Accepted Manuscript Oxidation of ethanethiol in aqueous alkaline solution by ferrate(VI): Kinetics, stoichiometry and mechanism Jing Wang, Tong Zheng, Chen Cai, Yanxiang Zhang, Huiling Liu PII: DOI: Reference:
S1385-8947(18)32205-8 https://doi.org/10.1016/j.cej.2018.11.005 CEJ 20307
To appear in:
Chemical Engineering Journal
Received Date: Revised Date: Accepted Date:
15 July 2018 30 October 2018 1 November 2018
Please cite this article as: J. Wang, T. Zheng, C. Cai, Y. Zhang, H. Liu, Oxidation of ethanethiol in aqueous alkaline solution by ferrate(VI): Kinetics, stoichiometry and mechanism, Chemical Engineering Journal (2018), doi: https:// doi.org/10.1016/j.cej.2018.11.005
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.
1
Oxidation of ethanethiol in aqueous alkaline solution by ferrate(VI): Kinetics,
2
stoichiometry and mechanism
3 Jing Wang a, Tong Zheng b,*, Chen Cai a, Yanxiang Zhang a, Huiling Liu c,*
4 5
a
6
of Technology, Harbin 150090, China
7
b
School of Environment, Harbin Institute of Technology, Harbin 150090, China
8
c
College of Environmental Science and Engineering, Tongji University, Shanghai
9
200092, China
State Key Laboratory of Urban Water Resource and Environment, Harbin Institute
10
11
E-mail:
[email protected] (Tong Zheng),
[email protected] (Huiling Liu).
Corresponding authors.
1
12 13
Abstract The oxidation of ethanethiol (C2H5SH), as a representative odorous pollutant, by
14
ferrate (Fe(VI)) in aqueous alkaline solution was investigated to understand its
15
kinetics, stoichiometry, and reaction pathways. The reaction kinetics and fraction of
16
different intermediates were found highly dependent on pH. The reaction behaved as a
17
5/2-order reaction in a pH range of 8.0–10.0, and a second-order reaction in a pH
18
range of 10.5–12.0. The rate constants decreased with increased pH, and the
19
speciation of Fe(VI) and C2H5SH during the reaction were further discussed. C2H5SH
20
was finally oxidized to nonvolatile and odorless sulfonic acid (C2H5SO3H) with the
21
intermediates of sulfinic acid (C2H5SO2H) and disulphide (C2H5SSC2H5). A plausible
22
mechanism involving two competing reactions for C2H5S• radicals was proposed to
23
explain the pH-dependent reaction order. In addition, kinetics and products of
24
C2H5SH degradation by Fe(VI) were compared with other oxidants (O3, chlorine,
25
ClO2, KMnO4, H2O2, and •OH). Our results provided a deep understanding of
26
C2H5SH degradation by Fe(VI) in aqueous alkaline solution.
27
Keywords: Ferrate(VI); Ethanethiol; Diethyl disulfide; Kinetics; Stoichiometry
2
28
1. Introduction
29
Odorous pollutants have gained increasing attention due to their offensive and
30
corrosive effects on the environment as well as the detrimental effect on human health.
31
Thiols, a group of typical organic sulfur compounds, are characterized by their
32
offensive odor that can be detected by human nose at very low concentrations [1-3].
33
Thiols are released from various sources such as sewage treatment works, landfill
34
sites, petroleum refining and other manufacturing processes [4-6]. Ethanethiol
35
(C2H5SH) as one of representative thiols is regarded as the smelliest substance with a
36
quite low odor threshold of ~8.7 × 10-3 ppb/v [7]. Low levels of C2H5SH in the air can
37
result in chronic harmful effects on lungs and nervous systems of human [8,9].
38
The chemical scrubbing technique has been proved to be an effective and reliable
39
process for odorous gas treatment where odorants could be absorbed to the liquid
40
phase and then oxidized rapidly by oxidants [10-12]. Ferrate(VI) (Fe(VI)) is
41
considered as a green oxidant due to its environmental benign characters and
42
integrated function of oxidation, disinfection and coagulation [13,14]. It is also a
43
selective oxidant exhibiting high reactivity with electron-rich groups of target
44
compounds including organic sulfur compounds, amines and phenols [15,16]. Fe(VI)
45
has shown potential advantages in sludge odor control for sewage treatment works
46
(e.g., thiosemicarbazide , thiourea dioxide and methyl mercaptan) [17]. Furthermore,
47
there is evidence that gaseous methyl mercaptan can be effectively degraded by the
48
electrochemically generated Fe(VI) in NaOH solution [18,19]. The rate constants of 3
49
thiols oxidation by Fe(VI) were reported in the range of ~104 M-1 s-1 at pH = 9 [20].
50
Therefore, C2H5SH with low molecular weight and weak acidity (pKa = 10.5 [21])
51
might be absorbed readily into the alkaline Fe(VI) solution, and then be oxidized with
52
an extremely fast reaction rate.
53
Significant progress has been made in exploring the kinetics and mechanism of
54
the reaction between Fe(VI) and organic pollutants in neutral and alkaline conditions.
55
Ordinarily, the role of pH in kinetics of the reaction is embodied in the variation of
56
rate constants with pH which could generally be quantitatively modeled [16,22,23].
57
However, the possible effect of pH on intermediates and products of the reaction
58
requires investigation as well. Reaction mechanisms involving one and two-electron
59
transfer as well as oxygen transfer steps have been proposed in the oxidation by
60
Fe(VI), depending on the nature of organic pollutants [20,24]. Nevertheless, more
61
information on radical formation and the intermediate species of Fe(VI) is necessary
62
for a deeper understanding of the reactions, which may assist with distinguishing
63
electron transfer and oxygen transfer steps [15,25]. Studies have been carried out on
64
the mechanism of thiols oxidation by Fe(VI), but individual thiols differ markedly in
65
their behavior toward Fe(VI) [20]. Methyl mercaptan oxidation by Fe(VI) might go
66
through a one-electron transfer step to produce Fe(V) and thiyl radical in the first step.
67
The final products were Fe(III) and SO42- with the cleavage of C-S bond [17]. The 2-
68
mercaptoethanesulfonic acid could be oxidized to sulfonic acid through two-electron
69
transfer steps, and Fe(VI) was finally reduced to Fe(III) with the initial product of 4
70
Fe(II) [26]. Cysteine oxidation occurred through two-electron transfer steps as well
71
producing cysteine sulfinic acid, but Fe(II) was detected as the only final product of
72
Fe(VI) [20,27]. Therefore, more studies are needed to further understand the oxidation
73
of thiols by Fe(VI). However, we are not aware of any studies that investigate on the
74
mechanism of C2H5SH oxidation by Fe(VI) in aqueous alkaline solution. Moreover,
75
dependence of reaction order and intermediates fraction on pH have not been reported
76
in such a reaction, which are not very common in the Fe(VI) reactions.
77
Hence, the objectives of this study are to (i) investigate the effect of pH on
78
reaction kinetics of C2H5SH with Fe(VI) in aqueous alkaline solution; (ii) determine
79
the reaction stoichiometry and identify the products of C2H5SH; (iii) propose a
80
plausible mechanism to explain the changes in reaction orders at different pH; and (iv)
81
compare the kinetics and products of C2H5SH degradation by Fe(VI) with other
82
known oxidants.
