You are on page 1of 8

Separation and Purification Technology 143 (2015) 19–26

Contents lists available at ScienceDirect

Separation and Purification Technology


journal homepage: www.elsevier.com/locate/seppur

Activated carbon fiber as heterogeneous catalyst of peroxymonosulfate


activation for efficient degradation of Acid Orange 7 in aqueous solution
Shiying Yang ⇑, Tuo Xiao, Jun Zhang, Youyuan Chen, Lei Li
Key Laboratory of Marine Environment and Ecology, Ministry of Education, Qingdao 266100, China
College of Environmental Science and Engineering, Ocean University of China, Qingdao 266100, China

a r t i c l e i n f o a b s t r a c t

Article history: Increasing attention has been paid to environmental friendly activation method of peroxymonosulfate
Received 10 September 2014 (PMS) in the field of advanced oxidation processes (AOPs). In this work, activated carbon fiber (ACF)
Received in revised form 14 January 2015 was utilized as an effective and green catalyst to activate peroxymonosulfate (PMS) for degradation of
Accepted 17 January 2015
aqueous organic pollutant. As results indicated, a very low dosage of ACF (0.3 g L 1) could efficiently acti-
Available online 22 January 2015
vate PMS to eliminate Acid Orange 7 (AO7), a probe compound, and the degradation of AO7 followed first
order kinetics. Surface chemistry analyses of ACF samples, including BET specific surface area, pH at the
Keywords:
point of zero charge, Boehm titration and FTIR spectroscopy, suggested that the basic sites on ACF surface
Activated carbon fiber
Peroxymonosulfate activation
were likely to be the active sites inducing the decomposition of PMS and the generation of radicals.
Catalytic activity Quenching studies confirmed that sulfate radicals (SO4 ) and hydroxyl radicals (HO), which formed on
Dye degradation the surface of ACF, were the primary oxidizing species. A possible pathway for AO7 degradation was pro-
Reaction mechanism posed by monitoring the temporal evolution of intermediates in the solution, with the use of GC–MS and
ionic chromatography (IC). It is suggested that the degradation process of AO7 was initiated by cleavage
of azo linkage and further undergoes ring-opening reactions.
Ó 2015 Elsevier B.V. All rights reserved.

1. Introduction the electron-transfer mediator, resulting in decomposition of


hydrogen peroxide (H2O2) [11] and PS [12]. Researchers have
Advanced oxidation processes (AOPs) are considered as one of found that contaminants can be successfully degraded in the AC/
the most effective methods to degrade organic contaminants due H2O2 system [13–16] or in the AC/PS system [17]. Our group pre-
to the generation of powerful oxidizing species. AOPs based on viously confirmed that granular activated carbon (GAC) can also
the hydroxyl radical (HO), which has a redox potential of 1.8– catalyze PMS to produce radicals for degradation of azo dye Acid
2.7 V vs. NHE (normal hydrogen electrode) [1], have been widely Orange 7 (AO7) [18].
studied and used in water or wastewater treatment [2–4]. In recent In the current work, we explore the application of activated car-
years, increasing attention has been paid to the sulfate radical bon fiber (ACF), which has higher adsorption rates in the gas or
(SO4 ). Compared to HO, SO4 demonstrated the same or even a liquid phase [19]. For a long time, ACF was exclusively used in
higher redox potential of 2.5–3.1 V vs. NHE [5] and has a great waste water treatment based on its function as adsorption materi-
ability to oxidize organic pollutants. Persulfate (PS) and peroxy- als [20,21], and in recent years it was also studied as catalyst sup-
monosulfate (PMS) is an oxidant which has been widely studied ports [22,23]. However, to the best of our knowledge, there are no
to generate SO4 with activators including UV [6], heat [7], base previous studies on the application of ACF for directly catalytic
[8] and transition metals [9,10]. PMS oxidation, although catalytic ozonation or Fenton oxidation
Porous carbon materials are excellent adsorbents and have been with ACF has been studied [4,24].
used since many years ago. Nowadays, the catalytic ability of acti- The objectives of this present study are, (a) to demonstrate the
vated carbon (AC) has attracted much scientific interests in the feasibility of ACF as a catalyst to activate PMS for degradation an
field of wastewater treatment. AC has many oxygen-containing azo dye Acid Orange 7 (AO7), which has commonly been used as
functional groups on the surface, which may act as active sites of a model compound for HO or SO4 based AOPs [2,4,17]; (b) to
investigate the reaction kinetics involving the influence of the con-
⇑ Corresponding author at: College of Environmental Science and Engineering, centration of PMS and the dosage of ACF; (c) to find the active sites
Ocean University of China, Qingdao 266100, China. Tel.: +86 532 66782780. on ACF surface which are responsible for the decomposition of
E-mail address: shiningpup@hotmail.com (S. Yang). PMS; (d) to identify the dominating oxidizing species generating

