Anda di halaman 1dari 72

Accepted Manuscript

Review

Chemistry of persulfates in water and wastewater treatment: a review

Stanisaw Wacawek, Holger V. Lutze, Klaudiusz Grbel, Vinod V.T. Padil,


Miroslav ernk, Dionysios.D. Dionysiou

PII: S1385-8947(17)31271-8
DOI: http://dx.doi.org/10.1016/j.cej.2017.07.132
Reference: CEJ 17398

To appear in: Chemical Engineering Journal

Received Date: 5 June 2017


Revised Date: 18 July 2017
Accepted Date: 19 July 2017

Please cite this article as: S. Wacawek, H.V. Lutze, K. Grbel, V.V.T. Padil, M. ernk, Dionysios.D. Dionysiou,
Chemistry of persulfates in water and wastewater treatment: a review, Chemical Engineering Journal (2017), doi:
http://dx.doi.org/10.1016/j.cej.2017.07.132

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.
Chemistry of persulfates in water and wastewater

treatment: a review

Stanisaw Wacawek 1*, Holger V. Lutze2,3,4, Klaudiusz Grbel5, Vinod V.T. Padil1,

Miroslav ernk1 and Dionysios. D. Dionysiou6*


1
Centre for Nanomaterials, Advanced Technologies and Innovation, Technical University of Liberec, Studentsk

1402/2, 461 17 Liberec 1, Czech Republic


2
University Duisburg-Essen, Faculty of Chemistry, Instrumental Analytical Chemistry, Universitaetsstrae 5, D-

45141 Essen, Germany

3
IWW Water Centre, Moritzstr. 26, D-45476 Mlheim an der Ruhr, Germany

4
Centre for Water and Environmental Research (ZWU) Universittsstr. 5 D-45141 Essen

5
Institute of Environmental Protection and Engineering, Department of Environmental Microbiology and

Biotechnology, University of Bielsko-Biala, Willowa 2, 43-309 Bielsko-Biaa, Poland

6
Environmental Engineering and Science Program, University of Cincinnati, 705 Engineering Research Center,

Cincinnati, Ohio, 45221-0012, United States of America

*corresponding authors: stanislaw.waclawek@tul.cz; dionysios.d.dionysiou@uc.edu

Abstract

Persulfate decontamination technologies either utilizing radical driven processes or direct

electron transfer are very powerful tools for the treatment of a broad range of impurities,

including halogenated olefins, BTEXs (benzene, toluene, ethylbenzene and xylenes),

perfluorinated chemicals, phenols, pharmaceuticals, inorganics and pesticides. Furthermore,

the reactivity of persulfates is extremely dependent on the related activation techniques and

the composition of the treated water matrix. Direct reactions of peroxydisulfate (PDS) or

1
peroxymonosulfate (PMS) are rather slow and mostly unsuitable for pollutant degradation.

However, PDS or PMS decompose at elevated temperatures under UV radiation, and

radiolysis treatment as well as in presence of reduced metal ions to form sulfate radicals

(SO4). (SO4)-based oxidation can also form secondary oxidants for instance carbonate

radicals, hydroxyl radicals, superoxide radicals or singlet oxygen which can influence both

transformation efficiency and product formation. The formation of such species is extremely

subjected on the water matrix composition and can hardly be predicted. One important aspect

in dealing with PDS or PMS is their analysis, which is often prone for interference by other

matrix components and hampered by the low stability of PDS and PMS in aqueous systems.

Numerous methods for analysis of PDS and PMS are available. The present work also

provides an overview on these methods.

Keywords: persulfates; peroxydisulfate; peroxymonosulfate; remediation; oxidation; sulfate

radical

1. Introduction

Over the last few decades, ubiquitous contamination with various inorganic and organic

substances has caused serious problems all over the world [1,2]. One reason is the misuse or

overuse of antibiotics e.g. for agricultural purposes, which enrich the population of multi-

resistant pathogens [3]. In many countries with a low human development index (HDI),

particularly in Africa and Asia, surface- and groundwater contamination can be a direct cause

of water scarcity. It is important, therefore, to develop cheap and efficient water and

wastewater treatment methods [4].

There are many different water and wastewater decontamination methods, which can be

universally divided into biological [5], physical [6], and chemical methods [7].

2
Oxidants used for environmental purposes can be divided into ones that possess a peroxide

bond and ones that do not. Oxidizing agents such as ozone and hypochlorite have long been

employed for water and wastewater treatment [8,9]. Oxidants possessing the O-O bond

(known as peroxide or the peroxo group), typically form free radicals which in turn may result

in pollutant degradation.

Peroxydisulfate (PDS) and peroxymonosulfate (PMS) are gaining increasing attention in

water and wastewater treatment [10-12] (Fig. 1). Many studies have shown that they are

capable of degrading highly toxic and persistent pollutants, e.g. polychlorinated biphenyls

(PCBs) [13,14], and are relatively cheap in comparison to other oxidants [15] (Table 1).

Fig. 1. Annual number of a) publications concerning PDS or PMS and b) citations. Source:

Web of knowledge (data as of July 2017).

This paper reviews the chemical and physical properties of persulfates in addition to their

determination and activation methods. Moreover, the direct oxidation reactions of persulfates

as well as oxidation with generated radicals are described.

3
2. Chemical and physical properties of persulfates

Peroxydisulfuric acid (H2S2O8) was first detected by Marcelin Berthelot (French chemist) in

1878 [16]. It can be formed in the electrolysis process of sulfate salt. The resulting PDS salt is

almost non-hygroscopic and has a good longevity. PDS could be found in the form of three

salts i.e. sodium, potassium and ammonia. Potassium PDS has very low solubility for in situ

remediation, and the application of ammonium PDS can cause residual ammonia and

secondary contamination. Hence, the sodium PDS salt (Na2S2O8) is the first choice for in situ

chemical oxidation (ISCO) treatment. It has a high solubility of 730 g kg-1 H2O (at 25 oC)

[17,18].

PDS is cheap in comparison to other oxidants used in situ (0.74 USD/kg), but is still more

expensive than hydrogen peroxide for large-scale applications. The bond dissociation energy

and bond length of [O3SO-OSO3]2- were respectively determined to be 140 kJ mol-1 and 1.497

[15,16,19].

Table 1

Various properties of common oxidants used for in situ remediation.

O-O bond- Average


Solubility in
dissociation estimated Price [USD Price [USD
Oxidant water at
energy [kJ lifetime in kg -1]2 mol-1]
-1
25 C [g L ]
mol ] -1
groundwater4

H2O2 213 soluble hours to days 1.5 0.05

O3 364 0.1 <1 hour 2.3 0.11

Potassium - 0.7-3 3 unstable in water 2000 396

4
ferrate at pH 9

PDS5 926 730 >5 months 0.74 0.18

PMS 3776 298 hours to days 2.21 1.36

1
PMS is herein considered as the Oxone salt (2KHSO5KHSO4K2SO4);
2
Prices per kg are taken from: [15]; [20]; http://www.labmanager.com/news/2010/04/new-

battelle-process-slashes-price-of-useful-but-expensive-chemical?fw1pk=2#.Vu2SzOZxBqF;

[21]
3
[22]
4
Depends on water hardness, transition metal concentration, dissolved organic carbon (DOC)

concentration, and many other factors


5
Sodium PDS
6
[23]

The remaining data was taken from [16,19]

Peroxymonosulfate (PMS; monopersulfate) has its origin from the Caros acid (H2SO5) also

known as persulfuric, peroxymonosulfuric, and peroxysulfuric acid [24]. It possesses reactive

oxygen closest to the hydrogen atom that carries the active power (Fig. 2b). Since the oxygen

is readily reactive, KHSO5 is usually found in a more stable form of a white triple salt,

2KHSO5KHSO4K2SO4 (potassium hydrogen PMS). Commercially, the product goes under

the trade-name of Oxone and has a high solubility in water, is safe to handle, hence it has a

good potential applicability [25-27]. In comparison to PDS, PMS has a shorter bond length

(1.46 ), which translates to a higher bond dissociation energy (377 kJ mol-1). In other words,

in theory, PMS requires more energy to produce radicals in the homolytic cleavage of the

peroxide bond.

5
Fig. 2. (a) Sodium PDS and (b) Oxone triple salt molecular structure (*potassium PMS

(KHSO5) - active part of Oxone; bond lengths were taken from [28]).

The cost of Oxone is the highest of all of the conventionally used oxidants in ISCO, although

it is still much cheaper than ferrates [21].

Persulfates are very stable in the solid state and remain stable for several months in pure

water. One of the advantages of PDS in comparison to the other commonly used ISCO

reagents (hydrogen peroxide or ozone) is its much higher stability i.e. smaller transportation

issues; elevated concentrations of PDS can be pumped into the contaminated area, and it

could be moved by the density driven dispersion through the treated soil [12].

Yen et al. [14] determined that PDS anions can persist in the soil system for over five months.

PMS does not show such a high persistence as PDS, although when being stored in a dry and

cold place, it loses only around 1% activity/month due to the oxygen and heat release

(http://www2.dupont.com/Oxone/en_US/index.html). Under the influence of elevated

temperatures (>300 C) it decomposes to SO2 and SO3. PMS is also not stable in water with

6
higher pH values (at pH 9 the stability reaches minimum), whereby concentration of its

protonated form (HSO5-) equals the concentration of the unprotonated one (SO52-) [29]. More

described information on the stability of PMS at various pH values can be found in Bouchard

et al. [29].

Peroxydisulfate and peroxymonosulfate are among the strongest oxidizing agents that are

applied in environmental remediation. The standard oxidation-reduction potential (ORP) for

the reduction of PDS anion (Eq. 1) equals to 2.01 V, hence it is higher than that of PMS i.e.

1.4 V (Eq. 2) [30,31].

S2O82- + 2H+ + 2e- 2HSO4- (1)

HSO5- + 2H+ + 2e- HSO4- + H2O (2)

In case persulfates are used for pollutant degradation, the radicals formed upon cleavage of

the peroxide bond are the most important. The radical formation can be initiated by different

means, i.e., photochemical or thermal cleavage of the peroxide bond or chemical reduction.

PDS typically decays in a radical pathways often resulting in sulfate radicals. However, PMS

may also decompose into sulfate radical and hydroxyl radical [32].

3. Activation mechanism

The key in PDS and PMS-based oxidation is formation of highly reactive species, which

themselves have a potential to degrade pollutants. This can be achieved mainly by thermal

[33-36], photolytic [37], sonolytic [38], radiolytic activation [39], as well as by the reactions

of PDS and PMS with iron oxide magnetic composites [40,41] including as well in situ

formed iron hydroxides [42] and quinones [43]. Furthermore, activation of persulfates by

hydroxide ions [44,45] in ozonation [46] or in presence of phenol [47] was described.

7
Moreover, sulfate radicals could be also generated by chemical reduction of PMS and PDS

using low valent transition metals as reductant [48,49] including also bimetallic or trimetallic

iron-based systems as well as iron scrap [50,51]. Photo and radiodissociation or thermal

activation of PMS precedes the evolution of both SO4- and OH [49,52]. The formation of

SO4- as the main oxidizing species can be performed only via transition-metal catalysis

(although the interchange of radicals in water solutions can be also notable and will be further

discussed in section 5.3). In the group of the transition element ions (Fe2+, Co 2+, Ni2+, Cu 2+,

Ru 3+, Ag+), divalent cobalt was determined to be the most efficient catalyst for PMS

according to Anipsitakis and Dionysiou [53] and Fernandez et al. [54]. As for PDS, similarly

to the Fenton reaction, Fe2+ is the most commonly used metal for the homogeneous catalysis.

The type and dose of the transition metal catalysis is of significant importance, since when it

is applied in excess, scavenging of sulfate radicals may become a problem [55], this problem

however can be solved by adopting the use of solid iron particles where the release of iron

species responsible for PDS activation occurred smoothly without the risk of sulfate radical

quenching as reported by Naim and Ghauch [51]. Ayoub and Ghauch [50] demonstrated as

well that PDS activation can be better sustained in solution especially while using bimetallic

and trimetallic iron-based systems making the process more efficient toward long term

application.

A detailed summary of the activation methods of PDS and PMS is presented in Tables 2 and

3, respectively.

8
Table 2

PDS activation methods.

Method Mechanism Predominant radical species Comments Reference

Sulfate radical/Hydroxyl radical Because of the low bond-dissociation energy, often

Heat Homolysis of peroxide bond (in presence of chloride / at low temperature increase can effectively cleave O-O [35,36,56]

elevated pH) bond

Often used =254 nm, Quantum yield 1.4 (depends


UV radiation Homolysis of peroxide bond Sulfate radical
on dissolved oxygen concentration) [57-59]

Homogenous:
One electron transfer Sulfate radical Often requires low pH [48]
Transition metals

Heterogeneous:

Transition metals [50,60]


One electron transfer Sulfate radical Preparation of catalyst is not economical
(mono, bi, and

trimetallic systems)

Chelated transition Stabilizes ferrous iron at neutral pH, widely used in


One electron transfer Sulfate radical [61]
metals situ

Base-catalysed hydrolysis of PDS to Sulfate radical/Hydroxyl


Alkaline pH Often pH>11 [44]
hydroperoxide, which later initiates radical/Superoxide radical

9
radical formation

Additional Fe3+ reduction on the cathode. Yuan et al.


Ferrous iron generated in the chemical
Electrolysis Sulfate radical/Hydroxyl radical [62] claim that OH radical contribution in this system [62]
and electrochemical corrosion of iron
is more significant

Peroxide bond of PDS is debilitated on

the defective edges and oxygen Reduced mesoporous carbon, carbon nanotubes,

groups (from which the carbonyl group and graphene oxide, displayed great catalytic
Nanocarbons Hydroxyl radical [15,63]
was found as the most active one) of properties, in contrary to nanodiamonds, fullerenes

carbocatalysts, than persulfate oxidizes and graphitic carbon nitride

adsorbed water or hydroxide ions

Low molecular weight, anionic


Other organics One electron transfer Sulfate radical [43,47]
organic compounds

Radiolysis Reaction with the solvated electron Sulfate radical Electron beam irradiation of aqueous solution [39]

Table 3

PMS activation methods.

Method Mechanism Predominant radical Comments Reference

10
species

Sulfate radical/ Higher temperatures are required to split the O-O bond, due to
Heat Homolysis of peroxide bond [49]
Hydroxyl radical a higher bond-dissociation energy in comparison to S2O82-

Sulfate radical/
UV radiation Homolysis of peroxide bond Often used =254 nm, Quantum yield 0.52 [49,52]
Hydroxyl radical

Homogenous: Often requires a low pH to have the metals in a desirable


One electron transfer Sulfate radical [49]
Transition metals oxidation state

Heterogeneous: [60]
One electron transfer Sulfate radical Preparation of catalyst is not economical
Transition metals

Base-catalysed hydrolysis of PMS to


Alkaline pH Superoxide radical [45]
hydrogen peroxide

One electron transfer from The contaminant degradation rate by an oxidant assisted with

Electrolysis electrochemically/chemically Sulfate radical electrolysis was observed in the following order: PMS > PDS [64]

produced Fe2+ > H2O2

Graphene demonstrated better catalytic performance than few

Nanocarbons One electron transfer Sulfate radical other allotropes of carbon, like: activated carbon (AC), [65]

graphite, graphene oxide, and carbon nanotubes

Sulfate radical/
Organics One electron transfer Polyimide as an electron donor [66]
Hydroxyl radical

11
Formation of an O3SO5 adduct Sulfate radical/
Ozone - [46,52]
that decomposes into radicals Hydroxyl radical

Radiolysis Homolysis of peroxide bond Sulfate radical - [49]

12
The application of PMS and PDS in oxidation treatment belongs to advanced oxidation

processes (AOPs) because of the free radicals presence [67].

AOPs can be either homogeneous or heterogeneous. Whilst homogenous AOPs are the most

efficacious at an acidic pH, heterogeneous advanced oxidation processes could function in a

wide pH spectrum [68,69]. Despite the fact that heterogeneous AOPs exhibit numerous

benefits in comparison to homogenous ones, the point that the catalyst creation is not easy nor

economical should also be considered [60] (however there are some exceptions of this rule

[51]).

