Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

Accepted Manuscript

Adsorption of natural organic matter and disinfection byproduct precursors from


surface water onto TiO2 nanoparticles: pH effects, isotherm modeling, and
implications for the use of TiO2 for drinking water treatment

Stephanie Gora, Susan Andrews


PII: S0045-6535(17)30144-3
DOI: 10.1016/j.chemosphere.2017.01.125
Reference: CHEM 18738

To appear in: ECSN

Received Date: 4 November 2016


Revised Date: 22 January 2017
Accepted Date: 24 January 2017

Please cite this article as: Gora, S., Andrews, S., Adsorption of natural organic matter and disinfection
byproduct precursors from surface water onto TiO2 nanoparticles: pH effects, isotherm modeling,
and implications for the use of TiO2 for drinking water treatment, Chemosphere (2017), doi: 10.1016/
j.chemosphere.2017.01.125.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

1 Adsorption of Natural Organic Matter and Disinfection Byproduct Precursors from

2 Surface Water on TiO2 Nanoparticles: pH Effects, Isotherm Modelling and Implications

3 for using TiO2 for Drinking Water Treatment

PT
5 Authors:

RI
6 Stephanie Goraa (corresponding author)

7 stephanie.gora@mail.utoronto.ca

SC
8

9 Susan Andrewsa

U
10 sandrews@civ.utoronto.ca
AN
11
a
12 Department of Civil Engineering
M

13 University of Toronto

14 35 St. George Street


D

15 Toronto, Ontario
TE

16 Canada

17 M5S 1A4
C EP
AC

1
ACCEPTED MANUSCRIPT

18 Adsorption of Natural Organic Matter and Disinfection Byproduct Precursors from

19 Surface Water on TiO2 Nanoparticles: pH Effects, Isotherm Modelling and Implications

20 for using TiO2 for Drinking Water Treatment

21 Stephanie Goraa and Susan Andrewsb

PT
a
22 Department of Civil Engineering, University of Toronto, 35 St. George

RI
23 Street, Toronto, Ontario, Canada M5S 1A4. Telephone: + 1 416 689 2439; fax: +1 416 978 3674;

24 e-mail: stephanie.gora@mail.utoronto.ca

SC
b
25 Department of Civil Engineering, University of Toronto, 35 St. George

26 Street, Toronto, Ontario, Canada M5S 1A4. Telephone: + 1 416 946 0908; fax: +1 416 978 3674;

U
27 e-mail: sandrews@civ.utoronto.ca
AN
28

29 Abstract
M

30 Titanium dioxide is a photocatalyst that can remove organic contaminants of interest to the drinking water

31 treatment industry, including natural organic matter (NOM) and disinfection byproduct (DBP) precursors.
D

32 The photocatalytic reaction occurs in two steps: adsorption of the contaminant followed by degradation of
TE

33 the adsorbed contaminant upon irradiation with UV light. The second part of this process can lead to the

34 formation of reactive intermediates and negative impacts on treated water quality, such as increased DBP
EP

35 formation potential (DBPfp). Adsorption alone does not result in the formation of reactive intermediates

36 and thus may prove to be a safe way to incorporate TiO2 into drinking water treatment processes. The
C

37 goal of this study was to expand on the current understanding of NOM adsorption on TiO2 and examine it
AC

38 in a drinking water context by observing NOM adsorption from real water sources and evaluating the

39 effects of the resulting reductions on the DBPfp of the treated water. Bottle point isotherm tests were

40 conducted with raw water from two Canadian water treatment plants adjusted to pH 4, pH 6 and pH 8 and

41 dosed with TiO2 nanoparticles. The DOC results were a good fit to a modified Freundlich isotherm. DBP

2
ACCEPTED MANUSCRIPT

42 precursors and liquid chromatography with organic carbon detection NOM fractions associated with DBP

43 formation were removed to some extent at all pHs, but most effectively at pH 4.

44

45 Keywords

PT
46 Titanium dioxide, advanced oxidation, natural organic matter, disinfection byproducts, adsorption,

47 drinking water

RI
48

SC
49 Highlights

50 • TiO2 photocatalysis with UVA light increased the DBPfp of a natural surface water at short

U
51 treatment times but reduced it at longer treatment times


AN
52 In the absence of irradiation, TiO2 nanoparticles adsorbed NOM and DBP precursors from natural

53 surface water


M

54 The adsorption of NOM, THM precursors and HAA precursors by TiO2 nanoparticles was

55 modelled using a simple isotherm model


D

56 • The adsorption of DOC, UV254, humic substances and DBP precursors by TiO2 nanoparticles
TE

57 occurred more readily at pH 4 than at pH 6 or pH 8


C EP
AC

3
ACCEPTED MANUSCRIPT

58 1. Introduction

59 Nanomaterials, defined as materials with any dimension in the nanoscale or having internal or surface

60 structure in the nanoscale (ISO, 2010), are increasingly being used in the fields of electronics, computing

61 and medicine. Some nanomaterials, including titanium dioxide (TiO2) nanoparticles, may also prove to be

PT
62 useful in environmental applications, including the treatment of drinking water and wastewater.

63

RI
64 1.1 Titanium dioxide for drinking water treatment

SC
65 The use of TiO2 for drinking water purification has been explored by many researchers, and at least two

66 companies have developed small-scale systems on the basis of TiO2 photocatalysis, but it has yet to be

U
67 widely applied for municipal drinking water treatment. Photocatalytic degradation of aqueous
AN
68 contaminants by TiO2 is generally thought to occur in two steps: adsorption and degradation. Adsorption

69 can occur in the absence of light, but degradation occurs only when TiO2 is irradiated. Upon irradiation,
M

70 reactive oxygen species (ROS), including the hydroxyl radical, and oxidative and reducing centres on the

71 surface of the nanoparticle degrade contaminants that have been adsorbed on the surface of the
D

72 photocatalyst. This two-step reaction is often described using the Langmuir–Hinshelwood mechanism
TE

73 (Malato et al., 2009). The two-stage nature of photocatalytic degradation differentiates it from other

74 advanced oxidation processes such as UV/H2O2 and O3/H2O2. Like other AOPs, however, many ROS
EP

75 formed when TiO2 is irradiated with UVA light are non-specific oxidants; hence, TiO2 photocatalysis has

76 the potential to provide concurrent disinfection and degradation of organic drinking water contaminants,
C

77 including taste and odour compounds and cyanotoxins (Fotiou et al., 2015), various pharmaceuticals
AC

78 (Avisar et al., 2013; Kanakaraju et al., 2014), and disinfection byproduct (DBP) precursors such as natural

79 organic matter (NOM) (Liu et al., 2008, Huang et al., 2007, Philippe et al., 2010). The photocatalytic

80 degradation process preferentially targets large aromatic NOM compounds, breaking them down into

81 smaller ones (Huang et al., 2007; Philippe et al., 2010). This can result in decreased membrane fouling

82 (Huang et al., 2007) and changes in DBP formation potential (DBPfp). The latter is of particular concern

4
ACCEPTED MANUSCRIPT

83 because although some of the degradation products of photocatalysis may be more likely to form DBPs

84 than the original compounds, others may have an equal or greater DBPfp, particularly at shorter treatment

85 times (Liu et al., 2008). As a result, an adsorption-based process may prove to be a safer and more

86 effective option for the removal of NOM from drinking water using TiO2.

