You are on page 1of 73

Accepted Manuscript

Review

Application of per oxymonosulfate and its activation methods for degr ada
tion of envir onmental or ganic pollutants: Review

Farshid Ghanbari, Mahsa Moradi

PII: S1385-8947(16)31475-9
DOI: http://dx.doi.org/10.1016/j.cej.2016.10.064
Reference: CEJ 15920

To appear in: Chemical Engineering Journal

Received Date: 13 June 2016


Revised Date: 14 October 2016
Accepted Date: 15 October 2016

Please cite this article as: F. Ghanbari, M. Moradi, Application of per oxymonosulfate and its activation methods
for degr adation of envir onmental or ganic pollutants: Review, Chemical Engineering Journal (2016), doi: http://
dx.doi.org/10.1016/j.cej.2016.10.064

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 Application of peroxymonosulfate and its activation methods for degradation

2 of environmental organic pollutants: Review

3 Farshid Ghanbari1a,b, Mahsa Moradi2c,d

4 a-Environmental Technologies Research Center, Ahvaz Jundishapur University of Medical


5 Sciences, Ahvaz, Iran.
6 b-Department of Environmental Health Engineering, School of Public Health, Ahvaz
7 Jundishapur University of Medical Sciences, Ahvaz, Iran.
8 c- Department of Environmental Health Engineering, School of Public Health, Shahid Beheshti
9 University of Medical Sciences, Tehran, Iran.
10 d-Department of Environmental Health Engineering, School of Medical Sciences, Tarbiat
11 Modares University, Tehran, Iran.
12

13 Abstract

14 The degradation of refractory organic compounds to harmless matters is one of the major

15 concerns of environmentalists. Advanced oxidation processes (AOPs) are promising

16 technologies producing the hydroxyl and sulfate radicals for pollutant degradation. Recently,

17 much attention has been paid to producing sulfate radicals by peroxymonosulfate (PMS) as

18 precursor for sulfate radical production. Nowadays, the use of PMS has acquired popularity

19 thanks to its high reactivity and also to its high potential in generating sulfate radical. Actually it

20 is becoming an alternative for hydrogen peroxide and persulfate. PMS is an unsymmetrical

21 oxidant which can be activated to produce both hydroxyl and sulfate radicals. Various methods

22 of PMS activation have been reported in literature including transitional metals (homogenous

23 and heterogeneous), ultraviolet, ultrasound, conduction electron, carbon catalysts and so on.

24 PMS activation has been broadly applied for a wide range of pollutants mostly in aqueous
1
Corresponding author: Tel.: +989122835398.
E-mail address: Ghanbari.env@gmail.com,
(F. Ghanbari)
Corresponding author: Tel.: +989126907929
2

E-mail address: Moradi.env@gmail.com,


(M. Moradi)

1
25 solution. A literature review is carried out on environmental application of PMS in degradation

26 of contaminants to clarify the performance of PMS. This review in detail describes the PMS

27 usage in remediation of environmental pollutants with focus on the different methods of

28 activation and the effect of main operational parameters on PMS-based processes. Moreover, the

29 identification of contribution of each radical is discussed based on quenching experiments and

30 electro spin resonance method. Finally, an overview on applying PMS in real wastewater and

31 other matrixes (air, soil and sludge) is conducted and some recommendations are proposed for

32 future studies.

33 Keywords: Peroxymonosulfate; sulfate radical; organic pollutants; activation methods; advanced

34 oxidation processes.

35 1. Introduction

36 Classically, advanced oxidation processes (AOPs) include in-situ generation of hydroxyl radical

37 with E° = 2.8 V for degradation of organic pollutants in polluted water. AOPs are known to be

38 the most effective method to oxidize bio-recalcitrant organic compounds [1,2]. Hydroxyl radical

39 is defined as a chemical species that has one unpaired electron on the oxygen atom. In fact, it is

40 characterized by one electron deficiency in comparison with OH- as a stable species [3]. There is

41 a variety of methods for generation of hydroxyl radical. Chemical [4], photochemical [5,6],

42 sonochemical [7] and electrochemical processes [8] can either directly or indirectly produce HO•

43 for reaction with organic compounds to destruct organic bonds. However, the use of chemical

44 oxidant is a conventional method for generating the free radicals. These oxidants are ozone (O3),

45 hydrogen peroxide (H2O2) and persulfate (S2O82-) which cannot be efficient for persistent

46 pollutants when are used on their own. Hence, they require activation for generation of free

2
47 radicals for the degradation of toxic and refractory contaminants. Table 1 represents redox

48 potential of common chemical oxidants [9].

49 Table 1. Oxidation potential of commonly used oxidants

Oxidant Oxidation potential (V)

Fluorine (F) 3.0

Hydroxyl radical (HO•) 2.8

Sulfate radical (SO4•-) 2.5-3.1

Ozone (O3) 2.1

Persulfate (S2O82-) 2.1

Peroxymonosulfate (HSO5-) 1.82

Hydrogen peroxide (H2O2) 1.8

Permanganate (MnO4-) 1.68

Chlorine dioxide (ClO2) 1.5

Chlorine (Cl2) 1.4

50

51 In the recent decade, a new chemical oxidant has been introduced by Anipsitakis and Dionysiou

52 for degradation of organic contaminants [10,11]. This inorganic oxidant is peroxymonosulfate

53 anion (HSO5-) that has been mostly used in synthesis of organic chemicals [12]. Moreover,

54 peroxymonosulfate (PMS) has been used as a chlorine-free additive in swimming pools for the

55 purpose of disinfection [13-15]. In fact, instead of looking for a replacement for chlorine

56 sanitizers, PMS acts as an oxidizer along with chlorine, with the usage of around 1-2 pounds per

57 10,000 gallons of pool water. In presence of 25 ppm PMS and 0.1 mg/L Co2+, 99.99% removal

58 of E. coli colonies was observed within 1 h of reaction [13]. PMS is an alternative to chlorine-

59 based bleaching agent which is used in the paper and pulp industry for delignification.

3
60 Peroxymonosulfate anion is derived from Caro’s acid or peroxysulfuric acid (H2SO5) that was

61 firstly described by Heinrich Caro [16]. Caro’s acid is a very reactive and powerful acid which is

62 quickly dissociated in water at neutral pH levels. Formerly, Caro’s acid had not been

63 successfully synthesized in form of stable mono-salts [16,17]. Potassium salt of

64 peroxymonosulfate is stabled in triple salt (2KHSO5·KHSO4·K2SO4) (Fig. 1) [18]. This salt is

65 marketed by Evonik and DuPont under trade names of Caroat and Oxone respectively. Oxone is

66 prepared through reaction of H2O2 solution with oleum and an alkali potassium compound.

67 Oxone is a white crystalline solid, which is very stable, non-toxic, cheap and readily soluble in

68 water (> 250 g/L at 20°C) [19,20]. PMS in water solutions is relatively stable in a way that only

69 5% of active oxygen dropped in 3 days [20]. Moreover, PMS is stable at pH less than 6 and at

70 pH=12. Minimum stability of PMS is observed at pH of 9 in which concentrations of HSO5- and

71 SO52- are equal (pKa2=9.4). Whereas PMS is hydrolyzed to H2O2 at pH < 1 [20-22].

72

73 Fig. 1 The molecular structure of Oxone salt (Figure reprinted with permission from [18]).

74 The peroxymonosulfate ion is a derivate of hydrogen peroxide in which one H-atom is replaced

75 by a SO3- group [23]. The distance of the O–O bond in peroxymonosulfate is 1.460 Å [24]. Some

76 specific characteristics of PMS are presented in Table 2.

77 Table 2. Characteristics of PMS

Peroxymonosulfate ion Properties

4
Structure

Molecule weight 113.0699 g/mol (614.74 as Oxone)

CAS Number 10058-23-8

Solubility in water (> 250 g/L at 20°C (based on Oxone)

Formula HSO5-

Other names Coroat, monopersulfate, Oxone, Curex

78

79 PMS is an unsymmetrical oxidant with redox potential of 1.82 V based on Eq. (1). It can

80 partially oxidize some organic compounds.

81 HSO  
  2H  2e → HSO  H O 1

82 Although PMS is thermodynamically a strong oxidant, its direct reaction with the majority of the

83 pollutants is too slow so that activation is required. PMS can automatically adjust the solution pH

84 with producing protons. Once it is activated, sulfate radical is generated playing an illustrious

85 role in degrading the pollutants. Besides, hydroxyl radical is another precious product of PMS

86 activation. Nowadays, sulfate radical-based AOPs are attracting much attention due to some

87 unique characteristics. Sulfate radical (SO4•-) is a powerful oxidant with high oxidation potential

88 (2.5-3.1 V) which can be generated by scission of peroxo-bond of either peroxymonosulfate

89 (PMS) or persulfate (PS) during their activation [25-27]. Unlike PMS, PS (S2O82-) is a symmetric

90 oxidant consisting of two groups of SO3- in O‒O bond with distance of 1.497 Å. Persulfate ions

91 are found in forms of sodium and potassium salts with high solubility and stability which are

92 commonly used in water and soil remediation [28,29].

5
93 Sulfate radical has a higher half-life than hydroxyl radical (30-40 µs vs 20 ns) mainly due to its

94 preference to react with organics during electron transfer while hydroxyl radical acts non-

95 selectively and participates in various reactions with equal preference [30-34]. Sulfate radical

96 based processes are known as very efficient promising processes for the degradation of

97 recalcitrant organics mainly due to significant oxidation ability of sulfate radical as well as slow

98 utilization of precursor oxidants. Applying sulfate radical, the reaction mechanism is normally

99 electron transfer. In case of hydroxyl radical, electron transfer along with hydrogen atom

100 abstraction is supposed to be the reaction mechanism which is less outstanding [35,36]. In order

101 to generate sulfate radical, PMS should be decomposed in presence of an activator. PMS

102 activators are categorized in these main classes: transitional metals, ultraviolet (UV), heat,

103 ultrasound (US), electron conduction and carbon-catalyst.

104 In recent years, three review articles and a book chapter have been published related to sulfate

105 radical application. In the book chapter [37] and one of the articles [30], authors have focused on

106 sulfate radical generation by PS and PMS activations. Heterogeneous cobalt based catalyst was

107 considered by another review article with focus on synthesizing and application of heterogeneous

108 catalyst for only cobalt based catalyst [35]. In the last review article, all aspects of heterogeneous

109 catalyst for main transitional metals (Co, Fe, Cu and Mn) and their combinations were reviewed.

110 Moreover, nonmetal carbon catalysts were also considered with details [38]. In present review,

111 we indicated all methods of PMS activation including homogenous and heterogeneous

112 transitional metal catalyst, metal-free heterogeneous catalysts, ultraviolet, ultrasound, conduction

113 electron and miscellaneous activation methods. In addition, PMS applications were reported in

114 treatment of real wastewaters, polluted air, soil and sludge.

115 2. Homogenous transitional metals for activation of PMS

6
116 Transitional metals have been extremely applied to activate the oxidant to produce free radicals

117 for the degradation of organic compounds. Fenton process involves a classic reagent (H2O2 +

118 Fe2+) which has been successfully used for degradation of various pollutants [39].

119 Fe  H O → HO•  Fe  OH  2

120 Based on Fenton reagent, a highly efficient method has been reported by Anipsitakis and

121 Dionysiou in which PMS was activated by cobalt ions to generate sulfate radical for degrading

122 the organic contaminants. Anipsitakis and Dionysiou showed that Co2+/PMS can be an

123 alternative method for degradation of organic compounds [11]. Excellent performance in neutral

124 pH is an applicable benefit of Co2+/PMS system in comparison with Fenton reagent. It should be

125 noted that in 1956, Ball and Edwards reported the decomposition of peroxymonosulfate by

126 cobalt ions for the first time [40].

Co  HSO
 → Co

 SO• 
 OH 3

127 Anipsitakis and Dionysiou have reported the activation of PMS by different transitional metals

128 for degradation of 2,4-Dichlorophenol (2,4-DCP). The results showed that PMS activation by

129 transitional metals for 2,4-DCP degradation followed this order: Ni2+< Fe3+ < Mn2+< V3+ < Ce3+<

130 Fe2+< Ru3+< Co2+[41]. The following reaction mechanisms of Co2+ with PMS are suggested [42]:

Co  H O → CoOH  H 4

CoOH  HSO •
 → SO  CoO  H O 5

CoO  2H → Co  H O 6

Co  HSO •
 → SO  Co

 H 7

Co  SO•
→ Co

 SO 
8

131 2SO• • •
 → SO O SO  9

SO• • •
O SO  → O  2SO 10

SO• • 
O SO  → O  S O 11

7
132 The above reactions are illustrated in Fig. 2 [43,44].

133

134 Fig. 2. Mechanism for the sulfate radical chain reaction (Figure reprinted with permission from
135 [43])
136
137 After these basic studies by Dionysiou’s group, several researchers have attempted to clarify

138 different aspects of PMS activation by transitional metals for degradation of organic pollutants.

139 PMS can be activated by ferrous ions for generating free radicals. Cobalt and ferrous ions are the

140 most common catalysts for activation of PMS. The reactions of PMS decompositions by Fe2+ are

141 presented by the following equations [45-49].

142 Fe  HSO
 → Fe

 SO• 
 OH 12

Fe  HSO •
 → SO  Fe

 H 13

Fe  SO•
→ Fe

 SO 
14

HSO •  •
  SO → SO  SO  H 15

HSO  •
  H O → SO  2HO  H 16

SO•  •
  2H O → SO  2HO  H 17

143 Based on the above reactions, transitional metals with higher oxidation state (Co3+ and Fe3+)

144 generate peroxymonosulfate radicals (SO5•-) while Co2+ and Fe2+ decompose PMS to SO4•- at the

145 first step. In other words, Co2+ and Fe2+ are superior for PMS activation to generate powerful

146 reactive radicals. The mechanism of PMS activation by homogenous metal catalysts is

147 schematically illustrated in Fig. 3.


8
148

149 Fig. 3 The scheme of PMS activation by transitional metals as homogeneous catalyst

150 2.1 The effect of PMS and transitional metal dosages

151 Similar with Fenton process, PMS and catalyst dosages should be in appropriate ratio to achieve

152 the optimum condition for maximum production of sulfate radical. According to Eqs. 14 and 15,

153 excessive amounts of PMS and transitional metals can act as scavenger for sulfate radical. In

154 fact, higher dosage of PMS and transitional metals can be the limiting agents for efficiency of

155 sulfate radical-based processes. Rastogi et al. stated that the optimum molar ratio between

156 ferrous ion and PMS was 1:1 while with increase or decrease of this ratio, scavenging effect of

157 excess catalyst and PMS occurred [47]. The ratios of 1:1 and 3:1 were frequently observed in

158 literature for PMS/transitional metal [50-55]. It was shown that the inhibitory effect of excessive

159 catalyst (Co2+) was more observed in comparison with excessive PMS [53]. Generally, oxidant

160 concentration was more than catalyst dosage in a way that molar ratio of 10000 for Oxone:Co2+

161 was reported for landfill treatment, this high molar ratio was attributed to high organic content in

162 high strength wastewater [56]. PMS concentration more than optimum condition decreased

163 function of the system which can be attributed to scavenging effect of excessive PMS producing

164 SO5•- with less reactivity (E0=1.1 V) in comparison with SO4•- [54]. In absence of homogeneous

9
165 catalyst, organic pollutant was not degraded for any PMS concentration indicating that sole

166 application of PMS cannot produce free radicals [57].

167 2.2 The effect of pH

168 In chemical oxidation processes, pH plays a key role in decomposition of oxidant to generate

169 free radicals. It also affects the formation of the species of transitional metals and their

170 availability for reaction with the oxidant. 2,4-DCP removal was investigated in Co2+/PMS and

171 Fenton processes at different pH values. Considering the results, it was confirmed that pH

172 significantly affected homogenous Fenton process while Co2+/PMS could provide a similar

173 efficiency in different pH values [11]. Huang et al. expressed that in acidic condition, the

174 decolorization rate was enhanced markedly as the solution pH increased from 3.5 to 5 while

175 decolorization decreased considerably as the pH of the solution increased from 6 to 8.4. At low

176 pH (pH<5), hydrogen ion (H+) scavenges sulfate and hydroxyl radicals according to Eqs. (18)

177 and (19), respectively [58]. This hypothesis has also been reported by Sun et al. [59].

HO•  H  e → H O 18

178 SO•  •
 H  e → HSO 19

179 Ji et al. indicated that acidic and basic conditions cannot decompose PMS into radical forms. At

180 high pH, the ability of catalyst (cobalt ion) was significantly reduced since Co(OH)2 precipitation

181 was formed. In addition, at high pH, the formation of HCO3−/CO32− as two well-known radical

182 scavengers can reduce degradation of the pollutant [42]. It has been asserted that complex

183 compounds of bicarbonate-Co and peroxyphosphate-Co could be produced in presence of

184 bicarbonate and dihydrogen phosphate ions behaving as activators for decomposition of PMS in

185 pH range of 5 to 7 [56]. Self-decomposition of PMS at pH of 9.0 and the precipitation of metals

186 at pH > 3.0 are rational reasons for good efficiency of the homogenous system at pH=3.0. [47].

187 The reduction of efficiency in PMS/Fe2+ at acidic pH is presumably related to the forming

10
188 (Fe(II)(H2O))2+ resulting in a drop in availability of free ferrous ion in the solution. The solution

189 pH was dramatically reduced after the addition of PMS. This reduction was attributed to the

190 presence of acidic bisulfate in Oxone salt [51].

191 2.3 Effect of non-target natural water constituents

192 Chloride ion (Cl-) is one of the major inorganic anions in water resource and almost all natural

193 waters. The chloride ion can scavenge free radicals produced in AOPs. The effect of Cl- has been

194 studied in HO•-based processes. This issue has also been considered in SO4•--based processes.

195 Chan and Chu [60] indicated that chloride ion had an inhibitory effect compared to nitrate,

196 acetate and sulfate anions. They expressed that the inhibitory effect of chloride was attributed to

197 consumption of the PMS and the sulfate radical in the process that consequently, led to the

198 formation of Cl• and Cl2 which were less reactive oxidants [60,61].

SO•   •
 Cl → SO  Cl 20

Cl•  Cl → Cl•


21

Cl• • 
 Cl → 2Cl  Cl 22

Cl•  Cl• → Cl 23

Cl  HSO 
 → SO  HOCl 24

2Cl  HSO 
  H → SO  Cl  H O 25

199 Moreover, bicarbonate ion had a detrimental effect on performance of PMS/Co2+ system.

200 Bicarbonate and carbonate quench free radicals of HO• or SO4•- to generate HCO3• and CO3•-

201 which are less reactive radical [42]. Wang et al. have studied the effect of chloride ion

202 concentration on Orange II decolorization by PMS/Co2+ (Fig. 4). Chloride ion had a dual effect

203 in PMS/Co2+ system. Chloride ion in low concentration (5 mM) decreased removal efficiency

204 while high concentration of chloride ion (500 mM) increased the rate constant by 4-15 fold. This

11
205 bilateral effect of Cl- on dyes decolorization was also observed in other halide ions (Br− and I−)

206 [62,63].

207

208 Fig. 4. Bleaching of Orange II (0.2 mM) solution containing Oxone (2.5 mM) and Co2+ (0.1
209 mM) in presence of various concentrations of chloride (Figure reprinted with permission from
210 [62]).
211
212 Nitrate ions showed scavenging effect in high concentration for sulfate radical [51,63].

213 Phosphate ions also demonstrated a scavenging effect for sulfate radical as well as hydroxyl

214 radical [58]. It was reported that the order of inhibitory effect of some anions followed S2O32-

215 >HCO3->CO32->NO2 ->HCOO->Cl->HPO42->NO3 ->H2PO4->SO42->CH3COO-. Indeed, anions

216 compete with organic compounds for reaction with sulfate radicals [64].

217 In addition, it was reported that presence of chloride ion in PMS/Co2+ system led to halogenation

218 of organic compounds in water. Chlorine and chloride radical atoms are electrophilic which can

219 be added to unsaturated bonds of organic compounds and produce AOX (adsorbable organic

220 halides) which are persistent to biodegradation. Actually, increase in Cl- concentration increases

221 the possibility of AOX formation or halogenation of organic compounds [63,65]. It should be

222 noted that the presence of Cl- increased toxicity of by-products in Co/PMS while it did not

223 change the toxicity of by-products of UV/H2O2 [66]. The presence of NOM in water increases

224 formation potential of halogenated disinfection byproducts in sulfate radical based system

12
225 through reconfiguration of NOM to phenolic intermediates which are susceptible to halogenation

226 [67]. The presence of Br- in PMS/Co2+ system resulted in formation of the Br-disinfection by-

227 products (Br-DBPs). Moreover, PMS could directly oxidize Br- for generation of bromine and

228 production of Br-DBPs consequently [68,69].

