Subscriber access provided by UNIV OF DURHAM
Environmental Processes
Nano-TiO2-Catalyzed Dehydrochlorination of 1,1,2,2Tetrachloroethane: Roles of Crystalline Phase and Exposed Facets Xuguang Li, Tong Li, Tong Zhang, Cheng Gu, Shourong Zheng, Haijun Zhang, and Wei Chen Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05479 • Publication Date (Web): 19 Mar 2018 Downloaded from http://pubs.acs.org on March 19, 2018
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 30
Environmental Science & Technology
1
Nano-TiO2-Catalyzed Dehydrochlorination of
2
1,1,2,2-Tetrachloroethane: Roles of Crystalline Phase and
3
Exposed Facets
4 5
Xuguang Li,† Tong Li,† Tong Zhang,† Cheng Gu,‡ Shourong Zheng,‡ Haijun Zhang,§ Wei Chen*,†
6 7
†
College of Environmental Science and Engineering, Ministry of Education Key Laboratory of
8
Pollution Processes and Environmental Criteria, Tianjin Key Laboratory of Environmental
9
Remediation and Pollution Control, Nankai University, Tianjin 300350, China
10
‡
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment,
11 12
Nanjing University, Nanjing, Jiangsu 210023, China §
School of Physics and Materials Science, Anhui University, Hefei, Anhui 230039, China
13 14
Manuscript prepared for Environmental Science & Technology
15 16
* Corresponding author: (Phone/fax) 86-22-85358169; (E-mail)
[email protected].
17
1
ACS Paragon Plus Environment
Environmental Science & Technology
18
TOC Art
19 20
2
ACS Paragon Plus Environment
Page 2 of 30
Page 3 of 30
21 22
Environmental Science & Technology
ABSTRACT Nano-scale titanium dioxide (nTiO2) is one of the most widely used metal oxide
23
nanomaterials. Once released into the environment, nTiO2 may catalyze abiotic transformation of
24
contaminants, and consequently affect their fate and effects. Here, we show that the overall
25
catalytic efficiency of nTiO2 for the dehydrochlorination reaction of 1,1,2,2-tetrachloroethane, a
26
commonly used solvent, depends on the crystalline phase and exposed facets of nTiO2, which
27
significantly affect the adsorption capacity and surface catalytic activity of nTiO2. Specifically,
28
under all three pH conditions tested (7.0, 7.5 and 8.0), the overall catalytic efficiency of eight
29
nTiO2 materials (as indicated by the surface-area-normalized reaction kinetic constants) followed
30
the order of rutile > anatase > TiO2(B). For anatase and TiO2(B) materials, the overall catalytic
31
efficiency increased with the increasing percentage of exposed {001} and {010} facets,
32
respectively. Crystalline phase and exposed facets significantly affected adsorption affinities of
33
nTiO2, likely by modulating surface hydrophobicity of nTiO2. Crystalline phase and exposed
34
facets also determined the activity of surface catalytic sites on nTiO2 by dictating the
35
concentration and strength of surface unsaturated Ti atoms, as the deprotonated hydroxyl groups
36
chemisorbed to these reactive Ti atoms served as bases to catalyze the base-promoted reaction.
37
3
ACS Paragon Plus Environment
Environmental Science & Technology
38 39
INTRODUCTION Nano-scale titanium dioxide (nTiO2) is one of the most widely used metal oxide
40
nanomaterials.1 For instance, anatase and rutile TiO2 are essential ingredients of paints,
41
cosmetics and food additives.1 TiO2(B) has growing applications in the fields of energy
42
conversion, such as solar cells, lithium batteries, and environmental photocatalysts.2,3 It was
43
estimated that the annual global production of TiO2 would reach approximately 2.5 million
44
metric tons by 2025.4 With the rapidly increasing production and use of nTiO2, its environmental
45
release and the associated implications (e.g., toxicity to aquatic organisms and effects on
46
bioavailability of environmental contaminants) have received much attention.5-9 One of the
47
potential environmental implications of nTiO2 is that once released into the environment, these
48
materials may affect the fate and effects of environmental contaminants by catalyzing their
49
transformation reactions, which consequently modifies the physicochemical properties and
50
biological effects of the contaminants. While numerous studies have been conducted to
51
understand the photocatalytic properties of nTiO2 (as photocatalysts for water and wastewater
52
treatments),10-12 the catalytic effects of nTiO2 on abiotic reactions that are more relevant to the
53
natural environments, such as hydrolysis, are not well understood.
54
A limited number of studies have shown that TiO2, like many other metal oxides, can
55
catalyze the hydrolysis reactions of certain organic contaminants (e.g., organophosphorus
56
pesticides and carboxylate esters),13-17 one of the most important types of abiotic transformation
57
reactions controlling the environmental fate of these contaminants.18,19 Even though
58
nTiO2-catalyzed hydrolysis may occur relatively more slowly compared to photocatalysis, this
59
process constantly occurs and is not limited by time (e.g., daytime) or location (e.g., photic zone)
60
and thus may dominate the degradation of organic contaminants in natural aquatic systems. It has
4
ACS Paragon Plus Environment
Page 4 of 30
Page 5 of 30
Environmental Science & Technology
61
been proposed that the metal atoms on the surface of metal oxides are likely reactive sites that
62
catalyze hydrolysis reactions.13,15,16,20 For instance, metal atoms can coordinate with organic
63
compounds or leaving groups and enhance the hydrolysis reactions.15,16 Metal atoms can also
64
induce deprotonation of metal-coordinated water molecules, generating metal-hydroxo species
65
that can serve as nucleophiles to catalyze hydrolysis reactions.15,16 Considering that the surface
66
of TiO2 contains both saturated six-coordinated Ti (Ti6c) and unsaturated five-coordinated Ti
67
(Ti5c) atoms,21 it is probably reasonable to assume that the unsaturated (and thus more reactive)
68
Ti5c atoms play the most important role in TiO2-mediated hydrolysis reactions of organic
69
compounds.
70
Naturally occurring crystalline phases of TiO2 include anatase, rutile, brookite, and
71
TiO2(B),22 and each crystalline phase may possess different percentages of exposed facets.23 For
72
example, anatase TiO2 generally possesses three fundamental low-index facets: {001}, {010}
73
and {101}.24 It has been shown that nTiO2 materials of different crystalline phases and exposed
74
facets differ significantly in photocatalytic reactivity. TiO2 nanomaterials of different crystalline
75
phases possess different energy band structures (i.e., band gap, conduction and valence band
76
positions), which in turn determine the light absorption and redox capacity of photogenerated
77
charge carriers.25 Meanwhile, the exposed facets of TiO2 can also influence the photocatalytic
78
performance, via differences in charge separation efficiency and adsorption of reactant
79
molecules.25-27 Moreover, theoretical calculations also showed that the adsorption energies of
80
organic compounds on nTiO2 materials vary with crystalline phases and exposed facets.28,29
81
Hence, we hypothesize that crystalline phase and exposed facets can also significantly affect the
82
efficiency of nTiO2 materials in catalyzing hydrolysis reactions, likely by influencing both the
83
adsorption affinities of nTiO2 for organic molecules and activity of surface catalytic sites,
5
ACS Paragon Plus Environment
Environmental Science & Technology
84 85
particularly, Ti5c atoms. The objective of this study was to understand how crystalline phase and exposed facets of
86
nTiO2 materials can affect their catalytic effects on environmentally relevant hydrolysis reactions.