PT
87

88 1.2 Natural organic matter

RI
89 The removal of NOM, a heterogeneous mixture of organic compounds found in most surface water

SC
90 sources, is an important objective of drinking water treatment because of its aesthetic, operational and

91 health effects. The latter are primarily linked to the role of NOM as a precursor to form both regulated

U
92 and unregulated DBPs. Operational concerns related to NOM include membrane fouling, competitive

93 adsorption to adsorbent materials intended for the removal of taste and odour compounds, interference
AN
94 with UV disinfection and consumption of chlorine during disinfection.

95 NOM is commonly quantified as total organic carbon (TOC) or dissolved organic carbon (DOC), bulk
M

96 parameters that measure the total mass of organic carbon compounds present in a water sample without
D

97 differentiating them from one another. UV-Vis absorbance at 254 nm (UV254) and fluorescence are also
TE

98 used to characterise NOM, though these methods are specific to NOM compounds containing aromatic

99 chromophores or fluorophores. More complex methods such as liquid chromatography with organic
EP

100 carbon detection (LC-OCD) have been developed that allow researchers to separate NOM into different

101 fractions based on size or chemical characteristics. The resulting fractions are then quantified using DOC
C

102 or UV254. LC-OCD separates NOM into five fractions based on size and/or chemical characteristics as
AC

103 follows: biopolymers, associated with membrane fouling (Wray et al., 2013); humic substances, which

104 have been linked to DBP formation (Wassink et al., 2011); building blocks; low-molecular-weight acids

105 (LMWAs); and low-molecular-weight neutrals (LMWNs) (Huber et al., 2011). A given NOM sample’s

106 potential to form regulated and unregulated DBPs can also be assessed more directly by measuring its

107 DBPfp by chlorinating water samples and measuring the DBPs formed under standardised conditions.

108

5
ACCEPTED MANUSCRIPT

109 1.3 Adsorption of NOM to TiO2

110 Studies by Mwaanga et al. (2014), Erhayem and Sohn (2014) and Kim and Shon (2007) have established

111 that pH and ionic strength affect the adsorption of NOM on TiO2 and have noted that larger, more

112 aromatic NOM compounds are adsorbed preferentially. The presence of bicarbonate, phosphate and

PT
113 nitrate has been shown to decrease NOM adsorption on TiO2 while the presence of magnesium and

114 calcium increases the likelihood of NOM adsorption (Erhayem and Sohn, 2014; Sun and Lee, 2012).

RI
115 Other studies have described the effects of individual ions on the agglomeration of TiO2 nanomaterials.

SC
116 Agglomeration decreases the overall surface area available for adsorption and as such is likely to have an

117 impact on the ability of TiO2 nanomaterials to adsorb NOM. Liu et al. (2013) reported that three types of

U
118 TiO2 nanomaterials were more likely to agglomerate under high-ionic strength conditions than at low-

119 ionic strength conditions. This finding is corroborated by those of Erhayem and Sohn (2014). The type of
AN
120 ions present in solution may also have an effect. Liu et al. (2013) observed greater increases in

121 agglomerate size when calcium was added to the water rather than sodium. They hypothesized that this
M

122 was because of the greater ability of Ca2+ to compress the electrical double layer surrounding the
D

123 nanomaterials relative to Na+. Greater compression of the electrical double layer results in less repulsion
TE

124 between individual nanoparticles and thus, greater agglomeration.

125 The presence of NOM increases the stability of nanomaterials in solution, though this effect is less
EP

126 pronounced in the presence of ions such as calcium (Liu et al., 2013; Zhang et al. 2009) and at high NOM

127 concentrations (Erhayem and Sohn, 2014). According to Zhang et al. (2009), NOM inhibits
C

128 agglomeration by increasing the overall negative charge of the particles, thus increasing the repulsive
AC

129 forces that keep them dispersed in solution.

130

131 1.4 Adsorption models

132 Adsorption processes are usually evaluated in the laboratory using adsorption isotherm models. The

133 Freundlich isotherm often fits well to empirical data and can be used to model heterogeneous systems

134 such as the adsorption of organic molecules to activated carbon (Summers et al., 1988). It can be used to

6
ACCEPTED MANUSCRIPT

135 describe multilayer adsorption, reversible adsorption and adsorbents with non-uniform adsorption sites

136 (Shahbeig et al., 2013).

137 Summers and Roberts (1988) found that a modified version of the Freundlich isotherm could be used to

138 describe the adsorption of NOM to activated carbon when experiments were conducted with a constant

PT
139 initial concentration of NOM and changing doses (D) of activated carbon. They also developed the

140 following equation (referred as the SR model in this paper) to express this relationship:

RI
141 = (1)

SC
142 where D has units of mg L−1 or g L−1. The SR model was used and extended upon by numerous

143 researchers including Karanfil et al. (1999), Li et al. (2002), Hyung and Kim (2008) and Qi et al. (2008)

144

U
to characterise the adsorption of NOM to activated carbon and carbon nanotubes. Erhayem and Sohn
AN
145 (2014) modelled the adsorption of Suwannee River humic acids, fulvic acids, and NOM to P25 TiO2

146 nanoparticles using the SR isotherm model. They noted that the SR adsorption constant (and thus the
M

147 extent of adsorption) increased at lower pH and at higher ionic strength.

148
D

149 1.5 Potential risks and opportunities associated with the use of TiO2 nanoparticles for water
TE

150 treatment

151 Conventional water treatment technologies such as coagulation and activated carbon are effective for
EP

152 NOM removal but, like all treatment technologies, have limitations. In North America, coagulation is

153 widely used to remove NOM and turbidity from drinking water, but it does not remove some recalcitrant
C

154 organics. It also creates a substantial amount of waste, often referred to as coagulation residuals.
AC

155 Activated carbon readily removes NOM and other organic compounds but eventually becomes exhausted,

156 and must undergo an expensive and energy intensive regeneration process if it is to be reused. Although

157 TiO2 is more well known for its photocatalytic properties, it also adsorbs NOM as a part of that process,

158 and is potentially regenerable onsite (Liu et al., 2014). As such, it might prove to be a useful alternative to

159 the existing treatment options.