R•  Cl• 
→ R − Cl  Cl 26

R − H  HOCl → R − Cl  H O 27

229 A variety of pollutants has been considered by researchers to investigate performance of

230 PMS/transitional metals system (Table 3).

231 Table 3. The homogeneous activation of PMS by transitional metals for different pollutants
Activator Pollutant Condition Removal References

efficiency

PMS=0.25 mM, Co2+=0.25 mM, pH=3, and 20 100% [50]


2+
Co Carbamazepine
min time

Molar ratio of oxidant to 95% [53]

Cu2+ Triclosan metal = 1:1, molar ratio of oxidant to

triclosan= 5:1, pH=7 and time=10 min

Molar ratio of oxidant to 95% [53]

Fe2+ Triclosan metal = 1:1, molar ratio of oxidant to

triclosan= 40:1, pH=7 and time=120 min

Molar ratio of oxidant to 95% [53]

Co2+ Triclosan metal = 1:1, molar ratio of oxidant to

triclosan=3:1, pH=7 and time=10 min

4-chloro-2- PMS = 2.0 mM, Co2+ = 0.08 mM, pH=3.9, >95% [63]

Co2+ nitrophenol/ 4- 4C2NP = 4-CP= 0.4 mM and time=120 min

chlorophenol

Co2+ 2,4,6-trichlorophenol PMS = 10.0 mM, Co2+ = 0.1 mM, pH=3.0, About 100% [70]

13
2,4,6-trichlorophenol = 0.2 mM and time=5

min

PMS = 0.44 mM, Fe2+ = 0.44 mM, pH=3 and About 100% [47]
2+
Fe 2-chlorobiphenyl
time= 30min

Oxone = 0.5 mM, Co2+ = 1 mM, Monuron = About 60% [71]


2+
Co Monuron
0.2 mM and time = 30min

Oxone = 5.9 mM, Co2+ = 0.015 mM, pH = 2.3 86-94% [72]


non-ionic surfactant
Co2+ Brij 35 = 680–2410 mg/L, and 24 h reaction
Brij 35
time

Oxone = 4.5 mM, oxone/Co2+ = 10000, COD = 57.5% [56]

Co2+ Landfill leachate pH=6.5, temperature = 30°C, time=300 min, Color = 83.3%

number of stepwise addition = 7

PMS = 0.4 mM, Co2+ =0.13 mM, pH = 3, basic 100% [54]


2+
Co basic blue 9
blue 9= 7 mg/L and time = 5 min

PMS=0.4 mM, Co2+=0.13 mM, pH=3, acid red 100% [54]


2+
Co acid red 183
183 = 160 mg/L and time =30 min

PMS = 1 mM, Fe2+ = 1 mM, pH= 3.0, Atrazine <50% [42]


2+
Fe Atrazine
= 50 µM and time =15 min

PMS = 1 mM, Co2+ = 10 µM, pH= 7.0, 100% [42]


2+
Co Atrazine
Atrazine = 50 µM and time = 45 min

Oxone = 2.5 mM, Co2+ =0.1 mM, Orange II = 100% [62]


2+
Co Orange II
0.2 mM, and time = 350 s

Molar ratio of SMX:PMS:Co at 1:10:10, About 100% [52]

Co2+ Sulfamethoxazole Sulfamethoxazole = 9 mg/L, pH = 7 and time =

5 min

Molar ratio of SMX:PMS:Co at 1:80:80, 95% (180 min) < [52]


Fe2+ Sulfamethoxazole Sulfamethoxazole = 9 mg/L and pH = 7 TOC = 55% (24

h)

14
Molar ratio of RhB: Fe2+:oxone = 1:10:10 , 100% [51]
2+
Fe Rhodamine B (RhB)
RhB = 0.2 mM, pH=3.51 and time = 90 min

N,N-diethyl-m- PMS = 250 µM, Fe2+ = 250 µM, pH = 7, 42% [73]


2+
Fe
toluamide (DEET) DEET= 25 µM and tim e= 80 min

Oxone= 1.0 mM. Fe2+=0.20 mM, pH = 3.0, 80% [45]


2+
Fe Methyl orange (MO)
MO= 0.10 mM and time =8 min

Mn2+=PMS= 1.244 mM, pH=7, 2,4-DCP = 24% [41]


2+
Mn 2,4-DCP
50.7 mg/L and time=4 h

PMS = 0.08 mM, Co2+ = 1 ppb, pH= 6, RBB = > 90% [58]
Reactive Black B
Co2+ 0.01 mM, temperature = 25°C and time= 30
(RBB)
min

PMS = 2.0 mM, Co2+ = 20 µM, pH = 3.0, DDT 100 [74]

Co2+ DDT = 1 mg/L, temperature = 30°C and time = 120

min

PMS = 0.2 mM, Co2+ = 0.5 µM, pH = 8.0, 96% [75]


Tetrabromobisphenol
Co2+ TBBPA = 9.2 µM, temperature = 20°C and
A (TBBPA)
time= 30 min

232

233 3. Heterogeneous transitional metals for PMS activation

234 In homogeneous systems, mass transfer limitation is insignificant and the available catalyst in the

235 aqueous solution freely reacts with PMS effectively. There are some limitations in application of

236 homogenous catalysts. In general, the recovery of homogenous catalysts is very difficult and

237 their separations require further steps that are not often technically or economically feasible [76].

238 Moreover, the stoichiometric amounts of catalysts for activations of high PMS concentration can

239 be almost high especially in the treatment of high strength wastewaters. Hence, their residuals in

240 effluents should be considered as an ulterior problem. In addition, sub-stoichiometric and over-

15
241 stoichiometric amounts of the catalysts affect the generation rate of sulfate and hydroxyl radicals.

242 Finally, transitional metals species are highly dependent on pH as they form the hydroxide

243 precipitates in alkaline pH and hydrated species in acidic pH which reduce availability of the

244 catalysts. The efficiency of the catalyst can be decreased by the interference of different agents,

245 the most important of which is the presence of various contaminants and inorganic compounds

246 that can be bonded to the catalysts. In case of cobalt ion, although low cobalt concentration is

247 used for activation of PMS, the most important problem is related to the potential environmental

248 and health impact of cobalt in aquatic environment. Cobalt may have serious health effects on

249 humans such as asthma like allergy, damage to the heart, thyroid and liver. Cobalt may also

250 make genetic changes in living cells [15,35,77]. Co(II) concentration of 0.05 mg/L has been

251 fixed for tolerance limit in drinking water [78]. Therefore, in order to achieve this concentration,

252 further processes are required. Indeed, the post treatment of residual cobalt in homogeneous form

253 increases the costs of the treatment process [79-81]. In this way, the catalyst is in solid phase

254 which can be readily separated from liquid phase and it is less harmful to the environmental [82].

255 In case of PMS activation, various heterogeneous catalysts have been used based on different

256 transitional metals (or their combinations) and nonmetal catalysts (carbon catalysts). The

257 mechanism of heterogeneous activation of PMS is schematically presented in Fig. 5.

16
258

259 Fig. 5 The scheme of heterogeneous activation of PMS

260 In this review, a variety of the heterogeneous catalysts used for activation of PMS is

261 investigated. In heterogeneous catalyst, the surface of catalyst plays an important role since

262 reactions occur at the surface and therefore mesoporous catalysts have been considered for

263 decomposition of PMS.

264 3.1 Cobalt oxides

265 Five types of cobalt oxide have been identified including CoO, CoO2, CoO(OH), Co2O3 and

266 Co3O4 [35]. Among them, CoO, Co2O3 and Co3O4 have been considered for activation of PMS to

267 degrade various pollutants. These oxides can be used along with various supports such as

268 carbon-based materials and zeolites or, used merely as nano-particles. However, Co3O4 nano-

269 particles exhibit excellent performance in case of PMS decomposition due to presence of Co(II)

270 and Co(III) in molecule structure. Therefore, the synthesis and application of Co3O4 nano-

271 particles as heterogeneous catalysts have been considered. Dionysiou’s group firstly explored

272 activity of CoO and Co3O4 (CoO‚Co2O3) for decomposition of PMS to oxidize 2,4-DCP. They

273 reported that at acidic conditions, Co from Co3O4 was slowly dissolved in solution, reaching 0.73

274 mg/L after 2 h of reaction time while dissolved cobalt ion was nearly 0.07 mg/L at neutral pH.

17
275 They believed that CoO component of Co3O4 (CoO.Co2O3) was responsible for leaching the

276 cobalt ion to solution since they had formerly declared that the CoO/Oxone system had been

277 homogeneous leading to significant dissolution of Co at acidic pH [83]. The stability of nano

278 Co3O4 in neutral condition was also reported by Chen. et al. [81]. It has been acclaimed that CoO

279 (involved in Co3O4) might also have contributed in the heterogeneous decomposition of Oxone

280 [84]. Wang et al. applied Co3O4 nanorods for activation of PMS to degrade phenol. PMS and

281 Co3O4 nanorods were not effective when they were used solely. It was observed that increase of

282 calcination temperature led to decrease of the catalytic activity of cobalt oxide [85].

283 Although Co3O4 nanoparticles exhibit good catalytic activity, they can easily agglomerate during

284 catalytic reaction, resulting in the reduction of catalytic performance [86]. Cobalt oxides and

285 metallic Co can be immobilized on various materials in order to improve catalytic activity (Fig.

286 6). Immobilization of cobalt species can reduce leaching of cobalt ions and prevent release of

287 nanoparticles to the environment [87]. Moreover, supported catalysts enhance stability of

288 catalyst and increase their reusability as consequence. It should be notified that supported

289 catalysts may simplify the separation of catalyst from solution.

290

291 Fig. 6 Supports used for immobilizing different cobalt species

292

293 3.1.1 Co-Metal oxide

18
294 In recent years, metal oxides have been considered as support for cobalt species in order to

295 activate PMS. Ding et al. used Co3O4–Bi2O3 nanocomposite oxides (CBO) for degradation of

296 different pollutants (methylene blue (MB), RhB, phenol and 2,4-dichlorophenol (2,4-DCP)). The

297 results showed that rate constant of CBO/PMS was 8.6 folds more than that of the Co3O4/PMS in

298 terms of MB removal. They reported that cobalt ion leaching was highly dependent to

299 calcination temperature and pH [88]. Stoyanova et al. immobilized cobalt and iron oxide on

300 MgO (CoFe2/MgO). They showed that the presence of Fe in the catalysts (both bulk and MgO

301 supported) negatively affected the efficiency which was referred to the oxidation state of Fe ions

302 in the synthesized Fe–Co mixed oxide systems. The ferric ion is an electron acceptor and cannot

303 directly activate PMS for generation of sulfate radical [89]. Cobalt implanted TiO2 (Co-TiO2)

304 and Co3O4/TiO2 have been used for activation of PMS. Compared to Co3O4/TiO2, the cobalt ion

305 implanted TiO2 nanocatalysts showed comparable catalytic activity with half the amount of

306 cobalt introduced into the catalysts. Co–TiO2 was more stable than Co3O4/TiO2 in terms of

307 reusability [90]. Various metal oxides (MgO, ZnO, Al2O3, ZrO2, P25 (TiO2)) have been exerted

308 for the activation of PMS. The presence of hydroxyl groups on the surface of MgO formed

309 CoOH+ intermediate which could promote decomposition of PMS. The order of activity of

310 supports was MgO > ZnO > TiO2 > ZrO2 > Al2O3 [91]. Bulk α-MnO2 and Co3O4 with low

311 activity could activate PMS for generation of sulfate while Co3O4/MnO2 nanoparticles

312 effectively activated PMS with 100% removal of phenol only in 20 min [92].

313 3.1.2 Co-carbon-based supports

314 In order to provide a large surface area for adsorption of organic compounds, carbon based

315 supports have been used for catalytic decomposition of PMS as an efficient heterogeneous

316 oxidation process. Carbon materials are attractive supports since carbon is non-metal, earth

19
317 abundant, chemically inert and electrically conductive element [93,94]. Activated carbon,

318 graphite, graphene, activated carbon fiber and fly ash were used for supporting the cobalt for

319 decomposition of PMS. A high synergistic effect was observed in application of Co3O4/graphene

320 oxide (GO) in comparison with sole applications of Co3O4 and GO for decomposition of PMS.

321 The Co–OH complexes formed were corresponded to activation of PMS [95]. In Co3O4/GO

322 system, increase in temperature improved efficiency through rupture of the O-O bond of PMS

323 and generation of sulfate radical consequently [96]. In addition, the use of graphene as support in

324 graphene/Co3O4 reduced considerably leaching of Co2+ compared to Co3O4 [97]. Expended

325 graphite (EG) was also chosen as support for cobalt oxide (Co3O4) to activate PMS. The results

326 showed that with increase of reuse number of EG/Co3O4, removal efficiency dropped while

327 leaching of cobalt ions was slightly increased [98]. In activated carbon fiber (ACF)/PMS, 35%

328 removal efficiency was obtained, while almost 100% decolorization was achieved in

329 Co@ACF/PMS which suggested the effective presence of cobalt ions onto ACF. The leaching

330 tests implied that Co ions were stabilized onto ACF since no amounts of Co ions were detected

331 in aqueous solution [79]. Co/Activated carbon (AC) was used for activation of Oxone. It was

332 determined that Co2O3 was the major form of Co species on AC. Co/AC was comparable with

333 the homogeneous process (Co2+/Oxone) [77]. Carbon xerogel (CX) [99] and carbon aerogel

334 (CA) [100] have been used as supports for Co. Cobalt impregnated CX showed high catalytic

335 activity in comparison with Co/CA and Co doped CX. It was observed that homogeneous

336 degradation of phenol using Co2+/Oxone was faster than heterogeneous Co/CA–Oxone.

337 Nanoflakes Co(OH)2 supported by reduced graphene oxide (rGO) as heterogeneous catalyst was

338 used for PMS activation. Complete pollutant removal was obtained during only 10 min which

339 was much higher than the efficiencies obtained by using carbon and oxide supported Co3O4

20
340 catalysts systems [101]. Zhou et al. prepared carbon microspheres supported cobalt catalysts

341 (Co/CS) for decomposition of phenol. Co/CS with calcination temperatures of 300, 400 and 500

342 could completely decompose 20 ppm phenol in 15, 5 and 10 min, respectively [90]. Metallic

343 cobalt (Co0) was also anchored on graphene nanosheets [102] and carbon nanotubes [103].

344 Similar with Fenton–like, zero valent cobalt releases Co2+ through aerobic oxidation and reaction

345 of Co with PMS. Moreover, Co3+ can be deposited on the zero valent cobalt surfaces to release

346 Co2+. The following reactions are based on reaction of zero valent cobalt with PMS.

Co $ → Co  2%  & = 0.28 ) 28

2 ≡ Co$  O  2H O → 2 ≡ Co  4OH  29

≡ Co$  2HSO
 → ≡ Co

 2SO• 
 2OH 30

≡ Co$  HSO
  2H → ≡ Co

 HSO
 H O 31

≡ Co$  2 ≡ Co → 3 ≡ Co 32

347 3.1.3 Other supports for cobalt

348 Various materials have been used as support for cobalt oxides and metallic cobalt including

349 zeolite, red mud, SBA-15 and so on. These materials often have high specific surface area, high

350 mechanical strength, excellent chemical stability, and low cost [104]. Table 4 displays different

351 supports for cobalt in heterogeneous catalysts to activate PMS. As can be seen, high performance

352 of supports and the reduction of leaching of cobalt ion make increase of the application of

353 supporters in heterogeneous catalysts.

354 Table 4. Various supports for cobalt in heterogeneous PMS activation for different pollutants

355 degradation

Removal TOC
Type of Amount of
Supporters Conditions efficiency removal References
pollutants leaching Co
(%) (%)

21
Phenol = 50 mg/L, Oxone/phenol

Co/SBA-15 Phenol mole ratio = 4:1, [Co/SBA-15- 98 98 85 µg/L [105]

400(5)] = 0.1 g/L, pH ∼7

CAF = 0.05 mM, PMS = 0.2 mM,

Co-MCM41 Caffeine catalyst = 200 mg/L, 100 13.23 - [106]

pH = 7.10, time = 15 min

0.2 ppm (in


Phenol = 25 mg/L, PMS= 2 g/L,
Co-ZSM-5 Phenol 100 - the third [107]
catalyst= 0.4 g/L, time=350 min
recycled test)

Phenol= 30 mg/L, Oxone= 2g/L,


Co/SBA-15 Phenol 100 - [108]
catalyst= 0.2 g/L, time=200 min

RhB = 30 mg/L, molar ratio of


<30 (4h
Oxone/RhB = 10/1, catalyst= 0.2
CoMg/SBA-15 RhB 100 reaction 80 µg/L [109]
g/L, time=5 min, T=25 ◦C, no
time)
solution pH adjustment

Phenol= 25 mg/L, Oxone = 0.4

Co-Red mud Phenol g/L, catalyst = 0.2 g/L, 100 - - [80]

time=50 min

Co-Zeolitic Imidazole RhB = 50 mg/L, PMS = 0.15 g/L,

Framework-67 (ZIF- RhB catalyst = 50 mg/L and time=50 About 100 - 0.8 mg/L [110]

67) min

0.05 ppm

(time=8min
Co3O4/polytetrafluoro Orange II = 0.05 mM, Oxone = 1
Orange II 100 - and in [111]
ethylene mM, pH=7.0 and time=<100 min
presence of

light)

Co/ natural Phenol= 30 mg/L, Oxone= 0.2

zeolites from Phenol g/L, catalyst= 0.4 g/L, T=25 ◦C 100 - - [112]

Indonesia (INZ) and time=300 min

Co/ natural Phenol Phenol= 30 mg/L, Oxone= 0.2 70 [112]

22
zeolites from g/L, catalyst= 0.4 g/L, T=25 ◦C

Australia (ANZ) and time=300 min

72.2% for

yolk−shell Co3O4@ 4-CP = 0.78 mM, PMS = 0.8 1:1


4-
Metal−Organic mM, catalyst PMS/4-
chlorophen 100 - [113]
Framework = 0.5 g/L, pH = 7.0, T = 298 K CP molar
ol (4-CP)
nanocatalyst and time = 30 min ratio (time

= 4.5 h)

RhB= 5.0 mg/L, PMS/RhB

CoFe/SBA-15 RhB molar ratio 20:1, catalyst=10 96 32.4 µg/L [114]

g/L, T=25 °C, Time=120 min

AO II= 20mg/L, PMS=1 g/L,

Fe3O4@C/Co AO II catalyst= 0.2 g/L, pH=4.0, 99% 0.08 mg/L [115]

T=25°C and time=40 min

Periodic mesoporous
Phenol=10 mg/L, PMS=6
aluminum
Phenol mM, catalyst=20 mg/L and 100 3.2 mg/L [116]
phosphonate
time=50 min
(PMAP)/Co

Nanospherical
Phenol=30 mg/L, Oxone=6
hexagonal
Phenol mM, catalyst=20 mg/L and 80 [117]
mesoporous silica
time=60 min
(HMS)/CoO

Phenol=25 mg/L, Oxone=2


Indonesia natural
Phenol g/L, catalyst= 0.4 g/L, T=25°C 100 [118]
zeolite/Co
and time=300 min

356

357 3.2 Transitional mixed metal spinels

23
358 The mixed metal oxides have mostly been used as heterogeneous catalyst for activation of

359 oxidant. Transitional metal spinels are efficient catalysts for the activation of oxidants under

360 formulation of AxB3-xO4 in which A and B can be metals such as Co, Fe, Cu, Mn, Zn, etc. [119].

361 However, these oxides are usually based on Co and Fe elements as main activators of PMS. In

362 this way, various catalysts have been synthesized and evaluated to activate PMS. In nano-ferrite

363 cobalt (Cox-3FexO4), the effect of cobalt ion on RhB decolorization was studied. The results

364 indicated that the presence of cobalt in the structure of nano ferrite is strongly appropriate for

365 RhB decolorization. Moreover, the decolorization rate increased with the increase of cobalt

366 content in the Cox-3FexO4 [120]. The sulfamethazine (SMZ) removal efficiency in CuCo2O4/PMS

367 process was 87.2% while the removal efficiencies were 51.1%, 11.3%, 12.5%, and 7.9% for

368 Co3O4, CuFe2O4, CuO, and Fe3O4, respectively. It has been shown that Co2+ was generated on

369 the surface of the catalyst which was the main responsible agent of generating the sulfate radical.