http://dx.doi.org/10.1016/j.seppur.2015.01.022
1383-5866/Ó 2015 Elsevier B.V. All rights reserved.
20 S. Yang et al. / Separation and Purification Technology 143 (2015) 19–26

from PMS decomposition and resulting in AO7 decay; and (e) to concentrated to 1.0 mL on a rotatory evaporator, and was kept at
give the possible pathway of AO7 degradation. 4 °C prior to GC analysis. The concentrated sample in CH2Cl2
(1 lL) was injected into a 6890 N/5975 GC/MS system for analysis.
DB-5MS capillary column with an inner diameter of 0.250 mm and
2. Material and methods
a length of 30 m was adopted in the separation system. The GC col-
umn was operated in a temperature programmed mode by main-
2.1. Materials
taining the temperature at 40 °C for 5 min, then increasing to
290 °C with an increment of 3 °C min 1 and finally holding at
Peroxymonosulfate (PMS), which is available as a triple salt of
290 °C for 10 min. Electron impact (EI) mode at 70 eV was used
sulfate commercially known as oxone (2KHSO5KHSO4K2SO4),
and the mass range scanned was 0–500 m/z. The substance analy-
was purchased from Shanghai Ansin Chemical Company, China.
sis was undertaken with reference to the software of NIST MS
AO7 and other chemicals were purchased from Shanghai Chemical
search 2.0.
Reagent Company and used without purification. Commercial acti-
The concentration of NH+4, NO2 , CH3COO and HCOO was per-
vated carbons (granular, about 2 mm in diameter and 4 mm in
formed with ionic chromatography (ICS-3000) and equipped with
length, SBET = 1050 m2 g 1, Vtotal = 0.49 cm3 g 1 and Vmicro = 0.26
cation exchange column of IonPac CS11-HC and anion exchange
cm3 g 1) were supplied by Shanxi Xinhua Chemical Plant.
column of IonPac AS12A.
Commercially viscose-based activated carbon fibers (felty, 10–20 lm
in diameter and 3–4 mm in thickness, SBET = 1080 m2 g 1,
Vtotal = 0.52 cm3 g 1 and Vmicro = 0.50 cm3 g 1) were obtained from 3. Results and discussion
Sutong Carbon Fiber Company, China. Before use they were both
washed with distilled water and subsequently dried at 105 °C for 3.1. Catalytic activity of ACF
24 h. The dried GAC and ACF samples were then kept in a desicca-
tor for later use and labelled as GAC0 and ACF0. Fig. 1a shows AO7 removal by PMS oxidation, GAC or ACF
adsorption, and GAC or ACF catalytic PMS oxidation. In 120 min,
2.2. Reactor and procedure the decolorization ratios of AO7 were 6%, 56% and 100% in the sys-
tem of PMS, ACF0 and the ACF0/PMS combined system, respec-
The experiments of AO7 removal by GAC or ACF adsorption, tively. Activation is necessary for application since PMS has a low
PMS oxidation, and GAC or ACF catalytic PMS oxidation were all reactivity. Obviously, there existed a synergistic effect in the
carried out in 500 mL flasks containing 250 mL of AO7 solution ACF0 and PMS combined system. That is to say, ACF has a great cat-
with an initial concentration of 100 mg L 1. The flasks were par- alytic activity of PMS activation for AO7 degradation.
tially immersed in a water bath fitted with a temperature control- Besides, it should also be noticed that the AO7 elimination rate
ler. The reaction mixture was stirred constantly at 25 ± 1 °C. Unless in the ACF0/PMS system is much higher than that in the GAC0/PMS
otherwise specified, the dosage of GAC or ACF was 0.3 g L 1, and system, in which only about 11% AO7 was removed in 120 min.
the mole ratio of PMS to AO7 was 20/1. The initial pH was adjusted This phenomenon may be attributed to the fact that ACF had a
to 5 by using 1 M NaOH or 1 M H2SO4. At a fixed interval, samples higher adsorption rate for AO7 (Fig. 1a). It is due to pore volume
were withdrawn and filtered before to be analyzed. and uniform microporosity [28] and a higher catalytic activity of
In the special experiments involving ACF pretreatment by PMS,
0.3 g L 1 ACF was immersed in 250 mL PMS solution of which the
initial concentration is 5.71 mM. The concentration was as same as
general oxidation experiments, and the reaction mixture was stir-
red constantly at 25 ± 1 °C. After specific time, ACF samples were
filtered out and washed with distilled water alternately for several
times, thereafter were dried at 105 °C for 3 h. After that these PMS-
oxidized ACF samples were also stored in a desiccator. It should be
noted that the experimental results labelled as ACF15, ACF30 and
ACF60 were the products of ACF pretreated with a PMS solution
(5.71 mM) for 15 min, 30 min and 60 min, respectively.