3.1 Homogeneous activation process

In the case of ISCO, activation should be slow to allow the long-term formation of radicals

and the treatment of heterogeneously distributed contaminants. For homogenous radical based

processes, one of the most common methods is to use chelated transition metals, which

significantly decreases the metal concentration required for the activation [70,71]. Rastogi et

al. [72] assessed the efficacy of several chelating agents i.e. ethylenediaminedisuccinate,

citrate, and pyrophosphate on ferrous iron activation of PMS and PDS under neutral pH

conditions. They found that PMS was the all-embracing oxidizing agent catalysed by all

ferrous iron-chelates, whereas citrate coupled with ferrous iron was the best chelate/iron

configuration.

3.2 Heterogeneous activation processes with metal catalysts

On the other hand, heterogeneous processes with material (often in nano-scale) synthesized

beforehand, have drawn much attention recently due to their excellent performance. In

addition, as mentioned above, cobalt has been found to be the very best homogenous activator

for PMS. Therefore, heterogeneous catalysts often also contain cobalt (especially the ones that

13
are used for peroxymonosulfate activation). Co 3O4 was used for instance by Chen et al. [73]

and Muhammad et al. [74] as a catalyst for PMS to degrade Acid Orange 7 (AO7) and phenol,

respectively (see also [75-78]). Interestingly, Zhang et al. [79] synthesized nanoscale Co3O4

and applied it in heterogeneous activation of PDS. The highest degradation rate of Orange G

(as a model compound) was at pH ~ 7, where the catalyst dissolution is very low. The authors

also confirmed that the hydroxyl and sulfate radicals were the main oxidants.

One of the major problems of the activation of persulfates with Co3O4 is an excessive quantity

of cobalt leaching. To improve catalytic capacity and decrease leaching, cobalt can be

attached to several supports (most often metal oxide) i.e. Al2O3, TiO2, SiO2, MgO or activated

carbon/carbon aerogel or graphene. For example, Yang et al. [78] confirmed that the TiO2

support increased the content of OH groups on the Co/TiO2 catalysts, enhancing the

generation of the Co-OH complex, that is believed to be a crucial stage in heterogeneous

activation of PMS (Eq. 3).

CoOH+ + HSO5- CoO+ + SO4- + H2O (3)

In another study, Yao et al. [80] used bimetallic oxides CoFe2O4 and CoFe2O4- on graphene to

catalyse PMS and thus abate phenol. Su et al. [81] successfully synthesized heterogeneous

Co xFe3-xO4 material and discovered that the iron-cobalt interactions are vital for effective

heterogeneous catalysis of PMS. Lin et al. [82] demonstrated an alternative path to synthesize

a magnetic nano-Co-graphene composite from carbonizing of a cobalt-based metal organic

frameworks and graphene oxide, and used it to activate PMS. In order to measure the long-

run catalytic activity, a 50-cycle decolorization of Acid yellow 17 was performed and the

efficacy of degradation remained at 97.6%, displaying its stability and efficient catalytic

activity. In their further study, a nanocomposite was prepared by the carbonization (one-step)

of a Co-based zeolitic imidazolate [83]. Higher cobalt/carbon nanocomposite loading, PMS

dose, acidic conditions and temperature greatly ameliorated the degradation of caffeine.

14
Following the change of the reaction rate after addition of different radical scavengers (i.e.,

methanol (MeOH) and tert-Butyl alcohol (TBA)), it was shown that the process involves

rather sulfate radicals than hydroxyl radicals. Shi et al. [84] demonstrated Co3O4/expanded

graphite, developed in a solvothermal synthesis, as an extremely effective heterogeneous

catalyst of PMS. They concluded that the abatement of Orange II in water is due to SO4-, and

100% removal can be reached after 8 minutes. Ding et al. [85] used cobalt and bismuth salts

and sodium hydroxide (as a precipitation agent) to create the Co3O4-Bi2O3 nanocomposite as a

heterogeneous catalyst for PMS. The nanocomposite demonstrated substantial catalytic

performance towards PMS for the removal of organic contaminants. Moreover, leaching of

cobalt was decreased to 43 g L-1, which is an improvement when compared to that of Co3O4

(158 g L-1) under equivalent testing conditions. Yang et al. [76] claimed that the

heterogeneous oxide CoFe2O4 was capable to steadily catalyse PMS for the abatement of 2,4-

dichlorophenol at pH 7. Furthermore, in this system, leaching of cobalt was reduced to a very

low extent. However, such a reduction in cobalt leaching requires the pH of the solution to be

close to 7.0, which is problematic in the applicability for treatment of water or wastewater due

to substantial acidification during activation [86]. Hence, it is still an important challenge to

develop an efficacious cobalt-carrying heterogeneous catalyst for PMS without causing a

secondary contamination. One way to overcome this problem was proposed by Rhadfi et al.

[87], who determined that the partial replacement of cobalt in Co 3O4 with Mn can be a

strategy for decreasing the quantity of the harmful Co element. Mn is more abundant in

nature, more environmentally friendly, and is twenty times cheaper than cobalt [88]. Wang et

al. [89] provided an example of the PMS catalytic mechanism using an iron oxide/C core shell

magnetic nanosphere supported by manganese oxide nanoparticle. They found that SO4- was

the primary radical for phenol degradation. The same authors also attempted to use 3D-

hierarchically structured MnO2 to activate PMS [90]. The catalytic mechanism of PMS

15
activation was determined therein by electron paramagnetic resonance (EPR (also known as

electron spin resonance, ESR)) spectra and showed that hydroxyl and sulfate radicals are

simultaneously produced in the activation processes, however SO4- played a more crucial role

in oxidation of carbolic acid. Similarly, Li et al. [91] synthesised three trivalent Mn

(oxyhydr)oxides, namely bixbyite, hausmannite and manganite, from which only manganite

showed good catalytic activity for PDS activation. Analogous to Wang et al. [90], it was also

shown that both SO4- and OH are present during the treatment, although the sulfate radicals

were selected as the predominant oxidative species in phenol oxidation.

Similarly to the homogeneous systems, iron-based compounds are frequently used for

persulfate activation due to their lower toxicity compared to cobalt. Tan et al. [92] reported a

nano-Fe3O4 catalyst in PMS activation process, whose stability decreased significantly from

the first to the third run. In another study, authors of a recent paper concerning heterogeneous

catalysis [60], fabricated CuFe2O4-Fe2O3 for the activation of peroxymonosulfate and

subsequent degradation of bisphenol A [93]. The performance of CuFe2O4-Fe2O3 was

compared with the alternative heterogeneous catalytic materials and shown to be superior in

comparison to CuFe2O4, CoFe2O4, CuBi2O4, CuAl2O4, Fe2O3, and MnFe2O4 (whereas the

catalyst order indicates their activity towards PMS). Moreover, it is possible to reuse

CuFe2O4-Fe2O3 catalyst (at least several times) without significant performance loss. In

addition, Zhang et al. [94] observed that the property of nano-Fe3O4 to activated PMS can be

recycled by hydroxylamine. Furthermore, the catalytic ability of Fe3O4 coupled with

hydroxylamine soared gradually in the ten consecutive runs.

Ghauchs group by working on bi and trimetallic systems has significantly contributed to the

heterogeneous catalyst in the field of peroxydisulfate [50]. They have determined inter alia

that the iron scrap of special properties can reduce the release of metal ions in solution to a

minimum and the repeated experiments did not show any efficiency loss [51]. Moreover, their

16
research on the reaction stoichiometric efficiencies (RSE) allows to better understand and

compare the reaction mechanism of different activation methods and its correlation to the

mineralization extent of the organic contaminant [95].

3.3 Heterogeneous activation with metal-free catalysts

Recently, catalytic materials with no metal content are gaining popularity due to the many

advantages, i.e. lack of secondary pollution, chemical stability, and that generally they are

considered as environmentally compatible [66]. It was confirmed that the functional groups of

carbonaceous materials that possess oxygen (especially the carbonyl group) can effectively

contribute to the activation of persulfates [96]. From the metal-free heterogeneous catalyst,

Sun et al. [65] found that the reduced graphene oxide (rGO) can serve as an effective

heterogeneous catalyst to activate PMS in order to form radicals. Zhang et al. [97] used

granular activated carbon (AC) as a green catalytic material for PMS to degrade Acid

Orange 7 in an aqueous solution. Also, Saputra et al. [98] reported that AC powder can be an

environmentally friendly catalyst for the efficient activation of PMS that later exhibited

excellent potential for phenol degradation mediated by sulfate radicals.

Lee et al. [63] discovered that carbon nanotubes could catalyse persulfates to form species

that are more reactive. Similarly, Duan et al. [15] evaluated the ability of various nanocarbons

to activate PDS for degradation of phenolic compounds, dyes, and intermediates formed

during their treatment process. As PDS activators, single wall carbon nanotubes, rGO, and

mesoporous carbon displayed great catalytic performances, while nanodiamonds, fullerenes,

and graphitic carbon nitride showed lower efficiencies. Furthermore, the carbo-catalysts

manifested much higher activity towards PDS activation in comparison to the universally

applied AC and metal oxides, i.e.: Fe3O4, Co3O4, MnO2, and CuO.

In a recent study, Andrew Lin and Zhang [99] showed that not only carbons can act as metal-

free peroxymonosulfate activators. They have demonstrated that the use of orthorhombic -

17
sulfur as a metal-free photo-catalyst for PMS (under irradiation with visible light) is efficient,

promising, and friendly for the environment.

Considering the fact that large amounts of solid waste that could be used for the activation of

persulfates are produced from various industries, there is no need to synthesize expensive

catalysis. Fly ashes derived from coal, biomass or oil combustion are major contributors to the

production of solid waste and could be used to prepare (Co)-based catalysts for PMS

activation [74]. It was found that the fly ash does not adsorb phenol, however it consists of

cobalt oxide, which can be used for the activation of PMS. In other studies, steel waste

powder [100] and the iron scrap from car rotary disc [51] were applied as activators for PDS.

The iron scrap should be cleaned under dry process where a layer of 3 microns is removed.

This layer contains excellent iron alloys formed during the friction process between the pad

and the disc. This material is widely available and not expensive.

Electron donor catalysis has various benefits but can also, as was stated before, become a new

source of contamination. In addition, it is often much less efficient, e.g. only one mole of

radical can be obtained from one mole of oxidant, in contrast to UV or heat activation. A

detailed study of the activation of oxidants with O-O bond by UV and elevated temperature

was performed by Yang et al. [19]. They found that the order for the decontamination efficacy

of peroxides activated by heat is: peroxydisulfate > peroxymonosulfate > hydrogen peroxide

and of UV (254 nm) activated peroxides: peroxydisulfate > hydrogen peroxide >

peroxymonosulfate (conditions: experiments were done in double distilled water;

concentration of AO7 (contaminant of concern) - 20 mg L-1; the molar ratio of hydrogen

peroxide/AO7 - 10/1). Surprisingly, they also determined that relatively common (in ground,

surface, and wastewaters) anions (i.e. HCO3-, HPO42-, Cl-, and CO32-) can initiate formation of

more reactive species from PMS. It has to be mentioned, that the efficiency of the above

processes are very dependent on the water matrix. In natural matrices the presence of chloride

18
and HCO3-/CO32- can greatly decrease the efficiency of sulfate radical based processes [101]

(however, it depends on the probe being treated; see below, section Matrix effects).

3.4 Alkaline activation

Another persulfate activation technique often used in situ is alkaline activation involving an

increase in pH (often >11) by dosing sodium hydroxide (NaOH) or potassium hydroxide

(KOH) solutions [12]. On the other hand, Cassidy et al. [102] concluded that in situ

stabilization amendments, including fields containing Ca(OH)2 and/or CaO, can effectively

activate PDS by increasing the pH, and in the case of CaO also the heat upon their reaction

with soil water (i.e. additional activation). Activated PDS decreased concentrations of BTEX

compounds and PAHs. The proposed mechanism of alkaline activation of PDS [44] relies on

the hydrolysis of PDS to hydrogen peroxide anion (HO2-) and subsequent reduction of PDS

by this anion with the production of sulfate and superoxide radicals (O2-). Very recently, Qi et

al. [45] proved a similar activation tendency for PMS, which can take place at high pH with

base-catalysed hydrolysis of PMS to hydrogen peroxide. It was discovered that O2 - was the

predominant radical species in this system, but the role of singlet oxygen was also found to be

significant.

3.5 Electro-activation

Several studies have evaluated the electro-activation of persulfates. Yuan et al. [62],

investigated the activation of PDS using ferrous iron produced in an electrolytic system. They

have determined that PDS is mainly decomposed by Fe2+ created in two steps (1) from the

corrosion of Fe0 (chemical and electro-chemical) and (2) from the cathodic reduction of Fe3+.

They also found OH to be the predominant radical in their system. Recently, Govindan et al.

[64] proved the efficacy of electrochemical activation for persulfates. Electrochemically

19
activated PMS was the best oxidant (from PDS, PMS and hydrogen peroxide) for

pentachlorophenol decontamination. Long and Zhang [103] stated that the electro-activation

of PDS can be effective for degradation of methylbenzene (toluene) from a surfactant flushing

solution. The results indicated that in the reduction process of ferric iron on the cathode, there

could be produced ferrous iron that could further activate PDS. This is in agreement with the

research work (mentioned before in this subsection) of Yuan et al. [62]. Also, simultaneous

use of electrolysis and UV process/goethite can enhance the radical formation from PMS and

PDS (Jaafarzadeh et al. [104]; Lin et al. [105]). According to Lin et al. [105], during the

electrolysis, goethite (-FeOOH) can serve as an additional catalyst for PDS due to

continuous generation of ferrous iron by the reduction and transformation of ferric iron on the

surface of goethite. Whereas Jaafarzadeh et al. [104] concluded that only with the

simultaneous use of electrolysis and UV, PMS (and electro-generated H2O2) can effectively

decolorize acid brown 14. Moreover, the great efficacy of electro-activation of PDS was

shown in Wacawek et al. [55]. This study has shown that pseudo first-order reaction rate

constant of lindane removal was (at the optimal conditions) 0.04 min-1, which is higher than

the one observed by Cao et al. [106] - 0.0083 min-1 (in the PDS/Fe2+ system) but lower than

that presented by Khan et al. [107] - 0.124 min-1 (in the PMS/Fe2+/UV system).

3.6 Other activation methods

Recent work reported by Cong et al. [108], focusing on the simultaneous use of PMS and

ozone for degradation of 4-chlorobenzoic acid, indicated that PMS may act in a similar way to

H2O2 in the promotion of OH production in ozonation. This theory was confirmed by Yang et

al. [46], who demonstrated that the reaction between PMS and ozone is chiefly responsible for

promoting ozone disappearance with a determined second-order rate constant of 2 104 M-1 s-
1
. Both sulfate and hydroxyl radicals were present in this system, which was confirmed by

20
chemical probes and their yields were found to be 0.43 0.1 and 0.45 0.1 per mol of ozone,

for SO4- and OH, respectively. The first step of the reaction is assigned to formation of an

adduct (-O3SOO- + O3 -O3SO5-), which can further decompose into more reactive radicals

(SO5- and O3-). The following reaction of SO5- with ozone is assumed to produce SO4-,

while O3- can convert to OH [8].

4. Determination methods

Numerous determination methods for PDS and PMS can be found in the literature. These

methods vary in their execution time and sensitivity of detection (see Table 4 and 5).

Tables 4 and 5 present a summary of PDS and PMS determination methods, respectively.

Table 4

PDS determination methods.

Time of
Method Substance used Limit of quantification Reference
measurement

Direct UV-
Spectrophotometry - - [58,109]
absorption

Titration Ce4+/Fe2+ <20 min >10-4 M [110]

Spectrophotometry KI/HCO3- 15 min 7 x 10-2 M [111]

Spectrophotometry H2SO4/NH4SCN ~40 min 2.1 x 10-4 M [112]

Methylene blue

Spectrophotometry under microwave 1 min 3 x 10-6 M [113]

radiation

N,N-diethyl-p-
Spectrophotometry 10 min 10-5 M [114]
phenylenediamine

Spectrophotometry Alcian blue 120 min 8.8 x 10-8 M [115]

Amperometry Electrodeposited - 10-5 M [116]

21
poly-brilliant

cresyl blue on

carbon nanotube

Electrode (glassy

carbon) with a

nanocomposite

Amperometry containing - 10-5 M [117]

ruthenium oxide

nanoparticles and

thionine

Electrode (glassy

carbon) with a

nanocomposite

Amperometry containing - 10-5 M [117]

ruthenium oxide

nanoparticles and

celestine blue

Prussian blue -

modified
Voltammetry - 5 x 10-5 M [118]
platinum disc

electrode

Mobile phase: 50

Ion chromatography mM KOH; flow 18 min 5.2 x 10-8 M [119]

rate: 1 mL/min

Table 5

PMS determination methods.