7
ACCEPTED MANUSCRIPT

160 TiO2 nanoparticles are not without health and environmental concerns. Inhalation is the route of exposure

161 of the greatest concern for human health (Shi et al., 2013), but the transport of nanoparticles in the

162 environment is coming under increasing scrutiny (Yang and Westerhoff, 2014). The adsorption of various

163 water components, including NOM, on TiO2 can facilitate the latter’s transport through water systems. To

PT
164 date, most of the studies on the interactions between NOM and TiO2 have aimed to elucidate these effects

165 to predict and minimise the impact of nanomaterials on the natural environment (e.g. Mwaanga et al.,

RI
166 2014; Erhayem and Sohn, 2014; Kim and Shon, 2007; and Liu et al., 2013)).

SC
167 This study aims to expand on the existing research into the adsorption of NOM by TiO2, most of which

168 have been conducted in a contaminant transport context. This study also explores the potential application

U
169 of TiO2 as an adsorbent in drinking water treatment by studying its behaviour in natural surface water

170 sources and measuring its ability to remove DOC, UV254 and DBP precursors.
AN
171

172 2. Materials and methods


M

173 2.1 Materials


D

174 Evonik Aeroxide P25 TiO2 nanoparticles were purchased from Sigma Aldrich (Canada) and used without
TE

175 further modification. THM and HAA standards were also purchased from Sigma Aldrich. DOC standards

176 were prepared by dissolving potassium hydrogen phthalate into milliQ water to create a 1000-mg L−1
EP

177 stock solution, which was then diluted as required. Raw water was obtained from the inlets of two water

178 treatment plants (WTPs) in Ontario, Canada, both of which use surface water supplies. The Peterborough
C

179 WTP is supplied by the Otonabee River, while the R.C. Harris WTP in Toronto draws its water from
AC

180 Lake Ontario, one of the largest lakes in North America, which provides water to over 9 million people in

181 Canada and the United States. All water samples were obtained ahead of any pre-chlorination point at the

182 WTP and characterised upon return to the laboratory. Measured ranges for five relevant water parameters

183 are provided in Table S1 in the supplemental file.

184 The water sources varied primarily in terms of their NOM content and aromaticity. The DOC of the

185 Otonabee River water was approximately 3–4 times higher than that of the Lake Ontario water, while its

8
ACCEPTED MANUSCRIPT

186 UV254 was approximately five times higher. The SUVA of the Otonabee River water ranged from 2.0 to

187 2.4 L/mg.m, while that of the Lake Ontario water ranged from 0.8 to 1.0 L/mg.m, indicating that the

188 NOM in the former is more aromatic than that in the latter.

189

PT
190 2.2 Analytical methods

191 DOC was determined using an O.I. Analytical Aurora 1030W TOC analyzer operating in the persulphate

RI
192 oxidation mode and UV absorption at 254 nm (UV254) was analysed on an Agilent 8453 UV-Vis

SC
193 analyzer. SUVA was calculated by normalizing UV254 by DOC. Alkalinity was measured using Standard

194 Method 2320 (APHA, 2005).

U
195 Duplicate raw and treated water samples were analysed using size exclusion LC-OCD as described by

196 Huber et al. (2011). The results of the analyses were processed using proprietary software (ChromCalc,
AN
197 DOC-LABOR, Karlsruhe, Germany).

198 The isoelectric point of the nanoparticles was determined by measuring the zeta potential of the
M

199 nanoparticles at different pH values. A series of samples containing 0.1 g L−1 TiO2 in 10 mM NaCl were
D

200 adjusted to pHs ranging from 2 to 9 using 0.1 N HCl or 0.1 N NaOH as per the method outlined by the
TE

201 Nanotechnology Characterization Laboratory (2009). The zeta potential of the samples was measured

202 using a Horiba Scientific Nanopartica SZ-100 Nanoparticle Analyzer. The size of the nanoparticle
EP

203 agglomerates formed at different pHs was determined using a Malvern MasterSizer 3000.

204 The uniform formation conditions (UFCs) method as described by Summers et al. (1996) was used to
C

205 assess the chlorine demand and DBPfp of the raw water and the water that had been treated with TiO2.
AC

206 The UFC test was designed to mimic the conditions commonly found in distribution systems in North

207 America. The chlorine demand and DBPfp tests were conducted on samples buffered with a borate

208 solution and adjusted to pH 8 with 1 N HCl or 1 N NaOH. Samples were stored in the dark at 20°C for 24

209 h, after which the free chlorine residual was measured using Standard Method 4500-G (APHA 2005). The

210 DBPfp samples were dosed with sufficient sodium hypochlorite to ensure that they would have a chlorine

211 residual of 1 ± 0.4 mg L−1 after the 24 h holding time. After 24 h, the trihalomethanes and haloacetic acids

9
ACCEPTED MANUSCRIPT

212 formed during the UFC tests were extracted according to Standard Method 6232 B and Standard Method

213 6251 B (APHA, 2005) and analysed on a Agilent 7890B GC-ECD [Standard Method 6232 B and

214 Standard Method 6251 B (APHA, 2005)]. Blanks and 20 µg L−1 check standards were analysed after

215 every 10 samples.

PT
216 All statistical analyses, including Tukey’s method for multiple comparisons to establish a point of

217 practical equilibrium and linear regression of the adsorption data to determine KF and KSR, were

RI
218 conducted at the 95% confidence level. Reported error values represent half of the calculated confidence

SC
219 interval unless otherwise specified.

220

U
221 2.3 Sample preparation AN
222 NOM degradation studies were conducted in a high-intensity UV reactor equipped with UV LED lamps

223 emitting UVA light at 365 cm with an average irradiance of 6.25 mW/cm2. Unchlorinated raw water from
M

224 the Peterborough Water Treatment Plant (Otonabee River water) was dispensed into three 50-mL batch

225 reactors, dosed with 0.25 g L−1 of P25 nanoparticles, allowed to mix in the dark for 1 min, and then
D

226 exposed to the LED light for times ranging from 0 to 60 min. The treated samples were chlorinated
TE

227 according to the UFC method and analysed for THMfp and HAAfp as described in Section 2.2.

228 The time required to reach a stable adsorption equilibrium between NOM and TiO2 nanoparticles was
EP

229 determined by adding 75 mL of raw unchlorinated water to duplicate 125 mL amber bottles and, when

230 necessary, adjusting the pH to 4, 6, or 8 with 1 N HCl or 1N NaOH. The bottles were dosed with 0.5 g L−1
C

231 of Evonik P25 TiO2 nanoparticles and mixed end-over-end in a box mixer for times ranging from 1 to 8 h.
AC

232 This time range was chosen because previous experiments (results not shown) had indicated that most

233 NOM adsorption occurred within 5 min and that equilibrium likely occurred between 1 and 4 h.

234 For the bottle point isotherm tests, eight 250 mL amber bottles were filled with 150 mL of raw water;

235 adjusted to pH 4, 6, or 8; dosed with 0, 0.01, 0.025, 0.05, 0.1, 0.25, 0.5, or 1 g L−1 of Evonik P25 TiO2

236 nanoparticles; and then mixed continuously in the dark for 4 h in an end-over-end box mixer. All

10
ACCEPTED MANUSCRIPT

237 experiments were run in triplicate. All replicate samples were analysed for DOC, UV254 and SUVA. The

238 samples from one replicate experiment were used to establish chlorine demand, and the remaining

239 samples were analysed for THMfp and HAAfp. The DOC, THMfp and HAAfp results of the bottle point

240 isotherm tests were evaluated for fit against the linearised Freundlich and SR models.

