the CMF membrane (b).

of ozone via formation of HO� radicals and/or simply degassing of ozone


as it depressurises (7 kPa) when exiting the membrane. These observa­
tions imply there is an important effect of degassing from both the
membrane and its module, however, quantification of the ozone losses
due to degassing is subject of further study. Therefore, the HO� radical
behaviour must be indirectly measured by pCBA as a probe to isolate the
unique effect of the membrane itself. The behaviour of HO� radical via
single pass tests will be examined later in this paper.
The Rct values (ratio of HO� radical exposure vs. ozone exposure)
calculated using the samples from 3 min and onwards for the beaker,
empty module and membrane are also shown in Table 2. For reference,
the original concentration data for pCBA over time is shown in Fig. A1
from the Supplementary Material. Under standard ozonation conditions
and ozone-based advanced oxidation processes, Rct values between 10 7
– 10 10 have been calculated (de Vera et al., 2017; Elovitz et al., 2000).
As expected, samples collected from the stirred feed beaker have the
lowest Rct value compared to either the empty module or membrane.
The lower rate of ozone decomposition in the stirred feed beaker
compared to the empty module and membrane are correlated to a lower
Rct value in the stirred feed beaker samples. However, the Rct value for
the membrane was slightly higher than that of the empty module. This is
due to the enhanced HO� radical formation by the ceramic membrane,
indicating the catalytic degradation of ozone to HO� radicals.
Fig. 2. Comparison of residual ozone concentration in recirculating water so­ However, the calculated decay rates of ozone for the membrane and
lution in a beaker, passing through membrane and empty module (0.5 μM empty module in Table 2 are not representative of the decomposition of
pCBA; 6.25 μM bromide; applied ozone: 10 mg/L; pump flow rate: 48 mL/min; ozone across the membrane/module alone. Since the ozonized tap water
flux: 72 L m 2⋅h 1; temperature: 22 � 1 � C; pH 7.5). solution is recirculated to the feed beaker, the decay rates in Table 2 are
representative of the combination of natural decomposition of ozone in
leaving via the permeate, the tap water was exposed to air before the stirred feed beaker along with the decomposition of ozone across the
returning to the feed beaker. Moreover, flowing water through a pump membrane/module. Agitation and repeated contact with air is a likely
and tubing sufficiently agitated the solution to cause ozone degassing factor in the loss of ozone. McClurkin et al. (2013) determined the
from solution. The differences in ozone consumption between the half-life of ozone as a function of air flow in an airtight cylinder
beaker and the experiments carried with the module might be due to (McClurkin et al., 2013). Three fans were used inside the cylinder to
reaction with tubing material or the module itself and degassing of
ozone. Looking more closely at specific times, between 2 and 8 min in
Fig. 2, the residual ozone concentration through the membrane is Table 2
noticeably lower than in the empty module. As the differences in re­ First order reaction rate constants k and Rct in water solution in a beaker, passing
sidual ozone concentration for the empty module and membrane is through the membrane and empty module (0.5 μM pCBA: 6.25 μM bromide;
applied ozone: 10 mg/L; pump flow rate: 48 mL/min, pH 7.5, temperature:
significant between 2 and 8 min, the first order linear reaction rate
22 � 1� C).
constants and Rct were calculated using data collected during this time
frame. Test k (s 1) Rct

Table 2 shows the first order linear reaction rate constants for the Stirred feed beaker 0.47 � 10 3
1.09 � 10 9
3 9
beaker, empty module and membrane. The differences observed for the Empty module 2.56 � 10 1.82 � 10
3 9
Metawater membrane 2.75 � 10 2.16 � 10
empty module and the membrane may be due to additional consumption