370 This study and the previous study confirmed that the presence of cobalt in spinel nanocatalysts

371 was a key issue for generation of sulfate radical. It was suggested that Cu2+ containing oxide was

372 more reactive for PMS activation. Therefore, Cu2+ in oxide or spinel surface is less likely to be

373 reduced to Cu+ in the first step when it is used as the catalyst. The sequencing reactions for

374 activation of PMS for CuCo2O4 are as follows [121]:

375 ≡ Co  HSO


 → ≡ Co

 SO•
  H 33

376 ≡ Co  HSO
 → ≡ Co

 SO• 
 OH 34

377 ≡ Cu − OH   HSO
 →≡ Cu

− OHOSO 
  OH 35

378 ≡ Cu − OHOSO
 →≡ Cu

− OH   SO•
36

379 ≡ Cu − OH   HSO


 →≡ Cu

− OOSO•
  H O 37

380 ≡ Cu − OOSO•
  2H O →≡ Cu

− OH   HSO
 O  2H 38

24
381 Ren et al. synthesized magnetic ferrospinel MFe2O4 (M = Co, Cu, Mn and Zn) for activation of

382 PMS. CoFe2O4 showed the highest catalytic performance. The removal efficiencies were 81.0%,

383 62.3%, 42.3%, and 30.0% for CoFe2O4/PMS, CuFe2O4/PMS, MnFe2O4/PMS and ZnFe2O4/PMS,

384 respectively. Moreover, they showed that system performance was strongly depended on

385 concentration of OH sites on catalyst [122]. In another study, the catalytic activity of Fe3−xMxO4

386 (M=Cr, Mn, Co and Ni) was assessed for degradation of Acid Orange II (AOII). With increase of

387 M content the degradation efficiency was increased except in case of Cr by which no change was

388 observed. [123]. Yang et al. have stated that Co (II) is a species of cobalt in CoFe2O4 that can

389 effectively activate PMS. Additionally, the strong Fe-Co in CoFe2O4 reduced leakage of Co

390 significantly [124]. Octahedral CoFe/CoFe2O4 submicron composite showed excellent

391 reusability thanks to its large size and structural stability. The presence of zero valent iron and

392 cobalt on CoFe2O4 enhanced the electron transfer between PMS and CoFe2O4 to facilitate free

393 radical generation [125]. Another spinel metal oxide, nano copper ferrite (CuFe2O4), exhibited a

394 great performance for activation of PMS [126-128]. In order to improve the performance,

395 CuFe2O4 supported AC was used for activation of PMS to degrade MB. Actually, in absence of

396 CuFe2O4, PMS influenced surface chemistry of AC which had a negative effect on MB

397 adsorption because the surface of AC was become acidic by PMS [129]. In CuFe2O4/PMS, Cu2+

398 leaching was 30 times lower (1.5 µg/L per 100 mg/L) than that in CuO. The rate constant of

399 organic pollutant degradation with CuFe2O4/PMS was 2.4 times higher than that with CuO/PMS.

400 In most studies, spinel metal oxide had better stability and activity in comparison with its

401 precursors (CuFe2O4 vs CuO and Fe2O3) [130,131]. The performance of CuFe2O4–Fe2O3 was

402 compared to other metal oxide catalysts. Their performances were in the following order:

403 CuFe2O4–Fe2O3 > CuFe2O4> CoFe2O4> CuBi2O4> CuAl2 O4 > Fe2O3 >MnFe2O4 [132]. Various

25
404 compositional CoxMn3-xO4 oxides were exerted for the activation of PMS for the decolorization

405 of RhB dye. The catalytic activity followed the order of CoMn2O4 > Co2MnO4 > Co0.5Mn2.5O4 >

406 Co3O4 > Co3O4 + Mn2O3 >Mn2O3. Applying leaching tests, it was revealed that the amounts of

407 Mn and Co ions were very low demonstrating that the activation of PMS had occurred at the

408 surface of the catalyst [133]. A mixed metal oxide spinel without Fe or Co was synthesized and

409 applied for degradation of pollutants in presence of PMS [134,135]. CuBi2O4 (CuB) was

410 deposited on different metal oxides. Catalytic activity of the catalysts was evaluated. The results

411 showed that CoO-CuB provided the most removal efficiency in comparison with FeO-CuB and

412 CuO-CuB. The order of the removal efficiency was CoO-CuB > Co3O4 >CuO-CuB > FeO-CuB

413 > Fe2O3 [135].

414 3.3 Iron-based heterogeneous catalyst

415 Iron-based catalysts have been extensively used for activation of hydrogen peroxide in Fenton-

416 based processes. Iron oxides (Magnetite (Fe3O4), maghemite (γ-Fe2O3), and hematite (α-Fe2O3))

417 and zero valent iron (ZVI) have been considered to activate peroxygens. Iron (Fe) is the fourth

418 most abundant element in the Earth’s crust which is also non-toxic and safe, thus the use of iron

419 for environmental application is justifiable and reasonable [136,137]. Moreover, among the

420 transitional metals, iron is a benign and low cost element and it is not a serious threat for human

421 health and the environment. Besides, magnetic properties of ZVI, Fe3O4 and γ-Fe2O3 received

422 much attention for activation of the oxidants from reuse point of view [138]. Various researchers

423 have utilized the nano-particles of iron along with some supports for activation of PMS. Most of

424 researchers have explained that iron can be a suitable alternative for cobalt because cobalt is a

425 toxic pollutant even in very low concentrations [136]. Magnetic carbon encapsulated nano iron

426 hybrids (nano Fe0/Fe3C@CS) [139] and nanoscaled zero valent iron (ZVI) encapsulated in

26
427 carbon spheres (nano-Fe0@CS) [140] were successfully used for PMS activation. These

428 composites were more effective compared to their constituents such as iron ions and oxides.

429 Fe2O3-montmorillonite exhibited high catalytic activity for PMS in which 85% of

430 dichlorophenol was mineralized during 2.5 h while its performance for H2O2 and acetic acid

431 were 70% and 50% within 3.5 h respectively. [141]. The porous Fe2O3 lonely revealed a stable

432 catalyst for PMS activation [142]. The dipicolinic acid-functionalized hematite has been applied

433 as PMS activator to remove BPA. This catalyst showed high efficiency which was comparable

434 with Co3O4 efficiency[143]. Fe@AFCs showed a high activity for decomposition of PMS in

435 comparison with H2O2. Reactive Red M-3BE (RR M-3BE) was effectively decolorized by

436 Fe@AFCs/PMS system. Fig. 7 presents the rate constants of Fe@AFCs/PMS and

437 Fe@AFCs/H2O2 and the efficiency of oxidant utilization (X) which the latter parameter is

438 defined as the number of moles of oxidant consumed per mole of the pollutant for 50%

439 degradation. As can be seen, the rate constant of Fe@AFCs/PMS was 0.3484 min-1 while this

440 value was 0.0500 min-1 for Fe@AFCs/H2O2 system. One can see that X value in Fe@AFCs/PMS

441 system was almost half of that in Fe@AFCs/H2O2 indicating more oxidation efficiency in case of

442 PMS [144].

443

27
444 Fig. 7 Comparison of apparent pseudo-first-order rate constants and oxidant consumption indices
445 for the Fe@ACFs/PMS and Fe@ACFs/H2O2 systems (Fe@ACFs: 2 g/L; [RR M-3BE]0: 50 mL;
446 [PMS]0: 0.5 mM; [H2O2]0: 0.5 mM; T = 50°C; pH 3.0) (Figure reprinted with permission from
447 [144]).
448
449 Bimetallic catalyst based on nano-X-Fe0 catalyst was utilized for activation of various

450 peroxygens in which X included Pd, Ag, Zn, Cu, Co, Cr, Mn, Cd and Ni. The presence of the X

451 increased removal efficiency in PMS/nano-X-Fe0. The order of best performance of catalysts

452 was: Co–Fe0 > Pd–Fe0 = Mn–Fe0 = Cd–Fe0 > Ni–Fe0 > Cr–Fe0 > Zn–Fe0 > Cu–Fe0 > Ag–Fe0

453 [145]. The magnetic Fe3O4/carbon supported manganese oxide nanoparticles (Mn/Fe3O4/carbon)

454 were considered for phenol degradation [146]. In this way, Fe3O4/Mn3O4/rGO [82] and

455 SiO2@Fe3O4 @MnO2 spheres [147] were also examined for PMS activation. Their catalytic

456 performance revealed that combination of the Fe3O4 and Mn3O4 with graphene or SiO2 could

457 provide higher catalytic activity than Fe3O4 and Mn3O4 or MnO2. An organometallic compound

458 (ferrocene) consisting of a Fe2+ atom was used for PMS activation. Cyclic Fe(II)/Fe(III) at center

459 of the catalyst was accountable for PMS activation in sulfate radical generation [148]. The

460 reducing agents of epigallocatechin-3-gallate [149] and NH2OH [150] were also used as an

461 accelerator of Fe(II)/Fe(III) cycle on surface of the iron oxides. These reducing agents

462 considerably enhanced PMS activation.

463 3.4. Other transitional metal

464 Besides cobalt and iron, copper and manganese oxides in nano-scale have been considered in

465 activation of PMS. Manganese oxides in nano size (MnO, Mn3O4, MnO2, Mn2O3) and their

466 composites with other materials are attractive for researchers in terms of stability, low cost, high

467 efficiency, multi-valence nature and being environmentally friendly [151-154]. Wang’s group

468 has focused on synthesizing manganese oxides in different oxidation state of Mn to activate PMS

469 [155-157]. They synthesized a series of manganese oxides (MnO, MnO2, Mn2O3 and Mn3O4) for

28
470 catalyzing PMS decomposition for phenol removal from aqueous solution. Mn oxides similar

471 with Mn+2 homogeneous systems, are able to generate SO4-• and SO5 -•.

472 HSO • 
  2MnO → Mn O  SO  OH 39

473 HSO •
  2Mn O → 3Mn O  SO  H 40

474 HSO •
  Mn O → 2MnO  SO  H 41

475 HSO •
  2MnO → Mn O  SO  H 42

476 According to the above reactions, MnO2 generates SO5-• which has less reactivity in comparison

477 with SO4-•. In fact, the reactivity of the Mn oxides depended on the capacity of manganese for

478 formation of different oxidation states (redox reaction of Mn2+/Mn3+ or Mn3+/Mn4+). Moreover,

479 “oxygen mobility” in the oxide lattice influences the reactivity of the catalyst. As shown in Fig.

480 8, Mn2O3 was the most effective oxide of the Mn oxides for PMS activation. The phenol

481 degradations were 100%, 90%, 66.4% and 61.5% for Mn2O3/PMS, MnO/PMS, Mn3O4/PMS and

482 MnO2/PMS respectively [158].

483

484 Fig. 8 Phenol removal efficiency in catalytic oxidation using a series of manganese oxides.
485 Reaction condition: [phenol] = 25 ppm, catalyst = 0.4 g/L, PMS = 2 g/L, and T = 25◦C (Figure
486 reprinted with permission from [158]).
487

29
488 Duan et al. synthesized manganese oxide octahedral molecular sieves (OMS-2) for degradation

489 of AO7 with PMS. This study showed that the responsible agent of decolorization was not

490 dissolved Mn2+ [159]. Wang et al. prepared three one-dimensional (1D) α-MnO2 nanostructures

491 (nanorods, nanotubes and nanowires) for activation of PMS. Nanowires provided the highest

492 performance which was attributed to higher surface area [160]. Different crystallographic phases

493 of MnO2 (α-, β-, and γ-MnO2) were also investigated. α-MnO2 showed the highest efficiency and

494 stability in comparison with other catalysts [161]. In an investigation, it was revealed that by

495 using α-MnO2/PMS, the obtained RhB and COD removals were 99% and 86%, respectively

496 [162]. Following the studies of Wang’s group, they also studied the effect of shape of α-Mn2O3

497 crystal on PMS activation. Cubic, truncated octahedra and octahedral shapes exhibited different

498 activities for phenol oxidation. The order of performance of the catalysts followed Mn2O3-cubic

499 > Mn2O3-octahedra > Mn2O3-truncated [163]. A sludge-hydrothermal carbonization/MnOx was

500 used for PMS activation. The performance of the catalyst was comparable with MnOx [164].

501 Cupric oxides (CuO and Cu2O) [165] and zero valent copper (ZVC, Cu0) [166,167] were

502 considered for activation of PMS. The analysis of CuO after and before reactions showed that the

503 proportions of Cu+ and Cu2+ were 25% and 75% in comparison with 100% proportion of Cu2+

504 before the reaction time[165]. The performance of Cu/ZSM5 was compared to CuO. It was

505 concluded that physical property played a key role in activation of PMS and the catalytic activity

506 of Cu/ZSM5 was much more than that of CuO regarding the surface areas of CuO and Cu/ZSM5

507 which were 1.52 and 354 m2/g respectively [168]. Catalytic activity of CuO was better than that

508 of Cu2O suggesting that the Cu2+-containing oxide was more effective for PMS activation [121].

509 The activity of CuFeO2 for PMS was evaluated. The presence of Cu(I) and Fe(III) showed a

510 synergistic effect which was probably due to reduction of Fe(III). The catalytic efficiency of

30
511 CuFeO2 was more than that of CuFe2O4 which was related to promoted reduction of solid Fe(III)

512 by the Cu(I) [169].

513 4. PMS activation by UV irradiation

514 Ultraviolent irradiation has been used for activation of various oxidants such as PS, H2O2,

515 percarbonate, ozone and chlorine to degrade a wide range of pollutants [170-172]. Recently, UV

516 has been successfully applied to decompose PMS for generation of sulfate and hydroxyl radicals.

517 Employing energy for PMS activation, the peroxide bond in PMS is the target of scission [30]. In

518 this approach, various forms of energy can be utilized such as heat, UV and ultrasound. Applying

519 UV radiation is benign and economic for the purpose of peroxide activation via destruction of

520 chemical bonds [173,174]. The energy of UV radiation varies between 300 kJ/Einstein in UV-A

521 radiation to 1200 kJ/Einstein in vacuum UV [175]. In case of PMS photo-activation, most

522 attempts have been focused on applying UV especially in wavelength of 254 nm. Actually, it has

523 been stated that using this wavelength of electromagnetic emission provides enough energy for

524 the scission of the peroxo band in PMS structure and PMS wavelength of energy absorbance is in

525 the UV range. UV/PMS process can degrade organic pollutants either directly by photolysis or

526 indirectly by sulfate and hydroxyl radicals [175,176].


/0
527 HSO 2 SO•
 1

 OH 43

528 SO•  •
 H O → H  SO  HO 44

/0
529 H O 12 H •  HO• 45

530 SO• • 
 OH  organic compounds → by − products  CO  H O  SO 46

531 Numerous studies have focused on activating PMS by UV; some of which are indicated here.

532 Sharma et al. investigated oxidative removal of bisphenol A (BPA) using UV-C/PMS .In their

533 work, applying 0.66 mM PMS brought about 96.7% BPA removal efficiency. Likewise, TOC

31
534 removal efficiency was 72.5% in 360 min reaction time [177]. Rao et al. [178] have stated that

535 Fe2+ could also be regenerated by applying UV based on the following equation:

[FeOH]  hv → Fe  HO• 47

536 Transitional metals in forms of ions and zero valent enhanced efficiency of UV/HSO5 - system

537 significantly through co-activation of PMS [179,180]. Verma et al. applied UV-C LED for the

538 activation of PMS. As they have stated, there are several advantages for applying LEDs as the

539 emission source such as supplying monochromatic emission at a specific wavelength and easy

540 handling regarding the compact design. Besides, mechanical stability is another advantage of

541 LED emission sources which involves less warm up time, lower replacement frequency and less

542 power requirement [181]. Some studies showed that visible light couldn’t activate PMS

543 [159,182]. The PMS activation by UV for the degradation of different contaminants is presented

544 in Table 5.

545 Table 5 Contaminants degradation using UV-activated PMS

Applied Removal
Contaminant Condition Highlight Reference
wavelength efficiency

Lindane= 17.15 µM, pH = 3, Fluorescence light was

Lindane 254 nm PMS= 500 µM, Fe2+ =500 92.2% (TOC) not effective in PMS [183]

µM and time= 180 min. decomposition

Diclofenac=0.63 mM,
∼80% (TOC) in 90
Diclofenac 254 nm Oxone=3.15 mM and T= 20 ∼100% (120 [184]
min
◦C min);

32
Presence of

AB14= 50 mg/L, PMS= 3 80.2% within 15 transitional metals

Acid brown 14 mM, UV light intensity=2.08 min; using 2 mM (Co2+, Fe2+ and Cu2+)
254 nm [173]
2 2+
(AB14) mW/cm , pH = 4, applied Fe increased to increased

current= 200 mA and 97% in 20 min. decolorization of

AB14 significantly.

[PMS]/[CIP] molar ratio =


Ciprofloxacin 254 nm Full degradation [185]
20, pH = 7 and time=60 min

BA degradation and

PMS decomposition
Benzoic acid BA=9.9 µM, PMS=100 µM,
254 nm ∼100% were significantly [186]
(BA) pH = 11.0 and time=10 min
enhanced at the pH

range of 8-11

Atenolol= 20 µm, PMS= 80

Atenolol 254 nm µm, pH = 7, T=23 °C and ∼88% [187]

time=30 min

Cylindrosperm 254 nm CYN=1 µM, UV fluence =40 100% 50 µg/L Fe2+ enhanced [179]

33
opsin mJ cm−2, PMS= 67 mg/L the process efficiency.

(CYN)

Sulfamonometh SMM= 5 mg/L, PMS=10


365 nm 96.78% [188]
oxine (SMM) mM, time=90 min

Dimethyl DMP =100 mg/L, PMS= 40 Electrical energy per

phthalate 254 nm mM, ∼95% order of removal=2.9 [174]

(DMP) pH, 3.0 and time=20 min kWh/m3.order

Atrazine= µM 4.64, PMS= Electrical energy per

92.80 µM, Fe2+ =8.95 µM, order of


Atrazine 253.7 nm 100% [5]
pH = 3, UV fluence= 480 removal=0.348

mJ/cm2 kWh/m3.order

2,4- 2,4-DCP= 100 mg/L PMS =1

dichlorophenol Solar mM, Co2+=4 µm and time=45 ∼100% [189]

(2,4-DCP) min

∼ 18% 2,4-DCP
2,4- 2,4-DCP= 0.307 mM,
300–400 in UV/PMS ∼ 95% in UV/PMS-
2+
dichlorophenol Co =4.6 mg/L, pH = 7 and [190]
nm system; 0.1Co/TiO2 system
(2,4-DCP) time= 1 h

34
Complete

Orange II=0.2 mM; Oxone= decolorization for Cu2+


Sunset
Orange II 20 mM, Fe2+ =0.6 mM and ∼100% and Mn2+ were [191]
lamp
time= 65 min obtained at 120 and

240 min respectively

Orange II =0.2 mM, Oxone=

Orange II Xe lamp 2.7 mM, Co2+= 0.1 mM and ∼100% [192]

time= 2 min

Addition of Fe2+ to

UV/PMS system

2,4,5- ∼84% 2,4,5-T in increased 2,4,5-T and


2, 4, 5-T =0.5 mM, PMS=
trichlorophenox 254 nm 180 min; ∼40% TOC removal [193]
2.5 mM, 2 UV lamps
yacetic acid (TOC) in 8 h efficiencies to almost

100% and 83%,

respectively.

Acid orange AO7=0.14 mM, PMS= 1.40


73% (TOC) [43]
(AO7) 365 nm mM and Co2+=14 µM

Anatoxin-a UV-C LED Anatoxin-a=1.5 µM, 98.6% With UV/PMS/Cu2+ [181]

35
(260 nm) [PMS]/[anatoxin-a] = 100:1 system, the

in ambient temperature and degradation of

pH = 6.4 anatoxin-a reached

more that 99% in 10

min of irradiation

time.

Cl- and CO32-

carbamazepine CBZ= 5 mg/L, PMS= 1 mM, increased CBZ


253.7 nm 76.2% [194]
(CBZ) pH = 11 and time=90 min degradation at

different degrees.

AB113=50 mg/L, Oxone


Acid Blue 113 82% TOC was
253.7 nm =4.2 mM, pH=5.8 and 100% [195]
(AB113) removed in 90 min
time=30 min

Triton X-45= 20 mg/L, TOC

=12 mg/L, PMS = 2.5 mM, 70% TOC removal


Triton X-45 253.7 nm 100% [196]
pH = 6.5, applied UV-C dose after 30 min

= 21 Wh/L and time =7 min

546

547 5. PMS Activation by ultrasound (US) and conduction electron

36
548 5.1. PMS Activation by US

549 Another way for activation of PMS is use of ultrasound (US) that has been reported by a few

550 researchers. Ultrasound with frequencies in the range 20–1000 kHz produces cavitation

551 phenomena that comprise the nucleation, growth, and collapse of bubbles in a liquid. These

552 collapsing bubbles have high temperature (5000 k) and high pressures (10 atm). Under this

553 condition, free radicals can be produced in solution. US irradiation has been ascertained that can

554 activate oxidants such as hydrogen peroxide, persulfate and recently, peroxymonosulfate [7,197-

555 200].

556 H O   → 2 HO• 48

557 S O    → 2SO•


49

558 HSO • •
   → SO  HO 50

559 Kurukutla et al. have demonstrated that PMS increases removal efficiency of sonolysis in RhB

560 degradation in comparison with persulfate and hydrogen peroxide [201]. Co2+/US/Oxone system

561 was conducted for amoxicillin degradation with COD parameter. COD removal efficiencies were

562 in the order of Oxone < Oxone /Co2+ < Oxone /US < Oxone /Co2+/US for the amoxicillin

563 solution [202,203]. Ultrasonic power increased the number of free radicals (SO4•- and HO•)

564 through increase in the number of active cavitation bubbles and fast activation of PMS [199].

565 Samarium doped ZnO (Sm/ZnO) was used for sonocatalytic degradation of phenazopyridine.

566 The degradation rates of phenazopyridine were 59% and 20% for Sm/ZnO/US and US/PMS

567 respectively. A combination of both systems (Sm/ZnO/PMS/US) could only degrade 69% of

568 phenazopyridine [204]. Although efficiency of US is considerable in activation of PMS, scale-up

569 of reactor and economic aspects are the main problems of sonochemical processes that reduce

570 use of US in large scale.

