1

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 35

Accepted Manuscript

Title: Degradation of carbon tetrachloride in aqueous solution


in the thermally activated persulfate system<!–<query
id="Q1"> “Your article is registered as a regular item and is
being processed for inclusion in a regular issue of the journal.
If this is NOT correct and your article belongs to a Special
Issue/Collection please contact p.d.thangavelu@elsevier.com
immediately prior to returning your corrections.”</query>–>

Author: Minhui Xu Xiaogang Gu Shuguang Lu Zhaofu Qiu


Qian Sui Zhouwei Miao Xueke Zang Xiaoliang Wu

PII: S0304-3894(14)01015-2
DOI: http://dx.doi.org/doi:10.1016/j.jhazmat.2014.12.031
Reference: HAZMAT 16466

To appear in: Journal of Hazardous Materials

Received date: 8-10-2014


Revised date: 10-12-2014
Accepted date: 18-12-2014

Please cite this article as: Minhui Xu, Xiaogang Gu, Shuguang Lu, Zhaofu Qiu, Qian
Sui, Zhouwei Miao, Xueke Zang, Xiaoliang Wu, Degradation of carbon tetrachloride
in aqueous solution in the thermally activated persulfate system, Journal of Hazardous
Materials http://dx.doi.org/10.1016/j.jhazmat.2014.12.031

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.
Degradation of carbon tetrachloride in aqueous solution in the thermally

activated persulfate system

Minhui Xu, Xiaogang Gu, Shuguang Lu*, Zhaofu Qiu, Qian Sui, Zhouwei Miao,

Xueke Zang, Xiaoliang Wu


State Environmental Protection Key Laboratory of Environmental Risk Assessment and Control

on Chemical Process, East China University of Science and Technology,

Shanghai 200237, China

PT
*Corresponding author. Tel.: +86 21 64250709, Fax: +86 21 64252737.

E-mail: lvshuguang@ecust.edu.cn (S. Lu)

RI
Highlights
SC
U
N

Carbon tetrachloride can be readily degraded in the thermally activated persulfate


A
M

system.

The predominant radical species responsible for CT degradation was identified.


D
TE

Effects of initial persulfate, carbon tetrachloride and solution matrix were studied.

Effects of solvents addition to the solution were studied.


EP

Abstract
CC

Thermal activation of persulfate (PS) has been identified to be effective in the


A

destruction of organic pollutants. The feasibility of carbon tetrachloride (CT)

degradation in the thermally activated PS system was evaluated. The experimental

results showed that CT could be readily degraded at 50℃ with a PS concentration of

0.5 M, and CT degradation and PS consumption followed the pseudo-first order


kinetic model. Superoxide radical anion (O2•-) was the predominant radical species

responsible for CT degradation and the split of C-Cl was proposed as the possible

reaction pathways for CT degradation. The process of CT degradation was accelerated

by higher PS dose and lower initial CT concentration. No obvious effect of the initial

pH on the degradation of CT was observed in the thermally activated PS system. Cl-,

HCO3-, and humic acid (HA) had negative effects on CT degradation. In addition, the

degradation of CT in the thermally activated PS system could be significantly

PT
promoted by the solvents addition to the solution. In conclusion, the thermally

RI
activated PS process is a promising option in in-situ chemical oxidation/reduction

SC
remediation for degrading highly oxidized organic contaminants such as CT that is
U
widely detected in contaminated sites.
N
A

Keywords:
M

Persulfate, Carbon tetrachloride, Thermal activation, Superoxide radical anion,


D

Groundwater
TE
EP

1. Introduction
CC

Chlorinated volatile organic compounds are widely used in the chemical industry

in the past several applications as lubricants, heat transfer fluids, plasticizers, cleaning
A

fluids, degreasing agents, spot removers, and cleaning solvents. Carbon tetrachloride

(CT) is one of the most commonly used among the chlorinated volatile compounds

and has been widely observed in the National Priority List (NPL) sites of USA [1]. CT
is an onerous contaminant due to its carcinogenicity, environmental persistence, and

relatively high solubility in water (805 mg/L at 20°C) [2]; it is commonly found in

groundwater and surface water at concentrations exceeding the maximum

contaminant level of 5 µg/L [3]. In situ chemical oxidation (ISCO) is a fast, cost

effective and viable remediation technology for the treatment of a wide range of lower

molecular weight organic contaminants, especially in groundwater [4]. As with

aqueous-based ISCO system, selections of oxidant and novel activators are the key for

PT
the chemical oxidation technology. Four main oxidants such as permanganate

RI
(MnO4-), Fenton's reagent (H2O2), ozone (O3), and persulfate (PS, S2O82-) are often

SC
used at ISCO sites. Permanganate can be directly injected into the subsurface, and no
U
other activation is needed. However, the reactivity of permanganate limits the range
N

of degradable contaminants, and permanganate can also be consumed by natural


A

organic matter. Ozone is reactive with a larger range of contaminants, but is limited
M

by its mass transfer, solubility and stability. Among the oxidants, PS and Fenton’s
D

reagent are two of the effective oxidants which are mostly being used for ISCO in the
TE

remediation of contaminated soil and groundwater [4,5]. H2O2 is activated by the


EP

catalyst to promote hydroxyl radical (HO•, E0 = 2.7 V) formation, which is


CC

instrumental in the destruction of contaminants. In addition, it can readily react with

organic contaminants, and no cases could be found where the excessive persistence of
A

H2O2. However, H2O2 is not a long-lived species and it rapidly decomposes in

contaminated sites. Especially, in recent years, PS-based ISCO remediation has been

proven to be the very promising technique for the treatment of refractory pollutants.
Compared with HO• mostly used in ISCO based on Fenton’s reagent, PS has been

recently proven to be a powerful oxidant that can produce sulfate radicals (SO4•-, E0 =

2.6 V) by various activation such as heat, transition metals, UV, or ultrasound

activation [6-9]. The PS anion is a strong, two-electron oxidizing agent with a redox

potential (E0 = 2.01 V), and it can be transported over a longer distance in the

subsurface. Under thermally-enhanced conditions (i.e., temperature of 40-100℃), PS

could be converted to a much powerful oxidant of SO4•- with an unpaired electron.