PT
241 All raw and treated samples were filtered through a 0.45-µm polyethersulphone (Pall) laboratory filter in

242 a standard vacuum filtration apparatus to remove particulate matter and TiO2 ahead of DOC, UV254,

RI
243 DBPfp and LC-OCD analysis.

SC
244

U
AN
M
D
TE
C EP
AC

11
ACCEPTED MANUSCRIPT

245 3. Results and discussion

246 3.1 DBP formation during photocatalysis

247 During the irradiation tests, the THMfp and HAAfp of the raw Otonabee River water were modestly

248 reduced by adsorption alone (15% and 10%) but both increased upon irradiation (Figure 1). The impact

PT
249 on THMfp was particularly dramatic. After only 5 min of irradiation, THMfp increased by 61% relative

250 to the control. After 30 min, THMfp began to decrease and after 60 min the THMfp of the treated water

RI
251 matched that of the control. The increase in HAAfp upon irradiation was smaller than that of THMfp and

SC
252 was reversed after 30 min of irradiation. After 60 min of irradiation, HAAfp was reduced by 35% relative

253 to the control.

U
300
AN
250
DBPfp (µg L-1)

200
M

150
THMfp
100 HAAfp
D

50
TE

0
Control 0 5 15 30 45 60
Irradiation Time (min)
EP

254

255 Fig. 1. THMfp and HAAfp of Otonabee River water treated with 0.25 g L−1 and irradiated by
C

256 high intensity UVA-LED light


AC

257 These results, which are similar to those obtained by Liu et al. (2008) and Philippe et al. (2009), illustrate

258 the risk associated with the use of photocatalysis for NOM and DBPfp reduction. Although both THMfp

259 and HAAfp were eventually reduced by the treatment, both increased at treatment times between 0 and 15

260 min. Shorter treatment times are desirable as they minimise the amount of space and energy required at

261 full scale. The modest reductions of THMfp and HAAfp through adsorption, however, suggest an

12
ACCEPTED MANUSCRIPT

262 alternative treatment option—could higher concentrations of TiO2 adsorb a sufficient amount of DBP

263 precursors to provide a viable reduction in overall DBPfp and do certain water quality conditions (e.g.

264 pH) favour the adsorption of DBP precursors to TiO2?

265

PT
266 3.2 NOM removal through adsorption – time series experiments

267 The practical adsorption equilibrium, defined as the point at which the 95% confidence of neighbouring

RI
268 means began to overlap and the slope of the line of mean concentration vs. time could no longer be

SC
269 distinguished from zero, was determined on the basis of the results of the DOC and UV254 time series

270 experiments conducted in each water source. Irrespective of the water type used or the parameter

U
271 observed, the results indicated that the majority of NOM adsorption to the P25 TiO2 particles occurred

272 within minutes and that a practical adsorption equilibrium was reached within 1 h (see Figures S1–S3 in
AN
273 the supplemental file). This is similar to results obtained by some researchers working with TiO2

274 materials, including P25 nanoparticles (Kim and Shon, 2007; Ng et al., 2014), though others have
M

275 suggested that a longer period of time might be required to reach full equilibrium (Mwaanga et al., 2014;
D

276 Erhayem, 2013).


TE

277 On the basis of the results of the time series experiments conducted in this study, all subsequent

278 equilibrium experiments were conducted with a 4-h adsorption period to ensure that all data were
EP

279 gathered at a point well beyond the practical point of equilibrium. The results of the time series

280 experiments also indicated that adsorption was most effective on pH 4, and that UV254 and SUVA were
C

281 more strongly affected by the treatment than DOC, hinting that aromatic NOM may have been
AC

282 preferentially adsorbed by the nanoparticles. The latter effect was more apparent in the Otonabee River

283 water because it had an initial raw water SUVA that was two times that of the Lake Ontario water.

284

285 3.3 Effects of pH and TiO2 dose on adsorption

286 Bottle point isotherm tests were conducted to further characterise the adsorption behaviour of the

287 nanoparticles at the three pH conditions as the dose of TiO2 was varied from 0.01 to 1 g L−1. As shown in

13
ACCEPTED MANUSCRIPT

288 Figure 2 and Figure S4 in the supplemental file, at equilibrium, the DOC and UV254 removals observed

289 in the Otonabee River and Lake Ontario samples were found to be pH dependent and in all cases, more

290 NOM was removed at pH 4 than at pH 6 and pH 8. Irrespective of pH, increasing the dose of TiO2 added

291 to the water resulted in a decrease in the amount of DOC remaining in the treated water. In both the

PT
292 Otonabee River (Figure 2A) and Lake Ontario water trials (Figure 2B), DOC removal increased quickly

293 as the TiO2 dose was increased from 0.01 to 0.25 g L−1, but slowed thereafter, though no definitive

RI
294 plateau was reached at any pH, suggesting that further increases in TiO2 dose beyond the maximum

applied in this study (1 g L−1) may have improved DOC removal even further. At pH 4 and 1 g L−1 of

SC
295

296 TiO2, the DOC of the Otonabee River water was reduced from 4.69 ± 0.12 to 1.10 ± 0.12 mg L−1 whereas

U
297 that of the Lake Ontario water was reduced from 1.64 ± 0.05 to 0.69 ± 0.04 mg L−1. DOC removal from
AN
298 both water sources was statistically significantly lower at pH 6 and pH 8 than at pH 4.

299
M
D
TE
C EP
AC

14
ACCEPTED MANUSCRIPT

6
A
5
DOC (mg L-1)

3 pH 4

PT
pH 6
2
pH 8

RI
1

0
0.0 0.2 0.4 0.6 0.8 1.0 1.2

SC
-1
TiO2 Dose (g L )
300

U
2.0
B
1.8
AN
1.6
1.4
DOC (mg L-1)

1.2
M

1.0 pH 4
0.8 pH 6
0.6
D

pH 8
0.4
0.2
TE

0.0
0 0.2 0.4 0.6 0.8 1
TiO2 Dose (g L-1)
EP

301

302 Fig. 2. Adsorption of DOC from raw unchlorinated water from Otonabee River water (A) and
C

303 Lake Ontario water (B) adjusted to pH 4, pH 6 and pH 8 and mixed with 0.5 g L−1 of P25
AC

304 TiO2 nanoparticles for 4 h

305 The UV254 results followed similar trends as DOC and are shown in Figure 4. UV254 was removed

306 more effectively than DOC, particularly from the Otonabee River water, where UV254 was reduced from

307 0.112 ± 0.002 to 0.013 ± 0.002 cm−1, approximately 88%, by a 1 g L−1 dose of P25 nanoparticles at pH 4.