2.3. Analytical methods

The decolorization of AO7 was monitored by measuring the


absorbance of samples at 484 nm (UNICO 2100 spectrophotome-
ter, UNICO (Shanghai) Instruments Co., Ltd., China) as a function
of reaction time.
The concentration of PMS was analyzed via an iodometric titra-
tion method [25]. The point of zero charge (pHpzc) for ACF was
determined according to a mass titration method [26]. The Boehm
titration method was used to determine the acidity and the basi-
city of ACF [27].
Specific surface area was determined by a ST-08A Specific Sur-
face Area Analyzer. Fourier transform infrared (FTIR) spectroscopy
analysis of the ACF was performed on a BRUKER TENSOR 27
spectrometer.
GC–MS was used to detect the organic intermediate products of
AO7 degradation. The sample at 120 min was extracted with Fig. 1. (a) The degradation of AO7 in different systems. (b) The comparison of PMS
CH2Cl2 (3 times, 20 mL each time). The extraction liquid was then decomposition ratios in the presence of ACF0 and GAC0.
S. Yang et al. / Separation and Purification Technology 143 (2015) 19–26 21

Fig. 2. UV–vis spectra of methanol extractions from used ACF in the ACF/PMS
system.

Fig. 4. (a) Effect of ACF dosage on the degradation of AO7 in the ACF/PMS system.
(b) The degradation rates of AO7 and PMS decomposition ratios vs. ACF dosages in
the ACF/PMS system.

Fig. 3. Effect of PMS concentration on the degradation of AO7 in the ACF/PMS


Table 1
system.
Surface characteristics of ACF0 and ACF60.

Sample SBET pHzpc Total acidity Total basicity


PMS decomposition than GAC (Fig. 1b). Apparently, ACF was supe- (m2 g 1
) (lmol g 1) (lmol g 1)
rior over GAC for activating PMS to generate some active species, ACF0 1080 7.23 330 643
which will degrade aqueous organic pollutant. ACF60 904 1.75 1928 352
Although 100% decolorization of AO7 was achieved in the ACF/
PMS system, it was still uncertain whether the removal of AO7 was
mainly due to oxidation other than adsorption, because some of 10/1 to 20/1 or 30/1. Possibly, this can be explained by the limited
AO7 molecular might be adsorbed on the surface of ACF. In order amount of active sites on the surface of ACF (0.3 g L 1) relative to
to answer this question, a simple experiment was conducted as fol- the PMS concentration. In the meantime, the excessive PMS may
lows: After 120 min, the ACF sample in the ACF/PMS system was scavenge sulfate radicals.
filtered out and then methanol was employed to extract the adsor- In order to confirm this assumption, effect of ACF dosage on the
bates on the surface of ACF. The recovery AO7 of extraction from AO7 degradation and PMS decomposition were studied, and the
ACF surface of this method is close to 61.3%. The results are pre- results are depicted in Fig. 4. As shown in Fig. 4a, AO7 was
sented in Fig. 2. The negligible absorbance at 484 nm of methanol degraded more and more quickly with increased ACF dosages.
extractions indicates that there is no obvious quantity of AO7 Fig. 4b displays the pseudo first-order kinetic constants (k0,AO7)
absorbed on ACF in the ACF/PMS system. In addition, the strong and PMS decomposition ratios as a function ACF dosage. Interest-
absorbance in the UV region, especially at 225 nm, 232 nm and ingly, as observed, both k0,AO7 and PMS decomposition ratios line-
252 nm, shows intermediate products of AO7 degradation, which arly increased with ACF dosages. The relationship between PMS
may be some aromatic compounds. In conclusion, it is believed decomposition ratios and ACF dosages is similar to the phenome-
that AO7 was almost fully oxidized other than adsorbed in the non that between PS decomposition and AC dosage, which was also
ACF/PMS system. obtained by other researchers [12].