PT
SO4•- is a more aggressive oxidizing agent than the PS anion [6]. Owing to its strong

RI
oxidizing potential, thermally activated PS system has been applied for the treatment

SC
of many pollutants, such as 1,1,1-trichloroethane, trichloroethylene, phenol,
U
methylene blue, antipyrine, and diuron [8,10,11,12,13,14]. Among the transition
N

metals, ferrous iron has been widely used to activate PS since it is relatively
A

inexpensive, nontoxic and effective [15]. Iron activated PS system must be carried out
M

under acidic condition since ferric ions would be precipitated as hydroxide at pH > 5.
D

This is undesirable in the practical application in which a weak acidic or neutral pH


TE

may be preferred. Moreover, the excessive Fe2+ can act as a scavenger of SO4•- and
EP

iron sludge generated during the treatment might block the transportation of reagents.
CC

The use of zero valent iron would overcome the disadvantages. Zero valent iron is

used as an alternative source of ferrous ion, and recycling ferric ion to ferrous ion
A

which can avoid the accumulation of excess ferrous iron and reduce the production of

iron sludge [16,17].

Thermally activated PS system has been carried out to effectively remove the
toxic chlorinated hydrocarbons such as trichloroethane and trichloroethylene in

oxidizing environments [8,10]. However, to the best of our knowledge, the

degradation of perchlorinated hydrocarbons in the thermally activated PS system has

never been studied. Therefore, the objectives of this study are (1) to investigate the

performance of CT degradation in the thermally activated PS system; (2) to identify

the main activated oxygen species responsible for CT degradation; and (3) to evaluate

the influence of the solution pH, solution matrix, and solvents addition on CT

PT
degradation performance.

RI
2. Materials and methods
SC
U
2.1. Materials.
N

Sodium persulfate (98.0%), tert-butyl alcohol (TBA, 99.0%), isopropanol (IPA,


A

99.7%), sodium nitrate (99.5%), sodium bicarbonate (99.5%), sodium chloride


M

(99.5%), methanol (99.9%), ethanol (99.9%), acetone (99.5%) and potassium iodide
D

(99.0%) were purchased from Shanghai Jingchun Reagent Co. Ltd. (Shanghai, China).
TE

Carbon tetrachloride (CT, 99.5%) and n-Hexane (97%) were purchased from
EP

Shanghai Lingfeng Chemical Reagent Co. Ltd. 1, 4-benzoquinone (C6H4O2, BQ,


CC

97%), and humic acid (HA, fulvic acid > 90%) were purchased from Aaladdin

Reagent Ltd Co. (Shanghai, China). Ultrapure water from a Milli-Q water process
A

(Classic DI, ELGA, Marlow, U.K.) was used for preparing aqueous solutions.

2.2. Experimental procedures.

All reactions were conducted in a series of 24 mL borosilicate vials capped with


polytetrafluoroethylene (PTFE) lined septa under controlled temperature (50℃). CT

and PS solutions were prepared by dissolving reagents at room temperature into

Milli-Q water to make stock solutions. Each stock solution was added to the

volumetric flask at desired concentration, then a series of reaction vials was fully

filled. Control tests were carried out in parallel without PS addition. At each

designated time the samples were removed from the reaction vials, chilled to 4℃ in

an ice bath for 5 min to quench the reaction and then analyzed. The initial pH in all

PT
tests was unadjusted except in the test for investigating the influence of pH. All

RI
experiments were conducted in duplicate and the mean values reported.

SC
In order to identify the dominating reactive oxygen species for CT degradation in
U
the thermally activated PS system, radical scavenger studies were performed. HO•
N

were scavenged in the system using tert-butyl alcohol (kHO• = 5.2 × 108 M-1s-1), which
A

is unreactive with SO4•- (kSO4•- ≤ 1 × 106 M-1s-1) [18,19]. SO4•- and HO• were
M

scavenged using isopropanol as no compound is available for scavenging SO4•- alone


D

(kOH• = 8.2 × 107 M-1s-1 and kSO4•- = 8.2 × 107 M-1s-1) [19,20]. To evaluate the role of
TE

the solvated electrons, nitrate was used as the scavenger of solvated electron that
EP

reacts rapidly with solvated electrons (ke = 9.7 × 109 M-1s-1) but not with other
CC

reductants [19]. In order to investigate the role of O2•-, benzoquinone was added to

the CT solution since benzoquinone is a scavenger of O2•- through rapid electron


A

transfer to generate benzoquinone radicals according to the Eq. 1 [21]:

BQ + O2•- →BQ•- + O2 k = 9.6 × 108 M-1s-1 (1)

2.3 Analytical methods.


Aqueous samples (1 mL) were analyzed following extraction with hexane (1 mL) for

3 min using a vortex stirrer and standing for 5 min for separation. The organic phase

(CT in hexane) was then transferred to a 2-mL GC vial with a plastic dropper. CT was

analyzed using a gas chromatograph (Agilent 7890A, Palo Alto, CA) equipped with

an electron capture detector, an autosampler (Agilent 7693), and an DB-VRX column

(60 m length, 250 µm i.d., and 1.4 µm thickness). The temperatures of the injector and

detector were 240℃ and 260℃, respectively, and the oven temperature was held

PT
constant at 100℃. The volatile organic intermediates formed in CT degradation

RI
experiments were identified by the EPA SW-846 Method 8260B using an automatic

SC
purge and trap (Tekmar Atomx, Mason, OH) coupled to a GC/MS (Agilent 7890/5975)
U
and with the same DB-VRX column. The concentration of PS was determined by a
N

spectrophotometric method using potassium iodide [22].


A
M

3. Results and Discussion


D

3.1. Performance of CT degradation in thermally activated PS system


TE

3.1.1. Degradation of CT in thermally activated PS system


EP

The removal of CT under 50℃ by PS and PS decomposition are presented in


CC

Fig. 1. The results show that 9.5% CT lost in 60 h indicating that no significant CT

loss due to thermolysis and/or volatilization in a batch reactor. However, more than
A

96% CT degraded by PS dose of 0.5 M in the thermally activated system in 60 h.

Statistically, the degradation of CT fitted well to the pseudo-first order kinetic model.