571 5.2 Conduction electron (e-) application in photocatalytic activation of PMS

37
572 PMS as an oxidant can be applied as electron acceptor in photocatalysis process. In this way,

573 both sulfate and hydroxyl radicals may be produced based on Eq. (52) [205,206]. The

574 peroxymonosulfate radical (SO5•-) is produced when the generated hole (h+) reacts with PMS

575 (Eq. (53)) [42]. Meanwhile, SO5•- can react with itself and produce sulfate radical based on Eq.

576 (54) [207].



Semiconductor  ℎD → %EF  ℎ 51

HSO  • • 
  eGH → OH  SO or HO  SO 52

HSO •
  ℎ → H  SO 53

SO• • •
  SO → 2SO  O 54

577 The conduction electron decomposes PMS and produces free radicals increasing degradation of

578 organic pollutant [205,208]. Fig. 9 represents the mechanism of photocatalysis in presence of

579 PMS as electron acceptor.

580

581 Fig. 9 the mechanism of PMS activation by conduction electron

582 Anderson et al. have applied PMS in NF-TiO2 photocatalysis system for degradation of atrazine

583 and amitrole. PMS compared to PS, had higher performance in photocatalysis process in atrazine

584 degradation especially in low concentration [209]. Anandan et al. investigated the presence of

585 various oxidants in photocatalysis process (Ag-TiO2). The presence of PMS in photocatalysis

38
586 process (Ag-TiO2) increased the degradation rate by seven-fold whereas presence of PS

587 increased the degradation rate by three-fold. H2O2 could not influence in enhancement of

588 degradation rate [210]. The performance of oxidants (PMS, PS and H2O2) in reduced graphene

589 oxide/TiO2-vis systems has been studied for decolorization of MB. The findings of oxidant effect

590 were in order of PS < PMS < H2O2 [211]. In PMS/ZnO/UV photocatalysis system, it was

591 exhibited that ZnO could activate PMS in absence of UV based on Eq. (55) [212].

ZnO  HSO
 → Zn

 HO•  SO•
 55

592 The degradation efficiency of nano Au-TiO2/UV system was 9-fold more than that of TiO2/UV

593 system. This behavior was related to the transfer of enhanced interfacial charge at the interface of

594 Au–TiO2 [213]. In Co-TiO2/UV/PMS system, advanced tests indicates that Co3O4 located on the

595 surface of the catalyst is the active site for PMS decomposition, and cobalt ion implanted in the

596 TiO2 lattice is a strong evidence for photocatalytic activity of Co–TiO2 in the visible light [214].

597 BiVO4/visible light was also effective for PMS activation. PMS decomposition in this

598 photocatalytic degradation resulted in formation of SO4•-, HO•, O2•- which were the main reactive

599 radicals [215]. The order of the degradation efficiencies was PMS > PS > H2O2 for

600 TiO2/UV/oxidant in AO7 degradation. The probable reason of this phenomenon was attributed to

601 more easily activated PMS compared to other two oxidants since PMS molecule is asymmetric,

602 with only one H replaced by SO3 [216]. ZnFe2O4 showed a dual role for degradation of the dye

603 since it acted as an activator in ZnFe2O4/PMS and a photocatalyst in ZnFe2O4/PMS/Vis [217]. In

604 this way, CoFe2O4/ZnO(CFZ)/UV was more effective in presence of PMS in comparison with

605 other oxidants. [218]. Recently, PMS has also been activated in the non-metal photocatalysis

606 process. The graphene-like carbon nitride (GCN) [219], polyimides [220], carbon nitride

607 materials [221] and orthorhombic α-sulfur [222] were excited by visible light irradiation to

608 activate PMS by conduction electron.


39
609 6. Metal-free heterogeneous catalysts

610 Carbon-based material (C) as a catalyst was also used for activation of PMS to generate free

611 radicals. These materials include activated carbon, graphene, nanodiamond, carbon nanotubes

612 and graphite exhibiting a catalytic behavior for activation of the oxidants [223]. As

613 environmentally friendly catalysts, these materials minimize the use of the metals and they can

614 be considered as promising alternatives to toxic and expensive metals [219,224].

615 HSO  •
  C → HO  SO  AC 56

616 HSO • 
  C → HO  SO  AC 57

HSO •
  C → H  SO  AC 58

617 Powder activated carbon (PAC) [225], granular activated carbon (GAC) [226] and activated

618 carbon fiber (ACF) have been used for PMS activation. The results showed that dye degradation

619 rate in the ACF/PMS system was much higher than that in the GAC/PMS. However, ACF had

620 an adsorption potential for pollutants which could justify increase of removal efficiency.

621 Moreover, the results of PMS decay demonstrated that PMS was clearly decomposed by AFC

622 whereas GAC could not affect the decomposition of PMS [227]. Ghanbari et al. simultaneously

623 applied AC and UV for activation of PMS. They stated that the contribution of UV was more

624 than that of AC in decomposition of PMS [228]. Nanocarbon structures have been used in

625 different dimensions. 1D single walled carbon nanotubes (SWCNTs), 2D graphene nanoplate

626 (GNP), 3D hexagonally ordered mesoporous carbon (CMK-3) and cubically-ordered mesoporous

627 carbon (CMK-8) were applied for catalytic oxidation of phenol with PMS. Amongst these

628 nanocarbon structures, CMK-3, CMK-8 and SWCNTs were efficient catalysts for activation of

629 PMS [229]. The modification of carbon catalyst surface can enhance activity of the catalyst

630 significantly since change in the catalyst surface directly influences adsorption capacity and

631 catalytic activity. AC modification by ammonia increased PMS activation through production of

40
632 pyrrolic group promoting electron transfer for PMS decomposition [230]. Nitrogen doping on

633 carbon catalyst boosted catalytic activity of the catalysts. In this way, N-doped graphene [231]

634 and nitrogen and sulfur co-doped CNT-COOH [232] acted as high efficient catalysts for PMS

635 decomposition. The combination of nitration and heat for modification of ACF increased

636 catalytic activity. Based on a qualitative analysis, the functional groups of pyridine,

637 imine/amine/amide and ‒NH2 have a promotional effect for PMS activation while ‒NO2 group

638 has an inhibitory function [233-235].

639 Although graphitic-carbon nitride (g-C3N4) has high nitrogen content (around 70% wt), the sole

640 use of g-C3N4 for PMS activation was unsuccessful while supported g-C3N4 on activated carbon

641 [236] and MCM-41 [237] exhibited a synergistic effect between support and g-C3N4 for PMS

642 activation which was related to O species in g-C3N4 as major active site in process of PMS

643 activation. Recently, some interesting studies have shown that the degradation mechanism by

644 nanodiamond, rGO [238] and N-doped single-walled carbon nanotube [239] is the non-radical

645 pathway. The results also exhibited that non-radical mechanism of these catalysts was based on

646 direct electron extracting of the organic pollutant.

647 7. Other methods for activation of PMS

648 Electrolysis (EC) has also been utilized to activate PMS for degradation of pollutants. In this

649 way, Lin and colleagues have used dimensionally stable anode (DSA) in EC/PMS/Fe3+ for

650 degradation of clofibric acid. They have explained that sulfate radical can be produced by

651 electron transfer reaction that can decompose 31% of PMS [240].

HSO   •
  % → OH  SO 59

652 This process is similar with Fered-Fenton process in which ferrous ion and H2O2 are chemically

653 introduced to electrochemical cell with high over-voltage anode [241]. It is worthwhile to

654 mention that ferrous ion is electrochemically produced for to activation of PMS in
41
655 electrocoagulation reactor [242]. However, the mechanism of PMS activation and mechanism of

656 pollutant removal are different with EC/PMS/Fe3+. Lou et al. have activated PMS by phosphate

657 buffer solution. They stated that this method could be a new way for activation of PMS

658 especially for phosphate-rich wastewater remediation [243]. Alkaline condition was also

659 considered for PMS activation in which sulfate radical [244], superoxide anion radical and

660 singlet oxygen [245] were dominant oxidizing agents. Unlike PS, activation of PMS by heat has

661 been rarely considered for degradation of pollutant. Among energy sources that are used for

662 activation of PMS, heat can reduce reaction time in comparison with other activators. PMS/heat

663 resulted in the highest degradation efficiency of 77% compared to 52% for PS and 2.5% for

664 H2O2 [246]. In another study, the order of decolorization efficiencies by heat activation was PS ≥

665 PMS > H2O2 [28]. PMS could be activated by benzoquinone (BQ) as an organic compound.

666 Oxidizing agent in the system was singlet oxygen instead of sulfate or hydroxyl radical [247].

667 Furthermore, degradation of dye has been considered by PMS/Cl- without addition of any

668 transitional metal. It should be noted that in this system, oxidizing agents were reactive chlorine

669 species. Indeed, PMS activated chloride ion to generate active chlorine for degradation of

670 organic compounds [248]. Nano-zero valent aluminum (nZVAl) can activate PMS through the

671 direct electron transfer from the nZVAl surface to PMS under acidic pH conditions according to

672 Eq. (60) [249].

Al$  3HSO
  3H → Al

 3SO•
 3H O tentative 60

673 PMS also produces sulfate radical in presence of ozone. First, SO52- reacts with O3 and SO82- is

674 produced as shown in Eq. (61). This product is decomposed in two parallel pathways based on

675 Eqs. (62) and (63). The reaction of SO5•- and O3 produces sulfate radical (Eq. (64)). On the other

676 hand, O3•- can generate hydroxyl radical according to Eqs. (65) and (66). In fact, PMS/O3 system

677 produces sulfate radical and hydroxyl radical simultaneously [250].


42
678 SO  
  O → SO 61

SO  → SO• •
  O 62

SO  → SO 
 2O 63

SO• •
  O → SO  2O 64

O•
 → O
•
 O 65

O•  H O → HO•  HO 66

679 8. Identification of sulfate and hydroxyl radicals

680 8.1 Electron paramagnetic resonance (EPR)

681 In order to identify the free radicals or unpaired electron spins, magnetic resonance spectroscopy

682 has widely been used in case of PMS decomposition to hydroxyl and sulfate radicals. Electron

683 paramagnetic resonance (EPR) or electron spin resonance (ESR) spectroscopy was utilized for

684 mechanism determination of the radicals. In this way, a spin trap (a nitrone or nitroso compound)

685 is employed for detection of free radicals in which the life of the radicals is too short. The most

686 common spin trap is 5,5-dimethyl-pyrroline N-oxide (DMPO) reacting with hydroxyl and sulfate

687 radicals resulting in the formation of a stable and distinguishable spin adduct which can be

688 detected by EPR technique. Wang’s group used EPR technique to identify the sulfate and

689 hydroxyl radicals in PMS/CMK-8 and PMS/SWCNT during 20 min reaction time. Both radicals

690 had been detected during degradation of phenol. With increase in reaction time amount of peaks

691 corresponding to DMPO-HO• and DMPO-SO4•- were significantly reduced especially in case of

692 PMS/SWCNT. These results confirmed that both hydroxyl and sulfate radicals contributed in

693 degradation of phenol [229]. In another study, it has been reported that PMS in absence of

694 pollutant (phenol) produces a strong EPR signal which can be attributed to 5,5-

695 dimethylpyrroline-(2)-oxyl-(1) (DMPOX). In fact, hydrolysis of PMS produces the hydrogen

696 peroxide as a by-product which in turn, generates hydroxyl radical [160].

43
HSO 
  H O → H O  HSO 67

H O → HO•  HO• 68

697 In presences of catalyst (α-MnO2) and phenol, both hydroxyl and sulfate radicals were detected

698 via the peaks of DMPO-HO• and DMPO-SO4•- adducts, respectively. This study illustrates that at

699 one minute reaction time, hydroxyl radical was only generated at the solution since DMPO-HO•

700 adduct (with hyperfine splitting constants of αN = αH = 14.9 G) was detected. With increasing the

701 reaction time to 10 min, the signal of DMPO-SO4•- adducts (with hyperfine splitting

702 constants of αN = 13.2 G, αN = 9.6 G, αN = 1.48 G and αH = 0.78 G) appeared along with the

703 signals of DMPO-HO• implying that both radicals were generated during reaction time [160].

704 Huang et al. believed that PMS firstly generated hydroxyl radicals and then their values

705 decreased since some of them reacted with PMS to generate sulfate radical [251].

2HSO • •
  2HO → 2SO  2H O  O 69

706 EPR was also used for determination of the scavenging effects of excessive PMS and catalyst

707 dosage (Fe@ACFs) [144]. In Co@ACFs/PMS [79] and cellulosic fibers-bonded cobalt

708 phthalocyanine (CFs-CoPc)/PMS [252] systems, EPR spectra showed that hydroxyl radical is the

709 main radical in these processes. A study showed that in degradation of BPA by Co2+/PMS/UV,

710 presence of UV increased the peak intensity of DMPO-HO• in comparison with Co2+/PMS

711 system which is attributed to side Eq. (70). They speculated that the SO4•− was more unstable in

712 the Co2+/PMS system which it preferred to be transformed into •OH at any pH [253].
KL
SO• • 
 H O 12 HO  SO  H 70

713 8.2 Scavenging tests

714 In case of PMS activation, hydroxyl radical and sulfate radical are generated in the solution.

715 Peroxymonosulphate radical (SO5•−) is also generated but its activity is negligible in comparison

716 with hydroxyl and sulfate radicals. Using EPR, radical species are identified whereas their
44
717 activities throughout the oxidation reactions are quite ambiguous [30]. In order to distinguish

718 whether the hydroxyl radical or sulfate radical is the dominant radical species, common types of

719 alcohols such as ethanol (EtOH), methanol (MeOH) and tert-butyl alcohol (TBA) are added to

720 the solution. This approach is categorized as a chemical probe method in which scientists take

721 the advantage of difference between reactions rates of the two mentioned categories of alcohols

722 (alpha hydrogen containing alcohols and alpha hydrogen-free alcohols) with HO• and SO4 -•

723 (Table 6 represents the second-order rate constants of reactions related to some commonly used

724 radical scavengers with hydroxyl and sulfate radicals). In this way, while studying the

725 degradation of organic compounds, alcohol competes with the organic compound to react with

726 the radical species; so that the degradation rate is expected to be decreased. EtOH which is an

727 alpha hydrogen containing alcohols is capable of scavenging both sulfate radical (1.6107 -

728 7.7107 M-1 s-1) and hydroxyl radical (1.2109 - 2.8109 M-1 s-1) with similarly high reaction

729 rates while TBA does not contain α hydrogen and has a much higher affinity to scavenge

730 hydroxyl radical (3.8108 - 7.6108 M-1 s-1) rather than sulfate radical (4.0105 - 9.1105 M-1

731 s-1) [41,103,106,254]. Indeed, rate constant of TBA reaction with hydroxyl radical is

732 approximately 1000 times greater than that observed with the sulfate radical [245]. Accordingly,

733 TBA is usually considered as the hydroxyl radical scavenger; while EtOH scavenges both

734 hydroxyl radical and sulfate radical with quite similar rate constants. Therefore, the difference

735 between degradation efficiencies in presence of either MeOH or EtOH and TBA represents the

736 sulfate radical function.

737 Table 6 Quenching agents for sulfate and hydroxyl radicals and related second order rate

738 constants (Table reprinted with permission from [255]).

Quenching agent Chemical Reaction rate constant (M-1s-1)

45
Formula Sulfate radical Hydroxyl radical

Phenol C6H6O 8.8×109 6.6×109

Anisole C7H8O 4×109 7.8×109

Benzoic acid C7H6O2 1.2×109 4.2×109

Methanol CH3OH 3.2×106 9.7×108

Nitrobenzene C6H5NO2 <106 (3-3.9)×109

Ethanol C2H6O (1.6-7.7) × 107 (1.2-2.8) × 109

Propanol C3H8O 6 ×107 2.8 × 109

Tert-butyl alcohol C4H10O (4-9.1) ×105 (3.8-7.6) ×108

739

740 In different studies conducted for degradation of pollutants, addition of ethanol revealed

741 significant scavenging effect while addition of TBA did not lead to a significant quenching

742 effect. This clearly implies that in the mentioned work, the dominant and responsible radical

743 species has been the sulfate radical [11,106,113,159]. In another study, authors elucidated the

744 contribution of reactive oxygen species by conducting the scavenging test in a base/PMS system.

745 It was observed that addition of neither TBA nor ethanol suppressed AO7 degradation indicating

746 that HO• and SO4-• have not been the reactive oxygen species responsible for AO7 degradation.

747 Nevertheless, it was found that addition of sodium azide could dramatically inhibit AO7

748 degradation. It is figured out that sodium azide acts as a unique scavenger for singlet oxygen

749 (rate constant = 2109 M-1 s-1). Nonetheless, singlet oxygen could not be considered as the only

750 reactive species since addition of sodium azide did not completely suppress the AO7

751 degradation. p-benzoquinone as a scavenger of superoxide anion radical (rate constant = 0.9-

752 1109 M-1 s-1), degradation of AO7 was completely inhibited endorsing the critical function of

46
753 the superoxide anion radical [245]. Fig. 10 displays the results of the quenching experiment in

754 Co0.75Fe2.25O4/PMS system for decolorization of RhB. As can be seen, in presence of phenol,

755 removal efficiency is significantly reduced compared to the control test (no quenching agent)

756 since phenol is a scavenger for both hydroxyl and sulfate radicals with high rate constant. On the

757 other hand, removal efficiency is slightly decreased in presence of TBA indicating that the

758 sulfate radical has been the predominant radical species for degradation of RhB in

759 Co0.75Fe2.25O4/PMS system [120].

760

761 Fig. 10. Effect of different radicals scavengers on the degradation of RhB at ambient
762 temperature. Experimental conditions: [RhB] = 0.014 mmol/L, [Oxone] = 1.0 mmol/L,
763 [Co0.75Fe2.25O4] = 0.05 g/L, pH=11.0. (Figure reprinted with permission from [120])
764
765 If addition of scavengers even EtOH does not lead to a significant decrease in degradation rate,

766 three assumptions can be brought 1. EtOH concentration has not been adequate to significantly

767 quench the radical species; 2. There might have been some non-radical pathways contributing in

768 the degradation of organic compounds; 3. The reaction rate of the radical species with organic

769 pollutant has been more than that with EtOH. Nevertheless, higher degradation inhibition as a

770 result of EtOH addition indicated that SO4-• has been more dominant than HO• [181].

771 Throughout literature, within most of the works carried out using PMS, sulfate radical has been

772 indicated as the major radical species diagnosed by alcohol-based scavenging tests. In other

47
773 words, conducting scavenging tests, hydroxyl radical has rarely been identified as the dominant

774 radical species in PMS-based systems [79]. Based on what has been stated in the relevant

775 studies, acidic pH levels support the domination of sulfate radical while on the other hand,

776 alkaline pH levels shift the reactions towards more production of hydroxyl radical which is

777 supposed to be due to the significant presence of hydroxide ions in alkaline conditions

778 [205,256,257].

779 9. PMS application for real wastewater treatment

780 Primary investigations have often focused on synthetic wastewaters. Although the study on real

781 wastewater treatment has always been difficult, it is invaluable since it represents the actual

782 conditions. The presence of several substances in real wastewater reduces efficiency of AOPs

783 because some anions and reductive agents scavenge SO4•- and HO•. Besides, they may interact

784 with the activation e.g. absorbance of UV in UV/PMS system. PMS-based processes have been

785 rarely used for real wastewaters. Landfill leachate has been treated by microwave/PMS [258],

786 wet air oxidation (WAO)/PMS (as a promoter) [259] and PMS/Co2+ [56]. PMS exhibited a

787 superiority compared to other oxidants (PS and H2O2). Compared to Fenton process with weak

788 efficiency in suspended solids (SS) and color removal, PMS/Co2+ provided 53.3% SS removal

789 and 83% color removal. Moreover, there was no sludge in PMS/Co2+ while Fenton oxidation

790 suffered from a huge sludge. The operation in neutral condition was another advantage of

791 PMS/Co2+ compared to Fenton process [56]. Fagier et al. exerted a combination of coagulation

792 and sulfate radical based processes for treatment of vinasse as high strength wastewater with

793 COD of 126 g/L and TOC of 48 g/L. Under condition of PMS:Fe2+ ratio of 1:2.5, 0.18 mM PMS

794 and pH=7, TOC removal was 33%. With increase of PMS dosage to 0.72 mM, TOC removal

795 was remarkably enhanced and reached to 69.9%. This study exhibited high efficiency of

48
796 PMS/Fe2+ system with low PMS dosage at neutral pH which can be considerable for industrial

797 scale [260]. Rodríguez-Chueca et al. used PMS to treat winery wastewater. The effects of

798 different transitional metals (Co2+, Fe2+, Cu2+, Ni2+, Ag+, Mg2+, Mn2+ and Zn2+) on PMS

799 activation were investigated and their results are presented in Fig. 11. Amongst the transitional

800 metals, Cu(II), Fe(II) and Co(II) exhibited strong activity in PMS activation. In presence of UV

801 radiation, higher COD removal was obtained. In addition, UV-A LEDs was the best source for

802 enhancement of the efficiency of PMS/transitional metal. To compare with the photo-Fenton

803 process, PMS/Co/UV system had two special superiorities including no production of sludge and

804 the operation at natural pH of the wastewater [261].