87
Eight nTiO2 materials—including three anatase TiO2 materials varying in exposed {001} facets
88
(referred to as anatase_L001, anatase_M001 and anatase_H001), three rutile TiO2 materials
89
varying mainly in particle size (rutile_1, rutile_2 and rutile_3), and two TiO2(B) materials
90
varying in morphology and exposed {010} facets (TiO2(B) nanorods and TiO2(B) nanowires,
91
referred to as TiO2(B)_rod and TiO2(B)_wire)—were selected to represent different variables of
92
the surface structures of nTiO2. The dehydrochlorination reaction of 1,1,2,2-tetrachloroethane
93
(TeCA) was selected as the model hydrolysis reaction,30,31 because TeCA is a widely used
94
chlorinated solvent32 and dehydrohalogenation reaction is one of the most environmentally
95
relevant pathways for the natural attenuation of chlorinated solvents.33 The overall catalytic
96
efficiencies of different nTiO2 materials were examined under three pH conditions. The results
97
were compared to understand how catalytic efficiency was influenced by crystalline phase and
98
exposed facets. Batch adsorption experiments, two-phase kinetic modeling and cation-mediated
99
reaction experiments were carried out to further understand the interactions between TeCA and
100
different nTiO2 materials.
101 102
MATERIALS AND METHODS
103
Materials. Anatase_L001 and the three rutile materials were purchased from Aladdin
104
Biological Technology (Shanghai, China). Anatase_M001, anatase_H001 and the two TiO2(B)
105
materials were synthesized using the hydrothermal methods reported in the literature.34,35 The
106
synthesis procedures are given in Supporting Information (SI).
6
ACS Paragon Plus Environment
Page 6 of 30
Page 7 of 30
Environmental Science & Technology
107
TeCA (> 99%) and trichloroethylene (TCE, > 99%) were purchased from Sigma–Aldrich
108
(St. Louis, USA). Stock solutions of TeCA and TCE were prepared in methanol, and stored at
109
-20 °C. The inorganic salts (KH2PO4, K2HPO4·3H2O, CuCl2, CdCl2 and PbCl2) were obtained
110
from Tianjin Chemical Reagent Co. (Tianjin, China).
111
Characterization of nTiO2 Materials. The shape and morphology of the nTiO2 materials
112
were characterized with scanning electron microscopy (SEM, Nova Nano SEM 230, Hillsboro,
113
USA). The Brunauer−Emmett−Teller (BET) specific surface areas of the nTiO2 materials were
114
determined by multipoint N2 adsorption–desorption using a Micromeritics ASAP2010
115
accelerated surface area and porosimetry system (Micromeritics Co., Norcross, USA). The
116
crystal structures of the nTiO2 materials were determined with X-ray diffraction (XRD) using a
117
D/Max-2500 diffractometer (Rigaku, Tokyo, Japan) with Cu Kα radiation. Raman spectra were
118
recorded with a Renishaw inVia Raman spectrometer (RM2000, London, UK) with an exciting
119
wavelength of 633 nm. Surface elemental compositions of the nTiO2 materials were determined
120
with X-ray photoelectron spectroscopy (XPS) (PHI 5000 VersaProbe, Tokyo, Japan). Fourier
121
transform infrared (FTIR) transmission spectra of the materials were obtained using a 110 Bruker
122
TENSOR 27 apparatus (Bruker Optics Inc., Karlsruhe, Germany). The relative hydrophobicity of
123
the nTiO2 materials was assessed by measuring n-dodecane–water partition coefficients (the
124
detailed sample preparation methods are given in SI)36 and the static water contact angle of the
125
materials. The contact angle measurements were carried out in the air using the sessile drop
126
method on a contact angle system OCA 20 (DataPhysics Instruments GmbH, Stuttgart,
127
Germany). The contact angles reported are mean values measured for 4 µL water droplets at five
128
positions of each material.
129
The types of nTiO2 surface acid sites were examined with FTIR spectra after pyridine
7
ACS Paragon Plus Environment
Environmental Science & Technology
130
adsorption37 using a Tensor 27 FTIR spectrometer (Bruker, Karlsruhe, Germany). The amount of
131
acid sites was estimated from temperature-programmed desorption of NH3 (NH3-TPD)38 carried
132
out using a Chemisorb 2720 (Micromeritics, Norcross, USA).
133
Reaction of TeCA in Homogeneous Aqueous Solution and in nTiO2 Suspensions.
134
Reaction kinetic experiments were conducted both in homogeneous aqueous solution (i.e.,
135
electrolyte solution in the absence of nTiO2 materials) and in nTiO2 suspensions. The electrolyte
136
solution that contained 0.05 M K2HPO4/KH2PO4 as a pH buffer and 0.01 M NaN3 as a
137
bio-inhibitor was added to a series of 40-ml amber glass bottles. The pH of the solution was
138
adjusted to 7.0, 7.5 or 8.0 with HNO3 or NaOH. The reaction matrices that contained nTiO2
139
material and the electrolyte solution were sonicated in a water bath for 30 min and then
140
equilibrated on an orbital shaker at 7 rpm for 24 h. Then, 20 µL of a TeCA stock solution (in
141
methanol) was injected into the solution with a micro-syringe to give an initial TeCA
142
concentration of 2.5 mg/L, and then the bottle was immediately filled with the phosphate buffer
143
solution, sealed, and left on an orbital shaker operated at 7 rpm and a constant temperature of
144
25 °C in the dark. At predetermined time intervals, 1.0 ml of the solution was withdrawn,
145
transferred to a clean glass vial containing a pH-3 buffer solution (0.05 M KH2PO4; pH adjusted
146
with HNO3) to terminate the reaction. The solution was filtrated through a 0.22-µm membrane
147
filter to remove nTiO2 materials, and then extracted with hexane (3:1, v:v). The hexane extract
148
was analyzed to determine the mass of TeCA and TCE (the reaction product) with an Agilent
149
6890N gas chromatography with electron capture detector (Agilent Technologies, Santa Clara,
150
USA) equipped with an HP-5 capillary column (30 m × 0.32 mm × 0.25 µm). The mass of TeCA
151
adsorbed to nTiO2 materials was less than 2% of the total mass (see the next section on
152
adsorption experiments) and thus was not quantified in all samples. Solution pH was also
8
ACS Paragon Plus Environment
Page 8 of 30
Page 9 of 30
Environmental Science & Technology
153
checked during sampling and was found to be essentially unchanged in all the experiments. Each
154
kinetic experiment was run in duplicate. Mass balance was within the range of 92-98% for all the
155
experiments (Figures S1-S4).
156
In the experiments involving divalent cations (Cu2+, Pb2+ and Cd2+), a stock solution of
157
CuCl2, PbCl2, and CdCl2 was added to a 3-morpholinopropanesulfonic acid (MOPS) buffer
158
solution containing nTiO2, to reach a predetermined concentration of Cu2+, Pb2+ and Cd2+. Then,
159
pH of the solution was adjusted to 7.0 with HNO3 or NaOH. All other procedures were identical
160
as those stated above. The effect of cation addition on the aggregation status of nTiO2 was
161
examined by dynamic light scattering (DLS) using a ZetaPALS (Brookhaven Instruments,
162
Holtsville, USA).