3.2. Reaction kinetics of AO7 removal 3.3. Active sites of ACF

The first-order loss of AO7 in solutions with a range of PMS/AO7 In order to explore how PMS oxidation caused the change of
molar ratios is illustrated in Fig. 3. When PMS/AO7 molar ratio ACF catalytic ability, surface textural and chemical properties of
increased from 5/1 to 10/1, there was an obvious enhancement different ACF samples were characterized. Data of specific surface
in the AO7 removal. However, the decay rate of AO7 can hardly area, pHpzc, total acidity and total basicity are summarized in
have an improvement when PMS/AO7 molar ratio increased from Table 1. Additionally, the FTIR spectrums of different ACF and
22 S. Yang et al. / Separation and Purification Technology 143 (2015) 19–26

Fig. 6. (a) Effect of PMS pretreatment on ACF catalytic ability. (b) Pseudo first-order
kinetics of AO7 removal by ACF catalytic PMS oxidation ([AO7]0 = 100 mg/L, ACF
dosage = 0.3 g/L, PMS/AO7 molar ratio = 20/1).

Fig. 5. (a) FTIR spectra for ACF0, ACF-Used, GAC-0. (b) FTIR spectra for ACF0, ACF15,
ACF30 and ACF60.
1558 cm 1 and 1111 cm 1 of them had pronounced increases. It
may be due to the formation or increase of lactone groups, which
have bands of both COC and C@O. From the above results, PMS pre-
GAC samples are also monitored as presented in Fig. 5. Major
treatment caused the increase of carboxyl and lactone on ACF sur-
absorption differences in the spectrums of samples are in five posi-
face, which are both acidic groups. The result of FTIR analysis was
tions: 3430 cm 1, 1710 cm 1, 1558 cm 1, 1400 cm 1 and
consistent with the measurement of the total acidity.
1111 cm 1. The absorption of 1558 cm 1 can be ascribed to stretch
Although it is difficult to find information about basic oxygen-
vibration of C@O [29], and 3430 cm 1 can be assigned to hydrogen-
containing groups in FTIR spectrums of ACF samples, an obvious
bonded OH [30]. The absorption in 1710 cm 1 and 1400 cm 1 was
decrease of the total basicity from ACF0 to ACF60 can be observed
due to the existence of COOH and C@OH stretching molds in phe-
from Table 1. It should be noted that the delocalized p-electrons of
nolic groups respectively. The peak at 1111 cm 1 indicated the
the basal planes were also assumed to have a basic nature [32].
involvement of OH or COC [31]. Fig. 5a showed that the hydro-
Above all, PMS pre-oxidation rendered ACF the increase of
gen-bonded OH of used ACF and GAC were poor while C@O and
acidic sites and the decrease of basic sites. The basic sites on the
CAOH increased compared to ACF-0. Furthermore, the FITR spec-
surface of ACF were possible active sites provoking the decomposi-
trum of ACF-Used demonstrated that there were no sign of AO7
tion of PMS and thereby the generation of radicals.
on the surface, which was consistent with the extraction
experiment.
After pre-oxidation by PMS, Fig. 6 showed that the treatment 3.4. Radical identification
has restrained the catalytic activity of ACF. The degradation rates
of ACF0, ACF15, ACF30, ACF60 are 5.8  10 2, 5.4  10 2, To identify the dominating oxidizing species responsible for
4.9  10 2, and 3.7  10 2 min 1 respectively. In a word, the AO7 decay in the ACF/PMS system, methanol (MA) and tert-butyl
longer pretreat time is, the more obvious inhibitory effect would alcohol (TBA) were initially selected to perform quenching studies.
emerge. MA is widely used as the scavenger of HO and SO4 , while TBA is an
In order to further understand the deep reasons of the restrain effective quenching agent for HO [33]. As depicted in Fig. 7, the
effect, we try to explore from the perspective of material proper- addition of either MA or TBA caused no effective decrease of AO7
ties. ACF exhibited the decrease of pHpzc from 7.23 to 1.75 and removal even with a high mole ratio to AO7 of 1000/1. Actually,
the sextuple increase of the total acidity, which possibly a reflec- prior to this study, similar results had been found in the GAC/PS
tion of the increase in acidic functional groups. The analysis of system [17]. It can be explained by the same fact that MA and
functional constituents by FTIR (Fig. 5b) proved this speculation. TBA as hydrophilic compounds are assumed not to be able to accu-
In contrast to ACF0, the OH absorptions of ACF15, ACF30 and mulate near or on the surface of GAC or ACF to a significant extent.
ACF60 at 3430 cm 1 only slightly decreased, and absorption at However, the oxidizing species in the ACF/PMS system may be
S. Yang et al. / Separation and Purification Technology 143 (2015) 19–26 23

Fig. 7. The degradation of AO7 in the ACF/PMS system in the presence of different
radical scavengers. (MA/AO7 molar ratio = 1000/1, TBA/AO7 molar ratio = 1000/1,
phenol/AO7 molar ratio = 50/1, NB/AO7 molar ratio = 50/1).