805

806 Fig. 11 Influence of different transitional metals in the removal of COD. Experimental
807 conditions: 2.5 mM PMS; 1 mM Mn+; pH 6.5; T = 323 K; 90 minutes (Figure reprinted with
808 permission from [261])
809
810 Real textile wastewater has been remediated by PMS/zero valent metal [167] and

811 sonophotocatalytic process along with Fe-doped Bi2O3 and PMS [262]. The average oxidation

812 state (AOS) of real textile wastewater was increased by PMS/ZVI and PMS/ZVC from -0.68 to

813 +1.7 and +0.8 respectively. These results indicated that PMS/zero valent metal system could be

49
814 an effective pretreatment for biological processes [167]. UVC/PMS following electrocoagulation

815 [182] and UVC/PMS/Co3O4 following permanganate and electro-Fenton [263] were applied to

816 treat pulp and paper wastewater. After PMS-based processes, biodegradability was improved in

817 terms of BOD5/COD and it reached > 0.4. Furthermore, it was observed that total dissolved

818 solids of the final effluent were significantly increased after PMS application.

819 10. Application of PMS for polluted air, soil and sludge

820 There has been much effort towards application of sulfate radical for oxidation of organic

821 pollutants in aqueous matrix while other matrixes e.g. air, soil and sludge have fewer been

822 considered specially in case of PMS application. Here, these few studies are figured out.

823 The application of PMS for the elimination of methyl mercaptan induced odor in a wet scrubber

824 pilot plant was studied. The authors explained one of the major removal mechanisms of methyl

825 mercaptan through mass transfer acceleration which had led to transfer of methyl mercaptan into

826 the aqueous solution. It was expressed that methyl mercaptan removal was high until PMS

827 depletion occurred. In alkaline condition, CH3S- was absorbed to the scrubbing solution which

828 was efficiently oxidized in presence of PMS. Therefore, CH3SH was not accumulated throughout

829 the scrubbing. One of the concerns in this study was partial decomposition of PMS which was

830 overcome through application of Na2SiO3 in the scrubbing solution. [264]. Inorganic gases of NO

831 and SO2 were also remediated in an impinging stream reactor by co-activation of PMS in

832 presence of Cu2+/Fe3+ and heat. SO2 was rapidly removed under most operating conditions

833 compared to NO. Sulfuric acid and nitric acid were the main by-products of this system.

834 Synergistic effect of transitional metals and heat increased generating sulfate and hydroxyl

835 radicals for oxidation of NO [265]. PMS/Co(II) system was evaluated for the treatment of diesel

836 fuel contaminated soil. In this way, Co(II) and Fe(II) were applied with various concentrations.

50
837 Highest degradation was obtained using CoCl2 as PMS activator. Besides, acidic conditions were

838 superior for the degradation of diesel in soil which was attributed to enhanced dissolution of

839 metal oxides within lower ranges of pH [266]. Degradation of 2-chlorobiphenyl in aqueous and

840 sediment matrixes was investigated applying Fe(II)/Fe(III) activated PMS. Regarding slower

841 reduction of ferric ion by peroxide, degradation rate was lower in Fe(III)/PMS system compared

842 to Fe(II)/PMS system. [47]. In a study carried out by Niu et al., disintegration of sludge was

843 evaluated by addition of PMS. Based on the obtained results, sludge particles were efficiently

844 broken leading to satisfactory disintegration of sludge. It was reported that extracellular

845 polymeric substances (EPS) was dramatically increased as a consequence of PMS addition.

846 Besides, the amount of polysaccharides in EPS increased with increase of PMS dosage. It was

847 also stated that humic-like fluorophores in slime EPS had a direct correlation with dosage of

848 PMS [267]. Toluene removal from anaerobically digested sludge was assessed using PMS and

849 PS. Application of each of the two oxidants in their highest dosages (3 g/L) and activated in the

850 highest applied temperature (55ºC) resulted in dramatic improvement of sludge dewaterability

851 [268]. Oxidation of soil washing wastewater was examined by applying two solar assisted AOPs

852 including Co/PMS/UV and photo-Fenton. Applying Co/PMS/UV, sodium dodecyl sulphate

853 (SDS) surfactant which had the responsibility of improving soil washing was removed

854 completely. It is really worthwhile to consider that the energy consumption required for complete

855 degradation of SDS in photo-Fenton process was about four times higher than that in

856 Co/PMS/UV. In this work, 51% COD removal was achieved using Co/PMS/UV which was

857 somehow less than that achieved using photo-Fenton (69%) [269].

858 11. Conclusion and prospective

51
859 The popularity of sulfate radical amongst scientists for the purpose of pollutants degradation

860 persuaded us to go for a comprehensive review on application of PMS as a promising source of

861 sulfate radical. Following the current investigations, AOPs are shifting towards use of sulfate

862 radical due to its unique characteristics some of which are high oxidation potential, applicability

863 in wide pH range, more stability than its opponent radical species (•OH) and being more

864 selective. To compare with other oxidants such as H2O2 and S2O82-, PMS is an effective and

865 promising oxidant for producing hydroxyl and sulfate radicals. PMS-based advanced oxidation

866 processes can open a new perspective of oxidative reactions to degrade the organic pollutants.

867 Activation methods are comprehensively discussed amongst which heterogeneous catalysts have

868 attracted much attention in case of PMS catalyzing since no metals would be released into the

869 aqueous solution. Moreover, heterogeneous catalysts can be used in several cycles especially in

870 case of magnetically heterogeneous catalysts. The understanding of behaviors of the supporters

871 and transitional metals in heterogeneous activation of PMS is necessity for which further efforts

872 and studies are required. Discovering novel methods for activating PMS have been recently

873 proliferated indicating that scientists and environmentalists are hopeful to PMS-based processes

874 as new tools for remediation of contaminated environments. However, the selectivity of

875 generating sulfate radical or hydroxyl radical in heterogeneous catalysts is a dilemma since the

876 performance of active sites for generation of free radicals is ambiguous. It is worthwhile to state

877 that coupling of PMS with other AOPs for synergistic effect needs more surveys for

878 identification of mechanisms and determination of predominant free radicals. The presence of

879 various metals in real industrial wastewater can justify application of PMS in industrial scales for

880 activation of PMS. Chemical sludge driven from the coagulation process in water treatment plant

881 is rich of iron which can be considered for activation of PMS. Nevertheless, inherent problem of

52
882 PMS is still remained which is related to sulfate ion release as the main by-product. This issue is

883 exacerbated by two other salts (K2SO4, KHSO4) which are applied along with PMS in Oxone salt

884 structure. Hence, the synthesis of a novel PMS salt without the two other salts is a necessity and

885 is a great favor. Here the solution might be the application of a subsequent process (ion-

886 exchange, membrane, etc.) after PMS-based processes with high PMS dosages for sulfate

887 removal. In addition, optimization of PMS dosage is a suitable strategy to avoid excessive PMS

888 thereby reducing the amounts of sulfate within the effluent. It should also be considered that in

889 comparison with common oxidants such as PS and H2O2, Oxone is produced in fewer amounts

890 and does have higher cost.

891 Finally, it should be stated that PMS-based processes are neoteric processes which still have a lot

892 to go through. Actually, their application in industrial scale requires many studies to stabilize

893 their position among other oxidative processes. Howbeit, due to desirable performance of PMS

894 in degrading the pollutants, it is not an exaggeration to find PMS as an alternative oxidant in

895 industrial scale.

896 Acknowledgment

897 No fund has been allocated to the current work. The authors would like to thank the anonymous

898 reviewers of Chemical Engineering Journal for invaluable comments. We also appreciate Prof.

899 D. Dionysiou (Co-Editor, Chemical Engineering Journal) for his consideration to our manuscript.

900 References

901 [1] P.V. Nidheesh, R. Gandhimathi, S.T. Ramesh, Degradation of dyes from aqueous solution by Fenton
902 processes: a review, Environ. Sci. Pollut. Res. 20 (2013) 2099-2132.
903 [2] W.H. Glaze, J.-W. Kang, D.H. Chapin, The Chemistry of Water Treatment Processes Involving
904 Ozone, Hydrogen Peroxide and Ultraviolet Radiation, Ozone Sci. Eng. 9 (1987) 335-352.
905 [3] J.L. Wang, L.J. Xu, Advanced Oxidation Processes for Wastewater Treatment: Formation of
906 Hydroxyl Radical and Application, Crit. Rev. Environ. Sci. Technol. 42 (2012) 251-325.

53
907 [4] V. Romero, S. Acevedo, P. Marco, J. Giménez, S. Esplugas, Enhancement of Fenton and photo-
908 Fenton processes at initial circumneutral pH for the degradation of the β-blocker metoprolol, Water Res.
909 88 (2016) 449-457.
910 [5] J.A. Khan, X. He, H.M. Khan, N.S. Shah, D.D. Dionysiou, Oxidative degradation of atrazine in
911 aqueous solution by UV/H2O2/Fe2+, UV//Fe2+ and UV//Fe2+ processes: A comparative study, Chem. Eng.
912 J. 218 (2013) 376-383.
913 [6] M. Ahmadi, F. Ghanbari, S. Madihi-Bidgoli, Photoperoxi-coagulation using activated carbon fiber
914 cathode as an efficient method for benzotriazole removal from aqueous solutions: Modeling, optimization
915 and mechanism, J. Photochem. Photobiol. A: Chem. 322–323 (2016) 85-94.
916 [7] N.N. Mahamuni, Y.G. Adewuyi, Advanced oxidation processes (AOPs) involving ultrasound for
917 waste water treatment: A review with emphasis on cost estimation, Ultrason. Sonochem. 17 (2010) 990-
918 1003.
919 [8] F. Ghanbari, M. Moradi, A comparative study of electrocoagulation, electrochemical Fenton, electro-
920 Fenton and peroxi-coagulation for decolorization of real textile wastewater: Electrical energy
921 consumption and biodegradability improvement, J. Environ. Chem. Eng. 3 (2015) 499-506.
922 [9] A. Babuponnusami, K. Muthukumar, A review on Fenton and improvements to the Fenton process for
923 wastewater treatment, J. Environ. Chem. Eng. 2 (2014) 557-572.
924 [10] G.P. Anipsitakis (2005). Cobalt/Peroxymonosulfate and Realated Oxidizing Reagents for Water
925 Treatment (Ph.D. thesis). University of Cincinnati, United States.
926 [11] G.P. Anipsitakis, D.D. Dionysiou, Degradation of Organic Contaminants in Water with Sulfate
927 Radicals Generated by the Conjunction of Peroxymonosulfate with Cobalt, Environ. Sci. Technol. 37
928 (2003) 4790-4797.
929 [12] M. Mirza-Aghayan, M. Molaee Tavana, R. Boukherroub, Direct oxidative synthesis of nitrones from
930 aldehydes and primary anilines using graphite oxide and Oxone, Tetrahedron Lett. 55 (2014) 5471-5474.
931 [13] G.P. Anipsitakis, T.P. Tufano, D.D. Dionysiou, Chemical and microbial decontamination of pool
932 water using activated potassium peroxymonosulfate, Water Res. 42 (2008) 2899-2910.
933 [14] J. Jang, M. Jang, W. Mui, C.A. Delcomyn, M.V. Henley, J.D. Hearn, Formation of Active Chlorine
934 Oxidants in Saline-Oxone Aerosol, Aerosol Sci. Technol. 44 (2010) 1018-1026.
935 [15] R. R. Solís, F.J. Rivas, O. Gimeno, Removal of aqueous metazachlor, tembotrione, tritosulfuron and
936 ethofumesate by heterogeneous monopersulfate decomposition on lanthanum-cobalt perovskites, Appl.
937 Catal., B 200 (2017) 83-92.
938 [16] L.A.C. Teixeira, J.P.M. Andia, L. Yokoyama, F.V. da Fonseca Araújo, C.M. Sarmiento, Oxidation
939 of cyanide in effluents by Caro’s Acid, Miner. Eng. 45 (2013) 81-87.
940 [17] H. Sun, S. Wang, Catalytic oxidation of organic pollutants in aqueous solution using sulfate radicals,
941 RSC Catal. 27 (2015) 209-247.
942 [18] S. Wacławek, K. Grübel, M. Černík, Simple spectrophotometric determination of monopersulfate,
943 Spectrochim. Acta, Pt. A: Mol. Biomol. Spectrosc. 149 (2015) 928-933.
944 [19] H. Hussain, I.R. Green, I. Ahmed, Journey Describing Applications of Oxone in Synthetic
945 Chemistry, Chem. Rev. 113 (2013) 3329-3371.
946 [20] B. Meunier, Metalloporphyrins as versatile catalysts for oxidation reactions and oxidative DNA
947 cleavage, Chem. Rev. 92 (1992) 1411-1456.
948 [21] E. Saputra (2013). Catalytic oxidation of toxic organics in aqueous solution for wastewater treatment
949 (Ph.D. Thesis). Curtin University, Australia.

54
950 [22] A.R. Chesney, C.J. Booth, C.B. Lietz, L. Li, J.A. Pedersen, Peroxymonosulfate Rapidly Inactivates
951 the Disease-Associated Prion Protein, Environ. Sci. Technol. 50 (2016) 7095-7105.
952 [23] M. Yu, A.L. Teel, R.J. Watts, Activation of Peroxymonosulfate by Subsurface Minerals, J. Contam.
953 Hydrol. 191 (2016) 33-43.
954 [24] J. Flanagan, W.P. Griffith, A.C. Skapski, The active principle of Caro's acid, HSO5-: X-ray crystal
955 structure of KHSO5.H2O, J. Chem. Soc., Chem. Commun. (1984) 1574-1575.
956 [25] X. Pang, Y. Guo, Y. Zhang, B. Xu, F. Qi, LaCoO3 perovskite oxide activation of peroxymonosulfate
957 for aqueous 2-phenyl-5-sulfobenzimidazole degradation: Effect of synthetic method and the reaction
958 mechanism, Chem. Eng. J. 304 (2016) 897-907.
959 [26] J. Zou, J. Ma, L. Chen, X. Li, Y. Guan, P. Xie, C. Pan, Rapid Acceleration of Ferrous
960 Iron/Peroxymonosulfate Oxidation of Organic Pollutants by Promoting Fe(III)/Fe(II) Cycle with
961 Hydroxylamine, Environ. Sci. Technol. 47 (2013) 11685-11691.
962 [27] J.-y. Pu, J.-q. Wan, Y. Wang, Y.-w. Ma, Different Co-based MOFs templated synthesis of Co3O4
963 nanoparticles to degrade RhB by activation of oxone, RSC Adv. 6 (2016) 91791-91797.
964 [28] S. Yang, P. Wang, X. Yang, L. Shan, W. Zhang, X. Shao, R. Niu, Degradation efficiencies of azo
965 dye Acid Orange 7 by the interaction of heat, UV and anions with common oxidants: Persulfate,
966 peroxymonosulfate and hydrogen peroxide, J. Hazard. Mater. 179 (2010) 552-558.
967 [29] M. Ahmadi, F. Ghanbari, Optimizing COD removal from greywater by photoelectro-persulfate
968 process using Box-Behnken design: assessment of effluent quality and electrical energy consumption,
969 Environ. Sci. Pollut. Res. 23 (2016) 19350–19361.
970 [30] B.-T. Zhang, Y. Zhang, Y. Teng, M. Fan, Sulfate Radical and Its Application in Decontamination
971 Technologies, Crit. Rev. Environ. Sci. Technol. 45 (2015) 1756-1800.
972 [31] D. Zhou, H. Zhang, L. Chen, Sulfur‐replaced Fenton systems: can sulfate radical substitute
973 hydroxyl radical for advanced oxidation technologies?, J. Chem. Technol. Biotechnol. 90 (2015) 775-779.
974 [32] H. Li, J. Wan, Y. Ma, Y. Wang, X. Chen, Z. Guan, Degradation of refractory dibutyl phthalate by
975 peroxymonosulfate activated with novel catalysts cobalt metal-organic frameworks: Mechanism,
976 performance, and stability, J. Hazard. Mater. 318 (2016) 154-163.
977 [33] K.S. Tay, N.S.B. Ismail, Degradation of β-blockers in water by sulfate radical-based oxidation:
978 kinetics, mechanism and ecotoxicity assessment, Int. J. Environ. Sci. Technol. 13 (2016) 2495-2504.
979 [34] K.-Y.A. Lin, B.-J. Chen, Prussian blue analogue derived magnetic carbon/cobalt/iron nanocomposite
980 as an efficient and recyclable catalyst for activation of peroxymonosulfate, Chemosphere 166 (2017) 146-
981 156.
982 [35] P. Hu, M. Long, Cobalt-catalyzed sulfate radical-based advanced oxidation: A review on
983 heterogeneous catalysts and applications, Appl. Catal., B 181 (2016) 103-117.
984 [36] S. Navalon, M. Alvaro, H. Garcia, Heterogeneous Fenton catalysts based on clays, silicas and
985 zeolites, Appl. Catal., B 99 (2010) 1-26.
986 [37] H. Sun, S. Wang, Chapter 6 Catalytic oxidation of organic pollutants in aqueous solution using
987 sulfate radicals, in: Catalysis: Volume 27, The Royal Society of Chemistry, 2015, pp. 209-247.
988 [38] W.-D. Oh, Z. Dong, T.-T. Lim, Generation of sulfate radical through heterogeneous catalysis for
989 organic contaminants removal: Current development, challenges and prospects, Appl. Catal., B 194
990 (2016) 169-201.
991 [39] J.J. Pignatello, E. Oliveros, A. MacKay, Advanced Oxidation Processes for Organic Contaminant
992 Destruction Based on the Fenton Reaction and Related Chemistry, Crit. Rev. Environ. Sci. Technol. 36
993 (2006) 1-84.

55
994 [40] D.L. Ball, J.O. Edwards, The Kinetics and Mechanism of the Decomposition of Caro's Acid. I, J.
995 Am. Chem. Soc. 78 (1956) 1125-1129.
996 [41] G.P. Anipsitakis, D.D. Dionysiou, Radical Generation by the Interaction of Transition Metals with
997 Common Oxidants, Environ. Sci. Technol. 38 (2004) 3705-3712.
998 [42] Y. Ji, C. Dong, D. Kong, J. Lu, New insights into atrazine degradation by cobalt catalyzed
999 peroxymonosulfate oxidation: Kinetics, reaction products and transformation mechanisms, J. Hazard.
1000 Mater. 285 (2015) 491-500.
1001 [43] X. Chen, X. Qiao, D. Wang, J. Lin, J. Chen, Kinetics of oxidative decolorization and mineralization
1002 of Acid Orange 7 by dark and photoassisted Co2+-catalyzed peroxymonosulfate system, Chemosphere 67
1003 (2007) 802-808.
1004 [44] J. Kim, J.O. Edwards, A study of cobalt catalysis and copper modification in the coupled
1005 decompositions of hydrogen peroxide and peroxomonosulfate ion, Inorg. Chim. Acta 235 (1995) 9-13.
1006 [45] Y. Zhiyong, W. Wenhua, S. Lin, L. Liqin, W. Zhiyin, J. Xuanfeng, D. Chaonan, Q. Ruiying,
1007 Acceleration comparison between Fe2+/H2O2 and Co2+/oxone for decolouration of azo dyes in
1008 homogeneous systems, Chem. Eng. J. 234 (2013) 475-483.
1009 [46] A. Rastogi (2008). Sulfate Radical-Based Environmental Friendly Chemical Oxidation Processes for
1010 Destruction of 2-Chlorobiphenyl (PCB) and Chlorophenols (CPs) (Master's thesis). University of
1011 Cincinnati, United States.
1012 [47] A. Rastogi, S.R. Al-Abed, D.D. Dionysiou, Sulfate radical-based ferrous–peroxymonosulfate
1013 oxidative system for PCBs degradation in aqueous and sediment systems, Appl. Catal., B 85 (2009) 171-
1014 179.
1015 [48] Y. Zhiyong, M. Bensimon, D. Laub, L. Kiwi-Minsker, W. Jardim, E. Mielczarski, J. Mielczarski, J.
1016 Kiwi, Accelerated photodegradation (minute range) of the commercial azo-dye Orange II mediated by
1017 Co3O4/Raschig rings in the presence of oxone, J. Mol. Catal. A: Chem. 272 (2007) 11-19.
1018 [49] Y. Zhiyong, L. Kiwi-Minsker, A. Renken, J. Kiwi, Detoxification of diluted azo-dyes at
1019 biocompatible pH with the oxone/Co2+ reagent in dark and light processes, J. Mol. Catal. A: Chem. 252
1020 (2006) 113-119.
1021 [50] R. Matta, S. Tlili, S. Chiron, S. Barbati, Removal of carbamazepine from urban wastewater by
1022 sulfate radical oxidation, Environ. Chem. Lett. 9 (2010) 347-353.
1023 [51] Y.R. Wang, W. Chu, Degradation of a xanthene dye by Fe(II)-mediated activation of Oxone process,
1024 J. Hazard. Mater. 186 (2011) 1455-1461.
1025 [52] P. Nfodzo, H. Choi, Sulfate Radicals Destroy Pharmaceuticals and Personal Care Products, Environ.
1026 Eng. Sci. 28 (2011) 605-609.
1027 [53] P. Nfodzo, H. Choi, Triclosan decomposition by sulfate radicals: Effects of oxidant and metal doses,
1028 Chem. Eng. J. 174 (2011) 629-634.
1029 [54] S.K. Ling, S. Wang, Y. Peng, Oxidative degradation of dyes in water using Co2+/H2O2 and
1030 Co2+/peroxymonosulfate, J. Hazard. Mater. 178 (2010) 385-389.
1031 [55] Y. Huang, Z. Wang, C. Fang, W. Liu, X. Lou, J. Liu, Importance of reagent addition order in
1032 contaminant degradation in an Fe(II)/PMS system, RSC Adv. 6 (2016) 70271-70276.
1033 [56] J. Sun, X. Li, J. Feng, X. Tian, Oxone/Co2+ oxidation as an advanced oxidation process: Comparison
1034 with traditional Fenton oxidation for treatment of landfill leachate, Water Res. 43 (2009) 4363-4369.
1035 [57] E.R. Bandala, Z. DomÍnguez, F. Rivas, S. Gelover, Degradation of atrazine using solar driven
1036 fenton-like advanced oxidation processes, J. Environ. Sci. Health Pt. B 42 (2007) 21-26.