Time of Limit of
Method Substance used Reference
measurement quantification

22
Titration Ce4+/Fe2+ <20 min 10-5 M [110]

Spectrophotometry KI/HCO3- 5 min 1.2 x 10-5 M [120]

Spectrophotometry Co2+/methyl orange 1 min 5 x 10-7 M [121]

N,N-diethyl-p-
Spectrophotometry 10 min 10-5 M [114]
phenylenediamine

Stationary phase: Super-Sep

anion-exchange column;

Mobile phase: 2 mM phthalic

acid with 5% (v/v)


Ion chromatography 11 min 4.5 x 10-5 M [122]
acetonitrile content, pH 3.0;

flow

rate 1.5 mL/min;

cond. 290 S/cm

Stationary phase: Varian

Polaris 3 C-18 A column;

Mobile phase:

HPLC methanol/phosphoric acid - ~5 x 10-7 M [123]

buffer; flow

rate 1.0 mL/min; UV at 260

nm

The first methods used for the determination of PDS and PMS were most probably iodometric

titrations [16]. Their spectrophotometric alternatives possess much higher sensitivity and do

not require as much reagent and time [111,120]. Determination of persulfates with the use of

dyes has gained popularity lately, although the method has been known for a long time [115].

Nevertheless, these methods possess very high sensitivity and short measurement time. More

information on the determination of persulfates with organic dyes can be found in Ding et al.

[124] and Zhang et al. [125]. In addition, the fast and accurate analysis of both PDS and PMS

can be performed with the use of ion chromatography and HPLC [119,122,123].

23
5. Persulfate decontamination technologies

5.1 Direct oxidation

Although reactions of non-catalysed persulfates occur at rates that are often slow, several

studies reported their direct oxidation processes. Probably the best known are the Elbs and

Boyland-Sims reactions, which rely on the nucleophilic displacement of peroxide oxygen

from the PDS ion [126]. In the Elbs reaction, phenolate anion (or a tautomer) acts as the

nucleophile, whereas in the Boyland-Sims reaction, the nucleophile is a neutral aromatic

amine. In both of these processes, there is no radical involvement (apart from the side

reactions).

In addition, during catalysed persulfate oxidation, persulfates can react with generated

radicals in several ways.

In Figure 3 the second-order reaction rate constants of persulfate reactions with several

species in an aqueous phase, can be observed.

24
Fig. 3. Comparison of the second-order reaction rate constants of persulfates with aqueous

species. Data taken from: Yang et al. [46]; Buxton et al. [127]; Restelli and Angeletti [128];

Herrmann et al. [57]; Davies et al. [129]; Roebke et al. [130]; Maruthamuthu and Neta [131];

Gilbert and Stell [132].

Further example of PDS oxidation not relying on the radical formation was presented by

Zhang et al. [21]. According to them, PDS weakly interacts with the surface of copper oxide

(under mild conditions) from which it can easily react with the contaminant of concern (2,4-

dichlorophenol). The reaction between PDS and the surface of CuO was assumed therein to

be the rate-limiting step.

Furthermore, Lei et al. [133] have shown that not only PDS can efficiently oxidize target

pollutants without the free radicals involvement. They have discovered that PMS can directly

react with cationic pigments without catalyst and in a broad pH range (2-12). Furthermore,

25
they concluded that Cl- anions improved the degradation efficacy of the target pollutants.

Radical quenching experiments and ESR studies revealed that the degradation of cationic

dyes by peroxymonosulfate does not rely on hydroxyl or sulfate radicals. The proposed

mechanism of decolorization relies on the complex formation between the dyes and PMS

followed by the electron transfer towards the PMS.

In addition, PMS may oxidize As(III) without an external activation as suggested by Wang et

al. [134]. They stated that the PMS can completely oxidize As(III) to As(V) within 24 hours,

which was possible even after addition of radical scavengers in high concentrations (1.6 M of

methanol).

5.2 Basic reactions in water matrices

Electron transfer reactions of oxygen yield reactive oxygen radicals [135]. Also, oxygens

role in SO4- systems was determined to be significant and was described carefully in a recent

paper by Xu et al. [136] and Ghauch et al. [42]. The step-by-step reduction of an oxygen

molecule to water followed by the formation of radicals during persulfate decomposition is

shown in Figure 4.

26
Fig. 4. Black arrows - scheme of oxygen molecule reduction to water; red arrows - radical

formation and behaviour in a persulfate system (dashed arrow - only valid for homolysis of

the PMS peroxide bond).

Superoxide disproportion may result in the formation of H2O2 which in turn can react to

generate OH (e.g., by chemical reduction). Furthermore, OH can also be generated together

with SO4- in reactions of PDS or PMS with transition metals [49] or in the photolysis of PMS

[137] and SO4- can also form OH in the reaction with OH- (k = 1.4 107 M-1 s-1, Herrmann et

al. [57], and reaction 4). Considering the concentration of OH- at normal conditions of water

and wastewater treatment (e.g., pH 6-8) the observed reaction rate constant of the reaction of

SO4- with OH- is in the range of 10 -1 to 10 2 M-1 s-1. However, reactions of SO4- with other

matrix components such as chloride or dissolved organic compounds are likely more

important at that pH-values (for relevant reaction rate constants see Table 6). Studies based on

27
radical scavenging analyses paired with the use of the ESR technique were performed to

identify the dominant radical species involved in persulfate-based oxidation [56,89,90].

Several compounds can be used for scavenging of SO4 and OH. For choosing the correct

scavenging agent, one has to consider reaction kinetics and product formation. It has to be

provided that only reactions under study are relevant and that products which could interfere

the reactions under study or analytical procedures are not formed.

TBA (khydroxyl radical = 6.0 10 8 M1 s1, ksulfate radical = 4.0 10 5 M1 s1) and ethanol (khydroxyl

radical = 1.2 10 9 M1 s1, ksulfate radical = 1.6 10 7 M1 s1) are the most often used chemical

probes to evaluate the relative contribution of SO4 and OH because of their different

reaction rate constants [138]. Also with the ESR technique, the radicals can be determined

qualitatively (rough estimation of the intensity of radicals in the spectra). The spin trap

compound, 5,5- dimethyl-1-pyrroline N-oxide (DMPO) reacts with radical intermediates in

oxidation systems and forms a unique radical adduct that is stable, relative to the radicals of

interest.

Indeed with these techniques, it was shown that in most cases SO4- prevails in acidic media

and OH predominates in alkaline media (e.g. tested at the pH of 5 and 11 by Fang et al.

[139,140] (at the pH of 9 both radicals were in similar quantity [140]). Similar conclusions

were obtained by Liang and Su, [138] by using various probes: TBA, phenol, and

nitrobenzene).

SO4- + OH- SO42- + OH k = 1.4 10 7 M-1 s-1 [57] (4)

Reactions of SO4- are very different from reactions of OH. SO4- are prone to react by

electron transfer reactions while OH favours H-abstraction and addition reactions. This can

explain that reaction rate constants of SO4- with pollutants have a larger spread than OH,

since most pollutants provide functional groups which can be attacked by OH. For more

28
information see [127] and [141]. For a more detailed explanation of radical identification, e.g.

with the use of radical scavengers, see [71,138-140].

5.3 Matrix effects: scavenging and secondary oxidants

In persulfate based oxidation processes, pollutants are mainly degraded by reactions of SO4-

which are formed by the different activation methods (see above). As in all radical based

processes, these radicals are largely consumed by main constituents of the water matrix. The

reaction of oxidants with water matrix constituents can be divided into three categories:

1. Scavenging of oxidants

2. Formation of secondary reactive species

3. Formation of by-products

In order to define how fast the reaction proceeds, second order rate constants and for

simplification of the experimental procedure, pseudo first order rate constants are often

used.

Table 6 shows second-order reaction rate constants of OH and SO4- with common anions and

natural organic matter (NOM) present in real matrices.

Table 6

Second-order reaction rate constants of hydroxyl and sulfate radicals with common anions

and NOM.

Rate constants
Radical Compound Reference
(M-1 s-1)

OH Cl- 3.0-4.3 x 109 [142,143]

SO4- Cl- 1.3-6.6 x 108 [144-146]

OH Br- 1.9 x 109 [147]

29
SO4- Br- 3.5 x 109 [148]

OH HCO3- n x 107 [149]

SO4- HCO3- 2.6-9.1 x 106 [109,150]

OH CO32- 4 x 108 [149]

SO4- CO32- 4.1 x 106 [151]

OH Humic acid 1.4 104* [152]

SO4- Humic acid 6.8 103* [152]

*(mg of C L-1)-1 s-1

Scavenging is mainly driven by dissolved organic matter (DOM), HCO3- and CO32- and in

some cases nitrite and reduced metal species such as Fe2+. DOM and reduced metal species

mainly consume oxidation capacity because the products formed in these reactions are not

reactive towards pollutants. For the reaction of the sulfate radical and hydroxyl radical with

humic acids the kinetic rate constant was measured to be 6.8 10 3 L mg C-1 s-1 and 1.4 104

L mg C-1 s-1 (mg C = mg carbon), respectively [152]. On the contrary, Luo et al. [32] observed

that the oxidation capacity of the hydroxyl radical and sulfate radical for humic acid

degradation (in the presence of Cl- and Br-) was nearly in the same order. This can be

explained by the below discussed conversion of SO4- into OH in presence of chloride. Hence

in both systems the same oxidant prevails, i.e., the hydroxyl radical. Especially Br- had a

negative impact on the efficiency of these processes, which is in agreement with the research

work of Yang et al. [153] (Although, in several studies a concentration of 1 mM of Br- and/or

Cl- seemed to enhance the removal of some compounds [51,59])

Electron transfer reactions of SO4- with phenolic moieties of DOM result in radical cations

which may rearranged in presence of water into carbon centred radicals in analogy to the

reaction of benzenes with SO4 - [154]. Carbon centred radicals react rapidly with oxygen ( 2

109 M-1 s-1 [155]). The ensuing peroxyl radicals eliminate O2- or HO2 or decompose in

30
bimolecular reactions [155]. Both pathways often result in carbonyls [155]. Halide ions also

react with SO4-. The reactivity of bromide towards SO4- is high (k = 3.5 109 M-1 s-1 [148])

and in pure water bromate is formed as a stable product [156], which is undesired due to its

carcinogenic potential (EU and US-EPA drinking water standard 10 g L-1). The presence of

organic matter, however, suppresses formation of bromate, indicating that no bromate is

formed in field water matrices [156]. Fang and Shang [157] came to a similar conclusion. Liu

et al. [158] investigated the transformation of Br- in a PMS (activated by cobalt) oxidation

process with the presence of phenol as a model compound imitating NOM. It was determined

that bromide can be efficiently converted to free bromine and bromine radicals. These

reactive species caused bromination of phenol and the formation of brominated disinfection

products including bromoform and bromoacetic acids and bromophenols as intermediates.

Brominated disinfection by-products were also degraded by surplus SO4-. Free bromine was

also created when cobalt was not present, indicating that Br- might be directly oxidized by

PMS. In another study these conclusions were confirmed and it was also found that the

brominated intermediates cannot be degraded in the absence of SO4- [159]. Similar work was

performed by Lu et al. [160], who determined that reactive bromine species formed in the

reaction with sulfate radicals can react with NOM and might form brominated products,

including brominated disinfection by-products. In addition, bromoacetic acid and bromoform

were formed in the presence of humic acids.

Iodide also reacts fast with SO4-, which may form iodate in analogy to bromide.

The products of the reaction of SO4- with chloride can result in chlorate formation in an H+

catalysed reaction [101]. However, at typical conditions of water treatment (pH > 7) no

chlorate formation was observed in UV/S2O82- [101]. The primary step in the reaction of SO4-

with chloride is an electron transfer yielding chlorine (Cl) atoms and sulfate. Cl forms a

H2OCl complex in water [146,161]. This complex can deprotonate and the resulting HOCl-

31
is in equilibrium with OH [101,146,161]. At pH 7 and in presence of 1 mM chloride, the

reaction of SO4- with chloride yields OH with a nearly 100% yield [101]. In case of chloride

concentrations in the lower mM range most of SO4- are scavenged by reactions with chloride

resulting in a conventional OH based AOP [101]. CO32- or HCO3- react fast with Cl and

dichloride (Cl2-) and thus, interrupt the OH forming reactions [101]. Since typical natural

waters contain chloride and HCO3-, the OH yield is likely much below the SO4- yield [101].

However, oxidation of HCO3- and CO32- gives rise to CO3-, which also exhibits a potential to

degrade pollutants (depending on their chemical structure and affinity with carbonate

radicals). Table 7 compiles rate constants of CO3- with several organic compounds. The

reaction rates are in the range of 106 109 M-1 s-1 and thus, on average slower compared to

SO4- or OH, which often react at second order rate constants in the range of 108 - 10 10 M-1 s-1

[149,162]. However, this may be compensated by the small reaction rate of CO3 - with the

scavengers in the water matrix, as will be explained below.

Table 7

Reaction rate constants of carbonate radicals with pollutants and matrix components and of

hydroxyl radicals and sulfate radicals with matrix components, apesticides.

Number Reactant Reaction rate constants Reference

Carbonate radical

DOC 40 and 280 L mg-1 s-1 [163,164]

1 Atrazinea 0.37 107 M-1 s-1 [164]

2 Fluometurona 0.4 0.3 107 M-1 s-1 [164]

3 Atratona 0.43 0.09 107 M-1 s-1 [164]

4 Tertbutryna 0.49 0.13 107 M-1 s-1 [164]

5 Fenurona 0.54-0.6 107 M-1 s-1 [164]

6 Prometryna 0.61 0.12 107 M-1 s-1 [164]

32
7 Irgarola 0.73 0.12 107 M-1 s-1 [164]

8 Ametryna 0.74 0.38 107 M-1 s-1 [164]

9 Diurona 0.8 0.2 107 M-1 s-1 [164]

10 Propanila 1.4 0.7 107 M-1 s-1 [164]

11 Monurona 1.5 0.4 107 M-1 s-1 [164]

12 Chlorotolurona 1.5-2.2 107 M-1 s-1 [164]

13 Isoproturona 2.5-3 107 M-1 s-1 [164]

14 4-cyanophenoxide 4.0 1.3 107 M-1 s-1 [164]

15 4-nitroaniline 7.7 3.4 107 M-1 s-1 [164]

16 Metoxurona 8.1-11 107 M-1 s-1 [164]

17 4-carboxyphenoxide 10.4 3.5 107 M-1 s-1 [164]

18 vanillinate 11.8 2.7 107 M-1 s-1 [164]

19 4-cyanoaniline 18 6 107 M-1 s-1 [164]

20 4-aminobenzenesulfonate 24 13 107 M-1 s-1 [164]

21 phenoxide 25 5.6 107 M-1 s-1 [164]

22 4-chlorophenoxide 35 12 107 M-1 s-1 [164]

23 3,4-dichloroaniline 41 9 107 M-1 s-1 [164]

24 4-hydroxyphenylacetate 60 20 107 M-1 s-1 [164]

25 4-chloroaniline 62 13 107 M-1 s-1 [164]

26 4-methylaniline 115 45 107 M-1 s-1 [164]

27 4-methylphenoxide 120 25 107 M-1 s-1 [164]

28 N,N-dimethylaniline 185 35 107 M-1 s-1 [164]

29 N-ethylaniline 220 40 107 M-1 s-1 [164]

30 N-methylaniline 255 45 107 M-1 s-1 [164]

31 aniline 255 18 61 13 107 M-1 s-1 [164]

32 4-methoxyphenoxide 310 60 107 M-1 s-1 [164]

Sulfate radical

DOC 6.8 103 L mg-1 s-1 [152]

HCO3- 9.1 106 M-1 s-1 [165]

Hydroxyl radical

33
DOC 2.5 104 L mg-1 s-1 [166]

HCO3- 8.5 106 M-1 s-1 [149]

The efficiency of pollutant degradation is depending on the scavenging rate of the water

matrix and the kinetics of the radical reaction with the pollutant. With the kinetics of

scavenger reactions and pollutant reactions at hand one can calculate the apparent degradation

rate (k) in a given water sample according to Katsoyiannis et al. [167]. Figure 5 shows k for

the compounds of Table 7 for reactions of CO3- assuming a water matrix with 1 mg L-1 DOC

and 1 mM HCO3-. k of CO3- is shown for the two reaction rate constants with DOC (Table

7). As will be explained below, HCO3- presumably does not contribute in scavenging of CO3-

. Figure 5 also shows k for the same water matrix in SO4- and OH reactions (dashed lines)

for a defined reaction kinetics towards trace compounds (dashed lines in Figure 5).