Fig. 10. (a) Concentration of NO2 and NH+4 (b) concentrations of formic acid and
acetic acid during AO7 degradation by the ACF/PMS system.
Fig. 8. UV–vis spectra change during the process of AO7 degradation.

As seen from Fig. 7, the addition of phenol or NB (with a mole


formed on the surface of ACF, as HO is formed on the surface of the ratio to AO7 of 50/1) significantly inhibited the performance of
GAC in the GAC/H2O2 system [14]. It was assumed that phenol and the ACF/PMS system with AO7 removal decreased from 100% to
nitrobenzene (NB) could move close to the surface of ACF because 36% and 65% in 120 min, respectively. Because phenol can react
of the interaction between the delocalized p electrons of the oxy- with both SO4 and HO with high rates (kSO4 /phenol = 8.8  109
gen-free Lewis basic sites and the free electrons of the dye mole- M 1 s 1, kHO /phenol = 6.6  109 M 1 s 1) [34,35], and that NB has
cule present in the aromatic rings. As a result, these two radical different reaction rate constants with SO4 and HO (kSO4 /NB < 106
inhibitors, phenol and NB, were also utilized to conduct quenching M 1 s 1, kHO /NB = 3.0  109 M 1 s 1) [36,37], the involvement of
experiments. both SO4 and HO was therefore confirmed. In others words, SO4

Fig. 9. GC–MS identified reaction intermediates during AO7 degradation by the ACF/PMS system.
24 S. Yang et al. / Separation and Purification Technology 143 (2015) 19–26

Fig. 11. Possible pathway of AO7 degradation in the ACF/PMS system.

and HO may share the responsibility for AO7 degradation. Further- absorbance at 229 and 310 nm was considered as evidence of aro-
more, it is concluded that the degradation of AO7 does not occur in matic fragment degradation in the dye molecule and its intermedi-
the bulk aqueous phase, but near or on the surface of ACF, since SO-4 ates [40].

and HO are formed on ACF surface and their half-lives are so To further identify the intermediate products, GC–MS analysis
short that they cannot move to bulk aqueous phase. was employed and the products identified are shown in Fig. 9.
And the concentration of NH+4, NO2 , CH3COO and HCOO as a
3.5. AO7 degradation pathway function of reaction time was also determined by IC as shown in
Fig. 10. As the result of Fig. 10, these four kinds of intermediate
To clarify the changes of molecular and structural characteris- products wiped out in 120 min and NO3 was not detected in the
tics of AO7 during the oxidation process in ACF-activated PMS oxi- experiments. In order to verify the inferences that ACF can adsorb
dation, samples were taken at time intervals of 0, 20, 40 and the intermediate products which generated by AO7, the experi-
80 min and representative UV–vis spectra changes as a function ments of adsorption of NO3 , CH3COO and HCOO by ACF were
of reaction time were observed. The corresponding spectra are proceeded. As Fig. S1 indicated that the ACF can adsorb low con-
shown in Fig. 8. The azo dye, AO7, basically consists of an azo link- centration of nitrate rapidly. In addition, the ACF can also adsorb
age (AN@NA), a benzene ring and naphthalene ring, all of which CH3COO and HCOO (Fig. S2).
exhibit different absorbance peaks. The maximum absorption Based on these results, a possible pathway for AO7 degradation
wavelength for AO7 was determined to be 484 nm, which accounts in the ACF/PMS system is proposed in Fig. 11. It is assumed that the
for the orange color of solution and can be attributed to the azo degradation process is initiated by the separation of azo links due
linkage. The two other bands in the ultraviolet region located at to the oxidative attack of SO4 and HO that generated from PMS by
229 and 310 nm are related to the naphthalene and benzene rings, the activation of ACF [41–44]. The free radicals render the forma-
respectively [38,39]. As the reaction proceeded, the visible band tion of sulfanilic acid and 1-amino-2-naphtol [45]. As sulfanilic
remarkably disappeared and this was mainly due to the fragmen- acid has high thermal stability, it can easily dissolve in water.
tation of the azo links by oxidation. In addition, the decay of the And the 1-amino-2-naphtol is sensitive to oxygen, so it can be
S. Yang et al. / Separation and Purification Technology 143 (2015) 19–26 25