83
2. Materials and methods
84
2.1. Chemicals and reagents
85
C2H5SH (GC grade, ≥ 99.5% purity), diethyl disulfide (C2H5SSC2H5, GC grade, >
86
99.0% purity) and sodium ethanesulfinate (C2H5SO2Na, 93% purity) were purchased
87
from Aladdin Chemical Inc. (Shanghai, China). Sodium ethanethiolate (C2H5SNa, 90%
88
purity) and ethanesulfonic acid (C2H5SO3H, 95% purity) were purchased from Sigma
89
Aldrich Chemical Inc. Potassium ferrate (K2FeO4) was synthesized according to the
90
wet method described by Li et al. [28]. Briefly, K2FeO4 was resulted from the 5
91
oxidation of ferric nitrate (Fe(NO3)3·9H2O) by potassium hypochlorite (KClO) in
92
strong alkaline solution, which was then further purified to obtain a higher purity.
93
Solid K2FeO4 was then dried and stored in a vacuum desiccators. The purity of solid
94
K2FeO4 was determined to be 95.6% by the chromite analytical method [29]. Further
95
details on the procedures of synthesis and purity analysis are provided in
96
supplementary material (Texts S1), as well as the characterization of K2FeO4 by X-ray
97
powder diffraction (XRD) (Text S2). Other chemicals used in the experiments were of
98
analytical grade and used without further purification. All aqueous solutions were
99
prepared with ultrapure water obtained from a Milli-Q water purification system (18
100
MΩ cm, Millipore, USA), and the pH was adjusted by adding either phosphoric acid
101
or NaOH solution.
102
The aqueous solution of Fe(VI) was freshly prepared by dissolving solid K2FeO4
103
in pure water or buffer solution, and subsequently filtrated through a 0.45μm
104
polyethersulfone (PES) syringe filter (Tianjin NAVIGATOR, China). Fe(VI)
105
concentration was determined at 510 nm with ε510 nm = 1150 M-1 cm-1 using a UV-Vis
106
spectrophotometer [30]. Aqueous solutions of C2H5SH were freshly prepared by
107
either dissolving solid C2H5SNa or adding C2H5SH to pure water just before use. The
108
stock solution of C2H5SSC2H5 for kinetic studies was prepared according to the
109
following steps. 0.5 mL C2H5SSC2H5 was added into 1 mL methanol, and then 0.25
110
mL tween 80 was added into the mixture to increase the aqueous solubility of
111
C2H5SSC2H5. After that, water was slowly added and the solution volume was 6
112
adjusted to 50 mL. C2H5SSC2H5 solution for products studies was prepared in
113
methanol.
114
2.2. Kinetics studies
115
Kinetics studies on the reaction of Fe(VI) with substrates (C2H5SH and
116
C2H5SSC2H5) were investigated using a stopped-flow spectrophotometer (SX 20)
117
from Applied Photophysics Ltd, UK at 25 ± 1 °C. The Fe(VI) solutions were prepared
118
in 5 mM Na2HPO4/1 mM Na2B4O7·10H2O buffer solutions (pH = 9.2). The aqueous
119
solutions of substrates were prepared in buffer solutions with a pH range of 7.0‒12.0.
120
20 mM phosphate buffer solution was used for pH = 7.0‒8.0. 10 mM borate buffer
121
solution was used for pH = 8.5‒12.0 together with 20 mM Na2HPO4. Phosphate was
122
used to control pH as well as to avoid the precipitation of Fe(III). The Fe(III)
123
precipitation not only interferes with the optical monitoring of the reaction but also
124
causes damage to the instruments. The experiments were performed under the pseudo-
125
order condition with substrates in at least 10-fold excess. The Fe(VI) concentrations
126
ranged from 0.09 to 0.14 mM. Fe(VI) solutions were mixed with equal volumes of
127
substrate solutions at the studied pH. Reaction rate constants were determined by
128
monitoring the decay of Fe(VI) absorbance as a function of time. Five experimental
129
runs were performed for each concentration of substrates to ensure the reliability of
130
the data. The pseudo-order rate constants reported in this study represented the
131
average values with the relative standard deviations less than 5%. In the control
132
experiments for Fe(VI) reaction with C2H5SH, the decrease in the concentration of 7
133
Fe(VI) in the given reaction time was less than 5% of the initial Fe(VI) concentration
134
at each studied pH. Therefore, the self-decomposition of Fe(VI) was ignored in
135
evaluating the rate constants. The pseudo first-order rate constants for the reaction
136
between Fe(VI) and C2H5SSC2H5 were corrected for the Fe(VI) decreases in the
137
presence of tween 80 (< 0.1%, v/v) at each studied pH. The modeling of the kinetic
138
data was performed with the software of Matlab R2016b.
139
2.3. Stoichiometry and products studies
140
Stoichiometric experiments were carried out at pH = 9.2 and 12.0 in 10 mM
141
borate/20 mM Na2HPO4 buffer solutions by mixing equal volumes of Fe(VI) and
142
C2H5SH solutions. Buffer solutions were used to maintain the pH of the reaction
143
mixtures. C2H5SH solutions were prepared at a constant concentration of 2.0 mM,
144
while Fe(VI) solutions ranged from 0 to 6.2 mM with different molar ratios of Fe(VI)
145
to C2H5SH in the reaction mixture. In the experiments without buffer, aqueous
146
solutions of Fe(VI) and C2H5SH were prepared in pure water at pH = 9.2 and 12.0 and
147
mixed immediately after preparation. Each reaction was conducted for 15 min at 20 ±
148
1 °C and the pH of the mixed solution was measured. The concentrations of C2H5SH
149
were determined according to the Ellman’s reagent method described in detail by
150
Riener et al. [31]. The concentrations of sulfonic ion (C2H5SO3-) and sulfinic ion
151
(C2H5SO2-) as the products of C2H5SH degradation were quantified by Dionex ICS-
152
2100 ion chromatography (IC) with an IonPac AS11-HC anion column (4 × 250 mm).
153
The final oxidation state of iron was determined using o-phenanthroline and 8
154
thiocyanate.
155
To identify other possible intermediates, the reaction solution was first extracted
156
with methyl tert-butyl ether (MTBE, GC grade, ≥ 99.9%), and then the extractant was
157
analyzed by gas chromatography/mass spectrometry (GC/MS) (Agilent 6890GC-
158
5975MSD) with a HP-5MS column (30 m × 0.25 mm). The injection mode was
159
splitless injection and a flow rate of the carrier gas (Helium) was 1.0 mL min-1. The
160
temperature of GC oven was set at 30 °C for 10 min, and then ramped up to 315 °C at
161
10 °C min-1. Electron ionization mode with an ionization energy of 70 eV was used in
162
mass spectrometer, and the scanned mass ranged from 20 to 400 m z-1 in a full scan
163
mode. The possible intermediates were analyzed and matched with the NIST standard
164
reference database. The concentration of C2H5SSC2H5 in the extractant was quantified
165
by GC/MS with selected ion monitor, and the extraction recovery was determined to
166
be ~95%.
167
3. Results and Discussion
168
3.1. Kinetics
169
The reaction rate of Fe(VI) with substrate is expressed and calculated by Eq. (1):
170
- d[Fe(VI)]/dt = k [Fe(VI)]m[S]n
171
where [Fe(VI)] and [S] are the concentrations of Fe(VI) and substrates (C2H5SH and
172
C2H5SSC2H5), respectively, M; k represents the apparent rate constant, M1-m-n s-1;
173
while m and n are the orders in the concentrations of Fe(VI) and substrate,
174
respectively.
(1)
9
175 176
The experiments were carried out with substrates in large excess. Eq. (1) could be rewritten as Eq. (2):
177
- d[Fe(VI)]/dt = k' [Fe(VI)]m
178
(2)
179
where k' = k [S]n, M1-m s-1.
180
The decreases in the concentrations of Fe(VI) with time at different pH are
181
shown in Fig. S3. In the pH range of 8.0–10.0, plots of ([Fe(VI)]0/[Fe(VI)]t)1/2 versus
182
time at different C2H5SH concentrations were found to be almost linear with positive
183
slopes and intercepts (intercept = 1) (R2 > 0.98, Fig. S4(a)). [Fe(VI)]0 and [Fe(VI)]t
184
represent the concentrations of Fe(VI) (M) at time zero and any time t, respectively.