56
1037 [58] Y.-H. Huang, Y.-F. Huang, C.-i. Huang, C.-Y. Chen, Efficient decolorization of azo dye Reactive
1038 Black B involving aromatic fragment degradation in buffered Co2+/PMS oxidative processes with a ppb
1039 level dosage of Co2+-catalyst, J. Hazard. Mater. 170 (2009) 1110-1118.
1040 [59] J. Sun, M. Song, J. Feng, Y. Pi, Highly efficient degradation of ofloxacin by UV/Oxone/Co2+
1041 oxidation process, Environ. Sci. Pollut. Res. 19 (2011) 1536-1543.
1042 [60] K.H. Chan, W. Chu, Degradation of atrazine by cobalt-mediated activation of peroxymonosulfate:
1043 Different cobalt counteranions in homogenous process and cobalt oxide catalysts in photolytic
1044 heterogeneous process, Water Res. 43 (2009) 2513-2521.
1045 [61] P. Wang, S. Yang, L. Shan, R. Niu, X. Shao, Involvements of chloride ion in decolorization of Acid
1046 Orange 7 by activated peroxydisulfate or peroxymonosulfate oxidation, J. Environ. Sci. 23 (2011) 1799-
1047 1807.
1048 [62] Z. Wang, R. Yuan, Y. Guo, L. Xu, J. Liu, Effects of chloride ions on bleaching of azo dyes by
1049 Co2+/oxone regent: Kinetic analysis, J. Hazard. Mater. 190 (2011) 1083-1087.
1050 [63] J. Zhou, J. Xiao, D. Xiao, Y. Guo, C. Fang, X. Lou, Z. Wang, J. Liu, Transformations of chloro and
1051 nitro groups during the peroxymonosulfate-based oxidation of 4-chloro-2-nitrophenol, Chemosphere 134
1052 (2015) 446-451.
1053 [64] Y. Yao, H. Chen, J. Qin, G. Wu, C. Lian, J. Zhang, S. Wang, Iron encapsulated in boron and
1054 nitrogen codoped carbon nanotubes as synergistic catalysts for Fenton-like reaction, Water Res. 101
1055 (2016) 281-291.
1056 [65] R. Yuan, S.N. Ramjaun, Z. Wang, J. Liu, Effects of chloride ion on degradation of Acid Orange 7 by
1057 sulfate radical-based advanced oxidation process: Implications for formation of chlorinated aromatic
1058 compounds, J. Hazard. Mater. 196 (2011) 173-179.
1059 [66] C. Fang, D. Xiao, W. Liu, X. Lou, J. Zhou, Z. Wang, J. Liu, Enhanced AOX accumulation and
1060 aquatic toxicity during 2,4,6-trichlorophenol degradation in a Co(II)/peroxymonosulfate/Cl− system,
1061 Chemosphere 144 (2016) 2415-2420.
1062 [67] J. Lu, W. Dong, Y. Ji, D. Kong, Q. Huang, Natural Organic Matter Exposed to Sulfate Radicals
1063 Increases Its Potential to Form Halogenated Disinfection Byproducts, Environ. Sci. Technol. 50 (2016)
1064 5060-5067.
1065 [68] W. Xie, W. Dong, D. Kong, Y. Ji, J. Lu, X. Yin, Formation of halogenated disinfection by-products
1066 in cobalt-catalyzed peroxymonosulfate oxidation processes in the presence of halides, Chemosphere 154
1067 (2016) 613-619.
1068 [69] K. Liu, J. Lu, Y. Ji, Formation of brominated disinfection by-products and bromate in cobalt
1069 catalyzed peroxymonosulfate oxidation of phenol, Water Res. 84 (2015) 1-7.
1070 [70] L. Xu, R. Yuan, Y. Guo, D. Xiao, Y. Cao, Z. Wang, J. Liu, Sulfate radical-induced degradation of
1071 2,4,6-trichlorophenol: A de novo formation of chlorinated compounds, Chem. Eng. J. 217 (2013) 169-
1072 173.
1073 [71] W. Chu, W.K. Choy, C.Y. Kwan, Selection of Supported Cobalt Substrates in the Presence of Oxone
1074 for the Oxidation of Monuron, J. Agric. Food. Chem. 55 (2007) 5708-5713.
1075 [72] M. Pagano, A. Volpe, G. Mascolo, A. Lopez, V. Locaputo, R. Ciannarella, Peroxymonosulfate–
1076 Co(II) oxidation system for the removal of the non-ionic surfactant Brij 35 from aqueous solution,
1077 Chemosphere 86 (2012) 329-334.
1078 [73] K.S. Tay, N.A. Rahman, M.R. Bin Abas, Chemical oxidation of N,N-diethyl-m-toluamide by sulfate
1079 radical-based oxidation: kinetics and mechanism of degradation, Int. J. Environ. Sci. Technol. 10 (2012)
1080 103-112.

57
1081 [74] W. Qin, G. Fang, Y. Wang, T. Wu, C. Zhu, D. Zhou, Efficient transformation of DDT by
1082 peroxymonosulfate activated with cobalt in aqueous systems: Kinetics, products, and reactive species
1083 identification, Chemosphere 148 (2016) 68-76.
1084 [75] Y. Ji, D. Kong, J. Lu, H. Jin, F. Kang, X. Yin, Q. Zhou, Cobalt catalyzed peroxymonosulfate
1085 oxidation of tetrabromobisphenol A: Kinetics, reaction pathways, and formation of brominated by-
1086 products, J. Hazard. Mater. 313 (2016) 229-237.
1087 [76] K. Pirkanniemi, M. Sillanpää, Heterogeneous water phase catalysis as an environmental application:
1088 a review, Chemosphere 48 (2002) 1047-1060.
1089 [77] P.R. Shukla, S. Wang, H. Sun, H.M. Ang, M. Tadé, Activated carbon supported cobalt catalysts for
1090 advanced oxidation of organic contaminants in aqueous solution, Appl. Catal., B 100 (2010) 529-534.
1091 [78] D.M. Manohar, B.F. Noeline, T.S. Anirudhan, Adsorption performance of Al-pillared bentonite clay
1092 for the removal of cobalt(II) from aqueous phase, Appl. Clay Sci. 31 (2006) 194-206.
1093 [79] Z. Huang, H. Bao, Y. Yao, J. Lu, W. Lu, W. Chen, Key role of activated carbon fibers in enhanced
1094 decomposition of pollutants using heterogeneous cobalt/peroxymonosulfate system, J. Chem. Technol.
1095 Biotechnol. (2015).
1096 [80] S. Muhammad, E. Saputra, H. Sun, H.-M. Ang, M.O. Tadé, S. Wang, Heterogeneous Catalytic
1097 Oxidation of Aqueous Phenol on Red Mud-Supported Cobalt Catalysts, Ind. Eng. Chem. Res. 51 (2012)
1098 15351-15359.
1099 [81] X. Chen, J. Chen, X. Qiao, D. Wang, X. Cai, Performance of nano-Co3O4/peroxymonosulfate
1100 system: Kinetics and mechanism study using Acid Orange 7 as a model compound, Appl. Catal., B 80
1101 (2008) 116-121.
1102 [82] B. Yang, Z. Tian, B. Wang, Z. Sun, L. Zhang, Y. Guo, H. Li, S. Yan, Facile synthesis of
1103 Fe3O4/hierarchical-Mn3O4/graphene oxide as a synergistic catalyst for activation of peroxymonosulfate
1104 for degradation of organic pollutants, RSC Adv. 5 (2015) 20674-20683.
1105 [83] G.P. Anipsitakis, E. Stathatos, D.D. Dionysiou, Heterogeneous activation of oxone using Co3O4, J.
1106 Phys. Chem. B 109 (2005) 13052-13055.
1107 [84] Y. Rao, M. Chen, M. Zhou, Degradation of Diclofenac by Co3O4-catalyzed Activation of Oxone
1108 Process, J. Adv. Oxid. Technol. 17 (2014) 145-151.
1109 [85] Y. Wang, L. Zhou, X. Duan, H. Sun, E.L. Tin, W. Jin, S. Wang, Photochemical degradation of
1110 phenol solutions on Co3O4 nanorods with sulfate radicals, Catal. Today 258, Part 2 (2015) 576-584.
1111 [86] J. Deng, S. Feng, K. Zhang, J. Li, H. Wang, T. Zhang, X. Ma, Heterogeneous activation of
1112 peroxymonosulfate using ordered mesoporous Co3O4 for the degradation of chloramphenicol at neutral
1113 pH, Chem. Eng. J. 308 (2017) 505-515.
1114 [87] K.-Y.A. Lin, B.-J. Chen, Magnetic carbon-supported cobalt derived from a Prussian blue analogue as
1115 a heterogeneous catalyst to activate peroxymonosulfate for efficient degradation of caffeine in water, J.
1116 Colloid Interface Sci. 486 (2017) 255-264.
1117 [88] Y. Ding, L. Zhu, A. Huang, X. Zhao, X. Zhang, H. Tang, A heterogeneous Co3O4-Bi2O3 composite
1118 catalyst for oxidative degradation of organic pollutants in the presence of peroxymonosulfate, Catal. Sci.
1119 Technol. 2 (2012) 1977-1984.
1120 [89] M. Stoyanova, I. Slavova, S. Christoskova, V. Ivanova, Catalytic performance of supported
1121 nanosized cobalt and iron–cobalt mixed oxides on MgO in oxidative degradation of Acid Orange 7 azo
1122 dye with peroxymonosulfate, Appl. Catal., A 476 (2014) 121-132.
1123 [90] Y. Zhu, S. Chen, X. Quan, Y. Zhang, Cobalt implanted TiO2 nanocatalyst for heterogeneous
1124 activation of peroxymonosulfate, RSC Adv. 3 (2013) 520-525.

58
1125 [91] W. Zhang, H.L. Tay, S.S. Lim, Y. Wang, Z. Zhong, R. Xu, Supported cobalt oxide on MgO: Highly
1126 efficient catalysts for degradation of organic dyes in dilute solutions, Appl. Catal., B 95 (2010) 93-99.
1127 [92] H. Liang, H. Sun, A. Patel, P. Shukla, Z.H. Zhu, S. Wang, Excellent performance of mesoporous
1128 Co3O4/MnO2 nanoparticles in heterogeneous activation of peroxymonosulfate for phenol degradation in
1129 aqueous solutions, Appl. Catal., B 127 (2012) 330-335.
1130 [93] Y. Zhuang, Q. Lin, L. Zhang, L. Luo, Y. Yao, W. Lu, W. Chen, Mesoporous carbon-supported
1131 cobalt catalyst for selective oxidation of toluene and degradation of water contaminants, Particuology 24
1132 (2016) 216-222.
1133 [94] K.-Y.A. Lin, Y.-C. Chen, C.-F. Huang, Magnetic carbon-supported cobalt prepared from one-step
1134 carbonization of hexacyanocobaltate as an efficient and recyclable catalyst for activating Oxone, Sep.
1135 Purif. Technol. 170 (2016) 173-182.
1136 [95] P. Shi, R. Su, S. Zhu, M. Zhu, D. Li, S. Xu, Supported cobalt oxide on graphene oxide: Highly
1137 efficient catalysts for the removal of Orange II from water, J. Hazard. Mater. 229–230 (2012) 331-339.
1138 [96] P. Shi, X. Wang, X. Zhou, Y. Min, J. Fan, W. Yao, Influence of calcination temperature on the
1139 catalytic performance of Co3O4/GO nanocomposites for Orange II degradation, RSC Adv. 5 (2015)
1140 34125-34133.
1141 [97] X.J. Zhou, P.H. Shi, Y.F. Qin, J.C. Fan, Y.L. Min, W.F. Yao, Synthesis of Co3O4/graphene
1142 composite catalysts through CTAB-assisted method for Orange II degradation by activation of
1143 peroxymonosulfate, J. Mater. Sci. Mater. Electron. 27 (2015) 1020-1030.
1144 [98] P. Shi, X. Dai, H. Zheng, D. Li, W. Yao, C. Hu, Synergistic catalysis of Co3O4 and graphene oxide
1145 on Co3O4/GO catalysts for degradation of Orange II in water by advanced oxidation technology based on
1146 sulfate radicals, Chem. Eng. J. 240 (2014) 264-270.
1147 [99] H. Sun, H. Tian, Y. Hardjono, C.E. Buckley, S. Wang, Preparation of cobalt/carbon-xerogel for
1148 heterogeneous oxidation of phenol, Catal. Today 186 (2012) 63-68.
1149 [100] Y. Hardjono, H. Sun, H. Tian, C.E. Buckley, S. Wang, Synthesis of Co oxide doped carbon aerogel
1150 catalyst and catalytic performance in heterogeneous oxidation of phenol in water, Chem. Eng. J. 174
1151 (2011) 376-382.
1152 [101] Y. Yao, C. Xu, S. Miao, H. Sun, S. Wang, One-pot hydrothermal synthesis of Co(OH)2 nanoflakes
1153 on graphene sheets and their fast catalytic oxidation of phenol in liquid phase, J. Colloid Interface Sci.
1154 402 (2013) 230-236.
1155 [102] Y. Yao, C. Xu, J. Qin, F. Wei, M. Rao, S. Wang, Synthesis of Magnetic Cobalt Nanoparticles
1156 Anchored on Graphene Nanosheets and Catalytic Decomposition of Orange II, Ind. Eng. Chem. Res. 52
1157 (2013) 17341-17350.
1158 [103] Y. Yao, H. Chen, C. Lian, F. Wei, D. Zhang, G. Wu, B. Chen, S. Wang, Fe, Co, Ni nanocrystals
1159 encapsulated in nitrogen-doped carbon nanotubes as Fenton-like catalysts for organic pollutant removal,
1160 J. Hazard. Mater. 314 (2016) 129-139.
1161 [104] M. Anbia, M. Rezaie, Synthesis of Supported Ruthenium Catalyst for Phenol Degradation in the
1162 Presence of Peroxymonosulfate, Water, Air, Soil Pollut. 227 (2016) 349.
1163 [105] L. Hu, X. Yang, S. Dang, An easily recyclable Co/SBA-15 catalyst: Heterogeneous activation of
1164 peroxymonosulfate for the degradation of phenol in water, Appl. Catal., B 102 (2011) 19-26.
1165 [106] F. Qi, W. Chu, B. Xu, Catalytic degradation of caffeine in aqueous solutions by cobalt-MCM41
1166 activation of peroxymonosulfate, Appl. Catal., B 134–135 (2013) 324-332.
1167 [107] P. Shukla, S. Wang, K. Singh, H.M. Ang, M.O. Tadé, Cobalt exchanged zeolites for heterogeneous
1168 catalytic oxidation of phenol in the presence of peroxymonosulphate, Appl. Catal., B 99 (2010) 163-169.

59
1169 [108] P. Shukla, H. Sun, S. Wang, H.M. Ang, M.O. Tadé, Co-SBA-15 for heterogeneous oxidation of
1170 phenol with sulfate radical for wastewater treatment, Catal. Today 175 (2011) 380-385.
1171 [109] L. Hu, F. Yang, W. Lu, Y. Hao, H. Yuan, Heterogeneous activation of oxone with CoMg/SBA-15
1172 for the degradation of dye Rhodamine B in aqueous solution, Appl. Catal., B 134–135 (2013) 7-18.
1173 [110] K.-Y.A. Lin, H.-A. Chang, Zeolitic Imidazole Framework-67 (ZIF-67) as a heterogeneous catalyst
1174 to activate peroxymonosulfate for degradation of Rhodamine B in water, J. Taiwan Inst. Chem. Eng. 53
1175 (2015) 40-45.
1176 [111] P. Raja, M. Bensimon, U. Klehm, P. Albers, D. Laub, L. Kiwi-Minsker, A. Renken, J. Kiwi, Highly
1177 dispersed PTFE/Co3O4 flexible films as photocatalyst showing fast kinetic performance for the
1178 discoloration of azo-dyes under solar irradiation, J. Photochem. Photobiol. A: Chem. 187 (2007) 332-338.
1179 [112] S. Muhammad, E. Saputra, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, Removal of Phenol Using
1180 Sulphate Radicals Activated by Natural Zeolite-Supported Cobalt Catalysts, Water, Air, Soil Pollut. 224
1181 (2013) 1-9.
1182 [113] T. Zeng, X. Zhang, S. Wang, H. Niu, Y. Cai, Spatial Confinement of a Co3O4 Catalyst in Hollow
1183 Metal–Organic Frameworks as a Nanoreactor for Improved Degradation of Organic Pollutants, Environ.
1184 Sci. Technol. 49 (2015) 2350-2357.
1185 [114] L. Hu, F. Yang, L. Zou, H. Yuan, X. Hu, CoFe/SBA-15 catalyst coupled with peroxymonosulfate
1186 for heterogeneous catalytic degradation of rhodamine B in water, Chin. J. Catal. 36 (2015) 1785–1797.
1187 [115] Z. Xu, J. Lu, Q. Liu, L. Duan, A. Xu, Q. Wang, Y. Li, Decolorization of Acid Orange II dye by
1188 peroxymonosulfate activated with magnetic Fe3O4@C/Co nanocomposites, RSC Adv. 5 (2015) 76862-
1189 76874.
1190 [116] Y.-P. Zhu, T.-Z. Ren, Z.-Y. Yuan, Co2+-loaded periodic mesoporous aluminum phosphonates for
1191 efficient modified Fenton catalysis, RSC Adv. 5 (2015) 7628-7636.
1192 [117] F. Yang, S. Zhou, H. Wang, S. Long, X. Liu, Y. Kong, A metal-assisted templating route (S0M+I-)
1193 for fabricating thin-layer CoO covered on the channel of nanospherical-HMS with improved catalytic
1194 properties, Dalton Trans. 45 (2016) 6371-6382.
1195 [118] E. Saputra, M.A. Budihardjo, S. Bahri, J.A. Pinem, Cobalt-exchanged natural zeolite catalysts for
1196 catalytic oxidation of phenolic contaminants in aqueous solutions, J. Water Process Eng. 12 (2016) 47-51.
1197 [119] D.S. Mathew, R.-S. Juang, An overview of the structure and magnetism of spinel ferrite
1198 nanoparticles and their synthesis in microemulsions, Chem. Eng. J. 129 (2007) 51-65.
1199 [120] S. Su, W. Guo, Y. Leng, C. Yi, Z. Ma, Heterogeneous activation of Oxone by CoxFe3−xO4
1200 nanocatalysts for degradation of rhodamine B, J. Hazard. Mater. 244–245 (2013) 736-742.
1201 [121] Y. Feng, J. Liu, D. Wu, Z. Zhou, Y. Deng, T. Zhang, K. Shih, Efficient degradation of
1202 sulfamethazine with CuCo2O4 spinel nanocatalysts for peroxymonosulfate activation, Chem. Eng. J. 280
1203 (2015) 514-524.
1204 [122] Y. Ren, L. Lin, J. Ma, J. Yang, J. Feng, Z. Fan, Sulfate radicals induced from peroxymonosulfate
1205 by magnetic ferrospinel MFe2O4 (M = Co, Cu, Mn, and Zn) as heterogeneous catalysts in the water, Appl.
1206 Catal., B 165 (2015) 572-578.
1207 [123] G. Wei, X. Liang, Z. He, Y. Liao, Z. Xie, P. Liu, S. Ji, H. He, D. Li, J. Zhang, Heterogeneous
1208 activation of Oxone by substituted magnetites Fe3−xMxO4 (Cr, Mn, Co, Ni) for degradation of Acid
1209 Orange II at neutral pH, J. Mol. Catal. A: Chem. 398 (2015) 86-94.
1210 [124] Q. Yang, H. Choi, S.R. Al-Abed, D.D. Dionysiou, Iron–cobalt mixed oxide nanocatalysts:
1211 Heterogeneous peroxymonosulfate activation, cobalt leaching, and ferromagnetic properties for
1212 environmental applications, Appl. Catal., B 88 (2009) 462-469.