The herbicide propanil (compound 10, Table 7) reacts with CO3- at a reaction rate constant of

1.4 107 M-1 s-1, arriving at k = 0.25 s-1, in case CO3- reacts with DOC at a rate of 40 L mg-1

s-1. The same k is achieved in OH based processes for pollutants with a much higher reaction

rate constant (k 9 109 M-1 s-1) and in SO4- based reactions with k 1 10 9 M-1 s-1 (dashed

lines in Figure 5). This is due to fact, that OH and SO4- have a high reactivity towards DOC

(k(SO4-) = 6.8 103 L mg-1 s-1 [152], and k(OH ) = 2.5 104 L mg-1 s-1 [163] and towards

HCO3- (k(SO4-) = 2.8 9.6 10 6 M-1 s-1 [165], and k(OH) 1 107 M-1 s-1 [149]. The

reaction of CO3- with organic matter is considerably lower (see above). Furthermore, HCO3 -

or CO32- is not important for pollutant degradation, because it presumably yields CO3-.

Hence, the overall scavenger rate of CO3- is considerably lower compared to SO4-. In case

the CO3- reacts with DOC with a rate constant of 280 L mg-1 s-1 the efficiency of CO3- for

pollutant degradation drops down and CO3- has to react with rate constants > 108 M-1 s-1 to

yield a faster degradation kinetics compared to OH. Compound 18 (vanilinate) represents

34
such a compound, assuming that OH does not react faster than 9 109 M-1 s-1 with vanilinate.

In case of SO4- which react slower with DOC, CO3- has to react even faster with pollutants to

achieve comparable k. Pollutants with electron rich moieties such as anilines reveal a fast

reactivity towards CO3-. It is conceivable that the degradation of electron rich compounds is

enhanced by HCO3-/CO32- in SO4- or OH based processes. This applies for compounds 15 to

32 (Table 7, phenols and anilines). Most pesticides listed in Table 7, however, reveal reaction

rate constants of < 10 7 M-1 s-1. Their degradation is probably mitigated in presence of HCO3-

/CO32- in a SO4- or OH based process, as observed for atrazine degradation in river water

[101] (cf. further discussion [101,164]).

k(carbonate radical) / M-1 s-1


0.00E+00 1.00E+09 2.00E+09 3.00E+09 4.00E+09
5 100
32
4.5 29 30, 31
26, 27 28
50
4
19-25
3.5
0
3 16-18
k / s-1

k / s-1
2.5 -50
15
2
10-14 -100
1.5 1-9 k(SO4 -) = 9 109 M -1 s-1

1 k(SO4 -) = 5 109 M -1 s-1 -150


0.5
k(OH) = 9 109 M-1 s-1
0 -200
0.00E+00 5.00E+07 1.00E+08 1.50E+08
k(carbonate radical) / M-1 s-1

Fig. 5. Calculated apparent first order kinetics (k) of pollutants degradation by carbonate-,

sulfate- and hydroxyl radicals in presence of 1 mg L-1 DOC and 1 mM HCO3-. Calculations

are based on reaction rate constants shown in Table 7. Circles: k(carbonate radicals)

assuming: k(CO3 -+ DOC) = 40 L mg-1 s-1, squares k(carbonate radicals): assuming k(CO3 -+

35
DOC) = 40 L mg-1 s-1, dashed lines k(sulfate radicals or hydroxyl radicals, as indicated in the

figure) for a given reaction rate constant towards a pollutant, as indicated in the figure.

It was also reported that reactive chlorine species, formed in the reaction of SO4 with Cl,

may also react with other matrix component and thus, form chlorinated products [144]. For

example, Fang et al. [140] reported that the total concentration of radicals was greatly

increased in the presence of chloride ion but the degradation efficiency of PCBs in this system

decreased. Yang et al. [153] also investigated the conversion of OH and SO4- to the halogen

radicals. Halogens reduced the abatement efficacy of cyclohexanecarboxylic acid and benzoic

acid (in the presence of seawater), which were chosen as the target pollutants. It was also

concluded therein that the activated PDS was more affected by Cl- than the activated

hydrogen peroxide system probably due to the fact that Cl- has a higher reactivity with sulfate

radicals than the hydroxyl radicals at pH of 7. Indeed OH is hardly effected by Cl- at typical

pH values of water treatment (6-8) since the reaction of OH with Cl- has a fast back reaction.

The oxidation of Cl- by OH is acid catalysed and only becomes important in OH reactions at

pH < 3 [168]. The degradation efficiency of cyclohexanecarboxylic acid was not altered by

the halogens, probably due to the high reactivity of alkenes with halogen radicals. This

statement was further confirmed in a study of Liu et al. [169] who found that in the presence

of chloride there could be a different mechanisms of oxidation than that of the SO4-, and

carbon isotope fractionation of trichloroethene (TCE) was used to prove this statement.

In another work, Xie et al. [159] focused on the generation of chlorinated intermediates in a

sulfate radical system, and they found that the formation of carbonaceous disinfection by-

products, i.e. haloacetic acid and chloroform, only increased a little, but the generation of

nitrogenous disinfection intermediates, i.e. haloacetonitriles and trichloronitromethane,

slightly decreased. On the contrary, Lu et al. [170] found that after treatment of surface water

36
with 0.1 M PDS for 48 hours, caused increase in concentration of compounds such as:

chloroform, trichloroacetic acid, and dichloroacetic acid by 16%, 37%, and 52%, respectively.

The above results indicate that before the application of persulfates in the field, bench scale

studies should be performed in detail due to the fact that the formation of dangerous by-

products depends highly on the treated water matrix properties.

To sum up, at typical conditions of drinking water treatment (i.e., pH 6-9) presence of SO4-

scavenged by chloride are largely transformed into OH. Reactive chlorine species are prone

to be scavenged by carbonate or bicarbonate, resulting in CO3 -. The fraction of radicals

turned into OH or CO3- is controlled by kinetics and concentration of the corresponding

reaction partners (further details on that reaction system see above and Lutze et al. [101]).

CO3- may become important for pollutant degradation due to their higher selectivity

compared to SO4- and OH, provided that CO3- readily react with such pollutants (k 107 M-1

s-1 or faster) (see also Figure 5 and Table 7).

5.4 Post-treatment toxicity assessment

Although persulfates treatment has many benefits, there are also several downsides that have

to be taken into consideration, i.e. contamination with sulfate salts or even worse newly

created hazardous compounds due to e.g. additional chlorination and/or bromination (as

discussed in subsection 5.2.2).

In addition, many researchers study toxicity in clean systems, although in actual matrix

conditions, the background DOC and other constituents may additionally generate toxicity.

For instance, in our recent study concerning the decontamination of groundwater polluted

with various chlorinated olefins (Wacawek and Kudlek, unpublished data), many new

substances (with a molecular mass larger than 250 g mol-1) were observed after UV/persulfate

treatment that were unnoticed after e.g. UV/hydrogen peroxide treatment. Therefore, during

37
bench-scale testing performed before field application, toxicity tests (that are relevant to the

treated site) should be performed in order to avoid any environmental side effects.

However, several authors reported that the toxicity of the matrix after persulfates treatment is

significantly lower. One example was provided by Zhang et al. [171] that evaluated toxicity

with Vibrio qinghaiensis sp. Q67 test after the PDS treatment of carbamazepine and

concluded that the acute toxicity has decreased together with the removal of the

pharmaceutical (the inhibitory effect of the solution after treatment decreased to 65% within

60 minutes). Also Temiz et al. [172] evaluated the toxicological safety (using two different

bioassays) of the zero-valent iron/PDS oxidation of Triton X-45 (TX-45). Vibrio fischeri and

Pseudokirchneriella subcapitata bioassays were used therein and the toxicity profiles of the

treated matrices significantly decreased from an original value of 66% relative inhibition to

21% (Vibrio fischeri) and from 16% relative inhibition to non-toxic values

(Pseudokirchneriella subcapitata). Vibrio fischeri has shown higher sensitivity to TX-45 and

its degradation products than the microalgae Pseudokirchneriella subcapitata.

Olmez-Hanci et al. [173] went even further and used three toxicity tests (Vibrio fischeri,

Daphnia magna, and Pseudokirchneriella subcapitata) and the Yeast Estrogen Screen

bioassay to assess the possible estrogenic and toxic properties of nonionic surfactant

octylphenol ethoxylate and its oxidation products. In the case of Vibrio fischeri and Daphnia

magna tests the inhibitory effect of nonionic surfactant octylphenol ethoxylate dropped

considerably after the application of PMS. However, treatment with PMS/UV-C generated

oxidation products that possessed highly toxic effect towards Pseudokirchneriella subcapitata

(opposite observations to Temiz et al. [172], concerning toxicity of TX-45 and its oxidation

by-products).

38
5.5 Decontamination of water and wastewater with free radicals generated in persulfate

systems

As shown in section 3, numerous methods of persulfate activation can be applied in water and

wastewater treatment technology in order to achieve the expected results. The

decontamination of water (subsection 5.5.1) and wastewater (subsection 5.5.2) with the

radicals generated from persulfates is described in detail below.

5.5.1 Water

SO4- are very reactive and able to degrade very persistent organic compounds.

Perfluorooctanoic acid (PFOA) e.g., that is unreactive towards OH, can be degraded in the

SO4- system according to many [174-179]. Recently, Qian et al. [180] determined the

degradation kinetics of PFOA in a UV/PDS system and proposed its degradation mechanism,

which relies on a sequential loss of CF2 units from it and its intermediates [179,180].

However, the degradation kinetics of the reaction SO4- with perflurocarbonic acid are very

slow (approximately 10 4 M-1 s-1). Since the reaction of SO4- with other matrix components is

much faster, a degradation of perfluorinated compounds is not feasible in SO4- based

oxidation.

SO4- can readily oxidize other organic pollutants such as 2,4-dichlorophenol [75], 2-

chlorobiphenyl [181], aniline [24], bisphenol A [182], calcon [183], Acid Orange 7 [19],

hexachlorocyclohexanes [55,184], Ponceau S [185], 4-fluorophenol [186], pentafluorobenzoic

acid [187], C.I. Reactive Black 5 [188], C.I. Basic Red 46 [189], methylene blue [190],

endosulfan [191], antipyrine [192], naproxen [95], chloramphenicol [59], ranitidine [51],

sulfamethoxazole [42,50], bisoprolol [36], ibuprofen [35], dimethyl phthalate [193] or

dimethylhydrazine [194] and chlorotriazine pesticides. Table 8 includes pseudo first-order

rate constants of various contaminants with generated from persulfates radicals.

39
Table 8

Decontamination of common water pollutants with radicals generated in persulfate systems.

Group of Model Main oxidative Pseudo first-order reaction rate


Comments Reference
compounds contaminant species constant

Possibly SO4- (not Trace quantities of hexachloroethane 7.6 x 10-3 min-1 (activation temperature of
Perchloroethene [144]
identified) as an intermediate. 50 oC)

Possibly SO4- (not Trace quantities of hexachloroethane 4.7 x 10-3 min-1 (activation temperature of
Trichloroethene [144]
identified) as an intermediate. 50 oC)

Chlorinated Isomerization between trans and cis-DCE


Possibly SO4- (not 1.6 x 10-3 min-1 (activation temperature of
olefins cis-dichloroethene presumably after formation of a single bonded [144]
identified) 50 oC)
intermediate.

Isomerization between trans and cis-DCE


trans- Possibly SO4- (not 3 x 10-3 min-1 (activation temperature of
presumably after formation of a single bonded [144]
dichloroethene identified) 50 oC)
intermediate.

Possibly SO4- (not Of all the BTEX compounds studied, benzene was 9.5 x 102 day-1 (activation temperature of
Benzene [195]
identified) most resistant to PDS oxidation. 20 oC; Oxidant/BTEX molar ratio 100/1)
BTEXs
Possibly SO4- (not 23.2 x 102 day-1 (activation temperature of
Toluene - [195]
identified) 20 oC; Oxidant/BTEX molar ratio 100/1)

41
Possibly SO4- (not 14.5 x 102 day-1 (activation temperature of
Ethylbenzene - [195]
identified) 20 oC; Oxidant/BTEX molar ratio 100/1)

Possibly SO4- (not 21.9 x 102 day-1 (activation temperature of


Xylene - [195]
identified) 20 oC; Oxidant/BTEX molar ratio 100/1)

Possibly SO4- 0.140.16 min1


Greater total organic carbon (TOC) removal was
Phenol and/or OH (not (UV activation parameters: = 254 nm; [196]
observed at elevated pH (11).
identified) Oxidant/Phenol molar ratio 168/1)

Phenols 0.025 min1


SO4- (scavenging Bisphenol A abatement with sulfate radicals was
(UV activation parameters: = 254 nm;
Bisphenol A tests were found to proceed via one electron transfer reaction [182]
40 W power; Io = 1.26 E s1;
performed) mechanism.
Oxidant/Bisphenol A molar ratio 3/1)

SO4- (scavenging
Degradation involved one electron transfer, 0.032 h1 (Activation temperature of 50
Diclofenac tests were [197]
o
hydroxylation, decarboxylation. C; Oxidant/Diclofenac molar ratio 10/1)
performed)

Pharmaceuti 0.087 min1


SO4- (scavenging Electron-transfer between sulfate radicals and
cals (UV activation parameters: = 254 nm; 9
Carbamazepine tests were carbamazepine was found to be the major [171]
W power; Oxidant/Carbamazepine molar
performed) mechanism.
ratio 10/1)

Bisoprolol SO4- and OH The formation of hydroxylated products through 8.5 10-2 min-1 (activation temperature of [36]

42
(scavenging tests hydroxylation was proposed. 60 oC; Oxidant/Bisoprolol molar ratio

were performed) 20/1)

SO4- and OH 29 10-2 min-1 (UV activation parameters:

Chloramphenicol (scavenging tests - = 254 nm; Oxidant/Chloramphenicol [59]

were performed) molar ratio 80/1)

10-3 sec-1
-
Possibly SO4 Approx. 96.4% of chloride ion was released after
(UV activation parameters: = 254 nm;
Lindane and/or OH (not the treatment, which was consistent with the TOC [107]
10 M = Fe2+; Oxidant/Lindane molar
identified) analysis.
ratio 73/1)

1.7 x 10-3 sec-1 (UV activation parameters:

= 254 nm; Oxidant/Atrazine molar ratio

50/1)
Pesticides Triazine
Manoj et al. [198] has determined
pesticides
bimolecular rate constants of the SO4- [32,152,198-
(atrazine, tert- -
Possibly SO4- reaction with triazines in the range of: 200]
butylazine,
and/or OH 4.61073109M1s1, whereas Lutze et
propazine)
al. [152]: 2.23.5 109 M1 s1; Khan et

al. [199]: 2.59 109 M1 s1 and 2.25

109 M1 s1 for the reaction with SO4- and

43

OH, respectively;

4.14 x 10-4 sec-1 (UV activation


SO4- and OH Degradation of endosulfan was began at the
Endosulfan parameters: = 254 nm; Oxidant/ [191]
(identified) endosulfans S=O group.
Endosulfan molar ratio 10/1)

The first step is an electron transfer from a benzene


14.8 10-2 min-1 (activation temperature
-
Possibly SO4 rings of MB to SO4-. After 20 min of reaction,
Methylene blue of 60 oC; Oxidant/Methylene blue molar [190]

and/or OH complete disappearance of all intermediates was
ratio 640/1)
noticed.

Reaction rates differed between UV/PMS and


0.175 min1 (UV activation parameters:
-
Other Possibly SO4 (not UV/PDS, possibly due to the production of the
Acid orange 7 = 254 nm; Oxidant/Acid orange 7 molar [19]
identified) hydroxyl radical during the photolysis of PMS and
ratio 10/1)
different quantum yields.

0.18 h1
SO4- accepts an electron from the carboxylate
PFOA SO4- and S2O8- (UV activation parameters: = 254 nm; [180]
group in a primary step.
Oxidant/PFOA molar ratio 33/1)

44
Moreover, Neppolian et al. [201] proved that these radicals can also be used for the removal

of inorganic pollutants in water. They investigated the UV/PDS oxidation of As(III) to the

less harmful As(V). It was established that humic acid had no effect on the reaction rate

(similar results to Wang et al. [134], although therein 1.6 M methanol was used as a

scavenger), even at 20 mg L-1. Yet, the continual addition of nitrogen considerably minimized

the rate of the reaction (by 20%), attributing this to the role of dissolved oxygen in the

reaction (at high concentrations of PDS this phenomenon was not observed, presumably due

to formation of oxygen in the reaction of peroxydisulfate with SO4-; [58]).