163
Adsorption Experiments of TeCA to nTiO2 Materials. Adsorption isotherms of TeCA to
164
the nTiO2 materials were obtained using the procedures reported in the literature.39 Briefly, a 1
165
g/L nTiO2 suspension (pH 5.0) was added to a series of 8-ml amber glass vials (the low pH was
166
used to ensure that abiotic transformation of TeCA would not occur). Next, a stock solution of
167
TeCA (in methanol) was added to the vials, and the vials were filled with the electrolyte solution,
168
capped and placed on an orbital shaker at 7 rpm for 5 d. Then, the suspensions were filtrated
169
through 0.22-µm membrane filters to remove nTiO2 materials, and extracted with hexane (3:1,
170
v:v) to determine the mass of TeCA and TCE. The concentrations of TeCA adsorbed to nTiO2
171
were calculated based on mass balance. All the adsorption experiments were run in triplicate.
172 173 174 175
Data Analysis. The reaction kinetics data of TeCA were fitted with the pseudo-first-order kinetic model: C/C0 = exp (−kobs ⋅ t)
(1)
where C0 (mg/L) is the initial TeCA concentration; C (mg/L) is the TeCA concentration at a 9
ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 30
176
given time t (h); and kobs (h-1) is the observed pseudo-first-order rate constant for the overall
177
TeCA degradation.
178 179 180
A two-phase kinetic model developed in our previous study30 was used to estimate the reaction constant of TeCA adsorbed to the surface of nTiO2 materials: ܥ({ = ܥଵି +
౩ ∙ూ ⋅ొ
) ⋅ ݁ [ ⋅(ିଵ)⋅௧] −
౩ ∙ూ ⋅ొ
భ
}భష
(2)
181
where ka (h−1) and ks (h−1) are the calculated pseudo-first-order rate constants for the dissolved
182
and nTiO2-adsorbed TeCA, respectively; KF (mg1-nLn/g) is the affinity coefficient of the
183
Freundlich isotherm; n (unitless) is the Freundlich linearity index; and CNM (g/L) is the
184
concentration of nTiO2 materials in the reaction system. The values of ka were obtained by fitting
185
the kinetics data in homogeneous aqueous solution (i.e., reaction in the absence of nTiO2
186
materials) with Equation 1.
187 188 189
RESULTS AND DISCUSSION Characteristics of Different nTiO2 Materials. The SEM images of the nTiO2 materials
190
(Figure S5) show the morphology of the materials. Anatase_L001 consisted of primarily
191
nanoparticles with 40-nm diameter, whereas anatase_M001 and anatase_H001 were nanoplates
192
with an average edge length of ~50 nm. The three rutile nTiO2 materials (rutile_1, rutile_2 and
193
rutile_3) were nanoparticles with diameters of 19.8 ± 6.9, 23.7 ± 7.9, and 30.2 ± 7.8 nm,
194
respectively. The average length of TiO2(B)_rod was ~200 nm, whereas that of TiO2(B)_wire
195
was ~2 µm.
196
The XRD patterns and Raman spectra (Figure 1a and 1b) show that all the nTiO2 materials
197
were in single crystal phase. According to the intensity ratios of the XRD peaks, the three
198
anatase TiO2 materials all had the strongest (101) peak (peak at 25.3°) (Figure 1a), indicating 10
ACS Paragon Plus Environment
Page 11 of 30
Environmental Science & Technology
199
preferential orientation growth in the {101} direction (JCPDS 21-1272). Nonetheless, the
200
intensity ratios for the (004) and (101) diffractions of anatase_M001 and anatase_H001 were
201
larger than that of anatase_L001 (Figure 1a, Table S1), indicating that these two materials
202
possessed higher percentage of exposed {001} facets than did anatase_L001.40 The percentages
203
of the {001} facets, estimated using the Raman spectra,41,42 were 7% for anatase_L001, 21% for
204
anatase_M001 and 38% for anatase_H001 (Table S2, Figure 1b). The three rutile TiO2 materials
205
all showed the strongest (110) peak (peak at 27.4°) (Figure 1a), indicating preferential
206
orientation growth in the {110} direction (JCPDS 21-1276). Additionally, the intensities of the
207
other peaks were similar for these three materials (Figure 1a, Table S1), indicating no significant
208
differences in exposed facets. Both TiO2(B) materials showed the strongest (110) peak (peak at
209
24.9°). However, the intensity ratios for the (010) and (110) diffractions were 0.82 for
210
TiO2(B)_rod and 0.79 for TiO2(B)_wire, remarkably larger than the standard value (0.30) in the
211
JCPDS card no. 74-1940 (Figure 1a, Table S1), suggesting that they both possessed a significant
212
amount of exposed {010} facets and TiO2(B)_rod exposed more {010} facets than
213
TiO2(B)_wire.
214
All the nTiO2 materials had comparable surface compositions in the same chemical states as
215
evidenced by the XPS spectra (Figure S6-8). The peaks with the binding energy of 458.9 and
216
464.5 eV were assigned to the Ti 2p3/2 and Ti 2p1/2 spin orbital splitting photoelectrons in the Ti4+
217
chemical state, and the O 1s signal revealed two peaks with binding energy of 530.6 and 532.2
218
eV, attributable to the lattice oxygen and hydroxyl species, respectively.43,44 The presence of
219
hydroxyl groups was also evident from the OH stretching vibrations between 3200 and 3600
220
cm-1 in the FTIR spectra (Figure S9).45
221
The concentrations of surface Ti5c atoms (C_Ti5c) of the nTiO2 materials were estimated by
11
ACS Paragon Plus Environment
Environmental Science & Technology
222
measuring the concentrations of surface Lewis acids (C_acid), since Lewis acid sites on TiO2
223
surface are normally attributed to Ti5c atoms.38 The pyridine adsorbed FTIR spectra (Figure S10a)
224
show that one strong peak at 1445 cm-1 was observed for the three rutile materials, whereas two
225
strong peaks—at 1445 and 1605 cm-1—were observed for the other five materials. It was
226
reported that the two peaks at 1445 cm-1 and 1605 cm-1 of pyridine-adsorbed FTIR spectra
227
correspond to the coordinated pyridine adsorbed on Lewis acid sites on nTiO2 materials.46 The IR
228
absorbance at 1545 cm-1, characteristic of pyridine chemisorbed at Brønsted acid sites, was not
229
observed, indicating the absence of Brønsted acid sites on the surfaces of the nTiO2 materials.46
230
Due to the absence of Brønsted acid sites on nTiO2 materials, the concentration of the Lewis acid
231
sites could be estimated from NH3-TPD, which quantifies the total concentrations of Lewis and
232
Brønsted acid sites.38 It is noteworthy that the peak areas of the NH3-TPD profiles of the nTiO2
233
materials were markedly different (Figure S10b), indicating that the nTiO2 materials possessed
234
different amounts of surface Ti5c. In general, the anatase materials contained higher
235
concentrations of surface Ti5c than rutile and TiO2(B). Among the three anatase materials, the
236
concentration of Ti5c increased with the percentage of exposed {001} facets, and TiO2(B)_rod
237
contained larger content of surface Ti5c than TiO2(B)_wire, likely due to their differences in
238
morphology and exposed facets (Table 1, Figure S11).