Under pseudo-first order kinetics, the rate of CT degradation is directly proportional


to the amount of un-degraded CT, which can be expressed as follow:

- dCi/dt = kobsCi (2)

Eq. 2 can be modified as Eq. 3:

ln(Ci/C0) = −kobst (3)

where Ci is the concentration of CT at time t, C0 is the initial concentration of CT, and

kobs is the pseudo-first order reaction rate constant, which can be calculated from the

slope of the line after plotting ln(C0/Ci) against time [23]. As shown in Fig. 1, kobs was

PT
calculated to be 0.034 h-1 and CT degradation performance could be well fitted to the

RI
pseudo-first order kinetic model, as indicated by high correlation coefficient of 0.97

as shown in Table 1.
SC
U
Huang et al. investigated the degradation of CT by sodium persulfate (1 g/L) at
N

40℃. The two replicate experiments indicated that PS in a deionized water matrix
A

was not able to degrade CT. The discrepancies in results between the early work and
M

this study were due to the different conditions applied in the tests. The degradation of
D

CT may be highly related to PS concentration in thermally activated PS system [24].


TE

Therefore, the concentration of PS may be the most key factor for degradation of CT.
EP

Fig. 1 also showed the PS decomposition during the degradation of CT in the


CC

whole 60 h reaction period. The decomposition of PS followed pseudo-first order

reaction kinetic model, as commonly reported in the literature [10,25,26,27]. The


A

values of kobs,PS obtained from this study for PS decomposition with or without CT are

roughly the same as 0.014 h-1, which was similar to the value reported by Johnson's

experiment [28], indicating that CT would not directly react with PS anions. The
reactive oxygen species generated from the PS decomposition may therefore be

deduced to be responsible for the degradation of CT.

PS decomposition pathways involve two possible reaction occurred

simultaneously. Under thermally enhanced condition, there is considerable evidence

that the peroxide bond of PS is broken to generate two SO4•-, as shown in Eq. 4. Then

a series of chain reactions will be initiated in the thermally activated PS system [29].

Another possible PS decomposition pathway is PS hydrolysis. In acid solution, PS

PT
hydrolysis to form HSO5- and this unstable intermediate will further decompose in

RI
acid solution to form H2O2 and HSO4-, as shown in Eqs. 5-6 [30]. Due to the

SC
formation of acid caused by the decomposition of PS, the solution pH decreased from
U
3.21 to 1.02 during the reaction period.
N

S2O82- + heat → 2SO4•- (4)


A

S2O82- + H2O → HSO5- + HSO4- (5)


M

HSO5- + 2H2O → H2O2 + HSO4- (6)


D

3.1.2. Identification of the reactive oxygen species responsible for CT


TE

degradation
EP

Once PS is activated, the resulting SO4•- could initiate a chain of reactions


CC

involving other radicals and oxidants generation (Eqs. 7-12) [31-35]. HO• can be

formed via Eqs. 7-8, and may participate in contaminant oxidation. HO• undergoes
A

hydrogen abstraction and addition more readily, while SO4•- is much prone to electron

transfer reactions. Because of a higher redox potential and the less-selective

contaminant destruction of HO•, the formation of HO• presents the potential to


increase the contaminant degradation rate and the possibility of reactions with a wider

variety of compounds. In addition, perhydroxyl radical (HO2•, E0 = 1.7 V) would be

generated by radical involving reactions between H2O2 and HO•/SO4•- (Eqs. 10-11).

Although HO2• is not a reductant and is only a weak oxidant [36], superoxide radical

anion (O2•-, E0 = -2.4 V) could be formed from further propagation reaction as shown

in Eq. 12. Teel and Watts studied the degradation of CT by modified Fenton's reagent

[37], and demonstrated that CT was degraded through a reductive mechanism. There

PT
are other sulfur based radicals that have been noted to be important in radical

RI
reactions with sulfite and peroxymonosulfate reaction systems, including SO5•- and

SC
SO3•- [38,39], but their occurrence or importance in PS based in-situ chemical
U
oxidation/reduction system is not yet clear.
N

All pH: SO4•- + H2O → HO• + H+ + SO42- (7)


A

Alkaline pH: SO4•- + OH- → HO• + SO42- (8)


M

HO• + HO• → H2O2 (9)


D

HO• + H2O2 → HO2• + H2O (10)


TE

SO4•- + H2O2 → HO2• + H+ + SO42- (11)


EP

HO2• ↔ O2•- + H+ pKa = 4.8 (12)


CC

CT degradation performances in the thermally activated PS system with different

scavengers are shown in Fig. 2. The degradation of CT was dramatically promoted by


A

the addition of tert-butyl alcohol or isopropanol, i.e. more than 98% CT was degraded

with the scavenger of tert-butyl alcohol or isopropanol in 3 h. Its irrelevance on

tert-butyl alcohol or isopropanol addition indicates that HO• and SO4•- were not main
reactants and other radicals generated simultaneously during the reaction were

magnified and could rather be responsible for the degradation of CT. This might be

related to the solvent shell surrounding the reactive species with the addition of

solvents, and it will be further discussed in the later context. HO2• is not likely to

significantly affect the degradation of CT due to its lower oxidation potential than

HO• [40]. Che and Lee found that oxidative degradation by HO• and/or HO2• [41],

could not be a main degradation pathway of CT in the pyrite Fenton system,

PT
suggesting reductive degradation may be the potential degradation pathway. Nitrate

RI
addition did not affect CT degradation, as shown in Fig. 2, indicating that solvated

SC
electrons are not likely responsible for CT degradation in thermally activated PS
U
systems. Hasan et al. [42] proposed a reductive pathway in the manganese
N

dioxide-catalyzed decomposition of H2O2 in which H2O2 decomposed to oxygen,


A

water, and a solvated electron. However, the result of nitrate addition suggested that
M

solvated electrons could not degrade CT in this study. The test with O2•- scavenger
D

benzoquinone resulted in only 7% CT loss, strongly indicating that O2•- contributed


TE

CT degradation in the thermally activated PS system. O2•- has also been well
EP

documented to play the dominant role in transforming of perchlorinated organic


CC

compounds such as CT and hexachloroethane [37,40,41,43].