Surprisingly, according to Diao et al. [202,203] SO4- coupled with nano zero-valent iron

(supported on bentonite, and used as a radical initiator), can be effective for the simultaneous

abatement of Cr(VI) and phenol from water. The reaction mechanism according to them

involved the removal of Cr(VI) mainly by reduction with nano zero-valent iron and phenol

removal mainly by the SO4- generated from the PDS.

Chlorinated olefins are ubiquitous contaminants, and although they can be degraded with

many biological [204] and less invasive chemical treatments (i.e. H2O2) [205], persulfates are

often used for their degradation in situ. Recently, Yan et al. [206] combined siderite-catalysed

H2O2 with PDS and effectively used it for the remediation of trichloroethene contamination

from groundwater. It was claimed therein, that in the absence of PDS (only catalysed

peroxide), most of the hydrogen peroxide was reduced within the first hour of the test,

resulting in non-efficient use of OH-radicals. After the addition of PDS, the decomposition

rate of H2O2 was mitigated due to a more sustainable release of OH-radicals. Furthermore, the

heat given by the decomposition reaction of H2O2 activated the PDS, and the generated sulfate

radicals were claimed to be the main oxidative species. In addition, dichloroacetic acid has

been detected as an intermediate. However, it has to be noted, that it was not explained therein

why the OH formation is not efficient in the absence and why it is efficient in the presence of

46
PDS. Xu et al. [207] also studied the abatement of trichloroethene (TCE) but in a thermally

activated peroxydisulfate system. Their results showed that TCE can be completely removed

from the solution within 9 minutes at 50 oC with an initial trichloroethene concentration of

0.15 mM and a PDS dose of 0.3 M, as a consequence of the active oxygen species formation

(SO4-, OH). Moreover, Zhao et al. [48] studied the simultaneous decontamination of 1,4-

dioxane, the inherent associate of TCE (frequently used as a solvent stabilizer for TCE), with

heat- and Fe2+ PDS activation. Analysis of carbon balance revealed that 96% and 93% of the

organic carbon was removed after the 1,4-dioxane abatement with and without activation

(addition of Fe2+), respectively.

Another commonly found and very toxic group of contaminants is pesticides, which

contribute to nine out of the twelve most hazardous and assiduous organic compounds defined

by the Stockholm Convention on Persistent Organic Pollutants (POPs) [208]. A very recent

study by Qin et al. [209] presented 1,1,1-trichloro-2,2-bis(p-chlorophenyl) ethane (DDT)

removal with Co2+ catalysed PMS. It was found that DDT was efficiently decomposed within

several hours, proportionally to PMS/Co2+ concentrations. The decontamination rates of DDT

were determined with pseudo-first-order reaction rate equations in several temperatures,

which enabled calculation of the activation energy (72 kJ mol-1). Several by-products of the

reaction were determined including: 4-chlorobenzoic acid, benzylalcohol,

dichlorobenzophenone, and the possible degradation pathway of DDT was suggested on the

foundation of the detected by-products. Zhu et al. [210] have also followed the DDT

degradation (and the reaction intermediates) but using PDS (activated by nanoscale zero-

valent iron) as a source of radicals. The degradation pathway was very similar to that of Qin

et al. [209] (in the PMS/cobalt system), although it should be noted that instead of 4-

chlorobenzoic acid, its dechlorinated version has been found. In addition, ESR results showed

simultaneous involvement of sulfate radicals and hydroxyl radicals in the degradation process.

47
Several authors examined atrazine degradation with persulfates. One group was Luo et al.

[32] who tested degradation of atrazine with three oxidants H2O2, PMS and PDS (UV

activated, 254 nm). The matrix effects, i.e. water hardness, Cl-, and NOM, were evaluated on

these three AOPs. It was determined that the concentrations of sulfate radicals and hydroxyl

radicals decrease with an increase of carbonate/bicarbonate concentrations. A detailed

description on the reactions of OH and SO4- with Cl- in presence of HCO3- is described in

section 5.3, see also Lutze et al. [152]. Main transformation products of atrazine in reactions

with OH and SO4- are desethyl-atrazine and desisopropyl atrazine [152,211,212] which are

similarly toxic as the parent compound [213]. The second order rate constant of different

chlorotriazine pesticides with SO4- was determined to be in the range of 1-5 109 M-1 s-1. The

primary dealkylation products still react fast 0.8-2 109 M-1 s-1. However, in case no alkyl

group is attached at the chlorotriazine ring (i.e., the chlorotriazine diamine) the reaction

becomes very slow for both OH (k < 107 M-1 s-1) [214] and SO4- (k 1.5 10 8 M-1 s-1) [152],

hence desethyldesisopropyl atrazine can be considered to survive oxidative treatment (cf.

mechanistic aspects of chlorotriazine pesticides degradation [152,211,212]. Research work of

Wacawek et al. [184] was one of the first evaluating the degradation efficiency of

hexachlorocyclohexane isomers by PMS. Cao et al. [106] and Khan et al. [107] provided

more detailed study focusing on the oxidation of one HCH isomer - lindane (-

hexachlorocyclohexane) by Fe2+ activated PDS and PMS, respectively. These studies revealed

that oxidation of HCH with activated persulfates is not only beneficial for the complete

removal of the parent compounds but also for obtaining complete mineralization.

Trichlorophenol was the main by-product detected in all of these studies (although there is no

agreement on the exact isomer generated). In addition, in a further study, Khan et al. [215]

determined the second-order rate constant of lindane with SO4- (1.3 109 M1 s1).

Kumierek et al. [216] investigated the degradation of 2,4-dichlorophenol and 2,4-

48
dichlorophenoxyacetic acid by ammonium PDS activated in several ways. 2,4-dichlorophenol

degraded faster and more efficiently in an alkaline environment (pH = 9.0), whereas 2,4-

dichlorophenoxyacetic degraded faster in an acidic environment (pH = 3.0). They have also

examined the synergistic activation of PDS with heat and ferrous iron to improve the

oxidation of 2,4-dichlorophenol and 2,4-dichlorophenoxyacetic acid. They were able to select

the optimal degradation conditions (molar ratio between PDS/Fe2+ - 1:2, temp. - 50 C),

where complete removal of contaminants was obtained after approximately 45 and 60

minutes, respectively.

As was mentioned in the introduction, the rapid emergence of resistant bacteria worldwide is

probably due to the overuse of medications that can later become a contamination of concern.

Several authors have focussed on the remediation of pharmaceuticals and since persulfates are

one of the newest ISCO reagents used, there are also several new studies describing their

reactivity towards pharmaceutical drugs. Monteagudo et al. [217] investigated ISCO of a

carbamazepine solution by PDS (at the same time) activated by UV, heat, Fe2+ ions, and

H2O2. Zhang et al. [171] determined the main by-products generated during the oxidation

process, including 10,11-epoxy-carbamazepine, acridine-9-carbaldehyde, acridine, and other

compounds with smaller molecular weight.

Trimethoprim and sulfamethoxazole in expired sulfamethoxazole tablets were the subject of

research conducted by Liu et al. [218], who studied their degradation by catalysed PDS

treatment. Zero valent-iron showed much better catalytic properties than alkaline activation,

which was completed after 0.5 hours, while full mineralization was achieved after 2 hours.

Also Ayoub and Ghauch [50] applied activated PDS for sulfamethoxazole degradation and

determined that the metallic iron-based particles (heterogeneous systems) are more efficient

than conventional Fe2+ fed systems (homogeneous systems) for the removal of this

pharmaceutical.

49
Chen et al. [197] examined the performance of thermally activated PDS on the degradation of

diclofenac in both water and polluted groundwater. The results implied that the degradation of

diclofenac could be fitted well to a pseudo 1st-order kinetic model, and that the rate constants

were larger at higher temperatures. Activation energy was also calculated and equalled to 158

kJ mol-1. The presence of a small dose of chloride (0-10 mM) enhanced the abatement of

diclofenac, whereas larger chloride addition (>10 mM) had opposite effect. HCO3-

demonstrated an insignificant effect on diclofenac elimination, while NOM, e.g., humic acids,

slightly inhibited diclofenac removal. The fast oxidation of diclofenac was further observed in

a groundwater sample from contaminated site. In addition, radical quenching tests revealed

that SO4- were the leading reactive species for diclofenac oxidation. Ibuprofen is also

considered as an emerging contaminant and its degradation in the activated PDS system was

carried out by Ghauch et al. [35]. The by-products were not detected throughout the treatment

process. It was concluded that this method could be an adequate approach for specific

treatment of small volumes of hot spot wastewater containing this contaminant (for example

hospital effluents). PDS was also tested on hospital effluent spiked with naproxen. This study

clearly demonstrated that thermally activated peroxydisulfate is a valid and efficient method

that can be used for the removal of dissolved pharmaceuticals in water and sewage water [95].

SO4- treatment also proved to be effective for highly contaminated mature landfill leachate.

Li et al. [219] used ferrous iron loaded AC as a heterogeneous PDS catalyst for its

pretreatment. The effects of the iron/PDS dose and initial pH on the abatement of the organic

pollution in the landfill leachate were determined. It was shown that the chemical oxygen

demand (COD) degradation rate exceeded 87.8% when simultaneous conditions were applied

i.e. Fe2+ dose of 127 mg L-1, PDS concentration of 0.5 M and initial pH of 3.0.

50
5.5.2 Wastewater and sludge

Several studies can be found in the literature that describe the use of persulfates in wastewater

treatment technologies. Their use is focused on exploiting their oxidative potential (removal

of contaminants) and improving the properties of sludge, i.e. dewaterability. Kronholm and

Riekkola [220] tried to answer the question whether potassium peroxydisulfate is a good

choice for wastewater oxidation below the critical temperature of water. The efficiency of

phenol, 1-naphthol, and 2,3-dichlorophenol oxidation in high-temperature (75-340 C)

pressurized (25-45 MPa) wastewater was investigated in an aqueous environment. The

removal percentages of phenol were good even at 115 C. Nonetheless, it has to be noted that

although this radical initiation method is not economically feasible and there were reported

many other methods for treatment of phenolic wastewater [221,222], the study of Kronholm

and Riekkola [220] has shown new alternative way for wastewater treatment. Recently, the

removal of COD from petrochemical wastewater and from real high-strength industrial

wastewater was studied by Babaei and Ghanbari [223] and Kattel et al. [224], respectively. In

both studies, the persulfates proved to be a viable alternative to the conventionally used

oxidants (i.e., percarbonate, hydrogen peroxide).

Generally, oxidative wastewater treatment options are considered as more efficient than the

conventional ones; however, one of the downsides of them is above-mentioned high cost of

the chemicals and energy involved. This drawback could be partially overcome by the

combination of various techniques (e.g. membrane ones), which could not only reduce the

amount of oxidant needed but also improve the second process, e.g. by decrease membrane

fouling [225].

Fagier et al. [226] presented the efficiency of coagulation-flocculation pretreatment coupled

with a SO4- oxidation process in the removal and mineralization of organic matter of

sugarcane vinasse. Ferric chloride (15 g L-1), a standard coagulation agent in wastewater

51
treatment plants (WWTP), was used and achieved a 70% TOC removal. The pretreated

vinasse subjected to a PDS/PMS oxidation process (activated by Fe2+) showed the highest

TOC removal efficiency at pH 7. Under the selected optimum conditions, approximately 70

and 49% TOC removal was achieved for PMS/Fe2+ and PDS/Fe2+, respectively. Also,

Rodrguez-Chueca et al. [227] investigated winery wastewater treatment using PMS coupled

with a transition metal and UV light. High COD and TOC removal efficacy (79% and 64%,

respectively) was observed under optimal conditions after three hours of treatment.

In another study, wastewater containing cytosine arabinoside (ara-C) was treated with PDS

and H2O2 activated by UV radiation [228]. It was found that addition of oxidants considerably

increased the ara-C removal effectiveness, and the TOC content in the wastewater declined

with longer oxidation period but the toxicity increased, surprisingly, mainly in UV/H2O2

system. Shu et al. [229] investigated the UV/PMS degradation of Acid Blue 113 containing

wastewater. They observed that there was no correlation between initial pH value and dye

removal efficiency but UV light intensity significantly affected the efficiency of TOC

removal.

Heterogeneous activation of PMS could also be used for the remediation of organic

contaminants in wastewater according to [27]. They concluded that a CoMn2O4 catalyst was

efficient for the heterogeneous activation of PMS and environmentally friendly. However, it

showed almost no catalytic activity to PDS and H2O2. They observed that Rhodamine B

degradation in wastewater was enhanced with an increase in reaction temperature (15-55 oC)

and inhibited with an increase in fulvic acid concentration (0-0.08 g L-1).

To date, a very limited number of papers have been published concerning the use of

persulfates for sludge disintegration. There are many investigations in the matter of sludge

disintegration by PDS and only a few focusing on the disintegration of activated sludge by

PMS. The methods often used for determining the degree of waste activated sludge (WAS)

52
disintegration include the measurement of soluble chemical oxygen demand (SCOD) and the

sludge volume index (SVI). Determining the SCOD can unveil the degree of polymer transfer

from the solid phase to the liquid phase, whereas the SVI is a measurement of the settleability

of the sludge, which can be measured in a 1000 mL measuring cylinder after 30 minutes of

sedimentation and expressed for a known initial sludge concentration. One of the first

references to persulfates being used for sludge disintegration can be found in a study by [230],

who observed that ferrous iron activated PDS has a positive effect on enhancing sludge

dewaterability with an 88.8% capillary suction time (CST) reduction within 1 minute. The

purpose of a CST test is to characterize the sludge dewaterability rapidly and easily. The time

the filtrate requires to travel a fixed distance in the filter paper is referred to as the capillary

suction time [231]. Similar results to Zhen et al. [232] were obtained by Shi et al. [233],

whereby the highest specific resistance to filtration (SRF; which is another often used method

for dewaterability assessment) and CST reduction efficiencies of 88.5 and 91.5%,

respectively, were acquired after PDS/Fe2+ oxidation. Electro-activated PDS, has also been

applied for sludge treatment [234] and it was concluded that the process can be potentially

applied to deal with wastewater from toluene nitration processes. In addition, it was

determined that 2,4-dinitrotoluene in wastewater under electro-activated PDS oxidation can

mainly be treated by virtue of SO4- descended from the reduction of PDS anions. Also, Zhen

et al. [235] performed electrolysis/PDS/Fe2+ oxidation to improve sludge dewaterability by

disrupting the protective barrier and cracking the entrapped cells, which resulted in releasing

the water inside extracellular polymeric substances and cells. Zhen et al. [232] found that a

combination of PDS and thermal processes (at a mild temperature) is efficient in enhancing

the dewaterability of sludge. They concluded that when the temperature is increased to 80 oC

in the presence of PDS, the flocs of waste activated sludge were drastically changed and that

this pretreatment resulted in the disruption of sludge flocs by degrading extracellular

53
polymeric substances. These results were confirmed in a recent study focusing on the

disintegration of sludge with heat activated PDS [236]. It was observed that organic matter

and polymer transfer from the solid phase to the liquid phase occurred. An increase in SCOD,

(almost a 15-fold increase over the WAS value) and a decrease in the SVI from 89.8 cm3 g-1

to 30.6 cm3 g-1 were also observed. A large issue for introduction of the presented method in

WWTP, concerns heat activation of persulfates. However, the temperature threshold sufficient

for PDS to rapidly form radicals (50-90 oC) can be reached using e.g. the heat generated in the

fermentation process, steam or hot air injection [237].

Probably the first reference to the use of PMS for chemical disintegration of waste activated

sludge was made in our recent papers [238,239]. Similarly to an earlier study [236], it was

concluded that heat application (50, 70 and 90 C) for PMS activation causes an increase in

the soluble COD value and protein concentration in the supernatant and positively influences

the SVI, which decreased from 89.8 to 17.2 mL g-1. Also, Niu et al. [240] and Liu et al. [241]

observed positive effects of WAS oxidation with PMS. Sludge disintegration was

characterised by a change in disintegration degree (DD), sludge particle size, and the

properties of extracellular polymeric substances.

Although thermally activated persulfates are efficacious for the disintegration and

improvement of sludge sedimentation properties [238], Zhen et al. [242] observed a possible

inhibitory effect on anaerobic digestion. On the contrary, Sun et al. [243] found that PDS

disintegration had a positive influence on the biogas yield. Therefore, the composition of the

sludge and the type of fermentation could be crucial for assessing the benefits of persulfates

for WAS disintegration.