239
Overall Catalytic Efficiency of nTiO2 in TeCA Hydrolysis is Dependent on Both
240
Crystalline Phase and Exposed Facets. Under all three pH conditions (7.0, 7.5 and 8.0) the
241
dehydrochlorination reaction of TeCA was enhanced in the presence of an nTiO2 material,
242
regardless of the specific type of nTiO2 used (Figure S12). For example, in the absence of nTiO2
243
(i.e., reaction in homogeneous aqueous solution), only 11% TeCA was transformed after 336 h at
244
pH 7.0, whereas in the presence of 100 mg/L nTiO2, transformation of TeCA increased to as
12
ACS Paragon Plus Environment
Page 12 of 30
Page 13 of 30
Environmental Science & Technology
245
much as 33%. Catalytic effects of nTiO2 were also observed at pH 7.5 and 8.0 (Figure S12).
246
Additionally, the reactivity of TeCA increased substantially with the increase of TiO2
247
concentration (Figure S13). TeCA dehydrochlorination in homogeneous aqueous solution
248
occurred through a β-elimination (E2) mechanism, in which hydroxide ion attacks the hydrogen
249
atom attached to the β-carbon, resulting in the breaking of a C-Cl bond and the formation of a
250
C=C bond.47 In the presence of TiO2, heterogeneous catalytic reaction of TeCA on the surface of
251
TiO2 (i.e., reaction of TiO2-adsorbed TeCA) occurred simultaneously.
252
The reaction kinetics followed the pseudo-first-order reaction kinetics reasonably, and the
253
fitted kinetic constants (Table S3) were used hereafter to compare the overall catalytic
254
efficiencies between different nTiO2 materials. It is evident that the overall catalytic efficiencies
255
of different nTiO2 materials differed from each other, especially at pH 7.0. Strikingly, the
256
BET-surface-area-normalized kobs values showed clear dependency on both crystalline phase and
257
exposed facets of nTiO2. Upon surface-area normalization, overall catalytic efficiencies of nTiO2
258
generally followed the order of rutile > anatase > TiO2(B), anatase_H001 > anatase_M001 >
259
anatase_L001 among the three anatase TiO2, and TiO2(B)_rod > TiO2(B)_wire between the two
260
TiO2(B) materials (Figure 2). Note that the three rutile materials differed mainly in particle size.
261
According to previous research,48 particle size influences catalytic efficiency of nTiO2 primarily
262
by affecting surface area. However, the variation among the kobs values of the three rutile
263
materials remained after surface-area-normalization (Figure 2), indicating that other factors also
264
affected the catalytic efficiency of the rutile materials. This might be attributable to the small
265
differences in exposed {101} and {211} facets among the three materials (Table S1).
266 267
Crystalline Phase and Exposed Facets Affect Adsorption Affinities of nTiO2. It is commonly accepted that the overall catalytic efficiency of a catalyst is largely affected by its
13
ACS Paragon Plus Environment
Environmental Science & Technology
268
adsorption affinity for target contaminants.49 Indeed, our study demonstrated the variable
269
adsorption affinities of the eight nTiO2 materials for TeCA (Figure S14), which likely have
270
contributed to the different reaction kinetics of TeCA catalyzed by these nanomaterials (Figure
271
S12). Strikingly, the surface-area-normalized adsorption coefficient, Kd (corresponding to an
272
equilibrium aqueous-phase TeCA concentration of approximately 2.5 mg/L), showed clear
273
dependence on both crystalline phase and exposed facets (Figure 3a). Specifically, the
274
normalized Kd values of the rutile TiO2 were significantly higher (3.2–5.3 times) than those of
275
the anatase TiO2 and TiO2(B), and among the three anatase TiO2 the normalized Kd increased
276
with the increasing percentage of exposed {001} facets (Figure 3a).
277
Adsorption of nonpolar, nonionic, hydrophobic organic contaminants (such as TeCA) to
278
metal oxides is driven primarily by the hydrophobic effect.50 Accordingly, a reasonable
279
explanation for the crystalline-phase- and facet-dependent adsorption affinity of nTiO2 for TeCA
280
is that the nTiO2 materials differed in crystalline phase and exposed facets possessed different
281
surface hydrophobicity. To verify this hypothesis, we compared the surface hydrophobicity of the
282
nTiO2 materials by measuring both their water contact angles51 and n-dodecane–water partition
283
coefficients (KDW).36 As expected, the relative adsorption affinities among the nTiO2 materials
284
correlated well with their surface hydrophobicity (Figures 3b and 3c). Overall, the relative
285
hydrophobicity of the rutile materials was higher than the anatase and TiO2(B) materials, as
286
shown by their larger water contact angles and KDW values (Table 1, Figure S15), leading to the
287
higher adsorption affinities of three rutile materials. Similarly, anatase materials with larger
288
percentage of exposed {001} facets had higher surface hydrophobicity and thus, higher
289
adsorption affinities. This demonstrated that crystalline phase and exposed facets determined the
290
adsorption of TeCA to nTiO2 by controlling the surface hydrophobicity of nTiO2.
14
ACS Paragon Plus Environment
Page 14 of 30
Page 15 of 30
Environmental Science & Technology
291
Crystalline Phase and Exposed Facets Determine Activity of Surface Catalytic Sites on
292
nTiO2. Even though the trend in overall catalytic efficiencies of nTiO2 appeared to be consistent
293
with the trend in their adsorption affinities, the variations among the eight nanomaterials were
294
not quantitatively comparable between the adsorption and hydrolysis experiments (Figures 2 and
295
3a). In the case of anatase versus rutile, the difference in surface-area-normalized adsorption
296
coefficients appeared to be much greater than the difference in surface-area-normalized kinetic
297
constants (Figures 2 and 3a). Similarly, the TiO2(B) and anatase materials showed similar
298
adsorption affinity (Figure 3a), whereas anatase generally exhibited greater overall catalytic
299
efficiency (Figure 2). To discern the additional factors controlling the overall catalytic efficiency
300
of TiO2, we compared the apparent reaction kinetic constants of TiO2-adsorbed TeCA (ks) among
301
different materials. Remarkably, the ks values, which indicate the activity of surface catalytic
302
sites, were also dependent on both crystalline phase and exposed facets (Figure 4). Anatase TiO2
303
exhibited higher ks values than both rutile TiO2 and TiO2(B), consistent with the abovementioned
304
effect of crystalline phase. Additionally, anatase materials with larger percentage of exposed
305
{001} facets had greater ks values, and the TiO2(B) material containing more {010} facets
306
exhibited relatively larger ks (Table S1 and Figure 4).