Unfortunately, no chlorinated intermediate byproducts were detected by GC-MS


A

analysis. This may be caused due to the similar reaction kinetics between the

formation and transformation of the chlorinated intermediate byproducts, or the low

concentration of the intermediate byproducts. CT degradation in the thermally


activated PS system is speculated to follow the pathway as below according to the

results reported in literatures: the C-Cl bond of CT may split and yield CCl3• and

chloride anions [44]. CCl3• may undergo proton abstraction to form CHCl3 or further

cleave C-Cl bond to generate dichlorocarbene [45]. Dichlorocarbene in solution may

hydrolyze and decompose to formyl chloride [46,47]. Aqueous formyl chloride would

decompose to CO, Cl-, and HCOO- [48,49,50].

3.2. Influence of various factors on CT degradation performance

PT
3.2.1. Effects of PS and initial CT concentration on CT degradation performance

RI
It has been well established that the concentration of oxidants plays an important

SC
role in the contaminate degradation process. As shown in Fig. 3(a), CT degradation
U
was highly PS dependent. 47.9%, 68.0% and 83.5% CT degradation (initial CT
N

concentration: 10 µM) achieved after 36 h at 0.1 M, 0.3 M and 0.5 M PS


A

concentration, respectively. As shown in Table 2, kobs of CT degradation increased


M

from 0.019 h-1 to 0.039 h-1 when the initial PS concentration increased, which was due
D

to more O2•- generated through a series of propagation reactions with increasing PS


TE

concentration. The result was consistent with previous studies that an increase of the
EP

oxidant concentration could enhance the degradation of target compound due to the
CC

promotion of the reactive oxygen species formation [40]. Liang and Bruell

investigated the thermally activated PS oxidation of trichloroethylene [6]. The


A

half-life of trichloroethylene decreased from 60.3 min to 25.0 min with the initial PS

concentration increase from 13.55 mM to 43.73 mM. Furman et al. studied the

degradation of hexachloroethane by heterogeneous birnessite-catalyzed


decomposition of H2O2 and found that hexachloroethane degradation increased as a

function of H2O2 concentration [43].

Interestingly, when the concentration of PS was 0.1 M, the CT degradation rate

dramatically decreased after 24 h. After this moment, the concentration of PS (78.3

mM) might not maintain enough to continuously generate O2•- as previous intensity.

This phenomenon indicated that the generation and intensity of O2•- was highly

dependent on the PS concentration.

PT
The variation of CT degradation efficiency in the thermally activated PS system

RI
with different initial CT concentrations, ranging from 5 µM to 1000 µM, was

SC
investigated and the results are shown in Fig. 3(b). While increasing the initial CT
U
concentration, the CT degradation rate was slightly reduced from 0.049 h-1 to 0.034
N

h-1 as shown in Table 2. When the initial concentration of CT was below 10 µM, the
A

CT degradation rates were almost the same. There was no significant concentration
M

effect on CT degradation at low initial CT concentrations, and this was due to the
D

generation of proper amount of O2•- with respect to CT. But when the initial CT
TE

concentration was higher than 10 µM, the CT degradation rate decreased due to the
EP

limited quantity of O2•- in the system. Che et al. discussed the degradation kinetics of
CC

trichloroethylene at different concentrations [51]. At low concentration of target

contaminant, excess amount of HO• generated relatively to trichloroethylene and


A

hence the reaction between HO• and H2O2 happened simultaneously in the pyrite

Fenton reaction. At high trichloroethylene concentration, HO• formed during the

reaction were used for trichloroethylene degradation mostly.


3.2.2. Effects of solution matrix on CT degradation performance

As the reactivity of oxidant in contaminated groundwater system might be

affected by the solution matrix, the influences of initial solution pH, Cl-, HCO3-, and

HA on CT degradation performance were investigated.

As shown in Fig. 4(a), the degradation kinetics in all investigated conditions were

similar, except a little superiority in unadjusted initial pH compared to others. When

the initial solution pHs were 3 and 7, the pHs quickly decreased to 2.56 and 2.66 in 1

PT
h. The initial pH of 12 decreased to 2.70 in 3 h. After 3 h of reaction, the pHs in all

RI
conditions reached close.

SC
Fig. 4(b) displayed the degradation of CT when the concentration of Cl- varied
U
from 1 mM to 100 mM. When the concentrations of Cl- were 1, 10 and 100 mM, the
N

CT degradation rates were 78.9%, 72.3% and 40.0%, respectively. High concentration
A

of Cl- in the system showed magnificent effect of inhibition. Cl- is known to compete
M

with CT for O2•- (Eq. 13), and hence reduce the degradation of CT [52]. The reactions
D

in Eqs. 15-17 describe the production of the free chlorine atom (Cl•) from reactions
TE

with SO4•- and HO• [53,54]. The Cl• is unstable in aqueous systems; if Cl-
EP

concentrations are sufficiently high, it would complex with a Cl- to form the Cl2•- (Eq.
CC

18). Cl2•- may further react with Cl2•-, Cl•, and O2•- to regenerate Cl- in solution (Eqs.

14,19,20), and therefore inhibit CT degradation [55]. Gu et al. investigated the


A

influence of Cl- on the degradation of trichloroethane in the thermally activated. They

found that Cl- had scavenging effects on radicals and then inhibited the degradation of

trichloroethane [56].
O2•- + Cl- → product (13)

O2•- + Cl2•- → 2Cl- + O2 (14)

SO4•- + Cl- → Cl• + SO42- (15)

HO• + Cl- ↔ ClOH•- (16)

ClOH•- + H+ ↔ Cl• + H2O (17)

Cl• + Cl- ↔ Cl2•- (18)

Cl2•- + Cl2•- → Cl2 + 2Cl- (19)

PT
Cl• + Cl2•- → Cl2 + Cl- (20)

RI
As shown in Fig, 4(c), CT degradation was reduced with the addition of HCO3-

SC
from 1 mM to 100 mM. HCO3- can directly react with O2•-, and the product CO3•- can
U
also react with O2•- to inhibit the degradation of CT (Eqs. 21-22) [57,58]. It can be
N

seen that the higher concentration of HCO3- showed weak inhibition compared to the
A

lower ones. The reason might be that HCO4- could be generated through the
M

equilibrium between HCO3- and H2O2 in aqueous solution [59]. The decomposition of
D

HCO4- could further generate other active oxygen species in situ such as O2•- and 1O2,
TE

which was benefit to CT degradation [60]. Mucka et al. investigated the radiation
EP

dechlorination of CT in the presence of HCO3-, and a slight negative effect of HCO3-


CC

on the dechlorination was observed [61].