308 UV254 is a measure of the aromaticity of the NOM present in water; hence, the results presented here

15
ACCEPTED MANUSCRIPT

309 suggest that aromatic NOM was preferentially removed over non-aromatic NOM during the adsorption

310 process.

311 As has been suggested by other researchers (Mwaanga et al., 2014), this pH dependence may be partially

312 explained through charge interactions. Zeta potential measurements conducted on the P25 nanoparticles

PT
313 indicated that they had an isoelectric point (IEP) between pH 6 and pH 6.5, consistent with the literature

314 (Kosmulski, 2009). Thus, at pH 4, they were positively charged; at pH 8, they were negatively charged;

RI
315 and at pH 6, they were approximately neutral. At pH 4, most NOM compounds would have been neutral

SC
316 or slightly negatively charged and both charge and hydrophobic interactions likely contributed to

317 adsorption. At pH 6, hydrophobic interactions between NOM compounds were likely the main

U
318 contributors to adsorption. At pH 8, both the nanoparticles and the NOM were negatively charged and

319 thus charge interactions would result in repulsion, rather than attraction, and any adsorption that occurred
AN
320 would have been attributable to hydrophobic interactions.

321 Nanoparticle agglomeration and its effects on surface area may also have contributed to the changes in
M

322 adsorption efficiency observed at different pHs. Agglomeration is most likely to occur when the pH is
D

323 near the isoelectric point/point of zero charge of the material in question because at this pH repulsive
TE

324 forces between individual particles are at a minimum (Liu et al., 2013). P25 has an IEP of approximately

325 6.5, which means that the nanoparticles were more likely to agglomerate at pH 6 than at pH 4 or pH 8.
EP

326 Indeed, the particle size distributions presented in Figure 3 indicate that the agglomerates formed by the

327 nanoparticles were larger in Otonabee River water adjusted to pH 6 than in Otonabee River water
C

328 adjusted to pH 4 or pH 8. In this study, better adsorption was observed at pH 4 than at pH 6.


AC

329 Agglomeration and the subsequent reduction in available surface area at pH 6 versus at pH 4 may have

330 contributed to the poorer adsorption observed at pH 6, whereas charge repulsion between NOM and the

331 TiO2 nanoparticles was more of a driver at pH 8.

16
ACCEPTED MANUSCRIPT

5
Percent

PT
pH 4
3 pH 6
2 pH 8

RI
1

SC
0
0 5 10 15 20
Diameter (µm)

U
332

333 Fig. 3. Size distribution of agglomerates of P25 nanoparticles in Otonabee River water adjusted
AN
334 to pH 4, pH 6 and pH 8

335
M

336 3.4 Modelling of adsorption isotherms


D

337 The DOC results of the bottle point isotherm tests were modelled using the linearised forms of the
TE

338 Freundlich and SR models, as shown in Figure 4 (SR model) and Figure S5 of the supplement

339 (Freundlich model). The isotherm parameters are summarised in Table 1 and Table S2 in the
EP

340 supplemental file.

341
C
AC

17
ACCEPTED MANUSCRIPT

342 Table 1 Summary of isotherm parameters for the adsorption of NOM from Otonabee River water

343 onto P25 TiO2 nanoparticles at pH 4, pH 6 and pH 8

Parameter Otonabee River Lake Ontario

pH 4 pH 6 pH 8 pH 4 pH 6 pH 8

PT
DOC

1/nSR 0.37 ± 0.03a 0.49 ± 0.04 0.47 ± 0.06 0.58 ± 0.05 0.57 ± 0.07 0.59 ± 0.10

RI
KSR (mg DOC/g TiO2)1−1/n 3.7 ± 1.1 1.3 ± 1.2 1.0 ± 1.2 1.5 ± 1.1 0.8 ± 1.2 0.5 ± 1.3

SC
R2 0.98 0.97 0.94 0.98 0.95 0.90

THMfp

U
1/nSR 0.39 ± 0.09 0.42 ± 0.10 0.42 ± 0.20 -b - -
AN
KSR (µg THMfp/g TiO2)1−1/n 27 ± 2 13 ± 2 8±4 - - -

R2 0.91 0.90 0.71 - - -


M

HAAfp

1/nSR 0.39 ± 0.15 0.48 ± 0.22 0.69 ± 0.29 - - -


D

KSR (µg HAAfp/g TiO2)1−1/n 8±2 4±3 1±5 - - -


TE

R2 0.81 0.76 0.76 - - -

344
EP

a
345 Error values represent half of the 95% confidence interval for each parameter.
b
346 Because the concentrations of THMs and HAAs detected in the raw Lake Ontario water after
C

347 chlorination at UFC conditions were near the analytical detection limit, THMfp and HAAfp isotherm
AC

348 experiments were not conducted with this water source.

349

350 At all pHs and in both water sources, the SR model was a better fit to the data, defined as a higher R2

351 value, than the Freundlich model. The SR model is generally thought to provide a more accurate fit for

18
ACCEPTED MANUSCRIPT

352 data from highly heterodisperse systems and when the isotherms are developed using variable doses of

353 adsorbent; hence, this result was not surprising.

354 Other researchers have observed that aromatic NOM and humic acids are preferentially adsorbed to P25

355 nanoparticles over other types of NOM (Erhayem and Sohn, 2014). Given that the two water sources

PT
356 differ mainly in terms of their NOM concentration and aromaticity, it is not surprising that the KSR values

357 obtained from the Lake Ontario tests were lower than those obtained from the Otonabee River tests. The

RI
358 higher 1/nSR values in the Lake Ontario tests also indicate that adsorption was less favourable in this water

SC
359 than in the Otonabee River water.

360 Within each water source, 1/nSR was nearly constant irrespective of pH and KSR was larger at pH 4 than at

U
361 pH 6 and pH 8, which once again indicates that the adsorption of NOM to P25 TiO2 nanoparticles was

362 more effective at pH 4 than at pH 6 or pH 8 in both water sources. The pH 6 and pH 8 confidence
AN
363 intervals for KSR overlapped in both water sources, perhaps suggesting that pH became less of a driver of

364 adsorption capacity when the pH of the water was equal to or higher than the IEP. These findings are in
M

365 agreement with those of other researchers (Mwaanga et al., 2014; Erhayem and Sohn, 2014; Sun and Lee,
D

366 2012), which have demonstrated that NOM adsorption to P25 TiO2 nanomaterials occurs more readily at
TE

367 low pH than at high pH. They are also in agreement with studies that have demonstrated that NOM

368 adsorption to TiO2 can be modelled using the Freundlich (Wiszniowski et al., 2002) and SR models
EP

369 (Erhayem and Sohn, 2014), though it should be noted that all the aforementioned studies made use of

370 standardised NOM or humic acid isolates (e.g. IHSS) in synthetic water matrices rather than natural water
C

371 sources.
AC

372 The KSR and 1/nSR results of this study and those of other TiO2 researchers are lower than those achieved

373 by other groups working with activated carbon and nanoscale carbon adsorbents, but not dramatically so.