4
K. Ibn Abdul Hamid et al. Journal of Environmental Management 253 (2020) 109655

obtain different flow rates of air. It was observed that the half-life
decreased exponentially as the air flow was increased by increasing
fan speed. Moreover, the experiments herein were conducted in such a
way that the membrane permeate was recirculated back to the feed
beaker and depressurised prior to feeding back to the membrane.
Therefore, to avoid agitation and to isolate the effect of the mem­
brane/module on the decomposition of ozone, single pass experiments
were conducted with tap water. This reduced the contact of the ozonized
sample with air and ensured adequate reaction time with the membrane
without active flow (agitation or pressure).
Table 3 shows an example of ozone decomposition for a holding time
inside the module with and without membrane of 1 min with an addi­
tional 40 s (1.6 min). This 40 s was determined by calibrating against the
breakthrough of the contents contained within the membrane as
described in Section 3.3. The results show that running the feed solution
through the empty module led to a lower concentration of ozone
compared to the expected decay in the beaker. Although, this drop in
ozone concentration was similar to when the membrane was installed,
indicating small differences in the effect on ozone decay between the
empty module and the membrane (Table 3). Meanwhile, the decrease in
ozone concentration with pCBA while being held with the membrane
was larger compared to the decrease in ozone with tert-butanol
(Table 3). This gives further evidence that catalytic decomposition of Fig. 3. Impact on ozone decomposition induced by the ceramic membrane for
different holding times shown by effective rate constants of ozone decomposi­
ozone occurred with the ceramic membrane via a HO� radical reaction.
tion in pCBA (0.5 μM) and tert-Butanol (50 mM) (6.25 μM bromide; flux:
More tests for comparing the ozone decomposition through the empty
72 L m 2⋅h 1; applied ozone: 10 mg/L; pump flow rate: 48 mL/min; tempera­
module with tert-butanol would be helpful to investigate this further. ture: 22 � 1 � C; pH 7.5).

4.2. Kinetic analysis of ozone decay in Tables 3 and i.e., ozone decomposition was slightly enhanced by HO�
radical generation due to the catalytic effect of membrane (Staehelin
To investigate the mechanisms of ozone decay more closely, the re­ and Hoigne, 1985). Moreover, the differences between the pCBA and
sults of ozone decay shown in Table 3 are analysed here. The first order tert-butanol rate constant decreased with increasing holding time (or
kinetic rate constants k were determined. The total ozone decomposition contact time with the membrane) and for a contact time of 7 min, no
rate during each contact time can be expressed by, differences were observed between the pCBA and tert-butanol experi­
ment. This is because, ozone decomposition is faster at higher
kTot ¼ kB þ kMemþMod (3) concentration.

where, kB and kMemþMod are the first order reaction rate constants (s 1) of
ozone inside the beaker and membrane (installed in its module) 4.3. Bromate formation
respectively. kB was calculated from the experiments (the natural
decomposition of ozone in beaker). kTot can be calculated from the In order to investigate the impact of CMF on bromate formation,
membrane decay results. Ozone decomposition across the membrane, bromide was added to the feed solution. Fig. 4 shows bromate forma­
kMemþMod can be determined by rearranging Equation (3). The calculated tions for an experiment performed with permeate recirculation to the
kMemþMod can be defined as the effective rate constant of the ceramic feed beaker for 15 min.
membrane in its module. As per Fig. 4, the bromate concentrations measured in the samples
Fig. 3 shows the calculated effective rate constants (kMemþMod) of taken at 15 min for the stirred beaker test (Fig. 2), empty module and
ozone decomposition for 1, 3, 5 and 7 min holding time. The rate con­ membrane were 4.38 μM, 1.63 μM and 0.8 μM, respectively. The amount
stant of ozone decomposition in the tert-butanol and pCBA solution of bromate formed in the stirred beaker test was much higher compared
through the membrane decreased for 1 min, 3 min, 5 min and 7 min to either the membrane and module tests, since the ozone was more
contact times. Fig. 3 also showed that the rate of ozone depletion was stable in the beaker test. Therefore, the exposure to bromide was higher
higher in the experiments carried out with pCBA compared to the ex­ and a higher bromate formation was observed. However, the formation
periments with tert-butanol. This result confirms the observation shown of bromate in the membrane test was found to be lower (50%) than that
of the empty module. The results therefore suggested that, bromate
Table 3 formation can be minimized using ceramic membranes. In previous
Comparison of residual ozone concentration in the feed beaker, the empty work by Ciba et al. (1995), it was found that the addition of a hetero­
module and the membrane for 1 min holding time. (Tap water with 6.25 μM geneous catalyst (TiO2) to ozone minimized the formation of bromate
bromide, 0.5 μM pCBA and 50 mM tert-butanol; pump flow rate: 48 mL/min; from brominated surface waters (Ciba et al., 1995). The results in this
flux: 72 L m 2⋅h 1; temperature: 22 � 1 � C; pH 7.5). study have identified possible effects including a slight ozone reduction
Test Residual ozone % reduction by catalysis and degassing, but these were very small relative to re­
concentration actions with ozone in bulk. The reduced bromate formation observed in
5 min 6 min our study may be due to a lower ozone exposure with bromide. The
40 s lower ozone exposure may also be due to the small catalytic formation of
Beaker 8.88 8.60 3.2 HO� radical at the surface of the membrane which in turn leads to a
Tap water with pCBA through empty module 8.87 8.04 9.4 lower formation of HOBr (Fischbacher et al., 2015). Further study on
Tap water with tert-butanol through 8.88 7.78 12.4 this was conducted with samples collected in the single-pass tests
membrane (Table 3).
Tap water with pCBA through membrane 8.88 7.41 16.6
Fig. 5 shows the corresponding bromate concentration results during