Sludge treatment with persulfates can also be focused on anaerobically digested sludge as

reported in a recent study [244] and in a recent article published in Nature: Scientific Reports

[245]. In addition to the enhancement of dewaterability, good efficiency of toluene removal

54
from anaerobically digested sludge could be observed after the treatment with persulfates

catalysed to form radicals with elevated temperatures from meso- or thermophilic digestion

[244].

6. Conclusions

Despite persulfates being efficacious substances for the remediation of water, wastewater and

sludge media, the decay of pollutants is extremely dependent on activation techniques and the

composition of the treated water matrix.

Non-activated persulfates react at rates that are often considered slow, but fast reactions

between the persulfates and free radicals generated from them can increase the significance of

these processes. Although non-catalysed persulfate reactions possess advantages, i.e. lower

cost, no secondary contamination due to the catalyst load and higher stability in the

subsurface, drawbacks i.e. often much slower reaction rates with contaminants and formation

of stable disinfection by-products, which reduce natural attenuation, favour the use of

activated persulfate reactions.

Activation can be achieved by versatile means such as heat, UV radiation, radiolysis or by

chemical methods, allowing to establish SO4 --based oxidation in very different fields such as

remediation, wastewater and drinking water treatment

Chemical activation of persulfates can be considered either a homogenous or a heterogeneous

reaction. Both of these radical initiation types have benefits and drawbacks, although

especially the high cost, chemical stability (leaching of catalyst constituents), long preparation

time, and chemical stability of heterogeneous catalysts in some cases still largely limits their

55
use in water treatment. In the subsurface the water quality parameters are of great importance.

The pH value and concentrations of halogens/natural organic matter were found to be

especially crucial. In most of the systems pH of ~3 was found to be beneficial for rapid

decontamination (although typically the pH of natural waters is not that low). However,

elevated pH values can enhance the reaction rates due to the additional activation (proven for

both persulfates) and the faster reaction of radicals with deprotonated compounds. Natural

organic matter and halogens are of relevance due to their radical scavenging and them being a

precursor to organic disinfection by-products. On the contrary, activation can be caused

incidentally by specific site conditions (e.g. large transition metal content or elevated pH).

In the catalysed persulfate systems the SO4- is often considered to be the main oxidative

species. However, depending on reaction conditions other reactive species such as hydroxyl

radical and superoxide radical were determined by many authors to play an important role,

too. In case SO4 - is converted to other reactive species, sometimes the unique features of SO4-

cannot be exploited. This is important to know in case very persistent compounds such as

chlorotriazine diamine or polyfluorinated compounds (e.g. PFOA) have to be degraded, which

are inert towards the conventional degradation methods

In the present review analytical methods for quantification of persulfates were presented and

assessed. Spectrophotometric methods are among the most commonly used for persulfate

determination; however, it should be noted that the liquid chromatographic methods are

among the most reliable with a low quantification limit.

Data on decontamination kinetics using activated persulfates show that such reactions are

regarded as extremely fast in comparison to biological or chemical reductive treatment

56
(However, it should be noted, that some compounds might be degraded only by the chemical

reduction). The pseudo-first order kinetic model (experiments conducted with an excess of

oxidant) is mostly used.

Persulfate decontamination technologies either with oxidation via direct electron transfer or

free radical driven processes were found to be very powerful tools for the remediation of a

wide range of contaminants, including chlorinated olefins, BTEXs, phenols, pharmaceuticals,

inorganics and pesticides. However, several disadvantages, i.e. pH changes, salinity of the soil

and creation of hazardous decontamination by-products should be seriously taken into account

before their application.

Acknowledgements

The work was supported by the project LO1201, the financial support of the Ministry of

Education, Youth and Sports in the framework of the targeted support of the National

Programme for Sustainability I and the OPR&DI project Centre for Nanomaterials,

Advanced Technologies and Innovation - CZ.1.05/2.1.00/01.0005. The authors also

acknowledge the assistance provided by the Research Infrastructure NanoEnviCz, supported

by the Ministry of Education, Youth and Sports of the Czech Republic under Project No.

LM2015073.

References

[1] A. Gosset, Y. Ferro, C. Durrieu, Methods for evaluating the pollution impact of urban
wet weather discharges on biocenosis: A review, Water Res. 89 (2016) 330354.
[2] P.J. Crutzen, S. Wacawek, Atmospheric chemistry and climate in the anthropocene /
Chemia atmosferyczna i klimat w antropocenie, Chemistry-Didactics-Ecology-
Metrology. 19 (2014) 928.
[3] J.L. Martinez, Environmental pollution by antibiotics and by antibiotic resistance
determinants, Environ. Pollut. 157 (2009) 2893902.

57
[4] T.A. Larsen, S. Hoffmann, C. Lthi, B. Truffer, M. Maurer, Emerging solutions to the
water challenges of an urbanizing world, Science. 352 (2016) 92833.
[5] T.M. Phillips, A.G. Seech, H. Lee, J.T. Trevors, Biodegradation of
hexachlorocyclohexane (HCH) by microorganisms, Biodegradation. 16 (2005) 363
392.
[6] D. Bass, N. Hastings, R. Brown, Performance of air sparging systems: a review of case
studies, J. Hazard. Mater. 72 (2000) 101119.
[7] D.W. Elliott, H.-L. Lien, W.-X. Zhang, Degradation of Lindane by zero-valent iron
nanoparticles, J. Environ. Eng. 135 (2009) 317324.
[8] C. von Sonntag, U. von Gunten, Chemistry of ozone in water and wastewater treatment,
London ; New York : IWA Pub., 2012.
[9] F.R. Spellman, Choosing Disinfection Alternatives for Water/Wastewater Treatment
Plants, CRC Press, 1999. https://books.google.com/books?id=Fnk-FkablroC&pgis=1
[10] K. Kaur, M. Crimi, Release of chromium from soils with persulfate chemical oxidation.,
Ground Water. 52 (2014) 74855.
[11] X. Xu, S. Li, Q. Hao, J. Liu, Y. Yu, H. Li, Activation of persulfate and its
environmental application, Int. J. Environ. Bioenergy. 1 (2012) 6081.
[12] R.L. Siegrist, M. Crimi, T.J. Simpkin, In Situ Chemical Oxidation for Groundwater
Remediation. Chapter 2: Fundamentals of ISCO using hydrogen peroxide, Springer;
2011 edition (March 22, 2011).
[13] G. Fan, L. Cang, G. Fang, W. Qin, L. Ge, D. Zhou, Electrokinetic delivery of persulfate
to remediate PCBs polluted soils: effect of injection spot., Chemosphere. 117 (2014)
4108.
[14] C.H. Yen, K.F. Chen, C.M. Kao, S.H. Liang, T.Y. Chen, Application of persulfate to
remediate petroleum hydrocarbon-contaminated soil: Feasibility and comparison with
common oxidants, J. Hazard. Mater. 186 (2011) 20972102.
[15] X. Duan, H. Sun, J. Kang, Y. Wang, S. Indrawirawan, S. Wang, Insights into
heterogeneous catalysis of persulfate activation on dimensional-structured nanocarbons,
ACS Catal. 5 (2015) 46294636.
[16] I.M. Kolthoff, I.K. Miller, The Chemistry of Persulfate. I. The kinetics and mechanism
of the decomposition of the persulfate ion in aqueous medium 1, J. Am. Chem. Soc. 73
(1951) 30553059.
[17] E.J. Behrman, D.H. Dean, Sodium peroxydisulfate is a stable and cheap substitute for
ammonium peroxydisulfate (persulfate) in polyacrylamide gel electrophoresis., J.
Chromatogr. B. Biomed. Sci. Appl. 723 (1999) 325326.
[18] S.G. Huling, B. Pivetz, In-situ chemical oxidation--engineering issue. EPA/600/R-
06/072, (2007).
[19] S. Yang, P. Wang, X. Yang, L. Shan, W. Zhang, X. Shao, R. Niu, Degradation
efficiencies of azo dye Acid Orange 7 by the interaction of heat, UV and anions with
common oxidants: Persulfate, peroxymonosulfate and hydrogen peroxide, J. Hazard.
Mater. 179 (2010) 552558.
[20] D.A. Reckhow, B. Langlais, D.R. Brink, AWWA Research Foundation., Compagnie
gnrale des eaux (Paris, France), Ozone in water treatment: application and
engineering: cooperative research report / American Water Works Association
Research Foundation, Compagnie generale des eaux; edited by Bruno Langlais, David
A. Reckhow, Deborah R. Brink, Lewis Publishers Chelsea, Mich, 1991.
[21] T. Zhang, Y. Chen, Y. Wang, J. Le Roux, Y. Yang, J.P. Crou, Efficient
peroxydisulfate activation process not relying on sulfate radical generation for water
pollutant degradation, Environ. Sci. Technol. 48 (2014) 58685875.

58
[22] A.G. Bailie, K. Bouzek, P. Lukek, I. Rouar, A.A. Wragg, Solubility of potassium
ferrate in 12 m alkaline solutions between 20C and 60C, J. Chem. Technol.
Biotechnol. 66 (1996) 35 - 40.
[23] S.W. Benson, Thermochemistry and kinetics of sulfur-containing molecules and
radicals, Chem. Rev. 78 (1978) 2335.
[24] H. Hussain, I.R. Green, I. Ahmed, Journey describing applications of oxone in synthetic
chemistry, Chem. Rev. 113 (2013) 332971.
[25] P.R. Shukla, S. Wang, H. Sun, H.M. Ang, M. Tad, Activated carbon supported cobalt
catalysts for advanced oxidation of organic contaminants in aqueous solution, Appl.
Catal. B Environ. 100 (2010) 529534.
[26] G. Lente, J. Kalmr, Z. Baranyai, A. Kun, I. Kk, D. Bajusz, M. Takcs, L. Veres, I.
Fbin, One - versus two - electron oxidation with peroxomonosulfate ion: reactions
with iron(II), vanadium(IV), halide ions, and photoreaction with cerium(III), Inorg.
Chem. 48 (2009) 17631773.
[27] Y. Yao, Y. Cai, G. Wu, F. Wei, X. Li, H. Chen, S. Wang, Sulfate radicals induced from
peroxymonosulfate by cobalt manganese oxides (CoxMn3xO4) for Fenton-Like reaction
in water, J. Hazard. Mater. 296 (2015) 128137.
[28] J. Spivey, K. Dooley, Y.-F. Han, Catalysis, The Royal Society of Chemistry, 2015.
[29] J. Bouchard, C. Maine, D.S. Argyropoulos, R.M. Berry, Kraft pulp bleaching using in-
situ dimethyldioxirane: mechanism and reactivity of the oxidants, Holzforschung. 52
(1998) 499505.
[30] P. Bajpai, Chapter Seven Peroxyacids bleaching, in: Environ. Benign Approaches
Pulp Bleach., Gulf Professional Publishing (2012) 167188.
[31] P.A. Block, R.A. Brown, D. Robinson, Novel activation technologies for sodium
persulfate In Situ chemical oxidation. Proceedings, Fourth International Conference on
Remediation of Chlorinated and Recalcitrant Compounds, Monterey, CA, USA. May
24-27. Monterey, (2004) 2A05.
[32] C. Luo, J. Ma, J. Jiang, Y. Liu, Y. Song, Y. Yang, Y. Guan, D. Wu, Simulation and
comparative study on the oxidation kinetics of atrazine by UV/H2O2, UV/HSO5-,
UV/S2O82-, Water Res. 80 (2015) 99108.
[33] Y. Ji, W. Xie, Y. Fan, Y. Shi, D. Kong, J. Lu, Degradation of trimethoprim by thermo-
activated persulfate oxidation: Reaction kinetics and transformation mechanisms, Chem.
Eng. J. 286 (2016) 1624.
[34] R.L. Johnson, P.G. Tratnyek, R.O.B. Johnson, Persulfate persistence under thermal
activation conditions, Environ. Sci. Technol. 42 (2008) 93509356.
[35] A. Ghauch, A.M. Tuqan, N. Kibbi, Ibuprofen removal by heated persulfate in aqueous
solution: A kinetics study, Chem. Eng. J. 197 (2012) 483492.
[36] A. Ghauch, A.M. Tuqan, Oxidation of bisoprolol in heated persulfate/H2O systems:
Kinetics and products, Chem. Eng. J. 183 (2012) 162171.
[37] R. Zhang, P. Sun, T.H. Boyer, L. Zhao, C.H. Huang, Degradation of pharmaceuticals
and metabolite in synthetic human urine by UV, UV/H2O2, and UV/PDS, Environ. Sci.
Technol. 49 (2015) 30563066.
[38] W.S. Chen, Y.C. Su, Removal of dinitrotoluenes in wastewater by sono-activated
persulfate, Ultrason. Sonochem. 19 (2012) 921927.
[39] J. Criquet, N. Karpel Vel Leitner, Electron beam irradiation of aqueous solution of
persulfate ions, Chem. Eng. J. 169 (2011) 258262.
[40] Y. Lei, C.-S. Chen, Y.-J. Tu, Y.-H. Huang, H. Zhang, Heterogeneous Degradation of
Organic Pollutants by Persulfate Activated by CuO-Fe3O4: Mechanism, Stability, and
Effects of pH and Bicarbonate Ions, Environ. Sci. Technol. 49 (2015) 68386845.

59
[41] Y. Ding, L. Zhu, N. Wang, H. Tang, Sulfate radicals induced degradation of
tetrabromobisphenol A with nanoscaled magnetic CuFe2O4 as a heterogeneous catalyst
of peroxymonosulfate, Appl. Catal. B Environ. 129 (2013) 153162.
[42] A. Ghauch, G. Ayoub, S. Naim, Degradation of sulfamethoxazole by persulfate assisted
micrometric Fe0 in aqueous solution, Chem. Eng. J. 228 (2013) 11681181.
[43] G. Fang, J. Gao, D.D. Dionysiou, C. Liu, D. Zhou, Activation of persulfate by quinones:
Free radical reactions and implication for the degradation of PCBs, Environ. Sci.
Technol. 47 (2013) 46054611.
[44] O.S. Furman, A.L. Teel, R.J. Watts, Mechanism of base activation of persulfate,
Environ. Sci. Technol. 44 (2010) 64236428.
[45] C. Qi, X. Liu, J. Ma, C. Lin, X. Li, H. Zhang, Activation of peroxymonosulfate by base:
Implications for the degradation of organic pollutants, Chemosphere. 151 (2016) 280-
288.
[46] Y. Yang, J. Jiang, X. Lu, J. Ma, Y. Liu, Production of sulfate radical and hydroxyl
radical by reaction of ozone with peroxymonosulfate: a novel advanced oxidation
process, Environ. Sci. Technol. 49 (2015) 73307339.
[47] M. Ahmad, A.L. Teel, R.J. Watts, Mechanism of persulfate activation by phenols,
Environ. Sci. Technol. 47 (2013) 58645871.
[48] L. Zhao, H. Hou, A. Fujii, M. Hosomi, F. Li, Degradation of 1,4-dioxane in water with
heat- and Fe2+-activated persulfate oxidation, Environ. Sci. Pollut. Res. 21 (2014) 7457
7465.
[49] G.P. Anipsitakis, D.D. Dionysiou, Transition metal/UV-based advanced oxidation
technologies for water decontamination, Appl. Catal. B Environ. 54 (2004) 155163.
[50] G. Ayoub, A. Ghauch, Assessment of bimetallic and trimetallic iron-based systems for
persulfate activation: Application to sulfamethoxazole degradation, Chem. Eng. J. 256
(2014) 280292.
[51] S. Naim, A. Ghauch, Ranitidine abatement in chemically activated persulfate systems:
Assessment of industrial iron waste for sustainable applications, Chem. Eng. J. 288
(2016) 276288.
[52] Y.H. Guan, J. Ma, X.C. Li, J.Y. Fang, L.W. Chen, Influence of pH on the formation of
sulfate and hydroxyl radicals in the UV/Peroxymonosulfate system, Environ. Sci.
Technol. 45 (2011) 93089314.
[53] G.P. Anipsitakis, D.D. Dionysiou, Degradation of organic contaminants in water with
sulfate radicals generated by the conjunction of peroxymonosulfate with cobalt,
Environ. Sci. Technol. 37 (2003) 47904797.
[54] J. Fernandez, P. Maruthamuthu, A. Renken, J. Kiwi, Bleaching and photobleaching of
Orange II within seconds by the oxone/Co2+ reagent in Fenton-like processes, Appl.
Catal. B Environ. 49 (2004) 207215.
[55] S. Wacawek, V. Anto, P. Hrabk, M. ernk, D. Elliott, Remediation of
hexachlorocyclohexanes by electrochemically activated persulfates, Environ. Sci.
Pollut. Res. 23 (2016) 765773.
[56] D. Zhao, X. Liao, X. Yan, S.G. Huling, T. Chai, H. Tao, Effect and mechanism of
persulfate activated by different methods for PAHs removal in soil, J. Hazard. Mater.
254255 (2013) 228235.
[57] H. Herrmann, A. Reese, R. Zellner, Time-resolved UV/VIS diode array absorption
spectroscopy of SOx- (x=3, 4, 5) radical anions in aqueous solution, J. Mol. Struct. 348
(1995) 183186.
[58] G. Mark, M.N. Schuchmann, H.-P. Schuchmann, C. von Sonntag, The photolysis of
potassium peroxodisulphate in aqueous solution in the presence of tert-butanol: a simple