185
The results indicated that the reaction was 3/2-order with respect to Fe(VI), i.e., m =
186
3/2. This was different from the common results that the reaction between Fe(VI) and
187
organic sulfur compounds showed a first-order dependence on each reactant [20,25].
188
It might be the first time that a fraction order with respect to Fe(VI) was obtained in
189
the reaction of Fe(VI) with thiols. However, in the pH range of 10.5–12.0, the
190
decreased absorbance of Fe(VI) as a function of time fitted well to single exponential
191
model, indicating that the reaction showed the first-order dependence on Fe(VI), i.e.,
192
m = 1 (R2 > 0.99, Fig. S4(b)). The k' values which were obtained from the linear
193
regression at different C2H5SH concentrations increased linearly with the C2H5SH
194
concentrations in both the pH ranges (R2 > 0.98, Fig. S5). This indicated that the
195
reaction was first-order in C2H5SH, i.e., n = 1. Moreover, the slopes n obtained from 10
196
the linear regression of logk' versus log[C2H5SH] also confirmed the first-order
197
dependence on C2H5SH (Table S1). Thus, two rate laws for the reaction of Fe(VI)
198
with C2H5SH could be expresses as Eqs. (3) and (4), respectively:
199
- d[Fe(VI)]/dt = k [Fe(VI)]3/2[C2H5SH]
200
- d[Fe(VI)]/dt = k [Fe(VI)][C2H5SH]
201
pH = 8.0–10.0
(3)
pH = 10.5–12.0
(4)
The values of k decrease nonlinearly from (5.29 ± 0.30) × 106 M-3/2 s-1 at pH =
202
8.0 to (1.58 ± 0.06) × 106 M-3/2 s-1 at pH = 10.0 (Fig. 1 and Table S1). This pH
203
dependence could be explained by the acid-base equilibrium of mono-protonated
204
Fe(VI) (HFeO4-) and C2H5SH (Eqs. (5) and (6)).
205
HFeO4 - ↔ H + + FeO42 -
pKa,
HFeO4 -
= 7.23 [32]
206 207
(5) C2H5SH ↔ H + + C2H5S -
pKa,
C2H5SH
= 10.5 [21]
208 209
(6) In the pH range of 8.0–10.0, two Fe(VI) species and two C2H5SH species could
210
react with each other (HFeO4- + C2H5SH, HFeO4- + C2H5S-, FeO42- + C2H5SH and
211
FeO42- + C2H5S-). The kinetic model described by Eqs. (7) and (8) was thus used to
212
stimulate the values of k at pH = 8.0–10.0. - d[Fe(VI)] dt = k [Fe(VI)]3/2 tot [C2H5SH]tot
213 = 214
∑kα ij
3/2 i
βj [Fe(VI)]3/2 tot [C2H5SH]tot
i = 1, 2 j = 1, 2
(7)
11
k
∑kα
= 215
ij
3/2 i
βj
i = 1, 2 j = 1, 2
(8) 216
where [Fe(VI)]tot = [HFeO4-] + [FeO42-], M; [C2H5SH]tot = [C2H5SH] + [C2H5S-], M;
217
αi and βj are the species fraction for Fe(VI) and C2H5SH, respectively; i and j are each
218
of the two Fe(VI) species and two C2H5SH species, respectively; and kij is the species-
219
species rate constant for the reaction of Fe(VI) species with C2H5SH species, M-3/2 s-1.
Measured k Model k
7
kapp (M-3/2 s-1)
10
106 5
10
k21FeO
4
3/2 2-
βC H SH 2
k12HFeO
5
3/2 -
4
k11HFeO
3/2 -
4
4
βC H S 2
-
5
βC H SH 2
5
10
103
7.5
8.0
8.5
9.0
9.5
10.0
pH
220 221
Fig. 1. Apparent rate constants k for the reaction of Fe(VI) with C2H5SH as a function
222
of pH at 25 ± 1 °C.
223
To further simplify the kinetic model, the reaction of FeO42- with C2H5S- was
224
neglected in model fitting due to its negligible contribution to the overall reaction.
225
The critical evaluation of the exclusion of FeO42- reaction with C2H5S- in the model
226
fitting is provided in supplementary material (Text S3). The values of the species-
227
species rate constants were estimated by the least-squares nonlinear regression of the 12
228
measured k values. The results fitted reasonably well with the experimental data (R2
229
> 0.99, solid line in Fig. 1). The estimated species-species rate constants were k11 =
230
(2.11 ± 0.23) × 107 M-3/2 s-1 for the reaction of HFeO4- with C2H5SH, k12 = (1.56 ±
231
0.11) × 1010 M-3/2 s-1 for the reaction of HFeO4- with C2H5S- and k21 = (1.77 ± 0.07) ×
232
106 M-3/2 s-1 for the reactions of FeO42- with C2H5SH, respectively. HFeO4- reacted
233
faster with protonated C2H5SH than FeO42-. Density functional theory calculations
234
showed that HFeO4- had higher oxidizing power than FeO42- owing to its lower
235
LUMO energy level in aqueous solution [33]. The fraction of HFeO4- decreased with
236
increased pH, which might be responsible for the decrease in the rate constants
237
between Fe(VI) and C2H5SH with the increased pH. Moreover, HFeO4- reacted with
238
C2H5S- approximately three orders of magnitude faster than with protonated C2H5SH.
239
This was in agreement with the previous investigations that thiolates, which were
240
orders of magnitudes more reactive as a nucleophile than the corresponding thiols,
241
were much more prone to be oxidized than protonated thiols [34]. Contributions of the
242
specific reactions to the apparent rate constants are shown in Fig. 1 (Dashed lines).
243
While the reaction of HFeO4- with C2H5SH dominated at neutral pH, both the
244
reactions of HFeO4- with C2H5S- and FeO42- with C2H5SH might dominate at pH =
245
7.7–8.6 and pH = 8.6–10.0, respectively.
246
As shown in Table S1, the values of k at pH = 10.5–12.0 were also pH dependent.
247
It slightly decreased from (1.30 ± 0.02) × 104 M-1 s-1 at pH = 10.5 to (0.93 ± 0.03) ×
248
104 M-1 s-1 at pH = 12.0, where FeO42- dominated the reaction. The reaction rates at 13
249
pH = 10.5–12.0 are influenced by two factors of (1) the fraction of C2H5S- with a
250
higher reaction rates than C2H5SH and (2) the oxidizing power of FeO42-. The
251
calculations of redox potential showed the decrease in electrode potential of FeO42-
252
/Fe(OH)3 couple with increased pH in alkaline conditions, which indicated that the
253
oxidizing power of FeO42- decreased as well [35]. Therefore, the decrease of the k
254
value as the increased pH at pH = 10.5–12.0 was attributed to the decreased oxidizing
255
power of FeO42- [25,35].
kapp (M-1 s-1)
40
Measured k Model k
30 20 10 0 7.0
7.5
8.0
8.5
9.0
9.5
10.0
10.5
pH
256 257
Fig. 2. Apparent rate constants k for the reaction of Fe(VI) with C2H5SSC2H5 as a
258
function of pH at 25 ± 1 °C.