5
K. Ibn Abdul Hamid et al. Journal of Environmental Management 253 (2020) 109655

Fievet et al., 2002; Labbez et al., 2003; Skluzacek et al., 2006). If a


charged membrane is in contact with an electrolyte solution, electro­
static repulsion will result in a lower concentration of ions with the same
charge as the membrane (referred to as co-ions) near the membrane
surface and within the membrane pores. On the other hand,
counter-ions, which have an opposite charge to that of the membrane
surface, are attracted to the membrane and the concentration of these
ions within the membrane and at its surface is higher than that in bulk
solution (Moslemi et al., 2012b). To exhibit a beneficial effect on
bromate removal by such mechanism, rejection of up to 68% has been
reported with a 5 kDa ceramic UF membranes having estimated pore
size of 0.004 μm (Moslemi et al., 2012b). However this dropped to
13.3% and lower in the presence of Ca2þ with concentrations starting at
0.4 mg/L. While Ca2þ ion influence on bromate rejection was subject of
their study, it was also acknowledged that minerals in general (including
the more dominant Naþ present in water supplies) will have similar
influence in shielding the surface and reducing the rejection effect on
bromate. Considering our test water had a mineral content
(EC ¼ 220 μS/cm) and a 25-fold larger pore size, we don’t expect the
membrane to participate in rejecting bromate by the charge repulsion
mechanism. Meanwhile the fundamentals effects of the surface proper­
ties of ceramic membrane (ion interactions with the surface and
Fig. 4. Bromate concentration during ozonation of a water sample passing zeta-potential) play an important role and is a worthy topic to explore
through ceramic membrane and empty module (0.5 μM pCBA; 6.25 μM bro­ any further beneficial bromate removal effects in detail.
mide; applied ozone: 10 mg/L; pump flow rate: 48 mL/min; pH 7.5; flux: This work attempted to identify the formation of HO� radicals during
72 L m 2⋅h 1; temperature: 22 � 1 � C, 15 min). ozonation of CMF. This has enabled to identify whether HO� radicals are
formed and relative to how much ozone is present, and any potential
the single-pass tests. The concentration of bromate formed in the influence on other mechanisms such as disinfection by-products. As the
membrane and empty module are compared with the concentration of degassing effect was a finding as part of the work, another dedicated
bromate formed in the feed water. It is interesting to observe that the work would be needed to isolate the effect of ozone degassing over other
concentrations of bromate in the permeate of the membrane is reduced ozone removal mechanisms such as chemical reaction including HO�
by 10 μg/L to 50 μg/L for every holding time. Whereas, the concentra­ radical formation. It would need close examination of the equipment
tion of bromate at the outlet of the empty module is essentially equal to used, and likely better undertaken on a pilot scale membrane system
that of the feed. The membrane therefore, appeared to slightly mitigate which uses realistic pumps, valves, piping and tanks.
the formation of bromate.
The reduction of bromate formation could also be explained by the 5. Conclusions
surface property of the ceramic membrane that causes the attraction and
repulsion of ions. The surface of ceramic membranes is typically Ceramic membranes are known to benefit from a pre-treatment of
charged. This is due to the amphoteric nature of the surface hydroxyl the feed water with ozone. However, the mechanism of ozone on the
groups on the metal oxide surface of the ceramic membrane (Tsuru, membrane surface is not well understood. This work found that a
2001). Previous studies have shown that due to their surface charge, commercial ceramic membrane provided by Metawater led to an in­
ceramic UF membranes can reject ions, even though the pore size of the crease of HO� radicals formation that induced an increased ozone decay.
membrane is much larger than the size of the ions (Combe et al., 1997; The contact of pipes and pump of the membrane rig and air exposure

Fig. 5. Bromate formation during ozonation of water sample passing through the ceramic membrane (a) and the empty module (b) (0.5 μM pCBA; 6.25 μM (500 μg/
L) bromide; applied ozone: 10 mg/L, pump flow rate: 48 mL/min; flux: 72 L m 2⋅h 1; temperature: 22 � 1 � C; pH 7.5).