374 In their original study, which was conducted with four NOM isolates and GAC doses similar to the TiO2

375 doses used in this study, Summers and Roberts (1988) observed KSR values ranging from 4.22 to 11.4 (mg

376 C/g GAC)1-1/n and 1/nSR values ranging from 0.254 to 0.347. Karanfil et al. (1999) evaluated the

377 adsorption of commercially available NOM isolates and NOM from natural water onto a series of

19
ACCEPTED MANUSCRIPT

378 commercially available and modified activated carbons. They observed KSR values ranging from 1.754 to

379 10.695 (mg C/g GAC)1-1/n in the natural water matrices. Hyung and Kim calculated KSR values ranging

380 from 5.471 to 13.088 (mg C/g MWNT)1-1/n and 1/nSR values ranging from 0.212 to 0.384 when they

381 evaluated the adsorption of commercially available NOM isolates onto multi-walled carbon nanotubes.

PT
100 A

RI
qe (mg DOC/g TiO2)

pH 4

SC
10 pH 6
pH 8

U
AN
1
1 10 100 1000

Ce/D (mg DOC/g TiO2)


M

382

100 B
D
qe (mg DOC/gTiO2)

TE

10
pH 4
pH 6
EP

pH 8
1
C

0
AC

0 1 10 100
Ce/D (mg DOC/g TiO2)
383

384 Fig.4. DOC data from Otonabee River water tests (A) and Lake Ontario water tests (B) fitted to

385 the Summers and Roberts (SR) model

386

387 3.5 Adsorption of DBP precursors

20
ACCEPTED MANUSCRIPT

388 Negligible amounts of THMs and HAAs were formed in the raw Lake Ontario water when it was

389 chlorinated according to the UFC method; hence, the adsorption of DBP precursors to the P25

390 nanoparticles was not explored for this water source. The raw water and treated samples prepared during

391 the Otonabee River adsorption tests were analysed for THMfp and HAAfp, and the results are shown in

PT
392 Figure 5.

393 As observed with DOC and UV254 removal, THMfp reduction through adsorption was pH dependent.

RI
394 Maximum THMfp reduction, 147 ± 16 µg L−1 to 38 ± 16 µg L−1 (74% reduction), was achieved at pH 4

SC
395 and a TiO2 dose of 1 g L−1. Less removal was achieved at this dose at pH 6 (153 ± 11 µg L−1 to 75 ± 11

396 µg L−1, 51% reduction) and pH 8 (154 ± 15 µg L−1 to 104 ± 15 µg L−1, 34% reduction).

U
397 HAA precursors were also removed through adsorption and this removal was pH dependent, though less
AN
398 so than for THM precursors. Figure 5B shows that at the highest concentration of TiO2 (1 g L−1) HAAfp

399 was reduced from 39 ± 7 µg L−1 to 19 ± 5 µg L−1 (50% reduction) at pH 4 and from 36 ± 2 µg L−1 to 23 ±
M

400 3 µg L−1 (40% reduction) at pH 6. HAAfp reduction at pH 8 was not statistically significant at the 95%

401 confidence level. Some of the variability in the HAA results can be explained by the fact that the UFC
D

402 test, which was used to evaluate the THMfp and HAAfp of the raw and TiO2-treated samples in this
TE

403 study, is conducted at pH 8, which does not favour the formation of HAAs. As a result, all the raw and

404 TiO2-treated samples had low HAAfp, making it difficult to isolate the effects of TiO2 adsorption on
EP

405 HAAfp removal, particularly at pH 8.


C
AC

21
ACCEPTED MANUSCRIPT

A
200
175
150
THMfp (µg L-1)

125 pH 4
100

PT
pH 6
75 pH 8
50

RI
25
0
0 0.25 0.5 0.75 1

SC
TiO2 Dose (g L-1)
406

U
50
B
AN
40
HAAfp (µg L-1)

30
M

pH 4
20 pH 6
pH 8
10
D

0
TE

0 0.25 0.5 0.75 1


Dose TiO2 (g L-1)
407
EP

408 Fig. 5. THMfp (A) and HAAfp (B) of Otonabee River water treated with increasing

409 concentrations of P25 TiO2 nanoparticles at pH 4, pH 6 and pH 8


C

410 The agreement between the THMfp and HAAfp datasets and the SR model are shown in Figure S6, and
AC

411 the isotherm parameters are summarised in Table 1. Although the R2 values of the THMfp and HAAfp

412 isotherms were lower than those of the DOC isotherms, the general trends indicate that, with the

413 exception of HAAfp at pH 8, TiO2 could remove significant amounts of THM and HAA precursors from

414 Otonabee River water through adsorption and that this adsorption could be modelled using the SR

415 isotherm model. Nonetheless, as a whole, the isotherm parameters for the two classes of DBPs should be

22
ACCEPTED MANUSCRIPT

416 approached with caution because they were developed using a small dataset that contained substantial

417 variation at low TiO2 doses. Additional experiments at higher TiO2 doses, using water sources with higher

418 concentrations of DBP precursors, and/or employing chlorination regimes more likely to result in THM

419 and HAA formation may help clarify the how well the SR model can predict the removal of DBP

PT
420 precursors from drinking water by TiO2 as well as the suitability of the model at different pHs.

421

RI
U SC
AN
M
D
TE
C EP
AC

23
ACCEPTED MANUSCRIPT

422 3.5 Effects of pH on adsorption of LC-OCD fractions

423 A selection of raw and TiO2-treated water samples from the Otonabee River experiment was analysed

424 using LC-OCD to determine whether any specific fractions were being removed during adsorption and

425 whether pH impacted the fractions adsorbed. As shown in Figure 6, the biopolymers and humic

PT
426 substances fractions were targeted for adsorption at all pHs but most effectively removed at pH 4. The

427 building blocks fraction was also removed to some degree at pH 4 but was essentially unaffected at pH 6

RI
428 and pH 8. The LMWA and LMWN fractions were not adsorbed at any pH. These results indicate that,

SC
429 consistent with the findings of other researchers (Erhayem and Sohn, 2014), large and aromatic NOM

430 compounds were preferentially adsorbed by TiO2 nanoparticles and help explain why U254, which is

U
431 associated with the humic substances fraction, was sometimes removed more effectively than overall

432 DOC.
AN
5
M

4
D

LMWN
DOC (mg L-1)

3
LMWA
TE

Building Blocks
2 Humic Substances
Biopolymers
EP

1
C

0
Raw Water pH 4 pH 6 pH 8
AC

433

434 Fig. 6. LC-OCD fractions present in raw unchlorinated Otonabee River water and water adjusted

435 to pH 4, pH 6 and pH 8 and mixed with 0.5 g L−1 of P25 TiO2 nanoparticles for 4 h

436

437

438 4. Conclusions
24
ACCEPTED MANUSCRIPT

439 The results of this study show that during adsorption aromatic NOM (as measured by UV254) is

440 preferentially removed over non-aromatic NOM and that the efficiency of NOM adsorption to TiO2 can

441 vary by water source. They also demonstrate that TiO2 nanoparticles preferentially adsorb larger NOM

442 molecules including the biopolymers and humic substances fractions. pH was shown to have a strong

PT
443 impact on the removal of NOM, including DBP precursors, from surface water by TiO2 nanoparticles.