259
The kinetics for the oxidation of C2H5SSC2H5, an intermediate of C2H5SH
260
degradation by Fe(VI), was investigated as well. The results indicated first-order
261
dependence on both of Fe(VI) and C2H5SSC2H5 (Fig. S6 and Eq. (9)). The values of k
262
for C2H5SSC2H5 which significantly decrease with the increased pH show similar pH
263
dependence as for C2H5SH (Fig. 2). The analysis of k values was performed using the 14
264
kinetic model of Eqs. (7) and (8), in which the reaction order in Fe(VI) was first-order
265
instead. Two reactions between two Fe(VI) species and C2H5SSC2H5 were considered
266
in model fitting (Text S3). The estimated species-species rate constants for the
267
reaction of HFeO4- with C2H5SSC2H5 and FeO42- with C2H5SSC2H5 were 58.47 (±
268
0.51) M-1 s-1 and 0.55 (± 0.16) M-1 s-1 (R2 > 0.99), respectively. It has been reported
269
that the reaction rate of organic sulfur compounds with Fe(VI) was controlled by the
270
degree of nucleophilicity of the sulfur atom [20,25]. The decreased nucleophilicity of
271
sulfur in C2H5SSC2H5 than C2H5SH was thus responsible for the much lower reaction
272
rate of C2H5SSC2H5. The rate constant for C2H5SSC2H5 with H2O2 was reported to be
273
(1.1 ± 0.4) × 10-4 M-1 s-1 at pH = 9.0 [21]. Hypochlorous acid (HOCl) was proved to
274
have a high reactivity with disulfides, and the rate constant for 3,3'-dithiodipropionic
275
acid with HOCl was (1.6 ± 0.6) × 105 M-1 s-1 at pH = 7.2–7.4 [36,37]. Overall, the
276
reaction rate of Fe(VI) with C2H5SSC2H5 was orders of magnitude higher than the rate
277
for H2O2 in alkaline solution, while probably much lower than the rate for HOCl in
278
neutral solution.
279
- d[Fe(VI)]/dt = k [Fe(VI)][C2H5SSC2H5]
280 281 282
(9) 3.2. Stoichiometry and Products The kinetic results indicated that the reaction of Fe(VI) with C2H5SH showed
283
overall 5/2-order behavior at pH = 8.0–10.0 but second-order behavior at pH = 10.5–
284
12.0. To investigate the possible impact of pH on the intermediates and products of 15
285
C2H5SH oxidation, the stoichiometric experiments were carried out at pH 9.2 and 12.0
286
which were in the pH range of 8.0–10.0 and 10.5–12.0, respectively. The degradation
287
of C2H5SH and its oxidized products by Fe(VI) at pH 9.2 and 12.0 in buffer solutions
288
are shown in Figs. 3 and 4. An increase in the amount of Fe(VI) resulted in a
289
nonlinear reduction of the C2H5SH concentration at both pH = 9.2 and 12.0. The
290
degradation of 1 mol C2H5SH needed ~1.2 mol Fe(VI) at pH = 9.2 but ~1.4 mol Fe(VI)
291
at pH = 12.0, although the final oxidation of C2H5SH had not been achieved. However,
292
2.2 mol Fe(VI) was found to be sufficient to completely convert 1 mol C2H5SH to its
293
final product of C2H5SO3H, but it required more than 2.4 mol Fe(VI) at pH = 12.0.
294
These results also indicated that Fe(VI) was not only consumed by C2H5SH but also
295
by other intermediates from its initial degradation reactions.
296
In the oxidation process, the formation of oxidized products varied with pH and
297
the amount of Fe(VI). At pH = 9.2, C2H5SO3H was gradually formed as a final
298
oxidized product (Fig. S7(a)), and its formation was slow at the beginning of the
299
reaction and then speeded up as the concentration of Fe(VI) raised up to 4.4 mM. In
300
the meantime, two relatively stable intermediates were formed before C2H5SO3H
301
formation, and they were determined to be C2H5SO2H and C2H5SSC2H5 (m/z = 122)
302
by IC and GC/MS, respectively (Fig. S8). As shown in Fig. 3, the amount of
303
C2H5SO2H formation significantly increased with the increased Fe(VI) concentration
304
of up to 2.4 mM, and then decreased with a further increased concentration of Fe(VI).
305
A same trend was also observed for the C2H5SSC2H5 formation. 16
306
To determine if C2H5SO3H could be further oxidized by Fe(VI) to its inorganic
307
form of SO42- at pH = 9.2 and ambient temperature, experiments were conducted by
308
mixing excess Fe(VI) with C2H5SO3H in buffer solutions for more than 12 hours. The
309
initial concentration of C2H5SO3H in the reaction mixture was kept constant at 0.1
310
mM and the concentrations of Fe(VI) were 0.6 and 1.0 mM, respectively. At both
311
conditions, the degradation of C2H5SO3H was found to be less than 5% and no SO42-
312
was obviously detected in the reaction mixture detected by IC. Therefore, C2H5SO3H
313
was probably the final oxidation product of C2H5SH at pH = 9.2 and ambient
314
temperature.
Species (mM)
2.0 pH = 9.2
1.5
1.0
(1)
C2H5SO2H
(2)
C2H5SSC2H5 (3) C2H5SO3H
0.0 0.0
(4)
S - Sum(1 + 2 + 3 + 4)
0.5
315
C2H5SH
1.0
2.0
3.0
4.0
5.0
6.0
7.0
Fe(VI) (mM)
316
Fig. 3. Degradation of C2H5SH and its oxidized products by Fe(VI) at pH = 9.2 in
317
buffer solution. Experimental conditions: [C2H5SH]0 = 2.0 mM, [Fe(VI)] = 0–6.2 mM.
318
At pH = 12.0, C2H5SO2H is a major product equivalent to almost 95% of the
319
consumed sulfur when the concentration of Fe(VI) is 2.8 mM, while C2H5SO3H is a
320
minor product equivalent to only 5% (Figs. 4 and S7(b)). The formation of C2H5SO2H 17
321
increased steadily with an increased Fe(VI) concentration of up to 2.8 mM, and then
322
demonstrated a slight decrease with a further increased Fe(VI). Furthermore,
323
C2H5SSC2H5 was initially formed and then gradually disappeared as the amount of
324
Fe(VI) increased. At both pH values, the sulfur mass balance in the reaction mixture
325
was maintained.
326
Experiments were also carried out at pH = 9.2 and 12.0 without buffer, in which
327
equal volumes of 5.0 mM Fe(VI) and 2.0 mM C2H5SH solutions were rapidly mixed.
328
At pH = 9.2, C2H5SO3H was the only products of C2H5SH. At pH = 12.0, C2H5SO2H
329
was the main product and C2H5SO3H was minor, which were similar with the results
330
obtained in buffer solutions. This indicated that the type of final products for C2H5SH
331
oxidation by Fe(VI) was not influenced by the existence of buffer. The final product
332
of Fe(VI) reduction was detected as Fe(III), in which Fe(OH)3 precipitated after the
333
reaction. Besides, the pH of the reaction mixture without buffer increased from the
334
initial value of 9.2 to 11.0 and from 12.0 to 12.1. 2.0
1.5
1.0
C2H5SH
(1)
1.5
C2H5SO2H
(2)
1.0
C2H5SSC2H5 (3)
0.5
C2H5SO3H
2.0
C2H5SSC2H5 (3)
0.0 0.0
0.5
0.0 0.0
335
Species (×10-2 mM)
Species (mM)
pH = 12.0
1.0
2.0
3.0
4.0
Fe(VI) (mM)
1.0
2.0
3.0
4.0
Fe(VI) (mM)
18
5.0
(4)
S - Sum(1 + 2 + 3 + 4)
336
Fig. 4. Degradation of C2H5SH and its oxidized products by Fe(VI) at pH = 12.0 in
337
buffer solution. Experimental conditions: [C2H5SH]0 = 2.0 mM, [Fe(VI)] = 0–4.7 mM.