oxidized in the air [46]. For these reasons, the GC–MS had not [8] O.S. Furman, A.L. Teel, R.J. Watts, Mechanism of base activation of persulfate,
Environ. Sci. Technol. 44 (2010) 6423–6428.
detected these two products. Then they are further oxidized, the
[9] G.P. Anipsitakis, D.D. Dionysiou, Radical generation by the interaction of
desulfurization lead sulfanilic acid to form 1,4-benzoquinone transition metals with common oxidants, Environ. Sci. Technol. 38 (2004)
[47]. At the same time the further oxidation of 1-amino-2-naphtol 3705–3712.
can product 1,2-naphthalenedione, 2H-1-benzopyran-2-one and [10] S.-Y. Oh, S.-G. Kang, P.C. Chiu, Degradation of 2,4-dinitrotoluene by persulfate
activated with zero-valent iron, Sci. Total Environ. 408 (2010) 3464–3468.
phthalic anhydride [48]. And the oxidation process of sulfanilic [11] M. Kimura, I. Miyamoto, Discovery of the activated-carbon radical AC+ and the
acid and 1-amino-2-naphtol can generate NH+4, NO2 . As proposed novel oxidation-reactions comprising the AC/AC+ cycle as a catalyst in an
in Fig. 9a, in the reaction the concentration of NH+4 is well above aqueous solution, Bull. Chem. Soc. Jpn. 67 (1994) 2357–2360.
[12] C.J. Liang, Y.T. Lin, W.H. Shih, Treatment of trichloroethylene by adsorption and
that of NO2 , and NO3 was not detected from the samples. We persulfate oxidation in batch studies, Ind. Eng. Chem. Res. 48 (2009) 8373–
can deduce that in the AO7 degradation process the N is mainly 8380.
transformed to NH+4. Then, all the organics of intermediate prod- [13] S. Khorramfar, N.M. Mahmoodi, M. Arami, H. Bahrami, Oxidation of dyes from
colored wastewater using activated carbon/hydrogen peroxide, Desalination
ucts undergo ring-opening reactions and lead to the formation of 279 (2011) 183–189.
short-chain aliphatic acids such as CH3COO and HCOO . Finally, [14] G. Wang, T. Wu, Y. Li, D. Sun, Y. Wang, X. Huang, G. Zhang, R. Liu, Removal of
these organics are mineralized into CO2 and H2O [45]. From ampicillin sodium in solution using activated carbon adsorption integrated
with H2O2 oxidation, J. Chem. Technol. Biotechnol. 87 (2012) 623–628.
Fig. 9b, both the concentration of CH3COO and HCOO increased [15] C.M. Domínguez, P. Ocón, A. Quintanilla, J.A. Casas, J.J. Rodriguez, Highly
firstly and then decreased. It can be interpreted that as the reaction efficient application of activated carbon as catalyst for wet peroxide oxidation,
went on, AO7 decolorizing reaction was nearly completed. The Appl. Catal. B: Environ. 140–141 (2013) 663–670.
[16] A. Anfruns, E. Garcıa-Suarezb, M. Montes-Moran, New insights into the
degradation of intermediate products would be the main reaction,
influence of activated carbon surface oxygen groups on H2O2 decomposition
so formic acid and acetic acid began to be degraded into CO2 and and oxidation of pre-adsorbed volatile organic compounds, Carbon 77 (2014)
H2O. Besides part of formic acid and acetic acid would be adsorbed 89–98.
on the surface of ACF. [17] S.Y. Yang, X. Yang, X.T. Shao, R. Niu, L.L. Wang, Activated carbon catalyzed
persulfate oxidation of azo dye Acid Orange 7 at ambient temperature, J.
Hazard. Mater. 186 (2011) 659–666.
[18] J. Zhang, X.T. Shao, C. Shi, S.Y. Yang, Decolorization of Acid Orange 7 with
4. Conclusions peroxymonosulfate oxidation catalyzed by granular activated carbon, Chem.
Eng. J. 232 (2013) 259–265.
ACF was successfully used to activate PMS to generate oxidizing [19] J.M. Valente Nabais, P.J.M. Carrott, M.M.L. Ribeiro Carrott, J.A. Menéndez,
Preparation and modification of activated carbon fibers by microwave heating,
species to efficiently degrade AO7 in aqueous phase at ambient Carbon 42 (2004) 1315–1320.
temperature. In the ACF/PMS system, catalytic oxidation made [20] Q.S. Liu, T. Zheng, P. Wang, J.P. Jiang, N. Li, Adsorption isotherm, kinetic and
the major contribution to the total AO7 removal rather than mechanism studies of some substituted phenols on activated carbon fibers,
Chem. Eng. J. 157 (2010) 348–356.
adsorption. Basic sites on the surface of ACF were possible active [21] X. Ma, F. Zhang, J. Zhu, Preparation of highly developed mesoporous activated
sites for PMS decomposition with radicals generated. SO4 and carbon fiber from liquefied wood using wood charcoal as additive and its
HO, formed on the surface of ACF, were mainly responsible for adsorption of methylene blue from solution, Biores. Technol. 164 (2014) 1–6.
[22] Y.Y. Yao, L. Wang, L.J. Sun, S. Zhu, Z.F. Huang, Y.J. Mao, W.Y. Lu, W.X. Chen,
AO7 degradation. Under the attack of radicals, the degradation pro-
Efficient removal of dyes using heterogeneous Fenton catalysts based on