60
1213 [125] H. Sun, X. Yang, L. Zhao, T. Xu, J. Lian, One-pot hydrothermal synthesis of octahedral
1214 CoFe/CoFe2O4 submicron composite as heterogeneous catalysts with enhanced peroxymonosulfate
1215 activity, J. Mater. Chem. A 4 (2016) 9455-9465.
1216 [126] Y.-H. Guan, J. Ma, Y.-M. Ren, Y.-L. Liu, J.-Y. Xiao, L.-q. Lin, C. Zhang, Efficient degradation of
1217 atrazine by magnetic porous copper ferrite catalyzed peroxymonosulfate oxidation via the formation of
1218 hydroxyl and sulfate radicals, Water Res. 47 (2013) 5431-5438.
1219 [127] T. Zhang, Y. Chen, T. Leiknes, Oxidation of Refractory Benzothiazoles with PMS/CuFe2O4:
1220 Kinetics and Transformation Intermediates, Environ. Sci. Technol. 50 (2016) 5864-5873.
1221 [128] X. Zhang, M. Feng, L. Wang, R. Qu, Z. Wang, Catalytic degradation of 2-phenylbenzimidazole-5-
1222 sulfonic acid by peroxymonosulfate activated with nitrogen and sulfur co-doped CNTs-COOH loaded
1223 CuFe2O4, Chem. Eng. J. 307 (2017) 95-104.
1224 [129] W.-D. Oh, S.-K. Lua, Z. Dong, T.-T. Lim, Performance of magnetic activated carbon composite as
1225 peroxymonosulfate activator and regenerable adsorbent via sulfate radical-mediated oxidation processes,
1226 J. Hazard. Mater. 284 (2015) 1-9.
1227 [130] T. Zhang, H. Zhu, J.-P. Croué, Production of Sulfate Radical from Peroxymonosulfate Induced by a
1228 Magnetically Separable CuFe2O4 Spinel in Water: Efficiency, Stability, and Mechanism, Environ. Sci.
1229 Technol. 47 (2013) 2784-2791.
1230 [131] Y. Xu, J. Ai, H. Zhang, The mechanism of degradation of bisphenol A using the magnetically
1231 separable CuFe2O4/peroxymonosulfate heterogeneous oxidation process, J. Hazard. Mater. 309 (2016) 87-
1232 96.
1233 [132] W.-D. Oh, Z. Dong, Z.-T. Hu, T.-T. Lim, A novel quasi-cubic CuFe2O4–Fe2O3 catalyst prepared at
1234 low temperature for enhanced oxidation of bisphenol A via peroxymonosulfate activation, J. Mater.
1235 Chem. A 3 (2015) 22208-22217.
1236 [133] Y. Yao, Y. Cai, G. Wu, F. Wei, X. Li, H. Chen, S. Wang, Sulfate radicals induced from
1237 peroxymonosulfate by cobalt manganese oxides (CoxMn3−xO4) for Fenton-Like reaction in water, J.
1238 Hazard. Mater. 296 (2015) 128-137.
1239 [134] W.-D. Oh, S.-K. Lua, Z. Dong, T.-T. Lim, A novel three-dimensional spherical CuBi2O4 consisting
1240 of nanocolumn arrays with persulfate and peroxymonosulfate activation functionalities for 1 H-
1241 benzotriazole removal, Nanoscale 7 (2015) 8149-8158.
1242 [135] W.-D. Oh, S.-K. Lua, Z. Dong, T.-T. Lim, Rational design of hierarchically-structured CuBi2O4
1243 composites by deliberate manipulation of the nucleation and growth kinetics of CuBi2O4 for
1244 environmental applications, Nanoscale 8 (2016) 2046-2054.
1245 [136] P. Xu, G.M. Zeng, D.L. Huang, C.L. Feng, S. Hu, M.H. Zhao, C. Lai, Z. Wei, C. Huang, G.X. Xie,
1246 Z.F. Liu, Use of iron oxide nanomaterials in wastewater treatment: A review, Sci. Total Environ. 424
1247 (2012) 1-10.
1248 [137] A.B. Cundy, L. Hopkinson, R.L.D. Whitby, Use of iron-based technologies in contaminated land
1249 and groundwater remediation: A review, Sci. Total Environ. 400 (2008) 42-51.
1250 [138] D. Chen, X. Ma, J. Zhou, X. Chen, G. Qian, Sulfate radical-induced degradation of Acid Orange 7
1251 by a new magnetic composite catalyzed peroxymonosulfate oxidation process, J. Hazard. Mater. 279
1252 (2014) 476-484.
1253 [139] Y. Wang, H. Sun, X. Duan, H.M. Ang, M.O. Tadé, S. Wang, A new magnetic nano zero-valent iron
1254 encapsulated in carbon spheres for oxidative degradation of phenol, Appl. Catal., B 172–173 (2015) 73-
1255 81.

61
1256 [140] H. Sun, G. Zhou, S. Liu, H.M. Ang, M.O. Tadé, S. Wang, Nano-Fe0 Encapsulated in Microcarbon
1257 Spheres: Synthesis, Characterization, and Environmental Applications, ACS Appl. Mater. Interfaces 4
1258 (2012) 6235-6241.
1259 [141] J. Virkutyte, R.S. Varma, Eco-Friendly Magnetic Iron Oxide-Pillared Montmorillonite for
1260 Advanced Catalytic Degradation of Dichlorophenol, ACS Sustain. Chem. Eng. 2 (2014) 1545-1550.
1261 [142] F. Ji, C. Li, X. Wei, J. Yu, Efficient performance of porous Fe2O3 in heterogeneous activation of
1262 peroxymonosulfate for decolorization of Rhodamine B, Chem. Eng. J. 231 (2013) 434-440.
1263 [143] W.-D. Oh, S.-K. Lua, Z. Dong, T.-T. Lim, High surface area DPA-hematite for efficient
1264 detoxification of bisphenol A via peroxymonosulfate activation, J. Mater. Chem. A 2 (2014) 15836-
1265 15845.
1266 [144] F. Gong, L. Wang, D. Li, F. Zhou, Y. Yao, W. Lu, S. Huang, W. Chen, An effective heterogeneous
1267 iron-based catalyst to activate peroxymonosulfate for organic contaminants removal, Chem. Eng. J. 267
1268 (2015) 102-110.
1269 [145] M.A. Al-Shamsi, N.R. Thomson, S.P. Forsey, Iron based bimetallic nanoparticles to activate
1270 peroxygens, Chem. Eng. J. 232 (2013) 555-563.
1271 [146] Y. Wang, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, Synthesis of magnetic core/shell carbon
1272 nanosphere supported manganese catalysts for oxidation of organics in water by peroxymonosulfate, J.
1273 Colloid Interface Sci. 433 (2014) 68-75.
1274 [147] S. Zhang, Q. Fan, H. Gao, Y. Huang, X. Liu, J. Li, X. Xu, X. Wang, Formation of Fe3O4@ MnO2
1275 ball-in-ball hollow spheres as a high performance catalyst with enhanced catalytic performances, J. Mater.
1276 Chem. A 4 (2016) 1414-1422.
1277 [148] K.Y.A. Lin, J.T. Lin, A.P. Jochems, Oxidation of amaranth dye by persulfate and
1278 peroxymonosulfate activated by ferrocene, J. Chem. Technol. Biotechnol. (2016) (In Press).
1279 [149] Q. Yi, L. Bu, Z. Shi, S. Zhou, Epigallocatechin-3-gallate-coated Fe3O4 as a novel heterogeneous
1280 catalyst of peroxymonosulfate for diuron degradation: Performance and mechanism, Chem. Eng. J. 302
1281 (2016) 417-425.
1282 [150] J. Zhang, M. Chen, L. Zhu, Activation of peroxymonosulfate by iron-based catalysts for orange G
1283 degradation: role of hydroxylamine, RSC Adv. 6 (2016) 47562-47569.
1284 [151] L. Zhang, L. Zhao, J. Lian, Nanostructured Mn3O4–reduced graphene oxide hybrid and its
1285 applications for efficient catalytic decomposition of Orange II and high lithium storage capacity, RSC
1286 Adv. 4 (2014) 41838-41847.
1287 [152] M. Wei, Y. Ruan, S. Luo, X. Li, A. Xu, P. Zhang, The facile synthesis of a magnetic OMS-2
1288 catalyst for decomposition of organic dyes in aqueous solution with peroxymonosulfate, New J. Chem. 39
1289 (2015) 6395-6403.
1290 [153] Y. Wang, Y. Xie, H. Sun, J. Xiao, H. Cao, S. Wang, 2D/2D nano-hybrids of γ-MnO2 on reduced
1291 graphene oxide for catalytic ozonation and coupling peroxymonosulfate activation, J. Hazard. Mater. 301
1292 (2016) 56-64.
1293 [154] J. Du, J. Bao, Y. Liu, H. Ling, H. Zheng, S.H. Kim, D.D. Dionysiou, Efficient activation of
1294 peroxymonosulfate by magnetic Mn-MGO for degradation of bisphenol A, J. Hazard. Mater. 320 (2016)
1295 150-159.
1296 [155] E. Saputra, S. Muhammad, H. Sun, A. Patel, P. Shukla, Z.H. Zhu, S. Wang, α-MnO2 activation of
1297 peroxymonosulfate for catalytic phenol degradation in aqueous solutions, Catal. Commun. 26 (2012) 144-
1298 148.

62
1299 [156] E. Saputra, H. Zhang, Q. Liu, H. Sun, S. Wang, Egg-shaped core/shell α-Mn2O3@α-MnO2 as
1300 heterogeneous catalysts for decomposition of phenolics in aqueous solutions, Chemosphere 159 (2016)
1301 351-358.
1302 [157] Q. Liu, X. Duan, H. Sun, Y. Wang, M.O. Tade, S. Wang, Size-Tailored Porous Spheres of
1303 Manganese Oxides for Catalytic Oxidation via Peroxymonosulfate Activation, J. Phys. Chem. C 120
1304 (2016) 16871-16878.
1305 [158] E. Saputra, S. Muhammad, H. Sun, H.-M. Ang, M.O. Tadé, S. Wang, Manganese oxides at
1306 different oxidation states for heterogeneous activation of peroxymonosulfate for phenol degradation in
1307 aqueous solutions, Appl. Catal., B 142–143 (2013) 729-735.
1308 [159] L. Duan, B. Sun, M. Wei, S. Luo, F. Pan, A. Xu, X. Li, Catalytic degradation of Acid Orange 7 by
1309 manganese oxide octahedral molecular sieves with peroxymonosulfate under visible light irradiation, J.
1310 Hazard. Mater. 285 (2015) 356-365.
1311 [160] Y. Wang, S. Indrawirawan, X. Duan, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, New insights into
1312 heterogeneous generation and evolution processes of sulfate radicals for phenol degradation over one-
1313 dimensional α-MnO2 nanostructures, Chem. Eng. J. 266 (2015) 12-20.
1314 [161] E. Saputra, S. Muhammad, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, Different Crystallographic
1315 One-dimensional MnO2 Nanomaterials and Their Superior Performance in Catalytic Phenol Degradation,
1316 Environ. Sci. Technol. 47 (2013) 5882-5887.
1317 [162] C. Liu, D. Pan, X. Tang, M. Hou, Q. Zhou, J. Zhou, Degradation of Rhodamine B by the α-
1318 MnO2/Peroxymonosulfate System, Water, Air, Soil Pollut. 227 (2016) 1-10.
1319 [163] E. Saputra, S. Muhammad, H. Sun, H.-M. Ang, M.O. Tadé, S. Wang, Shape-controlled activation
1320 of peroxymonosulfate by single crystal α-Mn2O3 for catalytic phenol degradation in aqueous solution,
1321 Appl. Catal., B 154–155 (2014) 246-251.
1322 [164] L. Xu, W. Liu, X. Li, S. Rashid, C. Shen, Y. Wen, Fabrication of MnOx heterogeneous catalysts
1323 from wet sludge for degradation of azo dyes by activated peroxymonosulfate, RSC Adv. 5 (2015) 12248-
1324 12256.
1325 [165] F. Ji, C. Li, L. Deng, Performance of CuO/Oxone system: Heterogeneous catalytic oxidation of
1326 phenol at ambient conditions, Chem. Eng. J. 178 (2011) 239-243.
1327 [166] P. Zhou, B. Liu, J. Zhang, Y. Zhang, G. Zhang, C. Wei, J. Liang, Y. Liu, W. Zhang, Radicals
1328 induced from peroxomonosulfate by nanoscale zero-valent copper in the acidic solution, Water Sci.
1329 Technol. (2016) (In Press).
1330 [167] F. Ghanbari, M. Moradi, M. Manshouri, Textile wastewater decolorization by zero valent iron
1331 activated peroxymonosulfate: Compared with zero valent copper, J. Environ. Chem. Eng. 2 (2014) 1846-
1332 1851.
1333 [168] F. Ji, C. Li, Y. Liu, P. Liu, Heterogeneous activation of peroxymonosulfate by Cu/ZSM5 for
1334 decolorization of Rhodamine B, Sep. Purif. Technol. 135 (2014) 1-6.
1335 [169] Y. Feng, D. Wu, Y. Deng, T. Zhang, K. Shih, Sulfate Radical-Mediated Degradation of
1336 Sulfadiazine by CuFeO2 Rhombohedral Crystal-Catalyzed Peroxymonosulfate: Synergistic Effects and
1337 Mechanisms, Environ. Sci. Technol. 50 (2016) 3119-3127.
1338 [170] M. Moradi, F. Ghanbari, Application of response surface method for coagulation process in
1339 leachate treatment as pretreatment for Fenton process: Biodegradability improvement, J. Water Process
1340 Eng. 4 (2014) 67-73.

63
1341 [171] M. Moradi, F. Ghanbari, E. Minaee Tabrizi, Removal of acid yellow 36 using Box–Behnken
1342 designed photoelectro-Fenton: a study on removal mechanisms, Toxicol. Environ. Chem. 97 (2015) 700-
1343 709.
1344 [172] H.R. Sindelar, M.T. Brown, T.H. Boyer, Evaluating UV/H2O2, UV/percarbonate, and UV/perborate
1345 for natural organic matter reduction from alternative water sources, Chemosphere 105 (2014) 112-118.
1346 [173] N. Jaafarzadeh, F. Ghanbari, M. Moradi, Photo-electro-oxidation assisted peroxymonosulfate for
1347 decolorization of acid brown 14 from aqueous solution, Korean J. Chem. Eng. 32 (2015) 458-464.
1348 [174] T. Olmez-Hanci, C. Imren, I. Kabdaşlı, O. Tünay, I. Arslan-Alaton, Application of the UV-C photo-
1349 assisted peroxymonosulfate oxidation for the mineralization of dimethyl phthalate in aqueous solutions,
1350 Photochem. Photobiol. Sci. 10 (2011) 408-413.
1351 [175] F.J. Beltran, Ozone-UV radiation-hydrogen peroxide oxidation technologies, in: M.A. Tarr (Ed.)
1352 Chemical Degradation Methods for Wastes and Pollutants: Environmental and Industrial Applications,
1353 Marcel Dekker, New York, USA, 2003, pp. 1-76.
1354 [176] N. Getoff, Purification of drinking water by irradiation. A review, J. Chem. Sci. 105 (1993) 373-
1355 391.
1356 [177] J. Sharma, I.M. Mishra, D.D. Dionysiou, V. Kumar, Oxidative removal of Bisphenol A by UV-
1357 C/peroxymonosulfate (PMS): Kinetics, influence of co-existing chemicals and degradation pathway,
1358 Chem. Eng. J. 276 (2015) 193-204.
1359 [178] Y. Rao, D. Xue, H. Pan, J. Feng, Y. Li, Degradation of ibuprofen by a synergistic
1360 UV/Fe(III)/Oxone process, Chem. Eng. J. 283 (2016) 65-75.
1361 [179] X. He, A.A. de la Cruz, D.D. Dionysiou, Destruction of cyanobacterial toxin cylindrospermopsin
1362 by hydroxyl radicals and sulfate radicals using UV-254 nm activation of hydrogen peroxide, persulfate
1363 and peroxymonosulfate, J. Photochem. Photobiol. A: Chem. 251 (2013) 160-166.
1364 [180] L.G. Devi, M. Srinivas, M.L. ArunaKumari, Heterogeneous advanced photo- Fenton process using
1365 peroxymonosulfate and peroxydisulfate in presence of zero valent metallic iron: A comparative study
1366 with hydrogen peroxide photo-Fenton process, J. Water Process Eng. 13 (2016) 117-126.
1367 [181] S. Verma, S. Nakamura, M. Sillanpää, Application of UV-C LED activated PMS for the
1368 degradation of anatoxin-a, Chem. Eng. J. 284 (2016) 122-129.
1369 [182] N. Jaafarzadeh, M. Omidinasab, F. Ghanbari, Combined electrocoagulation and UV-based sulfate
1370 radical oxidation processes for treatment of pulp and paper wastewater, Process Saf. Environ. Prot. 102
1371 (2016) 462-472.
1372 [183] S. Khan, X. He, H.M. Khan, D. Boccelli, D.D. Dionysiou, Efficient degradation of lindane in
1373 aqueous solution by iron (II) and/or UV activated peroxymonosulfate, J. Photochem. Photobiol. A: Chem.
1374 316 (2016) 37-43.
1375 [184] J. Rivas, O. Gimeno, T. Borralho, F. Beltrán, Influence of oxygen and free radicals promoters on
1376 the UV-254 nm photolysis of diclofenac, Chem. Eng. J. 163 (2010) 35-40.
1377 [185] M. Mahdi-Ahmed, S. Chiron, Ciprofloxacin oxidation by UV-C activated peroxymonosulfate in
1378 wastewater, J. Hazard. Mater. 265 (2014) 41-46.
1379 [186] Y.-H. Guan, J. Ma, X.-C. Li, J.-Y. Fang, L.-W. Chen, Influence of pH on the Formation of Sulfate
1380 and Hydroxyl Radicals in the UV/Peroxymonosulfate System, Environ. Sci. Technol. 45 (2011) 9308-
1381 9314.
1382 [187] X. Liu, T. Zhang, Y. Zhou, L. Fang, Y. Shao, Degradation of atenolol by UV/peroxymonosulfate:
1383 Kinetics, effect of operational parameters and mechanism, Chemosphere 93 (2013) 2717-2724.

64
1384 [188] Y. Pi, J. Feng, J. Sun, J. Sun, Facile, effective, and environment-friendly degradation of
1385 sulfamonomethoxine in aqueous solution with the aid of a UV/Oxone oxidative process, Environ. Sci.
1386 Pollut. Res. 20 (2013) 8621-8628.
1387 [189] E.R. Bandala, M.A. Peláez, D.D. Dionysiou, S. Gelover, J. Garcia, D. Macías, Degradation of 2,4-
1388 dichlorophenoxyacetic acid (2,4-D) using cobalt-peroxymonosulfate in Fenton-like process, J.
1389 Photochem. Photobiol. A: Chem. 186 (2007) 357-363.
1390 [190] Q. Yang, H. Choi, Y. Chen, D.D. Dionysiou, Heterogeneous activation of peroxymonosulfate by
1391 supported cobalt catalysts for the degradation of 2,4-dichlorophenol in water: The effect of support, cobalt
1392 precursor, and UV radiation, Appl. Catal., B 77 (2008) 300-307.
1393 [191] J. Fernandez, P. Maruthamuthu, J. Kiwi, Photobleaching and mineralization of Orange II by oxone
1394 and metal-ions involving Fenton-like chemistry under visible light, J. Photochem. Photobiol. A: Chem.
1395 161 (2004) 185-192.
1396 [192] J. Fernandez, V. Nadtochenko, J. Kiwi, Photobleaching of Orange II within seconds using the
1397 oxone/Co2+ reagent through Fenton-like chemistry, Chem. Commun. (2003) 2382-2383.
1398 [193] Y. Wang, W. Chu, Photo-assisted degradation of 2, 4, 5-trichlorophenoxyacetic acid by Fe (II)-
1399 catalyzed activation of Oxone process: the role of UV irradiation, reaction mechanism and mineralization,
1400 Appl. Catal., B 123 (2012) 151-161.
1401 [194] J. Deng, Y. Shao, N. Gao, S. Xia, C. Tan, S. Zhou, X. Hu, Degradation of the antiepileptic drug
1402 carbamazepine upon different UV-based advanced oxidation processes in water, Chem. Eng. J. 222
1403 (2013) 150-158.
1404 [195] H.-Y. Shu, M.-C. Chang, S.-W. Huang, Decolorization and mineralization of azo dye Acid Blue
1405 113 by the UV/Oxone process and optimization of operating parameters, Desalin. Water Treat. 57 (2016)
1406 7951-7962.
1407 [196] T. Olmez-Hanci, I. Arslan-Alaton, D. Dursun, B. Genc, D. Mita, M. Guida, L. Mita, Degradation
1408 and toxicity assessment of the nonionic surfactant Triton™ X-45 by the peroxymonosulfate/UV-C
1409 process, Photochem. Photobiol. Sci. 14 (2015) 569-575.
1410 [197] J. Lifka, B. Ondruschka, J. Hofmann, The Use of Ultrasound for the Degradation of Pollutants in
1411 Water: Aquasonolysis – A Review, Eng. Life Sci. 3 (2003) 253-262.
1412 [198] X. Zou, T. Zhou, J. Mao, X. Wu, Synergistic degradation of antibiotic sulfadiazine in a
1413 heterogeneous ultrasound-enhanced Fe0/persulfate Fenton-like system, Chem. Eng. J. 257 (2014) 36-44.
1414 [199] F. Soumia, C. Petrier, Effect of potassium monopersulfate (oxone) and operating parameters on
1415 sonochemical degradation of cationic dye in an aqueous solution, Ultrason. Sonochem. 32 (2016) 343-
1416 347.
1417 [200] J. Liu, J. Zhou, Z. Ding, Z. Zhao, X. Xu, Z. Fang, Ultrasound irritation enhanced heterogeneous
1418 activation of peroxymonosulfate with Fe3O4 for degradation of azo dye, Ultrason. Sonochem. 34 (2017)
1419 953-959.
1420 [201] A.B. Kurukutla, P.S.S. Kumar, S. Anandan, T. Sivasankar, Sonochemical Degradation of
1421 Rhodamine B Using Oxidants, Hydrogen Peroxide/Peroxydisulfate/Peroxymonosulfate, with Fe2+ Ion:
1422 Proposed Pathway and Kinetics, Environ. Eng. Sci. 32 (2015) 129-140.
1423 [202] S. Su, W. Guo, C. Yi, Y. Leng, Z. Ma, Degradation of amoxicillin in aqueous solution using
1424 sulphate radicals under ultrasound irradiation, Ultrason. Sonochem. 19 (2012) 469-474.
1425 [203] W. Guo, S. Su, C. Yi, Z. Ma, Degradation of antibiotics amoxicillin by Co3O4‐catalyzed
1426 peroxymonosulfate system, Environ. Prog. Sustain. Energy 32 (2013) 193-197.