307
As mentioned earlier, the unsaturated Ti5c atoms on the surface of nTiO2 are likely the most
308
important surface catalytic sites in TeCA dehydrochlorination. The amounts of surface reactive
309
Ti5c sites varied depending on crystalline phase and exposed facets (Figure S16). An interesting
310
observation was that the relative abundance of surface Ti5c among the nTiO2 materials (Figure
311
S11) appeared to follow the same trend as their ks values (Figure 4). Overall, reasonable
312
correlations were observed between the ks values and the estimated concentrations of Ti5c (C_Ti5c)
313
(Figure 5), indicating that Ti5c atoms likely served as the predominant active sites to catalyze the
15
ACS Paragon Plus Environment
Environmental Science & Technology
314
dehydrochlorination of TeCA. In aqueous solution, the surface of nTiO2 surface is saturated with
315
water molecules, and the Ti5c atoms (that are unsaturated and thus more reactive) would tend to
316
be coordinated by hydroxyl groups, which after deprotonation, can catalyze the base-promoted
317
dehydrochlorination of TeCA.30,52
318
It is noteworthy that the correlation between ks and C_Ti5c is more significant under higher
319
pH values (Figure 5). This was consistent with the mechanism that the deprotonated hydroxyl
320
groups chemisorbed to Ti5c were the primary catalytic moiety, in that, increasing pH would
321
facilitate deprotonation of the hydroxyl groups, resulting in greater catalytic activity. Moreover,
322
the catalytic efficiencies of nTiO2 in TeCA dehydrochlorination were considerably weakened in
323
the presence of divalent metal cations (i.e., Cu2+, Pb2+ and Cd2+) that bound to and likely
324
inactivated the hydroxyl groups on nTiO2 surface (Figure S17). The relative inhibiting effects
325
followed the order of Cu2+ > Pb2+ > Cd2+, consistent with the binding strength of these cations to
326
hydroxyl groups.53 This result was also in line with the role of the surface hydroxyl groups in
327
nTiO2-catalyzed hydrolysis reactions, even though the potential effects of cation-induced
328
aggregation on the concentrations of exposed surface reactive sites, i.e., hydroxyl groups on Ti5c,
329
could not be ruled out (for instance, the hydrodynamic diameter of anatase_L001 was 303.6 ±
330
25.4 nm in MOPS buffer, but increased to 392.8 ± 18.4 nm when the buffer contained 0.5 mM
331
Cu2+).
332
An interesting observation was that the C_Ti5c-normalized ks appeared to positively
333
correlate with the percentage of the Ti5c sites with low acidic activity (Table S4 and Figure S18;
334
the Ti5c sites with low acidic activity are operationally defined as acid sites with NH3 desorption
335
temperature below 300 °C in Figure S10b), possibly because their conjugate bases (i.e.,
336
deprotonated hydroxyl groups coordinated with Ti5c) were relatively strong and thus can 16
ACS Paragon Plus Environment
Page 16 of 30
Page 17 of 30
Environmental Science & Technology
337
efficiently catalyze the dechlorination process. This result not only further corroborated that the
338
surface hydroxyl groups chemisorbed to Ti5c were the active catalytic sites, but also underscored
339
that both the amount and strength of surface catalytic sites need to be considered in order to
340
accurately predict the catalytic efficiencies of nTiO2. For example, anatase_H001 showed large
341
ks value due to its high abundance of surface Ti5c sites (Figure 5), whereas the percentage of Ti5c
342
with low acidity on the surface of anatase_H001 was relatively small (Figure S10b and Table S4)
343
and as a result the ks per unit reactive sites (quantified as Ti5c concentrations) was low (Figure
344
S18).
345
Environmental Implications. Hydrolysis of contaminants is one of the most important
346
abiotic transformation processes controlling the environmental transport, fate and effects of
347
contaminants. The significant catalytic effects of nTiO2 materials observed in this study indicate
348
that when released into the environment, nTiO2 (and possibly other metal oxide nanomaterials)
349
may significantly affect hydrolysis reactions of organic contaminants. Evidently, overall catalytic
350
efficiencies of TiO2 nanomaterials are greatly affected by their intrinsic properties, including
351
crystalline phase and exposed facets that determine both the adsorption capacity of nTiO2 for
352
organic contaminants and the activity of surface catalytic sites. In particular, crystalline phase
353
and exposed facets control the abundance and activity of surface unsaturated Ti atoms and
354
therefore are essential parameters for predicting the kinetics of hydrolysis reactions occurring on
355
the surface of nTiO2. Findings of this study may have important implications for assessing the
356
environmental impacts of TiO2 nanomaterials, and may also shed light on the design of novel
357
catalytic nanomaterials for contaminant removal and environmental remediation via crystal
358
engineering.
359
17
ACS Paragon Plus Environment
Environmental Science & Technology
360
Acknowledgments. This project was supported by the National Science Fund for Distinguished
361
Young Scholars (Grant 21425729), the Ministry of Science and Technology of China (Grant
362
2014CB932001), and the National Natural Science Foundation of China (Grants 21237002). We
363
thank three anonymous reviewers for their valuable comments and suggestions on mechanistic
364
interpretation.
365 366
Supporting Information Available: Material synthesis procedures; tables summarizing intensity
367
ratios of XRD, calculated percentage of {001} facets, experimental parameters and fitted kinetic
368
constants, and relative abundance of the weak and strong acid sites; figures showing mass
369
balance data of kinetic experiments, SEM images, XPS spectra, FTIR spectra, pyridine-adsorbed
370
FTIR spectra and NH3-TPD profiles of nTiO2 materials, Ti5c concentrations, dehydrochlorination
371
kinetics of TeCA, effects of TiO2 content and divalent cations on dehydrochlorination kinetics,
372
adsorption isotherms of TeCA, photographs of water contact angles on surface of nTiO2
373
materials, schematic illustration of faceted TiO2 surfaces, and effects of acid strength on catalytic
374
efficiencies of nTiO2 materials. This information is available free of charge via the Internet at
375
http://pubs.acs.org.
376 377
Notes—The authors declare no competing financial interest.
378 379
REFERENCES
380
1. Macwan, D. P.; Dave, P. N.; Chaturvedi, S. A review on nano-TiO2 sol–gel type syntheses and
381
its applications. J. Mater. Sci. 2011, 46 (11), 3669-3686.
18
ACS Paragon Plus Environment
Page 18 of 30
Page 19 of 30
Environmental Science & Technology
382
2. Yang, D.; Liu, H.; Zheng, Z.; Yuan, Y.; Zhao, J.; Waclawik, E. R.; Ke, X.; Zhu, H. An
383
efficient photocatalyst structure: TiO2(B) nanofibers with a shell of anatase nanocrystals. J. Am.
384
Chem. Soc. 2009, 131 (49), 17885−17893.
385
3. Hua, X.; Liu, Z.; Bruce, P. G.; Grey, C. P. The Morphology of TiO2(B) Nanoparticles. J. Am.
386
Chem. Soc. 2015, 137 (42), 13612-13623.
387
4. Robichaud, C. O.; Uyar, A. E.; Darby, M. R.; Zucker, L. G.; Wiesner, M. R. Estimates of
388
upper bounds and trends in nano−TiO2 production as a basis for exposure assessment. Environ.
389
Sci. Technol. 2009, 43 (12), 4227−4233.
390
5. Lewicka, Z. A.; Benedetto, A. F.; Benoit, D. N.; Yu, W.; Fortner, J. D.; Colvin, V. L. The
391
structure, composition, and dimensions of TiO2 and ZnO nanomaterials in commercial
392
sunscreens. J. Nanopart. Res. 2011, 13 (9), 3607−3617.
393
6. Qiang, L.; Pan, X.; Zhu, L.; Fang, S.; Tian, S. Effects of nano−TiO2 on
394
perfluorooctanesulfonate bioaccumulation in fishes living in different water layers: Implications
395
for enhanced risk of perfluorooctanesulfonate. Nanotoxicology 2016, 10 (4), 471−479.
396
7. Long, T.; Saleh, N.; Tilton, R. D.; Lowry, G. V.; Veronesi, B. Titanium dioxide (P25)
397
produces reactive oxygen species in immortalized brain microglia (BV2): Implications for
398
nanoparticle neurotoxicity. Environ. Sci. Technol. 2006, 40 (14), 4346−4352.