60
actinometer for 254 nm radiation, J. Photochem. Photobiol. A Chem. 55 (1990) 157
168.
[59] A. Ghauch, A. Baalbaki, M. Amasha, R. El Asmar, O. Tantawi, Contribution of
persulfate in UV-254nm activated systems for complete degradation of chloramphenicol
antibiotic in water, Chem. Eng. J. 317 (2017) 10121025.
[60] W.-D. Oh, Z. Dong, T.-T. Lim, Generation of sulfate radical through heterogeneous
catalysis for organic contaminants removal: Current development, challenges and
prospects, Appl. Catal. B Environ. 194 (2016) 169-201.
[61] A. Rastogi, S.R. Al-Abed, D.D. Dionysiou, Sulfate radical-based ferrous-
peroxymonosulfate oxidative system for PCBs degradation in aqueous and sediment
systems, Appl. Catal. B Environ. 85 (2009) 171179.
[62] S. Yuan, P. Liao, A.N. Alshawabkeh, Electrolytic manipulation of persulfate reactivity
by iron electrodes for trichloroethylene degradation in groundwater, Environ. Sci.
Technol. 48 (2014) 656663.
[63] H. Lee, H.J. Lee, J. Jeong, J. Lee, N.B. Park, C. Lee, Activation of persulfates by carbon
nanotubes: Oxidation of organic compounds by nonradical mechanism, Chem. Eng. J.
266 (2015) 2833.
[64] K. Govindan, M. Raja, M. Noel, E.J. James, Degradation of pentachlorophenol by
hydroxyl radicals and sulfate radicals using electrochemical activation of
peroxomonosulfate, peroxodisulfate and hydrogen peroxide, J. Hazard. Mater. 272
(2014) 4251.
[65] H. Sun, S. Liu, G. Zhou, H.M. Ang, M.O. Tad, S. Wang, Reduced graphene oxide for
catalytic oxidation of aqueous organic pollutants, ACS Appl. Mater. Interfaces. 4 (2012)
54665471.
[66] Y. Tao, M. Wei, D. Xia, A. Xu, X. Li, Polyimides as metal-free catalysts for organic
dye degradation in the presence peroxymonosulfate under visible light irradiation, RSC
Adv. 5 (2015) 9823198240.
[67] Y. Deng, C.M. Ezyske, Sulfate radical-advanced oxidation process (SR-AOP) for
simultaneous removal of refractory organic contaminants and ammonia in landfill
leachate, Water Res. 45 (2011) 61896194.
[68] S. Ahmed, M.G. Rasul, W.N. Martens, R. Brown, M.A. Hashib, Heterogeneous
photocatalytic degradation of phenols in wastewater: A review on current status and
developments, Desalination. 261 (2010) 318.
[69] H. Zhang, H. Fu, D. Zhang, Degradation of C.I. Acid Orange 7 by ultrasound enhanced
heterogeneous Fenton-like process, J. Hazard. Mater. 172 (2009) 654660.
[70] A. Tsitonaki, B. Petri, M. Crimi, H. Mosbaek, R.L. Siegrist, P.L. Bjerg, In situ chemical
oxidation of contaminated soil and groundwater using persulfate: A review, Crit Rev
Env Sci Tec. 40 (2010) 5591.
[71] B.-T. Zhang, Y. Zhang, Y. Teng, M. Fan, Sulfate radical and its application in
decontamination technologies, Crit. Rev. Environ. Sci. Technol. 45 (2015) 1756-1800.
[72] A. Rastogi, S.R. Al-Abed, D.D. Dionysiou, Effect of inorganic, synthetic and naturally
occurring chelating agents on Fe(II) mediated advanced oxidation of chlorophenols,
Water Res. 43 (2009) 684694.
[73] X. Chen, J. Chen, X. Qiao, D. Wang, X. Cai, Performance of nano-
Co 3O4/peroxymonosulfate system: Kinetics and mechanism study using Acid Orange 7
as a model compound, Appl. Catal. B Environ. 80 (2008) 116121.
[74] S. Muhammad, E. Saputra, H. Sun, J. de C. Izidoro, D.A. Fungaro, H.M. Ang, M.O.
Tad, S. Wang, Coal fly ash supported Co 3O4 catalysts for phenol degradation using
peroxymonosulfate, RSC Adv. 2 (2012) 5645-5652.

61
[75] G.P. Anipsitakis, E. Stathatos, D.D. Dionysiou, Heterogeneous activation of Oxone
using Co 3O4, J. Phys. Chem. B. 109 (2005) 1305213055.
[76] Q. Yang, H. Choi, S.R. Al-Abed, D.D. Dionysiou, Iron-cobalt mixed oxide
nanocatalysts: Heterogeneous peroxymonosulfate activation, cobalt leaching, and
ferromagnetic properties for environmental applications, Appl. Catal. B Environ. 88
(2009) 462469.
[77] Q. Yang, H. Choi, Y. Chen, D.D. Dionysiou, Heterogeneous activation of
peroxymonosulfate by supported cobalt catalysts for the degradation of 2,4-
dichlorophenol in water: The effect of support, cobalt precursor, and UV radiation,
Appl. Catal. B Environ. 77 (2008) 300307..
[78] Q. Yang, H. Choi, D.D. Dionysiou, Nanocrystalline cobalt oxide immobilized on
titanium dioxide nanoparticles for the heterogeneous activation of peroxymonosulfate,
Appl. Catal. B Environ. 74 (2007) 170178.
[79] J. Zhang, M. Chen, L. Zhu, Activation of persulfate by Co3O4 nanoparticles for orange
G degradation, RSC Adv. 6 (2016) 758768.
[80] Y. Yao, Z. Yang, D. Zhang, W. Peng, H. Sun, S. Wang, Magnetic CoFe2O4graphene
hybrids: facile synthesis, characterization, and catalytic properties, Ind. Eng. Chem.
Res. 51 (2012) 60446051.
[81] S. Su, W. Guo, Y. Leng, C. Yi, Z. Ma, Heterogeneous activation of Oxone by Co xFe3-
xO4 nanocatalysts for degradation of rhodamine B, J. Hazard. Mater. 244245 (2013)
736742.
[82] K.-Y.A. Lin, F.-K. Hsu, W.-D. Lee, Magnetic cobaltgraphene nanocomposite derived
from self-assembly of MOFs with graphene oxide as an activator for
peroxymonosulfate, J. Mater. Chem. A. 3 (2015) 94809490.
[83] K.-Y.A. Lin, B.-C. Chen, Efficient elimination of caffeine from water using Oxone
activated by a magnetic and recyclable cobalt/carbon nanocomposite derived from ZIF-
67, Dalt. Trans. 45 (2016) 35413551.
[84] P.H. Shi, S.B. Zhu, H.G. Zheng, D.X. Li, S.H. Xu, Supported Co3O4 on expanded
graphite as a catalyst for the degradation of Orange II in water using sulfate radicals,
Desalin. Water Treat. 52 (2014) 33843391.
[85] Y. Ding, L. Zhu, A. Huang, X. Zhao, X. Zhang, H. Tang, A heterogeneous Co3O4
Bi2O3 composite catalyst for oxidative degradation of organic pollutants in the presence
of peroxymonosulfate, Catal. Sci. Technol. 2 (2012) 1977-1984.
[86] S. Sarkar, M.S. Adhikari, M. Banerjee, R.S. Konar, Thermal decomposition of
potassium persulfate in aqueous solution at 50C in an inert atmosphere of nitrogen in
the presence of acrylonitrile monomer, J. Appl. Polym. Sci. 35 (1988) 14411458.
[87] T. Rhadfi, J.Y. Piquemal, L. Sicard, F. Herbst, E. Briot, M. Benedetti, A. Atlamsani,
Polyol-made Mn3O4 nanocrystals as efficient Fenton-like catalysts, Appl. Catal. A Gen.
386 (2010) 132139.
[88] X. Zhai, W. Yang, M. Li, G. Lv, J. Liu, X. Zhang, Noncovalent hybrid of CoMn2O4
spinel nanocrystals and poly (diallyldimethylammonium chloride) functionalized carbon
nanotubes as efficient electrocatalysts for oxygen reduction reaction, Carbon N. Y. 65
(2013) 277286.
[89] Y. Wang, H. Sun, H.M. Ang, M.O. Tad, S. Wang, Synthesis of magnetic core/shell
carbon nanosphere supported manganese catalysts for oxidation of organics in water by
peroxymonosulfate, J. Colloid Interface Sci. 433 (2014) 6875.
[90] Y. Wang, H. Sun, H.M. Ang, M.O. Tad, S. Wang, 3D-hierarchically structured MnO2
for catalytic oxidation of phenol solutions by activation of peroxymonosulfate: Structure
dependence and mechanism, Appl. Catal. B Environ. 164 (2015) 159167.

62
[91] Y. Li, L.D. Liu, L. Liu, Y. Liu, H.W. Zhang, X. Han, Efficient oxidation of phenol by
persulfate using manganite as a catalyst, J. Mol. Catal. A Chem. 411 (2016) 264271.
[92] C. Tan, N. Gao, Y. Deng, J. Deng, S. Zhou, J. Li, X. Xin, Radical induced degradation
of acetaminophen with Fe3O4 magnetic nanoparticles as heterogeneous activator of
peroxymonosulfate, J. Hazard. Mater. 276 (2014) 452460.
[93] W.-D. Oh, Z. Dong, Z.-T. Hu, T.-T. Lim, A novel quasi-cubic CuFe2O4-Fe2O3 catalyst
prepared at low temperature for enhanced oxidation of bisphenol A via
peroxymonosulfate activation, J. Mater. Chem. A. 3 (2015) 2220822217.
[94] J. Zhang, M. Chen, L. Zhu, Activation of peroxymonosulfate by iron-based catalysts for
orange G degradation: role of hydroxylamine, RSC Adv. 6 (2016) 4756247569.
[95] A. Ghauch, A.M. Tuqan, N. Kibbi, Naproxen abatement by thermally activated
persulfate in aqueous systems, Chem. Eng. J. 279 (2015) 861873.
[96] Y. Wang, Z. Ao, H. Sun, X. Duan, S. Wang, Activation of peroxymonosulfate by
carbonaceous oxygen groups: experimental and density functional theory calculations,
Appl. Catal. B Environ. 198 (2016) 295302.
[97] J. Zhang, X. Shao, C. Shi, S. Yang, Decolorization of Acid Orange 7 with
peroxymonosulfate oxidation catalyzed by granular activated carbon, Chem. Eng. J. 232
(2013) 259-265.
[98] E. Saputra, S. Muhammad, H. Sun, S. Wang, Activated carbons as green and effective
catalysts for generation of reactive radicals in degradation of aqueous phenol, RSC Adv.
3 (2013) 21905-21910.
[99] K.-Y. Andrew Lin, Z.-Y. Zhang, -Sulfur as a metal-free catalyst to activate
peroxymonosulfate under visible light irradiation for decolorization, RSC Adv. 6 (2016)
1502715034.
[100] S.-Y. Oh, S.-G. Kang, Degradation of 2,4-dinitrotoluene by persulfate and steel waste
powder, Geosystem Eng. 13 (2010) 105110.
[101] H. V. Lutze, N. Kerlin, T.C. Schmidt, Sulfate radical-based water treatment in presence
of chloride: Formation of chlorate, inter-conversion of sulfate radicals into hydroxyl
radicals and influence of bicarbonate, Water Res. 72 (2015) 349360.
[102] D.P. Cassidy, V.J. Srivastava, F.J. Dombrowski, J.W. Lingle, Combining in situ
chemical oxidation, stabilization, and anaerobic bioremediation in a single application
to reduce contaminant mass and leachability in soil, J. Hazard. Mater. 297 (2015) 347
355.
[103] A. Long, H. Zhang, Selective oxidative degradation of toluene for the recovery of
surfactant by an electro/Fe2+/persulfate process, Environ. Sci. Pollut. Res. 22 (2015)
1160611616.
[104] N. Jaafarzadeh, F. Ghanbari, M. Moradi, Photo-electro-oxidation assisted
peroxymonosulfate for decolorization of acid brown 14 from aqueous solution, Korean
J. Chem. Eng. 32 (2015) 458464.
[105] H. Lin, Y. Li, X. Mao, H. Zhang, Electro-enhanced goethite activation of
peroxydisulfate for the decolorization of Orange II at neutral pH: Efficiency, stability
and mechanism, J. Taiwan Inst. Chem. Eng. 65 (2016) 390398.
[106] J.S. Cao, W.X. Zhang, D.G. Brown, D. Sethi, Oxidation of lindane with Fe(II)-activated
sodium persulfate, Environ. Eng. Sci. 25 (2008) 221228.
[107] S. Khan, X. He, H.M. Khan, D. Boccelli, D.D. Dionysiou, Efficient degradation of
lindane in aqueous solution by iron (II) and/or UV activated peroxymonosulfate, J.
Photochem. Photobiol. A Chem. 316 (2016) 3743.
[108] J. Cong, G. Wen, T. Huang, L. Deng, J. Ma, Study on enhanced ozonation degradation
of para-chlorobenzoic acid by peroxymonosulfate in aqueous solution, Chem. Eng. J.
264 (2015) 399403.

63
[109] L. Dogliotti, E. Hayon, Flash photolysis of persulfate ions in aqueous solutions. Study
of the sulfate and ozonide radical anions, J. Phys. Chem. 71 (1967) 25112516.
[110] P.E.A. Boudeville, Simultaneous determination of hydrogen peroxide,
peroxymonosulfuric acid, and peroxydisulfuric acid by thermometric titrimetry, Anal.
Chem. 55 (1983) 612615.
[111] C. Liang, C.F. Huang, N. Mohanty, R.M. Kurakalva, A rapid spectrophotometric
determination of persulfate anion in ISCO, Chemosphere. 73 (2008) 15401543.
[112] K.-C. Huang, R. a. Couttenye, G.E. Hoag, Kinetics of heat-assisted persulfate oxidation
of methyl tert-butyl ether (MTBE), Chemosphere. 49 (2002) 413420.
[113] L. Zhao, S. Yang, L. Wang, C. Shi, M. Huo, Y. Li, Rapid and simple
spectrophotometric determination of persulfate in water by microwave assisted
decolorization of Methylene Blue, J. Environ. Sci. (China). 31 (2015) 235239.
[114] S. Gokulakrishnan, A. Mohammed, H. Prakash, Determination of persulphates using
N,N-diethyl-p-phenylenediamine as colorimetric reagent: Oxidative coloration and
degradation of the reagent without bactericidal effect in water, Chem. Eng. J. 286
(2016) 223231.
[115] E. Villegas, Y. Pomeranz, J.A. Shellenberger, Colorimetric determination of persulfate
with alcian blue, Anal. Chim. Acta. 29 (1963) 145148.
[116] K.C. Lin, J.Y. Huang, S.M. Chen, Poly(brilliant cresyl blue) electrodeposited on multi-
walled carbon nanotubes modified electrode and its application for persulfate
determination, Int. J. Electrochem. Sci. 7 (2012) 91619173.
[117] M. Roushani, E. Karami, Electrochemical detection of persulfate at the modified glassy
carbon electrode with nanocomposite containing nano-ruthenium oxide/thionine and
nano-ruthenium oxide/celestine blue, Electroanalysis. 26 (2014) 17611772.
[118] M.F. De Oliveira, R.J. Mortimer, N.R. Stradiotto, Voltammetric determination of
persulfate anions using an electrode modified with a Prussian blue film, Microchem. J.
64 (2000) 155159.
[119] Z. Huang, C. Ni, F. Wang, Z. Zhu, Q. Subhani, M. Wang, Y. Zhu, Simultaneous
determination of peroxydisulfate and conventional inorganic anions by ion
chromatography with the column-switching technique, J. Sep. Sci. 37 (2014) 198203.
[120] S. Wacawek, K. Grbel, M. ernk, Simple spectrophotometric determination of
monopersulfate, Spectrochim. Acta - Part A Mol. Biomol. Spectrosc. 149 (2015) 928
933.
[121] J. Zou, J. Ma, X. Zhang, P. Xie, Rapid spectrophotometric determination of
peroxymonosulfate in water with cobalt-mediated oxidation decolorization of methyl
orange, Chem. Eng. J. 253 (2014) 3439.
[122] S. Ossadnik, G. Schwedt, Comparative study of the determination of
peroxomonosulfate, in the presence of other oxidants, by capillary zone electrophoresis,
ion chromatography, and photometry, Fresenius. J. Anal. Chem. 371 (2001) 420424.
[123] T. Zhang, H. Zhu, J.P. Crou, Production of sulfate radical from peroxymonosulfate
induced by a magnetically separable CuFe2O4 spinel in water: Efficiency, stability, and
mechanism, Environ. Sci. Technol. 47 (2013) 27842791.
[124] Y. Ding, L. Zhu, J. Yan, Q. Xiang, H. Tang, Spectrophotometric determination of
persulfate by oxidative decolorization of azo dyes for wastewater treatment., J. Environ.
Monit. 13 (2011) 30573063.
[125] J.Q. Zhang, J. Ma, J. Zou, H.Z. Chi, Y. Song, Spectrophotometric determination of
peroxymonosulfate anions via oxidative decolorization of dyes induced by cobalt, Anal.
Methods. 8 (2016) 973978.
[126] E.J. Behrman, The Persulfate Oxidation of Phenols and Arylamines (The Elbs and the
BoylandSims Oxidations), in: Org. React., John Wiley & Sons, Inc., 2004.