338
3.3. Plausible Mechanism
339
The plausible pathways of C2H5SH oxidation by Fe(VI) in aqueous alkaline
340
solution are shown in Fig. 6. Fe(VI) can oxidize thiols through a one-electron transfer
341
or two-electron transfer step to result in different products [20]. Reactions underwent
342
two-electron transfer steps would always yield Fe(II) as a reduced product from
343
Fe(VI), which were contrary to our experimental results. Furthermore, the observed
344
stoichiometric ratios of Fe(VI) to C2H5SH for complete oxidation of C2H5SH to
345
C2H5SO2H and C2H5SO3H were approximately 2.2 and 1.4, respectively. The results
346
were similar to the conclusion that the stoichiometric coefficient of Fe(VI) to organic
347
sulfur compound was 0.67 for a one oxygen-atom transfer when Fe(III) was the
348
reduced product [25]. Therefore, the initial step of C2H5SH oxidation by Fe(VI) might
349
take place to give C2H5S• radical and Fe(V) through a one-electron transfer step (Eq.
350
(10)). Moreover, the detection of C2H5SSC2H5 in the products of C2H5SH oxidation
351
provided additional support for the yield of C2H5S• radical due to the rapidly
352
recombination of alkylthiyl radicals (2kt = ~ 109 M−1 s−1 at pH = 5‒6 [38]). Then,
353
further oxidation of the C2H5S• radicals take place through two possible reactions
354
(Eqs. (11) and (12)), which are further reactions with Fe(VI) to form ethyl sulfenic
355
acid (C2H5SOH) and the dimerization of C2H5S• radicals. This two reactions are
19
356
assumed as the rate-controlling steps leading to the consumption of Fe(VI) and
357
oxidation of C2H5SH.
358
Fe(VI) + C2H5SH → C2H5S• + Fe(V)
359
k1
(10) k2
360
2C2H5S• → C2H5SSC2H5
(11)
361
C2H5S• + Fe(VI) + H2O → Fe(V) + C2H5SOH
k3
362
(12)
363
By applying the steady-state assumption to the concentration of C2H5S• radical,
364
Eq. (13) will be obtained.
365
k1[Fe(VI)][C2H5SH] = k2[C2H5S•]2 + k3[C2H5S•][Fe(VI)][H2O] (13)
366
where k1, k2 and k3 are the rate constants for the reactions of Eqs. (10)‒(12),
367
respectively, M-1 s-1; [Fe(VI)], [C2H5SH], [C2H5S•] and [H2O] represent the
368
concentrations for each of the reactants involved in the reactions of Eqs. (10)‒(12), M.
369 370
371
372 373 374
If k2[C2H5S•]2 ≫ k3[C2H5S•][Fe(VI)][H2O] is a reasonable assumption at pH = 8.0–10.0 when C2H5SH is in excess, Eq. (14) will be obtained. - d[Fe(VI)] dt = k3[C2H5S•][Fe(VI)][H2O] = k4[Fe(VI)]3/2[C2H5SH]1/2 (14) k1 1/2
where k4 = k3[H2O](k ) , M-1 s-1. 2 When [C2H5SH] ≫ [Fe(VI)], for any given concentration of Fe(VI), Eq. (14) reduces to Eq. (15), and this equation is identical with the experimental kinetic 20
375
376
expression for Fe(VI) at pH = 8.0–10.0. - d[Fe(VI)] dt = k5[Fe(VI)]3/2 (15)
377 378
k1 1/2
where k5 = k3[H2O]([C2H5SH]k ) , M-1/2 s-1. 2 If k2[C2H5S•]2 ≪ k3[C2H5S•][Fe(VI)][H2O] is a reasonable assumption at pH
379
= 10.5–12.0, Eq. (16) will be obtained, which is identical with the experimental
380
kinetic expression at pH = 10.5–12.0.
381
- d[Fe(VI)] dt = k3[C2H5S•][Fe(VI)][H2O] = k1[Fe(VI)][C2H5SH] (16)
382
As shown in Figs. 3 and 4, the amount of C2H5SSC2H5 formed at pH = 12.0 was
383
much less than the results at pH = 9.2 as the amount of Fe(VI) increased. It is
384
supposed that the increased pH is disadvantageous for the reaction of Eq. (11), which
385
is in accordance with the assumptions above. The reaction order with respect to
386
C2H5SH in Eq. (14) is inconsistent with our kinetic results that the reaction shows
387
approximate first-order dependence on C2H5SH at pH = 8.0–10.0. However, the
388
results as shown in Table S1 reveal that a reaction order with respect to C2H5SH tends
389
to decrease with a decreased pH value. This is in accordance with the results
390
speculated in Eqs. (14) and (16). At pH = 10.0–10.5, it is likely that the reaction is
391
mixed first and 3/2 order with respect to Fe(VI). No attempt has been made in this
21
392
study to analyze the data obtained in this intermediate pH region, and it needs a
393
further investigation.
394
It has been reported that sulfenic acid does not permit isolation except in very
395
special cases due to its very high reactivity [39], and therefore it would further be
396
oxidized rapidly to form C2H5SO2H. Fe(V) is approximately orders of magnitude
397
more reactive with compounds than Fe(VI) [40]. Hence, it is possible that once Fe(V)
398
formed, it could then be consumed immediately .
399
C2H5SSC2H5, still an odorous pollutant with its odor threshold value ranged from
400
0.05 to 16.5 ppb [2], can be degraded by Fe(VI) to C2H5SO2H. Reaction of Eq. (18)
401
was studied at pH = 9.2 in buffer solution by mixing Fe(VI) with C2H5SSC2H5 at
402
different molar ratios. The initial concentration of C2H5SSC2H5 in the reaction
403
mixture was kept constant at 0.4 mM and the concentrations of Fe(VI) ranged from 0
404
to 0.4 mM. Both of C2H5SO2H and C2H5SO3H were detected as the products of
405
C2H5SSC2H5. With the increased amount of Fe(VI), C2H5SO2H was initially formed
406
and gradually increased, and then C2H5SO3H was formed (Fig. 5). The oxidation of
407
C2H5SSC2H5 might occur in two possible pathways [21,36]. In the first pathway,
408
Fe(VI) oxidizes C2H5SSC2H5 by attacking on its S-S bond to yield C2H5SOH with a
409
cleavage of S-S bond, and then C2H5SOH further reacts with Fe(VI) rapidly to yield
410
C2H5SO2H. In the second pathway, Fe(VI) initially oxidizes C2H5SSC2H5 to diethyl
411
thiolsulfinate (C2H5SOSC2H5) or diethyl thiolsulfonate (C2H5SO2SC2H5) that is
412
analogous to the oxidation of cystine by Fe(VI) to thiosulfonate [27], and then further 22
413
attacks on the S-S bond to form C2H5SOH and C2H5SO2H. However, no trace of
414
C2H5SOSC2H5 and C2H5SO2SC2H5 was obviously detected in the reaction mixture by
415
GC/MS. It was speculated that the reaction might follow the first pathway.
C2H5SO2-
pH = 9.2
C2H5SO3-
Intensity (μS)
[Fe(VI)] = 0.40 mM
[Fe(VI)] = 0.24 mM
[Fe(VI)] = 0.08 mM [Fe(VI)] = 0 mM
2.0
416
2.5
3.0
3.5
4.0
4.5
5.0
5.5
6.0
Time (min)
417
Fig. 5. The IC chromatograms of products in the oxidization of C2H5SSC2H5 by
418
Fe(VI) at pH = 9.2 in buffer solution. Experimental conditions: [C2H5SSC2H5]0 = 0.4
419
mM, [Fe(VI)] = 0–0.4 mM, reaction time 30 min.