6
K. Ibn Abdul Hamid et al. Journal of Environmental Management 253 (2020) 109655

significantly reduced the residual ozone concentration. Meanwhile, an Hoign�e, J., Bader, H., 1983. Rate constants of reactions of ozone with organic and
inorganic compounds in water-I. Non-dissociating organic compounds. Water Res.
effective reduction of bromate formation by using ozonation followed by
17, 173–183.
the CMF was demonstrated. When ozonized water was recirculated to Ikhlaq, A., Brown, D.R., Kasprzyk-Hordern, B., 2015. Catalytic ozonation for the removal
the membrane module, bromate formation was reduced by 50% of organic contaminants in water on alumina. Appl. Catal. B Environ. 165, 408–418.
compared to the experiment conducted with an empty module. When Karnik, B.S., Davies, S.H.R., Chen, K.-C., Jaglowski, D.R., Baumann, M.J., Masten, S.J.,
2005. Effects of ozonation on the permeate flux of nanocrystalline ceramic
single pass experiments were performed, bromate mitigation was membranes. Water Res. 39, 728–734.
around 10%. The mechanism responsible for the mitigation of bromate Kim, J., Davies, S.H.R., Baumann, M.J., Tarabara, V.V., Masten, S.J., 2008. Effect of
was the suppression of the oxidation of Br and HOBr/BrO to BrO3 due ozone dosage and hydrodynamic conditions on the permeate flux in a hybrid
ozonation–ceramic ultrafiltration system treating natural waters. J. Membr. Sci. 311,
to the decomposition of ozone and therefore lower O3 exposure. From a 165–172.
practical point of view for application to water treatment, it can be Kurokawa, Y., Maekawa, A., Takahashi, M., Hayashi, Y., 1990. Toxicity and
concluded that although HO� radicals might react with organics present carcinogenicity of potassium bromate–a new renal carcinogen. Environ. Health
Perspect. 87, 309–335.
in the raw water, subtle effects such as reduction of bromate, could be Labbez, C., Fievet, P., Thomas, F., Szymczyk, A., Vidonne, A., Foissy, A.A., Pagetti, P.,
useful features of the unique catalytic effect of ceramic membranes. The 2003. Evaluation of the "DSPM" model on a titania membrane: measurements of
application of pCBA played a significant role in identifying HO� radical charged and uncharged solute retention, electrokinetic charge, pore size, and water
permeability. J. Colloid Interface Sci. 262, 200–211.
formation during ozone-CMF interaction, however further work is Legube, B., Karpel Vel Leitner, N., 1999. Catalytic ozonation: a promising advanced
needed to identify where the HO� radicals are formed (e.g. inside the oxidation technology for water treatment. Catal. Today 53, 61–72.
pores). Lehman, S.G., Liu, L., 2009. Application of ceramic membranes with pre-ozonation for
treatment of secondary wastewater effluent. Water Res. 43, 2020–2028.
Ma, J., Graham, N.J.D., 2000. Degradation of atrazine by manganese-catalysed
Acknowledgement ozonation - influence of radical scavengers. Water Res. 34, 3822–3828.
McClurkin, J.D., Maier, D.E., Ileleji, K.E., 2013. Half-life time of ozone as a function of air
The authors are grateful to the Collaborative Research Network movement and conditions in a sealed container. J. Stored Prod. Res. 55, 41–47.
Moslemi, M., Davies, S.H.R., Masten, S.J., 2011. Bromate formation in a hybrid
(CRN), Australia and National Centre of Excellence in Desalination, ozonation-ceramic membrane filtration system. Water Res. 45, 5529–5534.
Australia (NCEDA) for providing financial support of this project. The Moslemi, M., Davies, S.H., Masten, S.J., 2012. Empirical modeling of bromate formation
membrane was generously supplied by Metawater, Japan. The experi­ during drinking water treatment using hybrid ozonation membrane filtration.
Desalination 292, 113–118.