399
8. Battin, T. J.; Kammer, F. V. D.; Weilhartner, A.; Ottofuelling, S.; Hofmann, T.
400
Nanostructured TiO2: Transport behavior and effects on aquatic microbial communities under
401
environmental conditions. Environ. Sci. Technol. 2009, 43 (21), 8098−8104.
402
9. Giammar, D. E.; Maus, C. J.; Xie, L. Effects of particle size and crystalline phase on lead
403
adsorption to titanium dioxide nanoparticles. Environ. Eng. Sci. 2007, 24 (1), 85-95.
19
ACS Paragon Plus Environment
Environmental Science & Technology
404
10. Fujishima, A.; Zhang, X.; Tryk, D. A. TiO2 photocatalysis and related surface phenomena.
405
Surf. Sci. Rep. 2008, 63 (12), 515−582.
406
11. Ai, Z.; Gao, Z.; Su, K.; Ho, W.; Zhang, L. Aerosol flow synthesis of N, Si−codoped TiO2
407
hollow microspheres with enhanced visible−light driven photocatalytic performance. Cataly.
408
Commun. 2012, 29, 189−193.
409
12. Nalbandian, M. J.; Greenstein, K. E.; Shuai, D.; Zhang, M.; Choa, Y.; Parkin, G. F.; Myung,
410
N. V.; Cwiertny, D. M. Tailored synthesis of photoactive TiO2 nanofibers and Au/TiO2 nanofiber
411
composites: structure and reactivity optimization for water treatment applications. Environ. Sci.
412
Technol. 2015, 49 (3), 1654−1663.
413
13. Torrents, A.; Stone, A. T. Catalysis of picolinate ester hydrolysis at the oxide/water interface:
414
inhibition by adsorbed natural organic matter. Environ. Sci. Technol. 2002, 27 (12), 1060−1067.
415
14. Torrents, A.; Stone, A. T. Hydrolysis of phenyl picolinate at the mineral/water interface.
416
Environ. Sci. Technol. 1991, 25 (1), 143−149.
417
15. Smolen, J. M.; Stone, A. T. Metal (hydr) oxide surface−catalyzed hydrolysis of
418
chlorpyrifos−methyl, chlorpyrifos−methyl oxon, and paraoxon. Soil Sci. Soc. Am. J. 1998, 62 (3),
419
636−643.
420
16. Torrents, A.; Stone, A. T. Oxide surface−catalyzed hydrolysis of carboxylate esters and
421
phosphorothioate esters. Soil Sci. Soc. Am. J. 1994, 58 (3), 738−745.
422
17. Baldwin, D. S.; Beattie, J. K.; Coleman, L. M.; Jones, D. R. Phosphate ester hydrolysis
423
facilitated by mineral phases. Environ. Sci. Technol. 1995, 29 (6), 1706−1709.
424
18. Stumm, W.; Morgan, J.J. Aquatic chemistry: Chemical equilibria and rates in natural waters,
425
3rd ed. Wiley, New York, NY, USA, 1996.
20
ACS Paragon Plus Environment
Page 20 of 30
Page 21 of 30
Environmental Science & Technology
426
19. Schwarzenbach, R.P.; Gschwend, P.M.; Imboden, D.M. Environmental organic chemistry,
427
2rd ed. Wiley-Interscience, Hoboken, New Jersey, USA, 2003.
428
20. Jiang, W.; Furrer, G.; Kaufmann, S.; Schulin, R. Influence of clay minerals on the hydrolysis
429
of carbamate pesticides. Environ. Sci. Technol. 2001, 35 (11), 2226−2232.
430
21. Yang, H.; Sun, C.; Qiao, S.; Zou, J.; Liu, G.; Smith, S. C.; Cheng, H.; Lu, G. Anatase TiO2
431
single crystals with a large percentage of reactive facets. Nature 2008, 453 (7195), 638−641.
432
22. Bai, Y.; Seró, I. M.; Angelis, F. D.; Bisquert, J.; Wang, P. Titanium dioxide nanomaterials for
433
photovoltaic applications. Chem. Rev. 2014, 114 (19), 10095−10130.
434
23. Diebold, U. The surface science of titanium dioxide. Surf. Sci. Rep. 2003, 48 (5−8), 53−229.
435
24. Cao, Y.; Li, Q.; Li, C.; Li, J.; Yang, J. Surface heterojunction between (001) and (101) facets
436
of ultrafine anatase TiO2 nanocrystals for highly efficient photoreduction CO2 to CH4. Appl.
437
Catal. B 2016, 198, 378−388.
438
25. Henderson, M. A. A surface science perspective on tio photocatalysis. Surf. Sci. Rep. 2011,
439
66 (6-7), 185-297.
440
26. Yu, J.; Low, J.; Xiao, W.; Zhou, P.; Jaroniec, M. Enhanced photocatalytic CO₂-reduction
441
activity of anatase TiO₂ by coexposed {001} and {101} facets. J. Am. Chem. Soc. 2014, 136 (25),
442
8839-8842.
443
27. Liu, G.; Yang, H..; Wang, X..; Cheng, L.; Pan, J.; Lu, G.; Cheng, H. Visible light responsive
444
nitrogen doped anatase TiO2 sheets with dominant {001} facets derived from TiN. J. Am. Chem.
445
Soc. 2009, 131 (36), 12868−12869.
446
28. Luschtinetz, R.; Frenzel, J.; Milek, T.; Seifert, G. Adsorption of phosphonic acid at the TiO2
447
anatase (101) and rutile (110) surfaces. J. Phys. Chem.C 2009, 113 (14), 5730−5740.
21
ACS Paragon Plus Environment
Environmental Science & Technology
448
29. Chermahini, A. N.; Farrokhpour, H.; Zeinodini, A. Adsorption of some important tautomers
449
of 5−amino tetrazole on the (001) and (101) surfaces of anatase: Theoretical study. J. Mol. Struct.
450
2016, 1121, 203−214.
451
30. Chen, W.; Zhu, D.; Zheng, S.; Chen, W. Catalytic effects of functionalized carbon nanotubes
452
on dehydrochlorination of 1,1,2,2−tetrachloroethane. Environ. Sci. Technol. 2014, 48 (7),
453
3856−3863
454
31. Li, X.; Chen, W.; Zhang, C.; Li, Y.; Wang, F.; Chen, W. Enhanced dehydrochlorination of
455
1,1,2,2−tetrachloroethane by graphene−based nanomaterials. Environ. Pollut. 2016, 214,
456
341−348.
457
32. Vignona, L. C. Sorption and reactivity of 1,1,2,2−tetrachloroethane. Ph. D. Dissertation, Rice
458
University, Houston, TX, 2000.
459
33. Cooper, W. J.; Mehran, M.; Riusech, D. J.; Joens, J. A. Abiotic transformations of
460
halogenated organics. 1. Elimination reaction of 1,1,2,2−tetrachloroethane and formation of
461
1,1,2−trichloroethene. Environ. Sci. Technol. 1987, 21 (11), 1112−1114.