65
1427 [204] H. Eskandarloo, A. Badiei, M.A. Behnajady, G.M. Ziarani, Ultrasonic-assisted degradation of
1428 phenazopyridine with a combination of Sm-doped ZnO nanoparticles and inorganic oxidants, Ultrason.
1429 Sonochem. 28 (2016) 169-177.
1430 [205] M. Ahmadi, F. Ghanbari, M. Moradi, Photocatalysis assisted by peroxymonosulfate and persulfate
1431 for benzotriazole degradation: effect of pH on sulfate and hydroxyl radicals, Water Sci. Technol. 72
1432 (2015) 2095-2102.
1433 [206] S.K. Kuriechen, S. Murugesan, Carbon-Doped Titanium Dioxide Nanoparticles Mediated
1434 Photocatalytic Degradation of Azo Dyes Under Visible Light, Water, Air, Soil Pollut. 224 (2013) 1-8.
1435 [207] K.-Y.A. Lin, Z.-Y. Zhang, Metal-free activation of Oxone using one-step prepared sulfur-doped
1436 carbon nitride under visible light irradiation, Sep. Purif. Technol. 173 (2017) 72-79.
1437 [208] R.R. Solís, F.J. Rivas, M. Tierno, Monopersulfate photocatalysis under 365 nm radiation. Direct
1438 oxidation and monopersulfate promoted photocatalysis of the herbicide tembotrione, J Environ Manage
1439 181 (2016) 385-394.
1440 [209] J. Andersen, M. Pelaez, L. Guay, Z. Zhang, K. O'Shea, D.D. Dionysiou, NF-TiO2 photocatalysis of
1441 amitrole and atrazine with addition of oxidants under simulated solar light: Emerging synergies,
1442 degradation intermediates, and reusable attributes, J. Hazard. Mater. 260 (2013) 569-575.
1443 [210] S. Anandan, P. Sathish Kumar, N. Pugazhenthiran, J. Madhavan, P. Maruthamuthu, Effect of
1444 loaded silver nanoparticles on TiO2 for photocatalytic degradation of Acid Red 88, Sol. Energy Mater.
1445 Sol. Cells 92 (2008) 929-937.
1446 [211] H. Sun, S. Liu, S. Liu, S. Wang, A comparative study of reduced graphene oxide modified TiO2,
1447 ZnO and Ta2O5 in visible light photocatalytic/photochemical oxidation of methylene blue, Appl. Catal., B
1448 146 (2014) 162-168.
1449 [212] P. Shukla, I. Fatimah, S. Wang, H.M. Ang, M.O. Tadé, Photocatalytic generation of sulphate and
1450 hydroxyl radicals using zinc oxide under low-power UV to oxidise phenolic contaminants in wastewater,
1451 Catal. Today 157 (2010) 410-414.
1452 [213] N. Pugazhenthiran, S. Murugesan, P. Sathishkumar, S. Anandan, Photocatalytic degradation of
1453 ceftiofur sodium in the presence of gold nanoparticles loaded TiO2 under UV–visible light, Chem. Eng. J.
1454 241 (2014) 401-409.
1455 [214] Q. Chen, F. Ji, Q. Guo, J. Fan, X. Xu, Combination of heterogeneous Fenton-like reaction and
1456 photocatalysis using Co–TiO2 nanocatalyst for activation of KHSO5 with visible light irradiation at
1457 ambient conditions, J. Environ. Sci. 26 (2014) 2440-2450.
1458 [215] Y. Liu, H. Guo, Y. Zhang, W. Tang, X. Cheng, H. Liu, Activation of peroxymonosulfate by BiVO4
1459 under visible light for degradation of Rhodamine B, Chem. Phys. Lett. (2016).
1460 [216] X. Chen, W. Wang, H. Xiao, C. Hong, F. Zhu, Y. Yao, Z. Xue, Accelerated TiO2 photocatalytic
1461 degradation of Acid Orange 7 under visible light mediated by peroxymonosulfate, Chem. Eng. J. 193–194
1462 (2012) 290-295.
1463 [217] K. Zhu, J. Wang, Y. Wang, C. Jin, A.S. Ganeshraja, Visible-light-induced photocatalysis and
1464 peroxymonosulfate activation over ZnFe2O4 fine nanoparticles for degradation of Orange II, Catal. Sci.
1465 Technol. 6 (2016) 2296-2304.
1466 [218] P. Sathishkumar, N. Pugazhenthiran, R.V. Mangalaraja, A.M. Asiri, S. Anandan, ZnO supported
1467 CoFe2O4 nanophotocatalysts for the mineralization of Direct Blue 71 in aqueous environments, J. Hazard.
1468 Mater. 252–253 (2013) 171-179.

66
1469 [219] K.-Y.A. Lin, Y.-C. Chen, Accelerated decomposition of Oxone using graphene-like carbon nitride
1470 with visible light irradiation for enhanced decolorization in water, J. Taiwan Inst. Chem. Eng. 60 (2016)
1471 423-429.
1472 [220] Y. Tao, M. Wei, D. Xia, A. Xu, X. Li, Polyimides as metal-free catalysts for organic dye
1473 degradation in the presence peroxymonosulfate under visible light irradiation, RSC Adv. 5 (2015) 98231-
1474 98240.
1475 [221] Y. Tao, Q. Ni, M. Wei, D. Xia, X. Li, A. Xu, Metal-free activation of peroxymonosulfate by g-
1476 C3N4 under visible light irradiation for the degradation of organic dyes, RSC Adv. 5 (2015) 44128-
1477 44136.
1478 [222] K.-Y. Andrew Lin, Z.-Y. Zhang, α-Sulfur as a metal-free catalyst to activate peroxymonosulfate
1479 under visible light irradiation for decolorization, RSC Adv. 6 (2016) 15027-15034.
1480 [223] X. Duan, Z. Ao, D. Li, H. Sun, L. Zhou, A. Suvorova, M. Saunders, G. Wang, S. Wang, Surface-
1481 tailored nanodiamonds as excellent metal-free catalysts for organic oxidation, Carbon 103 (2016) 404-
1482 411.
1483 [224] X. Duan, H. Sun, Z. Ao, L. Zhou, G. Wang, S. Wang, Unveiling the active sites of graphene-
1484 catalyzed peroxymonosulfate activation, Carbon 107 (2016) 371-378.
1485 [225] E. Saputra, S. Muhammad, H. Sun, S. Wang, Activated carbons as green and effective catalysts for
1486 generation of reactive radicals in degradation of aqueous phenol, RSC Adv. 3 (2013) 21905-21910.
1487 [226] J. Zhang, X. Shao, C. Shi, S. Yang, Decolorization of Acid Orange 7 with peroxymonosulfate
1488 oxidation catalyzed by granular activated carbon, Chem. Eng. J. 232 (2013) 259-265.
1489 [227] S. Yang, T. Xiao, J. Zhang, Y. Chen, L. Li, Activated carbon fiber as heterogeneous catalyst of
1490 peroxymonosulfate activation for efficient degradation of Acid Orange 7 in aqueous solution, Sep. Purif.
1491 Technol. 143 (2015) 19-26.
1492 [228] F. Ghanbari, M. Moradi, F. Gohari, Degradation of 2,4,6-trichlorophenol in aqueous solutions using
1493 peroxymonosulfate/activated carbon/UV process via sulfate and hydroxyl radicals, J. Water Process Eng.
1494 9 (2016) 22-28.
1495 [229] S. Indrawirawan, H. Sun, X. Duan, S. Wang, Nanocarbons in different structural dimensions (0–
1496 3D) for phenol adsorption and metal-free catalytic oxidation, Appl. Catal., B 179 (2015) 352-362.
1497 [230] S. Yang, L. Li, T. Xiao, Y. Zhang, D. Zheng, Promoting effect of ammonia modification on
1498 activated carbon catalyzed peroxymonosulfate oxidation, Sep. Purif. Technol. 160 (2016) 81-88.
1499 [231] C. Wang, J. Kang, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, One-pot synthesis of N-doped
1500 graphene for metal-free advanced oxidation processes, Carbon 102 (2016) 279-287.
1501 [232] H. Liu, P. Sun, M. Feng, H. Liu, S. Yang, L. Wang, Z. Wang, Nitrogen and sulfur co-doped CNT-
1502 COOH as an efficient metal-free catalyst for the degradation of UV filter BP-4 based on sulfate radicals,
1503 Appl. Catal., B 187 (2016) 1-10.
1504 [233] H. Cao, L. Xing, G. Wu, Y. Xie, S. Shi, Y. Zhang, D. Minakata, J.C. Crittenden, Promoting effect
1505 of nitration modification on activated carbon in the catalytic ozonation of oxalic acid, Appl. Catal., B 146
1506 (2014) 169-176.
1507 [234] S. Yang, L. Li, T. Xiao, D. Zheng, Y. Zhang, Role of surface chemistry in modified ACF (activated
1508 carbon fiber)-catalyzed peroxymonosulfate oxidation, Appl. Surf. Sci. 383 (2016) 142-150.
1509 [235] H. Sun, C. Kwan, A. Suvorova, H.M. Ang, M.O. Tadé, S. Wang, Catalytic oxidation of organic
1510 pollutants on pristine and surface nitrogen-modified carbon nanotubes with sulfate radicals, Appl. Catal.,
1511 B 154–155 (2014) 134-141.

67
1512 [236] M. Wei, L. Gao, J. Li, J. Fang, W. Cai, X. Li, A. Xu, Activation of peroxymonosulfate by graphitic
1513 carbon nitride loaded on activated carbon for organic pollutants degradation, J. Hazard. Mater. 316 (2016)
1514 60-68.
1515 [237] H. Dong, M. Wei, J. Li, J. Fang, L. Gao, X. Li, A. Xu, Catalytic performance of supported g-C3N4
1516 on MCM-41 in organic dye degradation with peroxymonosulfate, RSC Adv. 6 (2016) 70747-70755.
1517 [238] X. Duan, Z. Ao, L. Zhou, H. Sun, G. Wang, S. Wang, Occurrence of radical and nonradical
1518 pathways from carbocatalysts for aqueous and nonaqueous catalytic oxidation, Appl. Catal., B 188 (2016)
1519 98-105.
1520 [239] X. Duan, Z. Ao, H. Sun, L. Zhou, G. Wang, S. Wang, Insights into N-doping in single-walled
1521 carbon nanotubes for enhanced activation of superoxides: a mechanistic study, Chem. Commun. 51
1522 (2015) 15249-15252.
1523 [240] H. Lin, J. Wu, H. Zhang, Degradation of clofibric acid in aqueous solution by an EC/Fe3+/PMS
1524 process, Chem. Eng. J. 244 (2014) 514-521.
1525 [241] E. Brillas, I. Sirés, M.A. Oturan, Electro-Fenton process and related electrochemical technologies
1526 based on Fenton’s reaction chemistry, Chem. Rev. 109 (2009) 6570-6631.
1527 [242] K. Govindan, M. Raja, S.U. Maheshwari, M. Noel, Analysis and understanding of amido black 10B
1528 dye degradation in aqueous solution by electrocoagulation with the conventional oxidants
1529 peroxomonosulfate, peroxodisulfate and hydrogen peroxide, Environ. Sci.: Water Res. Technol. 1 (2015)
1530 108-119.
1531 [243] X. Lou, L. Wu, Y. Guo, C. Chen, Z. Wang, D. Xiao, C. Fang, J. Liu, J. Zhao, S. Lu,
1532 Peroxymonosulfate activation by phosphate anion for organics degradation in water, Chemosphere 117
1533 (2014) 582-585.
1534 [244] B.-T. Zhang, W. Xiang, X. Jiang, Y. Zhang, Y. Teng, Oxidation of Dyes by Alkaline-Activated
1535 Peroxymonosulfate, J. Environ. Eng. 142 (2016) 10.1061/(ASCE)EE.1943-7870.0001084, 04016003.
1536 [245] C. Qi, X. Liu, J. Ma, C. Lin, X. Li, H. Zhang, Activation of peroxymonosulfate by base:
1537 Implications for the degradation of organic pollutants, Chemosphere 151 (2016) 280-288.
1538 [246] M.G. Antoniou, A.A. de la Cruz, D.D. Dionysiou, Degradation of microcystin-LR using sulfate
1539 radicals generated through photolysis, thermolysis and e− transfer mechanisms, Appl. Catal., B 96 (2010)
1540 290-298.
1541 [247] Y. Zhou, J. Jiang, Y. Gao, J. Ma, S.-Y. Pang, J. Li, X.-T. Lu, L.-P. Yuan, Activation of
1542 Peroxymonosulfate by Benzoquinone: A Novel Nonradical Oxidation Process, Environ. Sci. Technol. 49
1543 (2015) 12941-12950.
1544 [248] X.-Y. Lou, Y.-G. Guo, D.-X. Xiao, Z.-H. Wang, S.-Y. Lu, J.-S. Liu, Rapid dye degradation with
1545 reactive oxidants generated by chloride-induced peroxymonosulfate activation, Environ. Sci. Pollut. Res.
1546 20 (2013) 6317-6323.
1547 [249] I. Arslan-Alaton, T. Olmez-Hanci, S. Khoei, H. Fakhri, Oxidative degradation of Triton X-45 using
1548 zero valent aluminum in the presence of hydrogen peroxide, persulfate and peroxymonosulfate, Catal.
1549 Today (2016) (In press).
1550 [250] Y. Yang, J. Jiang, X. Lu, J. Ma, Y. Liu, Production of Sulfate Radical and Hydroxyl Radical by
1551 Reaction of Ozone with Peroxymonosulfate: A Novel Advanced Oxidation Process, Environ. Sci.
1552 Technol. 49 (2015) 7330-7339.
1553 [251] Z. Huang, H. Bao, Y. Yao, W. Lu, W. Chen, Novel green activation processes and mechanism of
1554 peroxymonosulfate based on supported cobalt phthalocyanine catalyst, Appl. Catal., B 154–155 (2014)
1555 36-43.

68
1556 [252] Z. Huang, Y. Yao, J. Lu, C. Chen, W. Lu, S. Huang, W. Chen, The consortium of heterogeneous
1557 cobalt phthalocyanine catalyst and bicarbonate ion as a novel platform for contaminants elimination based
1558 on peroxymonosulfate activation, J. Hazard. Mater. 301 (2016) 214-221.
1559 [253] Y.-F. Huang, Y.-H. Huang, Behavioral evidence of the dominant radicals and intermediates
1560 involved in Bisphenol A degradation using an efficient Co2+/PMS oxidation process, J. Hazard. Mater.
1561 167 (2009) 418-426.
1562 [254] D. Tang, G. Zhang, S. Guo, Efficient activation of peroxymonosulfate by manganese oxide for the
1563 degradation of azo dye at ambient condition, J. Colloid Interface Sci. 454 (2015) 44-51.
1564 [255] C. Liang, H.-W. Su, Identification of sulfate and hydroxyl radicals in thermally activated persulfate,
1565 Ind. Eng. Chem. Res. 48 (2009) 5558-5562.
1566 [256] R. Hazime, Q.H. Nguyen, C. Ferronato, A. Salvador, F. Jaber, J.M. Chovelon, Comparative study
1567 of imazalil degradation in three systems: UV/TiO2, UV/K2S2O8 and UV/TiO2/K2S2O8, Appl. Catal., B 144
1568 (2014) 286-291.
1569 [257] J.-C.E. Yang, H. Lan, X.-Q. Lin, B. Yuan, M.-L. Fu, Synthetic conditions-regulated catalytic
1570 Oxone efficacy of MnOx/SBA-15 towards butyl paraben (BPB) removal under heterogeneous conditions,
1571 Chem. Eng. J. 289 (2016) 296-305.
1572 [258] W. Zhang, S. Yang, R. Niu, X. Shao, L. Shan, X. Yang, P. Wang, Microwave-assisted COD
1573 removal from landfill leachate by hydrogen peroxide, peroxymonosulfate and persulfate, in:
1574 Bioinformatics and Biomedical Engineering (iCBBE), 2010 4th International Conference on, IEEE, 2010,
1575 pp. 1-4.
1576 [259] F.J. Rivas, F.J. Beltrán, F. Carvalho, P.M. Alvarez, Oxone-Promoted Wet Air Oxidation of Landfill
1577 Leachates, Ind. Eng. Chem. Res. 44 (2005) 749-758.
1578 [260] M.A. Fagier, E.A. Ali, K.S. Tay, M.R.B. Abas, Mineralization of organic matter from vinasse using
1579 physicochemical treatment coupled with Fe2+-activated persulfate and peroxymonosulfate oxidation, Int.
1580 J. Environ. Sci. Technol. 13 (2016) 1189-1194.
1581 [261] J. Rodríguez-Chueca, C. Amor, T. Silva, D.D. Dionysiou, G.L. Puma, M.S. Lucas, J.A. Peres,
1582 Treatment of winery wastewater by sulphate radicals: HSO5-/transition metal/UV-A LEDs, Chem. Eng. J.
1583 (2016) (In press).
1584 [262] G.K. Dinesh, S. Anandan, T. Sivasankar, Synthesis of Fe-doped Bi2O3 nanocatalyst and its
1585 sonophotocatalytic activity on synthetic dye and real textile wastewater, Environ. Sci. Pollut. Res. (2016)
1586 1-11.
1587 [263] N. Jaafarzadeh, F. Ghanbari, M. Ahmadi, M. Omidinasab, Efficient integrated processes for pulp
1588 and paper wastewater treatment and phytotoxicity reduction: Permanganate, electro-Fenton and
1589 Co3O4/UV/peroxymonosulfate, Chem. Eng. J. 308 (2017) 142-150.
1590 [264] S. Yang, Y. Li, L. Wang, L. Feng, Use of peroxymonosulfate in wet scrubbing process for efficient
1591 odor control, Sep. Purif. Technol. 158 (2016) 80-86.
1592 [265] Y. Liu, Y. Wang, Simultaneous Removal of NO and SO2 Using Aqueous Peroxymonosulfate with
1593 Coactivation of Cu2+/Fe3+ and High Temperature, AlChE J. (2016) http://dx.doi.org/10.1002/aic.15503
1594 (In press)
1595 [266] S.-H. Do, J.-H. Jo, Y.-H. Jo, H.-K. Lee, S.-H. Kong, Application of a peroxymonosulfate/cobalt
1596 (PMS/Co(II)) system to treat diesel-contaminated soil, Chemosphere 77 (2009) 1127-1131.
1597 [267] T. Niu, Z. Zhou, W. Ren, L.-M. Jiang, B. Li, H. Wei, J. Li, L. Wang, Effects of potassium
1598 peroxymonosulfate on disintegration of waste sludge and properties of extracellular polymeric substances,
1599 Int. Biodeterior. Biodegrad. 106 (2016) 170-177.

69
1600 [268] S. Wacławek, K. Grübel, P. Dennis, V.T.P. Vinod, M. Černík, A novel approach for simultaneous
1601 improvement of dewaterability, post-digestion liquor properties and toluene removal from anaerobically
1602 digested sludge, Chem. Eng. J. 291 (2016) 192-198.
1603 [269] E.R. Bandala, Y. Velasco, L.G. Torres, Decontamination of soil washing wastewater using solar
1604 driven advanced oxidation processes, J. Hazard. Mater. 160 (2008) 402-407.

1605

1606

1607

70
1608

1609

71
1610 Highlights

1611 • A to Z of peroxymonosulfate (PMS) activation methods are indicated and discussed.

1612 • PMS has shown a high ability in generating SO4•- for organic pollutants degradation.

1613 • PMS applications for real wastewaters and for air, soil and sludge are reviewed.

1614 • The advantages and limitations of PMS application are propounded.

1615 • Conventional and cutting-edge PMS-based processes are reported.

1616

1617

72

You might also like