O2•- + HCO3- → HO2- + CO3•- (21)


A

O2•- + CO3•- → CO32- + O2 (22)

Humic substances are the most abundant components of the colloidal and the

dissolved fraction of natural organic matter and they are characterized by a strong
binding capacity for both metals and organic pollutants, affecting their mobility and

bioavailability. Fig. 4(d) showed the CT degradation with the concentrations of HA

ranging from 1 mg/L to 10 mg/L. CT degradation slightly reduced with the increasing

concentration of HA in the thermally activated PS system which might be simply due

to the competing reaction for O2•-. Gu et al. also found that an adverse effect on

trichloroethane removal was observed with HA addition [56].

3.3. CT degradation in thermally activated PS system containing organic solvents

PT
CT degradation dramatically increased when the tert-butyl alcohol or isopropanol

RI
were added to the solution as shown in Fig. 2. This result indicated that the

SC
degradation of CT in the thermally activated PS system may be significantly
U
influenced by the solvents added to the solution. The same phenomenon was also
N

reported in Che and Lee's research when they studied the degradation of chlorinated
A

aliphatic compounds by Fenton reaction in pyrite suspension and found that the
M

degradation of CT was enhanced from 80% to 90% by the isopropanol addition in the
D

pyrite Fenton system [51]. Moreover, the addition of solvents having low polarity
TE

such as acetone, isopropanol, ethanol and methanol showed significant enhancements


EP

in CT degradation studied by Smith et al. [40] Although O2•- is a reactive radical for
CC

CT degradation in the thermally activated PS system, O2•- is characterized by low

reactivity in pure water (polar solvent) due to its high degree solvation, resulting in
A

the decrease of the lifetime and reactivity of O2•-.

To further investigate the effect of solvents reactivity of O2•- in the thermally

activated PS system, the reactions were repeated with the addition of equimolar
concentrations of four solvents. The reactions were conducted with 0.05 M PS and 1

M of each of the four solvents. As shown in Fig. 5, CT degradation significantly

increased with the addition of the four solvent (methanol, ethanol, isopropanol and

acetone). More than 99% of CT was degraded in the reaction time with the addition of

methanol, ethanol, and isopropanol, while 67.6% CT was degraded in 90 min with

acetone addition. The increased reactivity of O2•- in water solvent system is likely due

to the changes in the solvent shell surrounding O2•-. The data of Fig. 5 were obtained

PT
in reactions that contained a mixture of water and less polar compounds in the bulk

RI
solution, which likely resulted in a mixture of water and less polar molecules in the

SC
shell surrounding O2•-; the proportion of the less polar molecules in the solvent shell
U
might actually be higher than that in the bulk solution [62]. The mixed solvent shell
N

may have characteristics of both the less polar compound and water, resulting in an
A

increase in O2•- reactivity relative to purely aqueous solutions. It could be found in


M

Fig. 5 that there was a time lag in the reactions with the addition of solvents. The
D

reason might be due to the scavenging effect of the solvent to HO• and SO4•-,
TE

therefore reduced the reactions of Eqs. 8-11, and further inhibited the formation of
EP

O2•- at the beginning of the reaction. The lag time of CT degradation highly related to
CC

the empirical solvent polarity of the solvent, and shorter lag time was achieved in less

polar solvent (ETN: acetone < isopropanol < ethanol < methanol) as shown in Table 3.
A

In addition, each solvent had a different effect on the PS decomposition rate in the

thermally activated PS system. The residual PS concentrations were 39, 33 and 23

mM, when more than 99% of CT was degraded with the addition of isopropanol,
ethanol and methanol, respectively. The utilization efficiency of PS is higher in less

polar of the solvent. The solvent of less polar increases the reactivity of O2•-, and is

more efficient in CT degradation among isopropanol, ethanol, and methanol.

CT degradation performances with the different concentrations of solvents were

also investigated in the thermally activated PS system and the results are shown in Fig.

6. CT degradation increased as a function of acetone concentration in the thermally

activated PS system. Lower concentration of acetone increased the reactivity of O2•-,

PT
and led to faster CT degradation.

RI
4. Conclusion
SC
U
This study made an effort to explore CT degradation in a thermally activated PS
N

system in aqueous solution at 50℃. The experimental results showed that CT could
A

be readily degraded at 50℃ with a PS concentration of 0.5 M. CT degradation and


M

PS consumption followed a pseudo-first order kinetic model. Radical scavenger tests


D

showed that O2•- was the predominant radical species responsible for CT degradation,
TE

and the split of C-Cl was proposed as the possible reaction pathways for CT
EP

degradation. The process of CT degradation was accelerated by higher PS dose and


CC

lower initial CT concentration. No obvious effect of the initial pH on the degradation

of CT was observed in the thermally activated PS system due to the rapid pH decrease
A

after the start of the reaction. Cl-, HCO3-, and HA had negative effects on CT

degradation in which higher concentration of HCO3- (100 mM) showed less inhibitive

effect than lower ones. The degradation of CT in the thermally activated PS system
could be significantly promoted by the solvents addition to the solution, which

brought about an increase of O2•- reactivity in the solvent relative to purely aqueous

solutions. However, there was a lag time in CT degradation with the addition of

solvents, and the utilization efficiency of PS in less polar of the solvent is higher. The

above results strongly demonstrated that the thermally activated PS process is a

promising technique in in-situ chemical oxidation/reduction remediation for

degrading CT contaminated sites.

PT
Acknowledgment

RI
This study was financially supported by a grant from the National Environmental

SC
Protection Public Welfare Science and Technology Research Program of China (No.
U
201109013), the National Natural Science Foundation of China (No.41373094 and
N

No.51208199), the Shanghai Natural Science Funds (No. 12ZR1408000), China


A

Postdoctoral Science Foundation (No. 2013T60429), and the Fundamental Research


M

Funds for the Central Universities.