462
34. Li, C.; Koenigsmann, C.; Ding, W.; Rudshteyn, B.; Yang, K.; Regan, K. P.; Konezny, S. J.;
463
Batista, V. S.; Brudvig, G. W.; Schmuttenmaer, C. A.; Kim, J. H. Facet−dependent
464
photoelectrochemical performance of TiO2 nanostructures: An experimental and computational
465
study. J. Am. Chem. Soc. 2015, 137 (4), 1520−1529.
466
35. Zhen, M.; Guo, S.; Gao, G.; Zhou, Z.; Liu, L. TiO2−B nanorods on reduced graphene oxide
467
as anode materials for Li ion batteries. Chem. Commun. 2015, 51 (3), 507−510.
468
36. Walker, S.L., Hill, J.E., Redman, J.A., Elimelech, M. Influence of growth phase on adhesion
469
kinetics of Escherichia coli D21g. Appl. Environ. Microbiol. 2005, 71 (6), 3093-3099.
470
37. Ma, Y.; Xu, Q.; Zong, X.; Wang, D.; Wu, G.; Wang, X.; Li, C. Photocatalytic H2 production
22
ACS Paragon Plus Environment
Page 22 of 30
Page 23 of 30
Environmental Science & Technology
471
on Pt/TiO2–SO42- with tuned surface−phase structures: enhancing activity and reducing CO
472
formation. Energ. Environ. Sci. 2012, 5 (4), 6345−6351.
473
38. Yan, L.; Du, J.; Jing, C. How TiO2 facets determine arsenic adsorption and photooxidation:
474
Spectroscopic and DFT studies. Catal. Sci. Technol. 2016, 6 (7), 2419−2426.
475
39. Yang, K.; Xing, B. Sorption of phenanthrene by humic acid−coated nanosized TiO2 and ZnO.
476
Environ. Sci. Technol. 2009, 43 (6), 1845−1851.
477
40. Xiang, Q.; Lv, K.; Yu, J. Pivotal role of fluorine in enhanced photocatalytic activity of
478
anatase TiO2 nanosheets with dominant (001) facets for the photocatalytic degradation of acetone
479
in air. Appl. Catal. B 2010, 96 (3–4), 557–564.
480
41. Tian, F.; Zhang, Y.; Zhang, J.; Pan, C. Raman spectroscopy: A new approach to measure the
481
percentage of anatase TiO2 exposed (001) facets. J. Phys. Chem. C 2012, 116 (13), 7515–7519.
482
42. Wang, M.; Zhang, F.; Zhu, X.; Qi, Z.; Hong, B.; Ding, J.; Bao, J.; Sun, S.; Gao, C. DRIFTS
483
evidence for facet−dependent adsorption of gaseous toluene on TiO2 with relative photocatalytic
484
properties. Langmuir 2015, 31 (5), 1730–1736.
485
43. Erdem, B.; Hunsicker, R. A.; Simmons, G. W.; Sudol, E. D.; And, V. L. D.; Elaasser, M. S.
486
XPS and FTIR surface characterization of TiO2 particles used in polymer encapsulation.
487
Langmuir 2001, 17 (9), 2664−2669.
488
44. Zhu, J.; Yang, J.; Bian, Z.; Ren, J.; Liu, Y. M.; Cao, Y.; Li, H.; He, H.; Fan, K.
489
Nanocrystalline anatase TiO2 photocatalysts prepared via a facile low temperature nonhydrolytic
490
sol–gel reaction of TiCl4 and benzyl alcohol. Appl. Catal. B 2007, 76 (1), 82−91.
491
45. Karapati, S.; Giannakopoulou, T.; Todorova, N.; Boukos, N.; Dimotikali, D.; Trapalis, C.
492
Eco−efficient TiO2 modification for air pollutants oxidation. Appl. Catal. B 2015, 176−177,
493
578−585.
23
ACS Paragon Plus Environment
Environmental Science & Technology
494
46. Emeis, C. A. Determination of integrated molar extinction coefficients for infrared absorption
495
bands of pyridine adsorbed on solid acid catalysts. J. Catal. 1993, 141, 347−354.
496
47. Elsner, M.; Cwiertny, D. M.; Roberts, A. L.; Lollar, B. S. 1,1,2,2-tetrachloroethane reactions
497
with OH-, Cr(II), granular iron, and a copper-iron bimetal: Insights from product formation and
498
associated carbon isotope fractionation. Environ. Sci. Technol. 2007, 41 (11), 4111-4117.
499
48. He, F.; Li, J.; Li, T.; Li, G. Solvothermal synthesis of mesoporous TiO2: The effect of
500
morphology, size and calcination progress on photocatalytic activity in the degradation of
501
gaseous benzene. Chem. Eng. J. 2014, 237 (237), 312-321.
502
49. Ollis, D. F.; Pelizzetti, E.; Serpone, N. Destruction of water contaminants. Environ. Sci.
503
Technol. 1991, 25 (9), 1522−1529.
504
50. Ooka, C.; Yoshida, H.; Suzuki, K.; Hattori, T. Effect of surface hydrophobicity of
505
TiO2-pillared clay on adsorption and photocatalysis of gaseous molecules in air. Appl. Catal. A
506
2004, 260 (1), 47-53.
507
51. Wu, W.; Nancollas, G. H. Kinetics of heterogeneous nucleation of calcium phosphates on
508
anatase and rutile surfaces. J. Colloid Interf. Sci. 1998, 199 (2), 206-211.
509
52. Chen, W.; Li, Y.; Zhu, D.; Zheng, S.; Chen, W. Dehydrochlorination of activated
510
carbon−bound 1,1,2,2−tetrachloroethane: Implications for carbonaceous material−based
511
soil/sediment remediation. Carbon 2014, 78, 578−588.
512
53. Kinniburgh, D. G.; Milne, C. J.; Benedetti, M. F.; Pinheiro, J. P.; Filius, J.; Koopal, L. K.;
513
Riemsdijk, W. H. V. Metal ion binding by humic acid: Application of the NICA−Donnan model.
514
Environ. Sci. Technol. 1996, 30 (5), 1687−1698.
24
ACS Paragon Plus Environment
Page 24 of 30
Page 25 of 30
Environmental Science & Technology
Table 1. Selected Physicochemical Properties of the nTiO2 Materials
type of nTiO2
dominant
SABET 2
a
exposed facets
(m /g)
Anatase_L001
{101}
Anatase_M001
surface O b surface Ti
(at%)
b
C_acid (C_Ti5c) c Olatt/Ti
(at%)
ratio
57.4
30.1
7.4
60.2
29
9.8
{110}
31
Rutile_2
{110}
Rutile_3
OH
Olatt
83
12.5
{101}
35
Anatase_H001
{101}
Rutile_1
(µmol/m2)
b
KDW d
water contact angle (°)
total
weak
strong
1.87
2.1
1.1
1.0
0.04
17
32.4
1.86
6.3
1.4
4.9
0.09
18
57.4
32.8
1.75
6.6
0.8
5.8
0.09
20
13.2
56.4
30.4
1.86
0.9
0.8
0.1
0.56
35
28
12.6
57.2
30.2
1.89
0.9
0.4
0.5
0.49
33
{110}
27
12.8
57.0
30.2
1.89
0.8
0.7
0.1
0.33
25
TiO2(B)_rod
{110} & {010}
106
15.4
54.2
30.4
1.78
2.4
0.8
1.6
0.04
20
TiO2(B)_wire
{110} & {010}
185
17.2
52.4
30.4
1.72
0.7
0.2
0.5
0.03
19
a
SABET = surface area measured using the Brunauer–Emmett–Teller (BET) method.
b
Analyzed by X-ray photoelectron spectroscopy (XPS). The values represent the average of duplicate samples.
c
C_acid represents acid concentration (measured with NH3-TPD) normalized by surface area, which is approximately the concentration of surface Ti5c of
nTiO2 materials (C_Ti5c). d
KDW represents n-dodecane–water partition coefficient of nTiO2.