420
The C2H5SO2H is also labile and can further react with Fe(VI) to form
421
C2H5SO3H. The variation in fraction of products at different pH in excess of Fe(VI)
422
could be explained by the oxidizing ability of Fe(VI). The presence of oxygen atom
423
decreases the nucleophilicity of sulfur [20,25]. C2H5SO2H is thus a major oxidized
424
product at pH = 12.0. The appearance of C2H5SO3H in Fig. 4 also demonstrated a
425
much slower rate of the oxidation from C2H5SO2H to C2H5SO3H by Fe(VI) at pH =
426
12.0. As two main final products of C2H5SH degradation, both of C2H5SO2H and
427
C2H5SO3H are not classified either toxic or malodorous chemicals and can therefore 23
428
be discharged into the effluent after neutralization [41,42].
429
In our experiments, a molar ratio of Fe(VI) to C2H5SH for degradation of
430
C2H5SH at pH = 9.2 was determined to be ~1.2, which was slightly less than ~1.4 at
431
pH = 12.0. This result was in consistent with the stoichiometric coefficients predicted
432
by Eqs. (17) and (19). Furthermore, a molar ratio of Fe(VI) to C2H5SH for complete
433
oxidation of C2H5SH to C2H5SO3H was determined to be ~2.2 at pH = 9.2, which was
434
similar to the prediction by Eq. (20). The increase of pH in both reaction mixtures
435
with initial pH = 9.2 and 12.0 without buffer can be described by the two reactions of
436
Eqs. (19) and (20), respectively.
437
7HFeO 4- + 6C2H5SH + 10H2O → 7Fe(OH)3 + 5C2H5SO2H + 1/2C2H5SSC2H5
438
+ 7OH - (17)
439 440 441
2HFeO 4- + C2H5SSC2H5 + 4H2O → 2Fe(OH)3 + 2C2H5SO2H + 2OH (18) 4HFeO 4- + 3C2H5SH + 6H2O → 4Fe(OH)3 + 3C2H5SO2H + 4OH -
442 443
(19) 2HFeO 4- + C2H5SH + 3H2O → 2Fe(OH)3 + C2H5SO3H + 2OH -
444
(20)
24
445 446
Fig. 6. The plausible pathways of C2H5SH oxidation by Fe(VI) in aqueous alkaline
447
solution
25
448
3.4. Comparison with Other Oxidants
449
Table 1. Kinetics, major intermediates and products of C2H5SH with different
450
oxidants in aqueous solution at ambient temperature. pH
Kinetics
O3
4.7
3.0 × 105 M-1 s-1
C2H5SO3H
9.2
2.5 × 106 M-3/2 s-1
C2H5SSC2H5,
12.0
0.9 × 104 M-1 s-1
C2H5SO2H, C2H5SO3H
7.0
~ 107 M-1 s-1 a
6.0
4.0 × 104 M-1 s-1
8.0
4.0 × 106 M-1 s-1
KMnO4
7.0
25.9 M-1 s-1
H2O2
11.0
8.4 M-1 s-1
•OH
6.0
Fe(VI)
chlorine
ClO2
451
a
452
protein with HOCl at pH = 7.0.
453
Major intermediates and
Oxidant
products
C2H5SCl, C2H5SO2Cl C2H5SSC2H5, C2H5SO3H
Reference [41] this study
[37,43]
[44] C2H5SO3H C2H5SOH, C2H5SSC2H5, C2H5SO3H CH3COOH, SO42-
[45] [21,46] [47]
Speculated from the apparent rate constants of sulfur-containing amino acids and
The kinetics, major intermediates and products of C2H5SH degradation by Fe(VI)
454
are compared with other selective and nonselective oxidants (Table 1). As selective
455
oxidants, Fe(VI), O3 , ClO2 and chlorine have high reactivity with C2H5SH. C2H5SH
456
can be rapidly oxidized by Fe(VI) to C2H5SO2H and C2H5SO3H in alkaline solution.
457
The Fe(VI) is finally reduced to a non-toxic chemical of Fe(OH)3 which can act as an
458
effective coagulant in water and wastewater treatment .O3 is an electrophile with high
459
selectivity and unstable in water with a half-life time in the range of seconds to hours
460
depending on the water quality [48,49]. The oxidation of C2H5SH by O3 in acid 26
461
solution is fast resulting in the formation of C2H5SO3H, and the rate constant was
462
found to be independent on pH in the pH range of 0.85 to 4.7 [41]. The rate constant
463
of C2H5SH with ClO2 has been reported to increase exponentially with an increase in
464
pH and estimated as 4.0 × 104 M-1 s-1 and 4.0 × 106 M-1 s-1 at pH = 6 and 8,
465
respectively [44]. However, the dosage of ClO2 is restricted in water treatment
466
because of the blood poison of its reduction products. Reduced sulfur compound (e.g.,
467
sulfhydryl compound and disulfide) could easily be oxidized by chlorine with the rate
468
constants ranged from ~105 M-1 s-1 to ~107 M-1 s-1 in aqueous solution. Chlorine
469
initially oxidizes thiols to yield sulfenyl chloride as an intermediate. And then, the
470
formation of products depends on the chlorination condition, including disulfide,
471
sulfonic acid, and sulfonyl chloride [37]. However, chlorination has potential to form
472
toxic halogenated byproducts.
473
In comparison, the rate constants for KMnO4 and H2O2 are orders of magnitude
474
lower than for Fe(VI), O3 , ClO2 and chlorine. C2H5SH can be oxidized by KMnO4 to
475
C2H5SO3H, but the use of KMnO4 may cause secondary pollution. The rate constants
476
for the oxidation of low molecular weight thiols by H2O2 in aqueous solution were
477
reported in the range of ~10 M-1 s-1 [34]. And the rate constants calculated for thiolate
478
ions were similar and pKa-independent for considered low molecular weight thiols
479
[50]. Oxidation of C2H5SH by H2O2 was proved to be a nucleophilic substitution of
480
C2H5S- on H2O2 with the formation of transient C2H5SOH as the rate determining step
481
[34]. Then, the coupling product of C2H5SSC2H5 would undergo further but much 27
482 483
slower oxidation with H2O2 to give higher oxidation products [21,46]. •OH is a nonselective oxidant with high reactivity and reacts indiscriminately
484
with all kinds of matrix compounds in water [51,52]. It might be an only oxidant that
485
could oxidize C2H5SH to its inorganic form of SO42- in aqueous solution as shown in
486
Table 1. Ma et al. has shown that •OH generated from the modified β-PbO2 anode
487
could react with C2H5SH to produce ethyl radical (•C2H5) and thiyl radical (•HS) by
488
successively attacking on the C-S bond of C2H5SH. The radicals would further react
489
with •OH through a series of reactions and eventually convert to SO42- and
490
CH3COOH, which could be subsequently mineralized [47]. Generally, acidic and
491
neutral pH conditions are required for the generation of •OH in the reactions such as
492
Fenton reaction, photo-catalytic reaction as well as electrochemical system [52,53].
493
However, the acidic and neutral pH is negative for the mass transfer of C2H5SH from
494
gaseous phase to liquid phase in the chemical scrubbing process due to the weak
495
acidity of C2H5SH. With the integrated function of oxidation and coagulation, Fe(VI)
496
is relative more stable in alkaline condition than acidic condition. Therefore, Fe(VI)
497
may be a more suitable oxidant for C2H5SH degradation in the chemical scrubbing
498
technique.