D

Literature Cited:
TE

1. C.J. Lin, S.L. Lo, Y.H. Liou, Degradation of aqueous carbon tetrachloride by
EP

nanoscale zerovalent copper on a cation resin, Chemosphere 59 (2005)


CC

1299-1307.

2. K.H. Sweency, J.R. Fischer, Reductive degradation of halogenated pesticides,


A

U.S. Patent No. 3640821, 1972-2-8.

3. G.W. Reynolds, J.T. Hoff, R.W. Gillham, Sampling bias caused by materials used

to monitor halocarbons in groundwater, Environ. Sci. Technol. 24 (1990)


135-139.

4. A. Tsitonaki, B. Petri, M.L. Crimi, H. Mosbaek, R.L. Siegrist, P.L. Bjerg, In situ

chemical oxidation of contaminated soil and groundwater using persulfate: a

review, Crit. Rev. Env. Sci. Technol. 40 (2010) 55-91.

5. M.L. Crimi, J. Taylor, Experimental evaluation of catalyzed hydrogen peroxide

and sodium persulfate for destruction of BTEX contaminants, Soil Sediment

Contam. 16 (2007) 29-45.

PT
6. C.J. Liang, C.J. Bruell, Thermally activated persulfate oxidation of

RI
trichloroethylene: experimental investigation of reaction orders, Ind. Eng. Chem.

Res. 47 (2008) 2912-2918.


SC
U
7. C.J. Liang, C.J. Bruell, M.C. Marley, K. Sperry, Persulfate oxidation for in situ
N

remediation of TCE. I. Activated by ferrous ion with and without a


A

persulfate-thiosulfate redox couple, Chemosphere 55 (2004) 1225-1233.


M

8. X.G. Gu, S.G. Lu, Z.F. Qiu, Q. Sui, C.J. Banks, T. Imai, K.F. Lin, Q.S. Luo,
D

Photodegradation performance of 1,1,1-trichloroethane in aqueous solution: In


TE

the presence and absence of persulfate, Chem. Eng. J. 215 (2013) 29-35.
EP

9. F.F. Hao, W.L. Guo, A.Q. Wang, Y.Q. Leng, H.L. Li, Intensification of
CC

sonochemical degradation of ammonium perfluorooctanoate by persulfate oxidant,

Ultrason. Sonochem. 21 (2014) 554-558.


A

10. C.J. Liang, C.J. Bruell, M.C. Marley, K.L. Sperry, Thermally activated persulfate

oxidation of trichloroethylene (TCE) and 1,1,1-trichloroethane (TCA) in aqueous

systems and soil slurries, Soil Sediment Contam. 12 (2003) 207-228.


11. V.C. Mora, J.A. Rosso, D.O. Martire, M.C. Gonzalez, Phenol depletion by

thermally activated peroxydisulfate at 70℃, Chemosphere 84 (2011) 1270-1275.

12. C.Q. Tan, N.Y. Gao, Y. Deng, N. An, J. Deng, Heat-activated persulfate oxidation

of diuron in water, Chem. Eng. J. 203 (2012) 294-300.

13. S.A. Kordkandi, M. Forouzesh, Application of full factorial design for methylene

blue dye removal using heat-activated persulfate oxidation, J. Taiwan Inst. Chem.

E. 45 (2014) 2597-2604.

PT
14. C.Q. Tan, N.Y. Gao, Y. Deng, W.L. Rong, S.D. Zhou, N.X. Lu, Degradation of

RI
antipyrine by heat activated persulfate, Sep. Purif. Technol. 109 (2013) 122-128.

SC
15. J.Y. Zhao, Y.B. Zhang, X. Quan, S. Chen, Enhanced oxidation of 4-chlorophenol
U
using sulfate radicals generated from zero-valent iron and peroxydisulfate at
N

ambient temperature, Sep. Purif. Technol. 71 (2010) 302-307.


A

16. S.Y. Oh, S.G. Kang, P.C. Chiu, Degradation of 2,4-dinitrotoluene by persulfate
M

activated with zero-valent iron, Sci. Total Environ. 408 (2010) 3464-3468.
D

17. L.G. Devi, S.G. Kumar, K.M. Reddy, C. Munikrishnappa, Photo degradation of
TE

Methyl Orange an azo dye by advanced Fenton process using zero valent metallic
EP

iron: Influence of various reaction parameters and its degradation mechanism, J.


CC

Hazard. Mater. 164 (2009) 459-467.

18. P. Neta, V. Madhavan, H. Zemel, R. Fessenden, Rate constants and mechanism of


A

reaction of sulfate radical anion with aromatic compounds, J. Am. Chem. Soc.

99 (1977) 163-164.

19. G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical review of rate
constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl

radicals (•OH/•O-) in aqueous solution, J. Phys. Chem. Ref. Data 17 (1988)

513-886.

20. C.L. Clifton, R.E. Huie, Rate constants for hydrogen abstraction reactions of the

sulfate radical, SO4•-, alcohols, Int. J. Chem. Kinet. 21 (1989) 677-687.

21. J.M. Monteagudo, A. Durán, I.S. Martin, A. Carnicer, Roles of different

intermediate active species in the mineralization reactions of phenolic pollutants

PT
under a UV-A/C photo-Fenton process, Appl. Catal., B 106 (2011) 242-249.

RI
22. C.J. Liang, C.F. Huang, N. Mohanty, R.M. Kurakalva, A rapid spectrophotometric

SC
determination of persulfate anion in ISCO, Chemosphere 73 (2008) 1540-1543.
U
23. X.R. Xu, X.Z. Li, Degradation of azo dye Orange G in aqueous solutions by
N

persulfate with ferrous ion, Sep. Purif. Technol. 72 (2010) 105-111.


A

24. K.C. Huang, Z. Zhao, G.E. Hoag, A. Dahmani A, P.A. Block, Degradation of
M

volatile organic compounds with thermally activated persulfate oxidation,


D

Chemosphere 61 (2005) 551-560.