25
ACS Paragon Plus Environment
Environmental Science & Technology
(a)
Page 26 of 30
(b) A(101)
A(004)
I I II II
R(110)
R(101)
Intensity (a.u.)
Intensity (a.u.)
III R(211) IV
V
B(110)
10
20
30
B(010)
40
50
60
70
III IV V
VI
VI
VII
VII
VIII
VIII
100
80
200
300
400
500
600
700
Raman shift ( cm-1)
2θ ( Deg)
Figure 1. The XRD patterns (a) and Raman spectra (b) of the nTiO2 materials. The letters A, R and B refer to the anatase, rutile, and TiO2(B) phases, respectively. The numbers I, II, III, IV, V, VI, VII and VIII refer to anatase_L001, anatase_M001, anatase_H001, rutile_1, rutile_2, rutile_3, TiO2(B)_rod and TiO2(B)_wire, respectively.
26
ACS Paragon Plus Environment
Page 27 of 30
Environmental Science & Technology
(b)
(a) pH 7.0
kobs/SABET (g h
kobs/SABET (g h
pH 7.5 6.8e-5
-1
2.8e-5
m-2)
8.5e-5
-1
m-2)
3.5e-5
2.1e-5 1.4e-5 7.0e-6
3.4e-5 1.7e-5 0.0
An a
An a
ta An se_ L0 at 0 a An se_ 1 at M as 00 e_ 1 H0 Ru 0 1 til e_ 1 R ut ile _2 R Ti util O e_ 2 3 Ti (B)_ O 2 ( rod B )_ w ire
ta An se_ L0 at 0 a An se_ 1 at M0 as 0 e_ 1 H 00 R ut 1 ile _1 R ut ile _2 R Ti uti le O _3 2( Ti B)_ O r o 2( d B )_ w ire
0.0
5.1e-5
(c) pH 8.0 1.6e-4
kobs/SABET (g h
-1
m-2)
2.0e-4
1.2e-4 8.0e-5 4.0e-5
An a
An a
ta s
e_
L0
ta 0 s An e_ 1 at M0 as 01 e_ H 0 R 01 ut ile _1 R ut ile _2 R Ti uti O le_ 2( 3 Ti B)_ O 2 ( rod B) _w i re
0.0
Figure 2. Surface-area-normalized apparent pseudo-first-order kinetic constants (kobs) of different nTiO2 materials at pH 7.0 (a), 7.5 (b) and 8.0 (c). The kobs values were obtained using Eqn. 1. Each color of the columns represents a crystalline phase of nTiO2.
27
ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 30
(a)
2 Kd/SABET (L/m )
8.0e-3 6.4e-3 4.8e-3 3.2e-3 1.6e-3
An at
a An se_ L at as 001 An e_ a t M0 as 0 e_ 1 H0 01 R ut ile _1 R ut ile _2 R ut Ti i l e O _3 2 Ti (B) O _r od 2( B )_ w ire
0.0
(c) 8.0e-3
6.4e-3
6.4e-3
2 Kd/SABET (L/m )
2 Kd/SABET (L/m )
(b) 8.0e-3
4.8e-3 3.2e-3 1.6e-3
2 R = 0.9132
0.0
4.8e-3 3.2e-3 1.6e-3
2 R = 0.9710
0.0 15.0
20.0
25.0
30.0
35.0
40.0
0.0
0.1
Water ontact angle (o) Anatase_L001 Anatase_M001
0.2
0.4
0.5
0.6
KDW Anatase_H001 Rutile_1
Rutile_2 Rutile_3
TiO2(B)_rod TiO2(B)_wire
Figure 3. Surface-area-normalized adsorption coefficients (Kd) of different nTiO2 materials (a), and correlations between surface-area-normalized Kd and water contact angle (b) and n-dodecane–water partition coefficients (KDW) (c) of the nTiO2 materials. The Kd values correlate to an equilibrium aqueous-phase 1,1,2,2-tetrachloroethane concentration of approximately 2.5 mg/L. Each color of the columns represents a crystalline phase of nTiO2.
28
ACS Paragon Plus Environment
Page 29 of 30
Environmental Science & Technology
(b)
(a) 0.060
0.180
pH 7.0
pH 7.5
0.036
0.000
0.000 e_
se
ta s An a
An a
An a
ta
se
_L 00 1 An e_ M at 0 as 0 e_ 1 H 00 1 R ut ile R _1 ut ile _2 R u Ti O tile _3 2( B )_ Ti O ro 2( d B )_ w ire
0.012
L0 01 _M 0 ta se 01 _H 00 1 R ut ile R _1 ut ile _2 R u Ti t i l O e_ 2( 3 B ) Ti O _r o 2( B d )_ w ire
0.072
ta
0.024
0.108
An a
0.036
An a
-1 ks ((h )
0.144
ta s
-1 ks ((h )
0.048
(c) 0.240
pH 8.0
-1 ks ((h )
0.192 0.144 0.096 0.048
An a
An a
ta
se
_L 0 ta se 01 An _M at as 001 e_ H 00 1 R ut ile _ R ut 1 ile _2 R Ti uti le O _3 2( B )_ Ti O ro 2( d B )_ w ire
0.000
Figure 4. Reaction kinetic constants of 1,1,2,2-tetrachloroethane adsorbed to different nTiO2 materials (ks) at pH 7.0 (a), 7.5 (b) and 8.0 (c). The ks values were obtained using Eqn. 2. Each color of the columns represents a crystalline phase of nTiO2.
29
ACS Paragon Plus Environment
Environmental Science & Technology
(a)
(b)
0.060
0.180
pH 7.5
pH 7.0 0.048
0.144
-1 ks (h )
-1 ks (h )
Page 30 of 30
0.036 0.024 0.012
0.108 0.072 0.036
2 R = 0.7726
0.000
2 R = 0.7923
0.000 0.0
2.0e-3
4.0e-3
6.0e-3
8.0e-3
0.0
2.0e-3
4.0e-3
6.0e-3
8.0e-3
2 C_Ti5c (mmol/m )
2 C_Ti5c (mmol/m ) (c) 0.240
Anatase_L001 Anatase_M001 Anatase_H001 Rutile_1
pH 8.0
-1 ks (h )
0.192 0.144
Rutile_2 Rutile_3 TiO2(B)_rod
0.096
TiO2(B)_wire
0.048
2 R = 0.8698
0.000 0.0
2.0e-3
4.0e-3
6.0e-3
8.0e-3
2 C_Ti5c (mmol/m )
Figure 5. Correlations between the reaction kinetic constants of 1,1,2,2-tetrachloroethane adsorbed to different nTiO2 materials (ks) and the Ti5c concentrations on nTiO2 surface at pH 7.0 (a), 7.5 (b) and 8.0 (c). The ks values were obtained using Eqn. 2. Each color of the symbols represents a crystalline phase of nTiO2.
30
ACS Paragon Plus Environment