ments were conducted in the Curtin Water Quality Research Centre Moslemi, M., Davies, S.H., Masten, S.J., 2012. Rejection of bromide and bromate ions by
(CWQRC) at Curtin University. The authors are thankful for all these a ceramic membrane. Environ. Eng. Sci. 29, 1092–1096.
contributions. Munter, R., 2001. Advanced oxidation processes - current status and prospects. Proc.
Estonian Acad. Sci. Chem.
Myat, D.T., Roddick, F., Puspita, P., Skillman, L.C., Charrois, J.W.A., Kristiana, I.,
Appendix A. Supplementary data Uhl, W., Vasyukova, E.V., Roeszler, G., Chan, A., Zhu, B., Muthukumaran, S.,
Gray, S., Duke, M., 2018. Effect of oxidation with coagulation and ceramic
microfiltration pre-treatment on reverse osmosis for desalination of recycled
Supplementary data to this article can be found online at https://doi. wastewater. Desalination 431, 106–118.
org/10.1016/j.jenvman.2019.109655. Nawrocki, J., Fijołek, L., 2013. Effect of aluminium oxide contaminants on the process of
ozone decomposition in water. Appl. Catal. B Environ. 142–143, 533–537.
Nawrocki, J., Kasprzyk-Hordern, B., 2010. The efficiency and mechanisms of catalytic
References
ozonation. Appl. Catal. B Environ. 99, 27–42.
Neta, P., Dorfman, L.M., 1968. Pulse radiolysis studies. XIII. Rate constants for the
Bader, H., Hoign�e, J., 1981. Determination of ozone in water by the indigo method. reaction of hydroxyl radicals with aromatic compounds in aqueous solutions. Adv.
Water Res. 15, 449–456. Chem. Ser. 81, 222–230.
Ciba, N., Allemane, H., Prados, M.J., Luck, F., 1995. The impact of catalytic ozonation on Oneby, M.A., Bromley, C.O., Borchardt, J.H., Harrison, D.S., 2010. Ozone treatment of
bromate formation. In: Proceedings of the 12th Ozone World Congress, 1, secondary effluent at U.S. municipal wastewater treatment plants. Ozone Sci. Eng.
pp. 585–594. 32, 43–55.
Combe, C., Guizard, C.G., Aimar, P., Sanchez, V., 1997. Experimental determination of Ozekin, K., Westerhoff, P., Amy, G.L., Siddiqui, M., 1998. Molecular ozone and radical
four characteristics used to predict the retention of a ceramic nanofiltration pathways of bromate formation during ozonation. J. Environ. Eng. 124, 456–462.
membrane. J. Membr. Sci. 129, 147–160. Park, H., Kim, Y., An, B., Choi, H., 2012. Characterization of natural organic matter
David Yao, C.C., Haag, W.R., 1991. Rate constants for direct reactions of ozone with treated by iron oxide nanoparticle incorporated ceramic membrane-ozonation
several drinking water contaminants. Water Res. 25, 761–773. process. Water Res. 46, 5861–5870.
de Vera, G.A., Stalter, D., Gernjak, W., Weinberg, H.S., Keller, J., Farre, M.J., 2015. Pinkernell, U., von Gunten, U., 2001. Bromate minimization during ozonation:
Towards reducing DBP formation potential of drinking water by favouring direct mechanistic considerations. Environ. Sci. Technol. 35, 2525–2531.
ozone over hydroxyl radical reactions during ozonation. Water Res. 87, 49–58. Salhi, E., von Gunten, U., 1999. Simultaneous determination of bromide, bromate and
de Vera, G.A., Gernjak, W., Weinberg, H., Farr�e, M.J., Keller, J., von Gunten, U., 2017. nitrite in low μg l 1 levels by ion chromatography without sample pretreatment.
Kinetics and mechanisms of nitrate and ammonium formation during ozonation of Water Res. 33, 3239–3244.
dissolved organic nitrogen. Water Res. 108, 451–461. Schlichter, B., Mavrov, V.D., Chmiel, H., 2004. Study of a hybrid process combining
Dow, N., Murphy, D., Clement, J., Duke, M., 2013. Outcomes of the australian ozone/ ozonation and microfiltration/ultrafiltration for drinking water production from
ceramic membrane trial on secondary effluent. Water (AWA) 40, 7. surface water. Desalination 168, 307–317.