TE

25. K.C. Huang, R.A. Couttenye, G.E. Hoag, Kinetics of heat assisted persulfate
EP

oxidation of methyl tert-butyl ether (MTBE), Chemosphere 49 (2002) 413-420.


CC

26. R.H. Waldemer, P.G. Tratnyek, R.L. Johnson, J.T. Nurmi, Oxidation of

chlorinated ethenes by heat-activated persulfate: Kinetics and products, Environ.


A

Sci. Technol. 41 (2007) 1010-1015.

27. C.J. Liang, Z.S. Wang, C.J. Bruell, Influence of pH on persulfate oxidation of

TCE at ambient temperatures, Chemosphere 66 (2007) 106-113.


28. R.L. Johnson, P.G. Tratnyek, R.O. Johnson, Persulfate persistence under thermal

activation conditions, Environ. Sci. Technol. 42 (2008) 9350-9356.

29. D.A. House, Kinetics and mechanism of oxidations by peroxydisulfate, Chem.

Rev. 62 (1962) 185-203.

30. I.M. Kolthoff, I.K. Miller, The chemistry of persulfate. I. The kinetics and

mechanism of the decomposition of the persulfate ion in aqueous medium1, J.

Am. Chem. Soc. 73 (1951) 3055-3059.

PT
31. D.E. Pennington, A. Haim, Stoichiometry and mechanism of the

RI
chromium-peroxydisulfate reaction, J. Am. Chem. Soc. 90 (1968) 3700-3704.

SC
32. E. Hayon, A. Treinin, J. Wilf, Electronic spectra, photochemistry, and
U
autoxidation mechanism of the sulfite-bisulfite-pyrosulfite systems. The SO2-,
N

SO3-, SO4-, and SO5- radicals, J. Am. Chem. Soc. 94 (1972) 47-57.
A

33. R. Jain, M. Mathur, S. Sikarwar, A. Mittal, Removal of the hazardous dye


M

rhodamine B through photocatalytic and adsorption treatments, J. Environ.


D

Manage. 85 (2007) 956-964.


TE

34. R.J. Watts, B.C. Bottenberg, M.E. Jensen, T.H. Hess, A.L. Teel, Role of
EP

reductants in the enhanced desorption and transformation of chloroaliphatic


CC

compounds by modified Fenton’s reagent, Environ. Sci. Technol. 33 (1999)

3432-3437.
A

-
35. B.H. Bielski, D.E. Cabelli, R.L. Arudi, A. Ross, Reactivity of HO2/O2 radicals in

aqueous solution, J. Phys. Chem. Data 14 (1985) 1041-1100.

36. I.B. Afanas’ev, Superoxide Ion: Chemistry and Biological Implications, CRC
Press, Boca Raton, FL, 1989.

37. A.L. Teel, R.J. Watts, Degradation of carbon tetrachloride by modified Fenton’s

reagent, J. Hazard. Mater. 94 (2002) 179-189.

38. G.V. Buxton, S. McGowan, G.A. Salmon, J.E. Williams, N.D. Wood, A study of

the spectra and reactivity of the oxysulphur-radical anions involved in the chain

oxidation of S(IV): A pulse and g-radiolysis study, Atmos. Environ. 30 (1996)

2483-2493.

PT
39. G.P. Anipsitakis, D.D. Dionysiou, Degradation of organic contaminants in water

RI
with sulfate radicals generated by the conjunction of peroxymonosulfate with

SC
cobalt, Environ. Sci. Technol. 37 (2003) 4790-4797.
U
40. B.A. Smith, A.L. Teel, R.J. Watts, Identification of the reactive oxygen species
N

responsible for carbon tetrachloride degradation in modified Fenton’s systems,


A

Environ. Sci. Technol. 38 (2004) 5465-5469.


M

41. H.S. Che, W.J. Lee, Selevtive redox degradation of chlorinated aliphatic
D

compounds by Fenton reaction in pyrite suspension, Chemosphere 82 (2011)


TE

1103-1108.
EP

42. M.A. Hasan, M.I. Zaki, L. Pasupulety, K. Kumari, Promotion of the hydrogen
CC

peroxide decomposition activity of manganese oxide catalysts, Appl. Catal., A

181 (1999) 171-179..


A

43. O. Furman, D.F. Laine, A. Blumenfeld, A.L. Teel, K. Shimizu, I.F. Cheng, R.J.

Watts, Enhanced reactivity of superoxide in water-solid matrices, Environ. Sci.

Technol. 43 (2009) 1528-1533.


44. P. Neta, J. Grodkowski, A.B. Ross, Rate constants for reactions of aliphatic

carbon-centered radicals in aqueous solution, J. Phys. Chem. Ref. Data 25 (1996)

709-1050.

45. M.L. McCormick, P. Adriaens, Carbon tetrachloride transformation on the surface

of nanoscale biogenic magnetite particles, Environ. Sci. Technol. 38 (2004)

1045-1053.

46. M.L. McCormick, E.J. Bouwer, P. Adriaens, Carbon tetrachloride transformation

PT
in a model iron-reducing culture: relative kinetics of biotic and abiotic reactions,

RI
Environ. Sci. Technol. 36 (2002) 403-410.

SC
47. J.R. Pliego, W.B. deAlmeida, Reaction Paths for Aqueous Decomposition of CCl2,
U
J. Phys. Chem. 100 (1996) 12410-12413.
N

48. A.J. Swallow, 1978. In: Jennings, K.R., Cundall, R.B. (Eds.), Progress for
A

Reaction Kinetics. 9 No. 3/4, Pergamon Press, Oxford, pp. 195-366.


M

49. W. Choi, M.R. Hoffman, Kinetics and mechanism of CCl4 photoreductive


D

degradation on TiO2: the role of trichloromethyl radical and dichlorocarbene, J.


TE

Phys. Chem. 100 (1996) 2161-2169.


EP

50. A.J. Wagner, C. Vecitis, D.H. Fairbrother, Electron-stimulated chemical reactions


CC

in carbon tetrachloride/water (ice) films, J. Phys. Chem. B 106 (2002) 4432-4440.

51. H.S. Che, S.J. Bae, W.J. Lee, Degradation of trichloroethylene by Fenton reaction
A

in pyrite suspension, J. Hazard. Mater. 185 (2011) 1355-1361.