499
4. Conclusions
500
This study demonstrated that C2H5SH could be effectively degraded by Fe(VI) in
501
aqueous alkaline solution. The reaction showed overall 5/2-order behavior at pH =
502
8.0–10.0 and second-order behavior at pH = 10.5–12.0, and the rate constants 28
503
decreased with increased pH. C2H5SH could be finally oxidized to nonvolatile and
504
odorless C2H5SO3H. The stoichiometric ratio of Fe(VI) to C2H5SH for C2H5SH
505
oxidation to C2H5SO3H was determined as 2.0. The plausible mechanism of C2H5SH
506
degradation by Fe(VI) was proposed, in which the two competing reaction pathways
507
for C2H5S• radical were probably responsible for the changes in reaction order at
508
different pH. Compared with other selective and nonselective oxidants, Fe(VI) was
509
revealed to be an efficient and environmental friendly oxidant for C2H5SH
510
degradation. However, challenges still exist in the practical application of Fe(VI) due
511
to the high production cost of solid Fe(VI) products and the instability of Fe(VI)
512
solution. Therefore, the chemical scrubbing process with in situ generation and
513
application of Fe(VI) could be a suitable process to eliminate odors. The odorants
514
could first be absorbed to the liquid phase and then oxidized by the in situ generated
515
Fe(VI) quite rapidly. This may lead to the practical implementation of Fe(VI)
516
technology in odorous gas treatment.
517
Acknowledgements
518
This work was financially supported by State Key Laboratory of Urban Water
519
Resource and Environment (Harbin Institute of Technology) (2014DX08).
520 521
References
522
[1] A. Talaiekhozani, M. Bagheri, A. Goli, and M.R. Talaei Khoozani, An overview of principles
523
of odor production, emission, and control methods in wastewater collection and treatment
524
systems. J. Environ. Manage. 170 (2016) 186-206. 29
525 526
[2] E.C. Sivret, B. Wang, G. Parcsi, and R.M. Stuetz, Prioritisation of odorants emitted from sewers using odour activity values. Water Res. 88 (2016) 308-321.
527
[3] M.B. Jaber, B. Anet, A. Amrane, C. Couriol, T. Lendormi, P.L. Cloirec, G. Cogny, and R.
528
Fillières, Impact of nutrients supply and pH changes on the elimination of hydrogen sulfide,
529
dimethyl disulfide and ethanethiol by biofiltration. Chem. Eng. J. 258 (2014) 420-426.
530
[4] E. Agus, M.H. Lim, L. Zhang, and D.L. Sedlak, Odorous compounds in municipal
531
wastewater effluent and potable water reuse systems. Environ. Sci. Technol. 45 (2011) 9347-
532
9355.
533
[5] H. Tan, Y. Zhao, Y. Ling, Y. Wang, and X. Wang, Emission characteristics and variation of
534
volatile odorous compounds in the initial decomposition stage of municipal solid waste.
535
Waste Manage. 68 (2017) 677-687.
536
[6] C. Meusinger, A.B. Bluhme, J.L. Ingemar, A. Feilberg, S. Christiansen, C. Andersen, and
537
M.S. Johnson, Treatment of reduced sulphur compounds and SO2 by Gas Phase Advanced
538
Oxidation. Chem. Eng. J. 307 (2017) 427-434.
539
[7] W. Lu, Z. Duan, D. Li, L.M.C. Jimenez, Y. Liu, H. Guo, and H. Wang, Characterization of
540
odor emission on the working face of landfill and establishing of odorous compounds index.
541
Waste Manage. 42 (2015) 74-81.
542
[8] M. Schiavon, L.M. Martini, C. Corrà, M. Scapinello, G. Coller, P. Tosi, and M. Ragazzi,
543
Characterisation of volatile organic compounds (VOCs) released by the composting of
544
different waste matrices. Environ. Pollut. 231 (2017) 845-853.
545
[9] Y. Son, Decomposition of VOCs and odorous compounds by radiolysis: A critical review.
30
546 547 548
Chem. Eng. J. 316 (2017) 609-622. [10] T. Liu, X. Li, and F. Li, Development of a photocatalytic wet scrubbing process for gaseous odor treatment. Ind. Eng. Chem. Res. 49 (2010) 3617-3622.
549
[11] E. Vega, M.J. Martin, and R. Gonzalez-Olmos, Integration of advanced oxidation processes
550
at mild conditions in wet scrubbers for odourous sulphur compounds treatment.
551
Chemosphere 109 (2014) 113-119.
552
[12] J. Zeng, L. Hu, X. Tan, C. He, Z. He, W. Pan, Y. Hou, and D. Shu, Elimination of methyl
553
mercaptan in ZVI-S2O82- system activated with in-situ generated ferrous ions from zero
554
valent iron. Catal. Today 281 (2017) 520-526.
555 556
[13] V.K. Sharma, R. Zboril, and R.S. Varma, Ferrates: greener oxidants with multimodal action in water treatment technologies. Accounts Chem. Res. 48 (2015) 182-191.
557
[14] W. Yu, Y. Yang, and N. Graham, Evaluation of ferrate as a coagulant aid/oxidant
558
pretreatment for mitigating submerged ultrafiltration membrane fouling in drinking water
559
treatment. Chem. Eng. J. 298 (2016) 234-242.
560
[15] V.K. Sharma, L. Chen, and R. Zboril, Review on high valent FeVI (Ferrate): a sustainable
561
green oxidant in organic chemistry and transformation of pharmaceuticals. ACS Sustain.
562
Chem. Eng. 4 (2016) 18-34.
563
[16] Y. Lee, S.G. Zimmermann, A.T. Kieu, and U. von Gunten, Ferrate (Fe(VI)) Application for
564
Municipal Wastewater Treatment: A Novel Process for Simultaneous Micropollutant
565
Oxidation and Phosphate Removal. Environ. Sci. Technol. 43 (2009) 3831-3838.
566
[17] C. He, X. Li, and V.K. Sharma, Elimination of sludge odor by oxidizing sulfur-containing
31
567 568 569 570 571 572 573
compounds with ferrate(VI). Environ. Sci. Technol. 43 (2009) 5890-5895. [18] L. Ding, H. Liang, and X. Li, Oxidation of CH3SH by in situ generation of ferrate(VI) in aqueous alkaline solution for odour treatment. Sep. Purif. Technol. 91 (2012) 117-124. [19] E. Yang, J. Shi, and H. Liang, On-line electrochemical production of ferrate (VI) for odor control. Electrochim. Acta 63 (2012) 369-374. [20] V.K. Sharma, G.W.L. III, and F.J. Millero, Mechanisms of oxidation of organosulfur compounds by ferrate(VI). Chemosphere 82 (2011) 1083-1089.
574
[21] C. Feliers, L. Patria, J. Morvan, and A. Laplanche, Kinetics of oxidation of odorous sulfur
575
compounds in aqueous alkaline solution with H2O2. Environ. Technol. 22 (2001) 1137-1146.
576
[22] J. Shin, U. von Gunten, D.A. Reckhow, S. Allard, and Y. Lee, Reactions of ferrate(VI) with
577
iodide and hypoiodous acid: kinetics, pathways, and implications for the fate of iodine during
578
water treatment. Environ. Sci. Technol. 52 (2018) 7458-7467.
579 580 581 582 583 584
[23] J. Shin, D. Lee, T. Hwang, and Y. Lee, Oxidation kinetics of algal-derived taste and odor compounds during water treatment with ferrate(VI). Chem. Eng. J. 334 (2018) 1065-1073. [24] V.K. Sharma, Oxidation of inorganic compounds by ferrate(VI) and ferrate(V): one-electron and two-electron transfer steps. Environ. Sci. Technol. 44 (2010) 5148-5152. [25] V.K. Sharma, Ferrate(VI) and ferrate(V) oxidation of organic compounds: Kinetics and mechanism. Coordin. Chem. Rev. 257 (2013) 495-510.
585
[26] J.F. Read, E.K. Adams, H.J. Gass, S.E. Shea, and A. Theriault, The kinetics and mechanism
586
of oxidation of 3-mercaptopropionic acid, 2-mercaptoethanesulfonic acid and 2-
587
mercaptobenzoic acid by potassium ferrate. Inorg. Chim. Acta 281 (1998) 43-52.