Elovitz, M.S., von Gunten, U., 1999. Hydroxyl radical/ozone ratios during ozonation Skluzacek, J.M., Isabel Tejedor, M., Anderson, M.A., 2006. An iron-modified silica
processes. I. The Rct concept. Ozone: Sci. Eng. 21, 239–260. nanofiltration membrane: effect of solution composition on salt rejection.
Elovitz, M.S., von Gunten, U., Kaiser, H.-P., 2000. Hydroxyl radical/ozone ratios during Microporous Mesoporous Mater. 94, 288–294.
ozonation processes. II. The effect of temperature, pH, alkalinity, and DOM Song, J., Zhang, Z., Zhang, X., 2017. A comparative study of pre-ozonation and in-situ
properties. Ozone Sci. Eng. 22, 123–150. ozonation on mitigation of ceramic UF membrane fouling caused by alginate.
Fievet, P., Labbez, C., Szymczyk, A., Vidonne, A., Foissy, A.A., Pagetti, J., 2002. J. Membr. Sci. 538, 50–57.
Electrolyte transport through amphoteric nanofiltration membranes. Chem. Eng. Sci. Staehelin, J.N., Hoigne, J., 1985. Decomposition of ozone in water in the presence of
57, 2921–2931. organic solutes acting as promoters and inhibitors of radical chain reactions.
Fischbacher, A., L€oppenberg, K., von Sonntag, C., Schmidt, T.C., 2015. A new reaction Environ. Sci. Technol. 19, 1206–1213.
pathway for bromite to bromate in the ozonation of bromide. Environ. Sci. Technol. Tsuru, T., 2001. Inorganic porous membranes for liquid phase separation. Separ. Purif.
49, 11714–11720. Methods 30, 191–220.
Gracia, R.d., Cortes, S., Sarasa, J., Ormad, M.P., Ovelleiro, J.L., 2000. Heterogeneous USEPA, 2002. National primary drinking water regulations: long term 1 enhanced
catalytic ozonation with supported titanium dioxide in model and natural waters. surface water treatment rule. Final Rule. Fed Regist 67, 1811–1844.
Ozone Sci. Eng. 22, 461–471. von Gunten, U., 2003. Ozonation of drinking water: Part I. Oxidation kinetics and
Gracia, R.d., Cort�es, S., Sarasa, J., Ormad, P., Ovelleiro, J.L., 2000. Catalytic ozonation product formation. Water Res. 37, 1443–1467.
with supported titanium dioxide. The stability of catalyst in water. Ozone Sci. Eng. von Gunten, U., 2003. Ozonation of drinking water: Part II. Disinfection and by-product
22, 185–193. formation in presence of bromide, iodide or chlorine. Water Res. 37, 1469–1487.
Han, Q., Wang, H., Dong, W., Liu, T., Yin, Y., 2014. Suppression of bromate formation in von Gunten, U., Hoigne, J., 1994. Bromate formation during ozonization of bromide-
ozonation process by using ferrate(VI): batch study. Chem. Eng. J. 236, 110–120. containing waters: interaction of ozone and hydroxyl radical reactions. Environ. Sci.
Hausmann, A., Duke, M.C., Demmer, T., 2013. Principles of Membrane Filtration, Technol. 28, 1234–1242.
Membrane Processing. Blackwell Publishing Ltd., pp. 17–51 von Gunten, U., Oliveras, Y., 1998. Advanced oxidation of bromide-containing waters:
bromate formation mechanisms. Environ. Sci. Technol. 32, 63–70.

7
K. Ibn Abdul Hamid et al. Journal of Environmental Management 253 (2020) 109655

Wei, D., Tao, Y., Zhang, Z., Liu, L., Zhang, X., 2016. Effect of in-situ ozonation on ceramic Zhao, G., Lu, X., Zhou, Y., Gu, Q., 2013. Simultaneous humic acid removal and bromate
UF membrane fouling mitigation in algal-rich water treatment. J. Membr. Sci. 498, control by O3 and UV/O3 processes. Chem. Eng. J. 232, 74–80.
116–124. Zhu, B., Hu, Y., Kennedy, S., Milne, N., Morris, G., Jin, W., Gray, S., Duke, M., 2011. Dual
WHO, 1996. Guidelines for Drinking-Water Quality, Health Criteria and Other function filtration and catalytic breakdown of organic pollutants in wastewater using
Supporting Information. World Health Organization, Geneva. ozonation with titania and alumina membranes. J. Membr. Sci. 378, 61–72.
WHO, 2009. Bromide in drinking-water. Guidelines Drink. Water Qual. 15.

You might also like