52. C.A. Long, B.H. Bielski, Rate of reaction of superoxide radical with

chloride-containing species, J. Phys. Chem. 84 (1980) 555-557.


53. G.V. Buxton, M. Bydder, G.A. Salmon, The reactivity of chlorine atoms in

aqueous solution II: The equilibrium SO4•- + Cl- SO42- + Cl•, Phys. Chem. Chem.

Phys. 1 (1999) 269-273.

54. X.Y. Yu, J.R. Barker, Hydrogen peroxide photolysis in acidic aqueous solutions

containing chloride ions. I. Chemical mechanism, J. Phys. Chem. A 107 (2003)

1313-1324.

55. S. Navaratnam, B.J. Parsons, A.J. Swallow, Some reactions of the dichloride

PT
anion radical, Radiat. Phys. Chem. 15(1980) 159-161.

RI
56. X.G. Gu, S.G. Lu, L. Li , Z.F. Qiu, Q. Sui, K.F. Lin, Q.S. Luo, Oxidation of 1,1,1-

SC
trichloroethane stimulated by thermall activated persulfate, Ind. Eng. Chem. Res.
U
50 (2011) 11029-11036.
N

57. K.H. Schmidt, Electrical conductivity techniques for studying the kinetics of
A

radiation-induced chemical reactions in aqueous solutions, Int. J. Radiat. Phys.


M

Chem. 4 (1972) 439-468.


D

58. D. Behar, G. Czapski, I. Duchovny, Carbonate radical in flash photolysis and


TE

pulse radiolysis of aqueous carbonate solutions, J. Phys. Chem. 74 (1970)


EP

2206-2210.
CC

59. J. Flanagan, D.P. Jones, W.P. Griffith, A.C. Skapski, A.P. West, On the existence

of peroxocarbonates in aqueous solution, J. Chem. Soc., Chem. Commun. 1


A

(1986) 20-21.

60. J.M. Lin, M.L. Liu, Chemiluminescence from the decomposition of

peroxymonocarbonate catalyzed by gold nanoparticles, J. Phys. Chem. B 112


(2008) 7850-7855.

61. V. Mucka, V. Cuba, M. Pospisil, R. Silber, Radiation dechlorination of some

chlorinated hydrocarbons particularly of carbon tetrachloride in the presence of

HCO3-- or NO3--ions, Appl. Catal., A: General 271 (2004) 195-201.

62. C. Reichardt. Solvents and solvent effects in organic chemistry, 3rd ed.;

Wiley-VCH: New York, 2003.

PT
RI
Table 1. CT degradation performance in the thermally activated PS system
Operational conditions kobs (h-1) Correlation coefficient,R2 kobs,PS (h-1) R2(PS)

[PS] = 0.5 M -
SC- 0.014 0.93
U
[PS] = 0.5 M, [CT] = 1 mM 0.034 0.97 0.014 0.93
N
A
M

Table 2. Kinetic parameters of CT degradation performance under various conditions


D

PS(M)/CT(µM) Solution matrix kobs (h-1) R2


TE

0.1/10 - 0.019 0.98


0.3/10 - 0.031 0.99
EP

0.5/10 - 0.039 0.99


CC

0.5/1 - 0.049 0.96


0.5/100 - 0.039 0.99
A

0.5/1000 - 0.034 0.97


0.5/10 Initial pH = 3 0.037 0.98
0.5/10 Initial pH = 7 0.040 0.96
0.5/10 Initial pH = 12 0.033 0.96
0.5/10 Cl- = 1 mM 0.038 0.96
0.5/10 Cl- = 10 mM 0.034 0.95
0.5/10 Cl- = 100 mM 0.012 0.89
0.5/10 HCO3- = 1 mM 0.036 0.95
0.5/10 HCO3- = 10 mM 0.033 0.91
0.5/10 HCO3- = 100 mM 0.037 0.97
0.5/10 HA = 1 mg/L 0.038 0.98
0.5/10 HA = 5 mg/L 0.036 0.97
0.5/10 HA = 10 mg/L 0.032 0.96

PT
RI
Table 3. CT degradation performance with the addition of solvent

ETN (from low to high) Acetone SC


Isopropanol Ethanol Methanol
U
× 103 10.73 0.91 0.59 0.37
N

Lag time (min) 10 16 31 68


A
M
D

Figure Captions
TE

Fig. 1. CT degradation and PS decomposition in the thermally activated PS system.


EP

(50℃, [PS] = 0.5 M, [CT] = 1 mM)


CC

Fig. 2. Effects of scavengers on the degradation of CT. (50℃, [PS] = 0.5 M, [CT] =
A

10 µM, [TBA] = [IPA] = 0.5 M, [NaNO3] = [BQ] = 0.05 M)

Fig. 3. Effect of a) PS concentration, and b) initial CT concentration on the

degradation of CT. (50℃, [PS] = 0.5 M, [CT] = 10 µM)

Fig. 4. Effect of solution matrix on the degradation of CT. a) initial pH; b) Cl-; c)
HCO3-; 4) HA. (50℃, [PS] = 0.5 M, [CT] = 10 µM)

Fig. 5. CT degradation perofrmance in the thermally activated PS system in the

presence of four solvents. (50℃, [PS] = 0.05 M, [CT] = 10 µM, [Methanol] =

[Ethanol] = [IPA] = [Acetone] = 1 M)

Fig. 6. Effect of acetone dosage on the degradation of CT (50℃, [PS] = 0.05 M, [CT]

= 10 µM)

PT
Fig. 1.

RI
SC
U
N
A
M
D
TE
EP

Fig. 2
CC
A
PT
RI
Fig. 3(a)
SC
U
N
A
M
D
TE
EP
CC
A

Fig. 3(b)
PT
RI
Fig. 4(a)
SC
U
N
A
M
D
TE
EP
CC
A

Fig. 4(b)
PT
RI
Fig. 4(c)
SC
U
N
A
M
D
TE
EP
CC

Fig. 4(d)
A
A

Fig. 6
Fig. 5
CC
EP
TE
D
M
A
N
U
SC
RI
PT
A
CC
EP
TE
D
M
A
N
U
SC
RI
PT

You might also like