444 Specifically, more adsorption occurred at low pH than at higher pH. The poorer adsorption observed at

RI
445 pH 6 and pH 8 may be related to both agglomeration and charge repulsion at higher pH, with the former

SC
446 dominating at pH 6 and the latter at pH 8.

447 A modified version of the Freundlich isotherm model (SR model) was found to provide an excellent fit to

U
448 the DOC data gathered in this study. The resulting isotherm parameters were within but at the low end of

449 the range usually observed during NOM adsorption to GAC and carbon nanomaterials, indicating that,
AN
450 particularly at neutral pH, the TiO2 nanoparticles were less effective than the adsorbents currently used in

451 drinking water plants. Unlike TiO2, however, GAC is generally expensive and energy intensive to
M

452 regenerate and the regeneration must usually be conducted offsite, whereas TiO2 is potentially
D

453 regenerable and reusable in place. The THMfp and HAAfp datasets were also fitted to the SR model, with
TE

454 generally positive results. The results presented in this paper show that TiO2 adsorption is a viable way to

455 remove NOM and DBP precursors from drinking water and that this removal can be modelled using
EP

456 simple isotherm models. The results also suggest that researchers hoping to design adsorption-based TiO2

457 processes should keep in mind that pH adjustment might be required to optimise performance.
C

458
AC

459 Acknowledgements

460 The authors would like to acknowledge the training provided by Jim Wang, the laboratory assistance

461 provided by Yijun (Jessie) Gai and Michelli Park, and the support of the Drinking Water Research Group

462 at the University of Toronto. The authors are also grateful to Dr. Monica Tudorancea and Dr. Sigrid

463 Peldzsus (University of Waterloo) for performing LC-OCD analyses. Funding was provided by the

25
ACCEPTED MANUSCRIPT

464 National Science and Engineering Research Council of Canada and the Ontario Ministry of Training,

465 Colleges and Universities.

466

PT
RI
U SC
AN
M
D
TE
C EP
AC

26
ACCEPTED MANUSCRIPT

467 References

468 American Public Health Association, 2005. Standard Methods for the Examination of Water and

469 Wastewater, 21st ed., Washington D.C., APHA

470 Avisar, D., Horovitz, I., Lozzi, L., Ruggieri, F., Baker, M., Abel, M-L, Mamane, H., 2013. Impact of

PT
471 water quality on removal of carbamazepine in natural waters by N-doped TiO2 photo-catalytic thin film

472 surfaces, Journal of Hazardous Materials, 244-245, 463-471

RI
473 Chowdhury, Z.K., Summers, R.S., Westerhoff, G.P., Leto, B.J., Nowack, K.O., Corwin, C.J., 2013.

SC
474 Activated Carbon: Solutions for Improving Water Quality, Passantino, L.B. (Ed.), Denver, USA,

475 American Water Works Association

U
476 Erhayem, M., 2013. Effect of naturally occurring organic matter (NOOM) type and source on NOOM
AN
477 adsorption onto titanium dioxide nanoparticles under varying environmental conditions, Thesis, Florida

478 Institute of Technology, USA


M

479 Erhayem, M. and Sohn, M., 2014. Stability studies for titanium dioxide nanoparticles upon adsorption of

480 Suwannee River humic and fulvic acids and natural organic matter, Science of the Total Environment,
D

481 468-469, pp. 249-257


TE

482 Fotiou, T., Triantis, T.M., Kaloudis, T., Hiskia, A., 2015. Evaluation of the photocatalytic activity of TiO2

483 based catalysts for the degradation and mineralization of cyanobacterial toxins and water off-odor
EP

484 compounds under UV-A, solar, and visible light, Chemical Engineering Journal, 261, 17-26

485 Huang, X., Leal, M., Li, Q., 2008. Degradation of natural organic matter by TiO2 photocatalytic oxidation
C

486 and its effect on fouling of low-pressure membranes, Water Research, 42, 1142-1150
AC

487 Huber, S.A., Balz, A., Abert, M., Pronk, W., 2011. Characterisation of aquatic humic and non-humic

488 matter with size-exclusion chromatography-organic carbon detection and organic nitrogen detection (LC-

489 OCD-OND), Water Research, 45, 879-888

490 Hyung, H., Kim, J-H, 2008. Natural organic matter (NOM) adsorption to multi-walled carbon nanotubes:

491 Effect of NOM characteristics and water quality parameters, Environmental Science and Technology, 42,

492 4416-4421

27
ACCEPTED MANUSCRIPT

493 International Organization for Standardization, 2010. Nanotechnologies – Methodology for the

494 classification and categorization of nanomaterials, ISO/TR 11360:2010(E)

495 Kanakaraju, D., Glass, B.D., Oelgemoller, M., 2014. Titanium dioxide photocatalysis for pharmaceutical

496 wastewater treatment, Environmental Chemistry Letters, 12, 27-47

PT
497 Karanfil, T., Kitis, M., Kilduff, J.E., Wigton, A., 1999. Role of granular activated carbon surface

498 chemistry on the adsorption of organic compounds 2, Environmental Science and Technology, 33, 3225-

RI
499 3233

SC
500 Li, F., Yuasa, A., Ebie, K., Azuma, Y., Hagishita, T., Matsui, Y., 2002. Factors affecting the adsorption

501 capacity of dissolved organic matter onto activated carbon: Modified isotherm analysis, Water Research,

U
502 36, 4994-4604

503 Liu, S., Lim, M., Fabris, R., Chow, C., Drikas, M., Amal, R., 2008. TiO2 photocatalysis of natural organic
AN
504 matter in surface water: Impact on trihalomethane and haloacetic acid formation potential, Environmental

505 Science and Technology, 42, 6218-6223


M

506 Liu, S., Lim, M., and Amal, R. 2014. TiO2-coated natural zeolite: Rapid humic acid adsorption and
D

507 effective photocatalytic regeneration, Chemical Engineering Science, 105, 46-52


TE

508 Liu, W., Sun, W., Borthwick, A., and Ni, J., 2013. Comparison on aggregation and sedimentation of

509 titanium dioxide titanate nanotubes and titanate nanotubes-TiO2: Influence of pH, ionic strength, and
EP