32
588
[27] J.F. Read, S.A. Bewick, and C.R. Graves, The kinetics and mechanism of the oxidation of S-
589
methyl-L-cysteine, L-cystine and L-cysteine by potassium ferrate. Inorg. Chim. Acta 303
590
(2000) 244-255.
591 592 593 594
[28] C. Li, X.Z. Li, and N. Graham, A study of the preparation and reactivity of potassium ferrate. Chemosphere 61 (2005) 537-543. [29] J.M. Schreyer, G.W. Thompson, and L.T. Ockerman, Oxidation of chromium(lll) with potassium ferrate(VI). Anal. Chem. 22 (1950) 1426-1427.
595
[30] Z. Luo, M. Strouse, J. Jiang, and V.K. Sharma, Methodologies for the analytical
596
determination of ferrate(VI): a Review. J. Environ. Sci. Health, Part A Toxic/Hazard. Subs.
597
Environ. Eng. 46 (2011) 453-460.
598 599 600 601
[31] C.K. Riener, G. Kada, and H.J. Gruber, Quick measurement of protein sulfhydryls with Ellman's reagent and with 4,4'-dithiodipyridine. Anal. Bioanal. Chem. 373 (2002) 266-276. [32] M. Feng, and V.K. Sharma, Enhanced oxidation of antibiotics by ferrate(VI)-sulfur(IV) system: Elucidating multi-oxidant mechanism. Chem. Eng. J. 341 (2018) 137-145.
602
[33] R. Sarma, A.M. Angeles-Boza, D.W. Brinkley, and J.P. Roth, Studies of the di-iron(VI)
603
intermediate in ferrate-dependent oxygen evolution from water. J. Am. Chem. Soc. 134
604
(2012) 15371-15386.
605
[34] A. Zeida, R. Babbush, M.C. González Lebrero, M. Trujillo, R. Radi, and D.A. Estrin,
606
Molecular basis of the mechanism of thiol oxidation by hydrogen peroxide in aqueous
607
solution: challenging the SN2 paradigm. Chem. Res. Toxicol. 25 (2012) 741-746.
608
[35] Y. Wang, H. Liu, and G. Liu, Oxidation of diclofenac by potassium ferrate (VI): reaction
33
609
kinetics and toxicity evaluation. Sci. Total Environ. 506-507 (2015) 252-258.
610
[36] D.I. Pattison, and M.J. Davies, Absolute rate constants for the reaction of hypochlorous acid
611
with protein side chains and peptide bonds. Chem. Res. Toxicol. 14 (2001) 1453-1464.
612
[37] M. Deborde, and U. von Gunten, Reactions of chlorine with inorganic and organic
613
compounds during water treatment - kinetics and mechanisms: a critical review. Water Res.
614
42 (2008) 13-51.
615 616 617 618 619 620
[38] Morton Z. Hoffman, and E. Hayon, Pulse radiolysis study of sulfhydryl compounds in aqueous solution. J. Phys. Chem. 77 (1973) 990-996. [39] G. Capozzi, and G. Modena, Oxidation of thiols. in: S. Patai, (Ed.), The chemistry of the thiol group, John Wiley & Sons, Ltd., 1974, pp. 785-839. [40] V.K. Sharma, D.B. O'Connor, and D.E. Cabelli, Sequential One-Electron Reduction of Fe(V) to Fe(III) by Cyanide in Alkaline Medium. J. Phys. Chem. B 105 (2001) 11529-11532.
621
[41] K. Kirchner, and W. Litzenburger, Oxidising scrubbing of gas using ozone-reaction and
622
absorption kinetics of the ozone-ethyl mercaptan system. Chem. Eng. Sci. 37 (1982) 948-950.
623
[42] Y.G. Adewuyi, Oxidation of biogenic sulfur compounds in aqueous media: kinetics and
624
environmental implications. in: E.S. Saltzman, and W.J. Cooper, (Eds.), Biogenic Sulfur in
625
the Environment, ACS Symposium Series, American Chemical Society: Washington, DC,
626
1989, pp. 529-559.
627
[43] M.G. Conti-Ramsden, N.W. Brown, and E.P.L. Roberts, Towards an odour control system
628
combining slurry sorption and electrochemical regeneration. Chem. Eng. Sci. 79 (2012) 219-
629
227.
34
630
[44] J.R. Kastner, K.C. Das, C. Hu, and R. McClendon, Effect of pH and temperature on the
631
kinetics of odor oxidation using chlorine dioxide. J. Air Waste Manage. 10 (2003) 1218-1224.
632
[45] Y. Liu, X. Zhang, J. Dai, and H. Xu, Kinetics on ethanethiol oxidation by potassium
633 634 635 636 637 638 639 640 641 642 643
permanganate in drinking water. Environ. Sci. 29 (2008) 1261-1265. [46] D.W. Giles, J.A. Cha, and P.K. Lim, The aerobic and peroxide-induced coupling of aqueous thiols-I. kinetic results and engineering significance. Chem. Eng. Sci. 41 (1986) 3129-3140. [47] X. Ma, Z. Wu, M. Zhou, and J. Ding, Electrochemical scission of C-S bond in ethanethiol on a modified β-PbO2 anode in aqueous solution. Sep. Purif. Technol. 109 (2013) 72-76. [48] U. von Gunten, Ozonation of drinking water: Part I. Oxidation kinetics and product formation. Water Res. 37 (2003) 1443-1467. [49] M. Mehrjouei, S. Müller, and D. Möller, A review on photocatalytic ozonation used for the treatment of water and wastewater. Chem. Eng. J. 263 (2015) 209-219. [50] C.C. Winterbourn, and D. Metodiewa, Reactivity of biologically important thiol compounds with superoxide and hydrogen peroxide. Free Radical Bio. Med. 27 (1999).
644
[51] Y. Lee, and U. von Gunten, Oxidative transformation of micropollutants during municipal
645
wastewater treatment: Comparison of kinetic aspects of selective (chlorine, chlorine dioxide,
646
ferrateVI, and ozone) and non-selective oxidants (hydroxyl radical). Water Res. 44 (2010)
647
555-566.
648
[52] M. Feng, Z. Wang, D.D. Dionysiou, and V.K. Sharma, Metal-mediated oxidation of
649
fluoroquinolone antibiotics in water: A review on kinetics, transformation products, and
650
toxicity assessment. J. Hazard. Mater. 344 (2018) 1136-1154.
35
651
[53] G. Wang, Q. Chen, Y. Liu, D. Ma, Y. Xin, X. Ma, and X. Zhang, In situ synthesis of
652
graphene/WO3 co-decorated TiO2 nanotube array photoelectrodes with enhanced
653
photocatalytic activity and degradation mechanism for dimethyl phthalate. Chem. Eng. J. 337
654
(2018) 322-332.
655
Graphical abstract
C2H5SH + Fe(VI) pH-dependent kinetics
C2H5SH
Fe(VI)
C2H5SSC2H5 C2H5S •
Fe(VI)
C2H5SO3H
C2H5SOH pH = 10.5 – 12.0
pH = 8.0 – 10.0
d[Fe(VI)] k [Fe(VI)][C2H5SH] dt
d[Fe(VI)] k [Fe(VI)]3/2[C2H5SH] dt
656 657 658
Highlights
659
C2H5SH can be degraded rapidly by Fe(VI) in aqueous alkaline solution.
660
The reaction kinetics and fraction of intermediates are highly dependent on pH.
661
The reaction shows a 3/2-order dependence on Fe(VI) at pH = 8.0~10.0.
662
Nonvolatile and odorless C2H5SO3H is the final product of C2H5SH degradation.
663
A proposed mechanism successfully explains the pH-dependent reaction order.
664
36