cess of AO7 is initiated by cleavage of azo linkage and further activated carbon fibers with enhanced activity, Chem. Eng. Sci. 101 (2013)
undergoes ring-opening reactions to cause mineralization. It is 424–431.
practical significance of ACF, an environment-friendly catalyst, dis- [23] Z. Huang, H. Bao, Y. Yao, Novel green activation processes and mechanism of
peroxymonosulfate based on supported cobalt phthalocyanine catalyst, Appl.
plays efficient catalytic activity of PMS activation. Catal. B: Environ. 154–155 (2014) 36–43.
[24] X.F. Qu, J.T. Zheng, Y.Z. Zhang, Catalytic ozonation of phenolic wastewater with
activated carbon fiber in a fluid bed reactor, J. Colloid Interf. Sci. 309 (2007)
Acknowledgement 429–434.
[25] I.M. Kolthoff, I.K. Miller, The chemistry of persulfate. I. The kinetics and
Financial support from the National Natural Science Foundation mechanism of the decomposition of the persulfate ion in aqueous medium, J.
Am. Chem. Soc. 73 (1951) 3055–3059.
of China (No. 21107101) is fully acknowledged. [26] J.P. Reymond, F. Kolenda, Estimation of the point of zero charge of simple and
mixed oxides by mass titration, Powder Technol. 103 (1999) 30–36.
[27] H.P. Boehm, Some aspects of the surface chemistry of carbon blacks and other
Appendix A. Supplementary material carbons, Carbon 32 (1994) 759–769.
[28] J. Yang, P. Juan, Z. Shen, Removal of carbon disulfide (CS2) from water via
adsorption on active carbon fiber (ACF), Carbon 44 (2006) 1367–1375.
Supplementary data associated with this article can be found, in [29] G. Socrates, Infrared Characteristic Group Frequencies, second ed., Wiley,
the online version, at http://dx.doi.org/10.1016/j.seppur.2015.01. Chichester, 1994.
022. [30] J.M. O’Reilly, R.A. Mosher, Functional groups in carbon black by FTIR
spectroscopy, Carbon 21 (1983) 47–51.
[31] M.S. Solum, R.J. Pugmire, M. Jagtoyen, F. Derbyshire, Evolution of carbon
References structure in chemically activated wood, Carbon 33 (1995) 1247–1254.
[32] M.A. Montes-Morán, D. Suárez, J.A. Menéndez, E. Fuente, On the nature of basic
sites on carbon surfaces: an overview, Carbon 42 (2004) 1219–1224.
[1] G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical review of rate
[33] J.Y. Zhao, Y.B. Zhang, X. Quan, S. Chen, Enhanced oxidation of 4-chlorophenol
constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl
using sulfate radicals generated from zero-valent iron and peroxydisulfate at
radicals (OH/O) in aqueous solution, J. Phys. Chem. Ref. Data 17 (1988) 513–
ambient temperature, Sep. Purif. Technol. 71 (2010) 302–307.
886.
[34] M.E. Lindsey, M.A. Tarr, Inhibition of hydroxyl radical reaction with aromatics
[2] H. Zhang, Y.L. Li, X. Zhong, X.N. Ran, Application of experimental design
by dissolved natural organic matter, Environ. Sci. Technol. 34 (2000) 444–449.
methodology to the decolorization of Orange II using low iron concentration of
[35] J. Ziajka, W. Pasiuk-Bronikowska, Rate constants for atmospheric trace
photoelectro-Fenton process, Water Sci. Technol. 63 (2011) 1373–1380.
organics scavenging SO4 in the Fe-catalysed autoxidation of S(IV), Atmos.
[3] H. Yao, P. Sun, D. Minakata, Kinetics and modeling of degradation of ionophore
Environ. 39 (2005) 1431–1438.
antibiotics by UV and UV/H2O2, Environ. Sci. Technol. 47 (2013) 4581–4589.
[36] P. Neta, V. Madhavan, H. Zemel, R. Fessenden, Rate constants and mechanism
[4] L. Wang, Y. Yao, Z. Zhang, Activated carbon fibers as an excellent partner of
of reaction of SO4 with aromatic compounds, J. Am. Chem. Soc. 99 (1977)
Fenton catalyst for dyes decolorization by combination of adsorption and
163–164.
oxidation, Chem. Eng. J. 251 (2014) 348–354.
[37] R.G. Zepp, Nitrate-induced photooxidation of trace organic chemicals in water,
[5] P. Neta, R.E. Huie, A.B. Ross, Rate constants for reactions of inorganic radicals in
Environ. Sci. Technol. 21 (1987) 443–450.
aqueous solution, J. Phys. Chem. Ref. Data 17 (1988) 1027–1284.
[38] A. Azam, A. Hamid, Effects of gap size and UV dosage on decolorization of C.I.
[6] T.K. Lau, W. Chu, N.J.D. Graham, The aqueous degradation of butylated
Acid Orange 7 by UV/H2O2 process, J. Hazard. Mater. 133 (2006) 133–167.
hydroxyanisole by UV/S2O28 : study of reaction mechanisms via dimerization
[39] X. Zhang, Y. Wang, G. Li, J. Qu, Oxidative decomposition of azo dye C.I. Acid
and mineralization, Environ. Sci. Technol. 41 (2007) 613–619.
Orange 7 (AO7) under microwave electrodeless lamp irradiation in the
[7] C. Tan, N. Gao, Y. Deng, N. An, J. Deng, Heat-activated persulfate oxidation of
presence of H2O2, J. Hazard. Mater. 134 (2006) 183–189.
diuron in water, Chem. Eng. J. 203 (2012) 294–300.
26 S. Yang et al. / Separation and Purification Technology 143 (2015) 19–26