510 natural organic matter, Colloids and Surfaces A: Physicochemical Engineering Aspects, 434, 319-328

511 Loosli, F., Vitorazi, L., Berret, J-F, and Stoll, S., 2015. Towards a better understanding on agglomeration
C

512 mechanisms and thermodynamic properties of TiO2 nanoparticles interacting with natural organic matter,
AC

513 Water Research, 80, 139-148

514 Kim, S-H and Shon, H.K., 2007. Adsorption characterization for multi-component organic matters by

515 titanium oxide (TiO2) in wastewater, Separation Science and Technology, 42, 1775-1792

516 Kosmulski, M., 2009. Compilation of PZC and IEP of sparingly soluble metal oxides and hydroxides

517 from literature, Advances in Colloid and Interface Science, 152, 14-25

28
ACCEPTED MANUSCRIPT

518 Malato, S., Fernandez-Ibanez, P., Maldonado, M.I., Blanco, J., Gernjak, W., 2009. Decontamination and

519 disinfection of water by solar photocatalysis: Recent overview and trends, Catalysis Today, 147, 1-59

520 Mwaanga, P., Carraway, E.R., Schlautman, M.A., 2014. Preferential sorption of some natural organic

521 matter fractions to titanium dioxide nanoparticles: influence of pH and ionic strength, Environmental

PT
522 Monitoring and Assessment, 186, 8833-8844

523 Nanotechnology Characterization Laboratory, 2009. NCL Method PCC-2: Measuring Zeta Potential of

RI
524 Nanomaterials, National Cancer Institute, U.S. National Institutes of Health

SC
525 Ng, M., Kho, E.T., Liu, S., Lim, M., Amal, R., 2014. Highly adsorptive and regenerative magnetic TiO2

526 for natural organic matter (NOM) removal in water, Chemical Engineering Journal, 246, 196-203

U
527 Philippe, K.K. Hans, C. MacAdam, J., Jefferson, B., Hart, J., and Parsons, S.A., 2010. Photocatalytic

528 oxidation, GAC, and biotreatment combinations: An alternative for the coagulation of hydrophilic rich
AN
529 waters?, Environmental Technology, 31, 1423-1434

530 Qi, S. Schideman, L.C., 2008. An overall isotherm for activated carbon adsorption of dissolved organic
M

531 matter in water, Water Research, 42, 3353-3360


D

532 Shahbeig, H., Bagheri, N., Ghorbanian, S., Hallajisani, A., Poorkarimi, S., 2013. A new adsorption
TE

533 isotherm model of aqueous solutions on granular activated carbon, World Journal of Modelling and

534 Simulation, 9, 243-254


EP

535 Shi, H., Magaye, R., Castranova, V., and Zhao, J., 2013. Titanium dioxide nanoparticles: A review of

536 current toxicological data, Particle and Fibre Toxicology, 10:15


C

537 Summers, R. and Roberts, P., 1988. Activated Carbon Adsorption of Humic Substances: Heterodisperse
AC

538 Mixtures and Desorption, Journal of Colloid and Interface Science, 122, 367-381

539 Summers, R.S., Hooper, S.M., Shukairy, H.M., Solarik, G., Owen, D., 1996. Assessing DBP yield:

540 Uniform formation conditions, Journal of the American Water Works Association, 88, 80-93

541 Sun, D.D. and Lee, P.F., 2012. TiO2 microsphere for the removal of humic acid from water: Complex

542 adsorption mechanisms, Separation and Purification Technology, 91, 30-37

29
ACCEPTED MANUSCRIPT

543 Wassink, J.D., Andrews, R.C., Peiris, R.H., Legge, R.L., 2011. Evaluation of fluorescence excitation-

544 emission and LC-OCD as methods of detecting removal of NOM and DBP precursors by enhanced

545 coagulation, Water Science and Technology: Water Supply, 11, 621

546 Wiszniowski, J., Robert, D., Surmacz-Gorska, J., Miksch, K., and Weber, J-V, 2002. Photocatalytic

PT
547 decomposition of humic acids on TiO2, Part I: Discussion of adsorption and mechanism, Journal of

548 Photochemistry and Photobiology A: Chemistry, 153, 267-273

RI
549 Wray, H.E., Andrews, R.C., Bérubé, P.R., 2013. Surface shear stress and membrane fouling when

SC
550 considering natural water matrices, Desalination, 330, 22-27

551 Yang, Y. and Westerhoff, P., 2014. Presence in, and Release of, Nanomaterials from Consumer Products,

U
552 Nanomaterials, Advances in Experimental Medicine and Biology, 811 (Capco, D.G. and Chen, Y.,

553 editors), Springer Science + Business Media, Berlin


AN
554 Zhang, Y., Chen, Y., Westerhoff, P., and Crittenden, J., 2009. Impact of natural organic matter and

555 divalent cations on the stability of aqueous nanoparticles, Water Research, 43, 4249-4257
M

556
D
TE
C EP
AC

30
ACCEPTED MANUSCRIPT

557 Figure Captions

558 Fig. 1. THMfp and HAAfp of Otonabee River water treated with 0.25 g L−1 and irradiated by

559 high intensity UVA-LED light

560 Fig. 2. Adsorption of DOC from raw unchlorinated water from Otonabee River water (A) and

PT
561 Lake Ontario water (B) adjusted to pH 4, pH 6 and pH 8 and mixed with 0.5 g L−1 of P25

562 TiO2 nanoparticles for 4 h

RI
563 Fig. 3. Size distribution of agglomerates of P25 nanoparticles in Otonabee River water adjusted

SC
564 to pH 4, pH 6 and pH 8

565 Fig.4. DOC data from Otonabee River water tests (A) and Lake Ontario water tests (B) fitted to

U
566 the Summers and Roberts (SR) model
AN
567 Fig. 5. THMfp (A) and HAAfp (B) of Otonabee River water treated with increasing

568 concentrations of P25 TiO2 nanoparticles at pH 4, pH 6 and pH 8


M

569 Fig. 6. LC-OCD fractions present in raw unchlorinated Otonabee River water and water adjusted

570 to pH 4, pH 6 and pH 8 and mixed with 0.5 g L−1 of P25 TiO2 nanoparticles for 4 h
D
TE
C EP
AC

31
ACCEPTED MANUSCRIPT

Highlights

• TiO2 photocatalysis with UVA light increased the DBPfp of a natural surface water at short

treatment times but reduced it at longer treatment times

• In the absence of irradiation TiO2 nanoparticles adsorbed NOM and DBP precursors from natural

PT
surface water

RI
The adsorption of NOM, THM precursors, and HAA precursors by TiO2 nanoparticles was

modeled using a simple isotherm model

SC
• The adsorption of DOC, UV254, humic substances, and DBP precursors by TiO2 nanoparticles

occurred more readily at pH 4 than at pH 6 or pH 8

U
AN
M
D
TE
C EP
AC

You might also like