[40] H. Li, Y. Gong, Q. Huang, H. Zhang, Degradation of Orange II by UV-assisted [45] H. Zhao, Y. Sun, L. Xu, J. Ni, Removal of Acid Orange 7 in simulated wastewater
advanced Fenton process: response surface approach, degradation pathway, using a three-dimensional electrode reactor: removal mechanisms and dye
and biodegradability, Chem. Res. 52 (2013) 15560–15567. degradation pathway, Chemosphere 78 (2010) 46–51.
[41] S. Zhang, H. Yu, Y. Zhao, Kinetic modeling of the radiolytic degradation of Acid [46] M. Kudlich, M.J. Hetheridge, H.J. Knackmuss, A. Stolz, Autoxidation reactions of
Orange 7 in aqueous solutions, Water. Res. 39 (2005) 839–846. different aromatic o-aminohydroxynaphthalenes that are formed during the
[42] R. Yuan, S.N. Ramjaun, Z. Wang, J. Liu, Effects of chloride ion on degradation of anaerobic reduction of sulfonated azo dyes, Environ. Sci. Technol. 33 (1999)
Acid Orange 7 by sulfate radical-based advanced oxidation process: 896–901.
implications for formation of chlorinated aromatic compounds, J. Hazard. [47] X. Zhong, L. Xiang, S. Royer, S. Valange, V. Barrault, H. Zhang, Degradation of
Mater. 196 (2011) 173–179. C.I. Acid Orange 7 by heterogeneous Fenton oxidation in combination with
[43] D. Zhou, L. Chen b, C. Zhang, A novel photochemical system of ferrous sulfite ultrasonic irradiation, J. Chem. Technol. Biotechnol. 86 (2011) 970–977.
complex: kinetics and mechanisms of rapid decolorization of Acid Orange 7 in [48] X. Chen, X. Wang, H. Xiao, C. Hong, F. Zhu, Y. Yao, Z. Xue, Accelerated TiO2
aqueous solutions, Water. Res. 57 (2014) 87–95. photocatalytic degradation of Acid Orange 7 under visible light mediated by
[44] H.V. Lutze, N. Kerlin, T.C. Schmidt, Sulfate radical-based water treatment in peroxymonosulfate, Chem. Eng. J. 193–194 (2012) 290–295.
presence of chloride: Formation of chlorate, inter-conversion of sulfate
radicals into hydroxyl radicals and influence of bicarbonate, Water. Res., in
press. http://dx.doi.org/10.1016/j.watres.2014.10.006.

You might also like