64
[127] G. V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical Review of rate
constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals in
aqueous solution, J. Phys. Chem. Ref. Data. 17 (1988) 513-520.
[128] G. Restelli, G. Angeletti, eds., Physico-Chemical Behaviour of Atmospheric Pollutants,
Springer Netherlands, Dordrecht, 1990.
[129] M.J. Davies, B.C. Gilbert, R.O.C. Norman, Electron spin resonance. Part 67. Oxidation
of aliphatic sulfides and sulfoxides by the sulfate radical anion and of aliphatic radicals
by the peroxydisulfate anion, J. Chem. Soc. Perkin Trans. 2 Phys. Org. Chem. (1984)
503509.
[130] W. Roebke, M. Renz, A. Henglein, Pulsradiolyse der anionen S2O82 und HSO5 in
Wssriger Lsung, Int. J. Radiat. Phys. Chem. 1 (1969) 3944.
[131] P. Maruthamuthu, P. Neta, Radiolytic chain decomposition of peroxomonophosphoric
and peroxomonosulfuric acids, J. Phys. Chem. 81 (1977) 937940.
[132] B.C. Gilbert, J.K. Stell, Mechanisms of peroxide decomposition. An ESR study of the
reactions of the peroxomonosulfate anion (HOOSO3-) with titanium(III), iron(II), and -
oxygen substituted radicals, J. Chem. Soc. Perkin Trans. 2 Phys. Org. Chem. (1990)
12811288.
[133] Y. Lei, C.-S. Chen, J. Ai, H. Lin, Y.-H. Huang, H. Zhang, Selective decolorization of
cationic dyes by peroxymonosulfate: non-radical mechanism and effect of chloride,
RSC Adv. 6 (2016) 866871.
[134] Z. Wang, R.T. Bush, L.A. Sullivan, C. Chen, J. Liu, Selective oxidation of arsenite by
peroxymonosulfate with high utilization efficiency of oxidant, Environ. Sci. Technol. 48
(2014) 39783985.
[135] W. Stumm, J.J. Morgan, Aquatic chemistry: Chemical equilibria and rates in natural
waters, John Wiley & Sons, Inc., New York, (1995).
[136] X. Xu, G. Pliego, J.A. Zazo, J.A. Casas, J.J. Rodriguez, Mineralization of naphtenic
acids with thermally-activated persulfate: The important role of oxygen, J. Hazard.
Mater. 318 (2016) 355362.
[137] H. Herrmann, On the photolysis of simple anions and neutral molecules as sources of O-
/OH, SOx- and Cl in aqueous solution., Phys. Chem. Chem. Phys. 9 (2007) 393564.
[138] C. Liang, H.-W. Su, Identification of Sulfate and Hydroxyl Radicals in Thermally
Activated Persulfate, Ind. Eng. Chem. Res. 48 (2009) 55585562.
[139] G.D. Fang, D.D. Dionysiou, D.M. Zhou, Y. Wang, X.D. Zhu, J.X. Fan, L. Cang, Y.J.
Wang, Transformation of polychlorinated biphenyls by persulfate at ambient
temperature, Chemosphere. 90 (2013) 15731580.
[140] G.D. Fang, D.D. Dionysiou, Y. Wang, S.R. Al-Abed, D.M. Zhou, Sulfate radical-based
degradation of polychlorinated biphenyls: Effects of chloride ion and reaction kinetics,
J. Hazard. Mater. 227228 (2012) 394401.
[141] P. Neta, V. Madhavan, H. Zemel, R.W. Fessenden, Rate constants and mechanism of
reaction of sulfate radical anion with aromatic compounds, J. Am. Chem. Soc. 99
(1977) 163164.
[142] A.E. Grigorev, I.E. Makarov, A.K. Pikaev, Formation of Cl2- in the bulk solution
during the radiolysis of concentrated aqueous solutions of chlorides, Khimiya Vysokikh
Ehnergij 18 (1987) 123-126.
[143] G.G. Jayson, B.J. Parsons, A.J. Swallow, Some simple, highly reactive, inorganic
chlorine derivatives in aqueous solution. Their formation using pulses of radiation and
their role in the mechanism of the Fricke dosimeter, J. Chem. Soc. Faraday Trans. 1. 69
(1973) 1597-1607.

65
[144] R.H. Waldemer, P.G. Tratnyek, R.L. Johnson, J.T. Nurmi, Oxidation of chlorinated
ethenes by heat-activated persulfate: Kinetics and products, Environ. Sci. Technol. 41
(2007) 10101015.
[145] K.-J. Kim, W.H. Hamill, Direct and indirect effects in pulse irradiated concentrated
aqueous solutions of chloride and sulfate ions, J. Phys. Chem. 80 (1976) 23202325.
[146] W.J. McElroy, A laser photolysis study of the reaction of sulfate(1-) with chloride and
the subsequent decay of chlorine(1-) in aqueous solution, J. Phys. Chem. 94 (1990)
24352441.
[147] D. Zehavi, J. Rabani, Oxidation of aqueous bromide ions by hydroxyl radicals. Pulse
radiolytic investigation, J. Phys. Chem. 76 (1972) 312319.
[148] J.L. Redpath, R.L. Willson, Chain reactions and radiosensitization: model enzyme
studies, Int. J. Radiat. Biol. Relat. Stud. Physics, Chem. Med. 27 (1975) 389398.
[149] G.V. Buxton, N.D. Wood, S. Dyster, Ionisation constants of OH and HO2 in aqueous
solution up to 200C. A pulse radiolysis study, J. Chem. Soc. Faraday Trans. 1 Phys.
Chem. Condens. Phases. 84 (1988) 1113-1121.
[150] R.E. Huie, C.L. Clifton, Temperature dependence of the rate constants for reactions of
the sulfate radical, SO4-, with anions, J. Phys. Chem. 94 (1990) 85618567.
[151] S. Padmaja, P. Neta, R.E. Huie, Rate constants for some reactions of inorganic radicals
with inorganic ions. Temperature and solvent dependence, Int. J. Chem. Kinet. 25
(1993) 445455.
[152] H. V. Lutze, S. Bircher, I. Rapp, N. Kerlin, R. Bakkour, M. Geisler, C. Von Sonntag,
T.C. Schmidt, Degradation of chlorotriazine pesticides by sulfate radicals and the
influence of organic matter, Environ. Sci. Technol. 49 (2015) 16731680.
[153] Y. Yang, J.J. Pignatello, J. Ma, W.A. Mitch, Comparison of halide impacts on the
efficiency of contaminant degradation by sulfate and hydroxyl radical-based advanced
oxidation processes (AOPs), Environ. Sci. Technol. 48 (2014) 23442351.
[154] R.O.C. Norman, P.M. Storey, P.R. West, Electron spin resonance studies. Part XXV.
Reactions of the sulphate radical anion with organic compounds, J. Chem. Soc. B.
(1970) 10871095.
[155] C. von Sonntag, H.-P. Schuchmann, Peroxyl radicals in aqueous solutions., John Wiley
& Sons, Inc., Chiehester, England, 1997.
[156] H. V. Lutze, R. Bakkour, N. Kerlin, C. von Sonntag, T.C. Schmidt, Formation of
bromate in sulfate radical based oxidation: Mechanistic aspects and suppression
bydissolved organic matter, Water Res. 53 (2014) 370377.
[157] J.Y. Fang, C. Shang, Bromate formation from bromide oxidation by the UV/persulfate
process, Environ. Sci. Technol. 46 (2012) 89768983.
[158] K. Liu, J. Lu, Y. Ji, Formation of brominated disinfection by-products and bromate in
cobalt catalyzed peroxymonosulfate oxidation of phenol, Water Res. 84 (2015) 17.
[159] P. Xie, J. Ma, W. Liu, J. Zou, S. Yue, Impact of UV/persulfate pretreatment on the
formation of disinfection byproducts during subsequent chlorination of natural organic
matter, Chem. Eng. J. 269 (2015) 203211.
[160] J. Lu, J. Wu, Y. Ji, D. Kong, Transformation of bromide in thermo activated persulfate
oxidation processes, Water Res. 78 (2015) 18.
[161] G. V. Buxton, M. Bydder, G. Arthur Salmon, K.J. Radford, C. Yamanaka, Reactivity of
chlorine atoms in aqueous solution Part 1 The equilibrium ClMNsbd+Cl- Cl2-, J. Chem.
Soc. Faraday Trans. 94 (1998) 653657.
[162] A.B. Ross, P. Neta, Rate constants for reactions of inorganic radicals in aqueous
solution, U.S. Dept. of Commerce, National Bureau of Standards, U.S. Govt. Print. Off.,
1979. https://books.google.pl/books?id=l34-AAAAIAAJ.

66
[163] R.A. Larson, R.G. Zepp, Reactivity of the carbonate radical with aniline derivatives,
Environ. Toxicol. Chem. 7 (1988) 265274.
[164] S. Canonica, T. Kohn, M. Mac, F.J. Real, J. Wirz, U. Von Gunten, Photosensitizer
method to determine rate constants for the reaction of carbonate radical with organic
compounds, Environ. Sci. Technol. 39 (2005) 91829188.
[165] P. Neta, R.E. Huie, A.B. Ross, Rate constants for reactions of inorganic radicals in
aqueous solution, J. Phys. Chem. Ref. Data. 17 (1988) 10271284.
[166] D.M. Schwarzenbach, R.P., Gschwend, P.M. and Imboden, Photochemical
transformation reactions. in: Environ. Org. Chem., John Wiley & Sons, Inc., New York,
(1993) 436484.
[167] I.A. Katsoyiannis, S. Canonica, U. von Gunten, Efficiency and energy requirements for
the transformation of organic micropollutants by ozone, O3/H2O2 and UV/H2O2, Water
Res. 45 (2011) 38113822.
[168] U. von Gunten, Ozonation of drinking water: Part II. Disinfection and by-product
formation in presence of bromide, iodide or chlorine, Water Res. 37 (2003) 14691487.
[169] Y. Liu, A. Zhou, Y. Gan, X. Li, Variability in carbon isotope fractionation of
trichloroethene during degradation by persulfate activated with zero-valent iron: Effects
of inorganic anions., Sci. Total Environ. 548549 (2016) 15.
[170] J. Lu, W. Dong, Y. Ji, D. Kong, Q. Huang, Natural organic matter exposed to sulfate
radicals increases its potential to form halogenated disinfection byproducts., Environ.
Sci. Technol. 50 (2016) 5060-5067.
[171] Q. Zhang, J. Chen, C. Dai, Y. Zhang, X. Zhou, Degradation of carbamazepine and
toxicity evaluation using the UV/persulfate process in aqueous solution, J. Chem.
Technol. Biotechnol. 90 (2015) 701708.
[172] K. Temiz, T. Olmez-Hanci, I. Arslan-Alaton, Zero-valent iron-activated persulfate
oxidation of a commercial alkyl phenol polyethoxylate., Environ. Technol. 37 (2016)
17571767.
[173] T. Olmez-Hanci, I. Arslan-Alaton, D. Dursun, B. Genc, D.G. Mita, M. Guida, L. Mita,
Degradation and toxicity assessment of the nonionic surfactant TritonTM X-45 by the
peroxymonosulfate/UV-C process., Photochem. Photobiol. Sci. 14 (2015) 569575.
[174] H. Hori, A. Yamamoto, E. Hayakawa, S. Taniyasu, N. Yamashita, S. Kutsuna, H.
Kiatagawa, R. Arakawa, Efficient decomposition of environmentally persistent
perfluorocarboxylic acids by use of persulfate as a photochemical oxidant, Environ. Sci.
Technol. 39 (2005) 23832388.
[175] Y.C. Lee, S.L. Lo, P. Te Chiueh, Y.H. Liou, M.L. Chen, Microwave-hydrothermal
decomposition of perfluorooctanoic acid in water by iron-activated persulfate oxidation,
Water Res. 44 (2010) 886892.
[176] Y. Lee, S. Lo, J. Kuo, C. Hsieh, Decomposition of perfluorooctanoic acid by
microwaveactivated persulfate: Effects of temperature, pH, and chloride ions, Front.
Environ. Sci. Eng. 6 (2011) 1725.
[177] Y.C. Lee, S.L. Lo, J. Kuo, Y.L. Lin, Persulfate oxidation of perfluorooctanoic acid
under the temperatures of 20-40C, Chem. Eng. J. 198199 (2012) 2732.
[178] Y.C. Lee, S.L. Lo, J. Kuo, C.P. Huang, Promoted degradation of perfluorooctanic acid
by persulfate when adding activated carbon, J. Hazard. Mater. 261 (2013) 463469.
[179] C.S. Liu, C.P. Higgins, F. Wang, K. Shih, Effect of temperature on oxidative
transformation of perfluorooctanoic acid (PFOA) by persulfate activation in water, Sep.
Purif. Technol. 91 (2012) 4651.
[180] Y. Qian, X. Guo, Y. Zhang, Y. Peng, P. Sun, C.-H. Huang, J. Niu, X. Zhou, J.C.
Crittenden, Perfluorooctanoic acid degradation using UV-persulfate process: modeling
of the degradation and chlorate formation, Environ. Sci. Technol. 50 (2016) 772781.

67
[181] Y. Wang, C.S. Hong, Effect of hydrogen peroxide, periodate and persulfate on
photocatalysis of 2-chlorobiphenyl in aqueous TiO2 suspensions, Water Res. 33 (1999)
20312036.
[182] J. Sharma, I.M. Mishra, D.D. Dionysiou, V. Kumar, Oxidative removal of bisphenol A
by UV-C/peroxymonosulfate (PMS): Kinetics, influence of co-existing chemicals and
degradation pathway, Chem. Eng. J. 276 (2015) 193204.
[183] M.K. Sahoo, B. Sinha, M. Marbaniang, D.B. Naik, R.N. Sharan, Mineralization of
Calcon by UV/oxidant systems and assessment of biotoxicity of the treated solutions by
E. coli colony forming unit assay, Chem. Eng. J. 181 (2012) 206214.
[184] S. Wacawek, V. Anto, P. Hrabk, M. ernk, Remediation of hexachlorocyclohexanes
by cobalt-mediated activation of peroxymonosulfate, Desalin. Water Treat. 57 (2016)
26274-26279.
[185] M.K. Sahoo, M. Marbaniang, B. Sinha, D.B. Naik, R.N. Sharan, UVC induced TOC
removal studies of Ponceau S in the presence of oxidants: Evaluation of electrical
energy efficiency and assessment of biotoxicity of the treated solutions by Escherichia
coli colony forming unit assay, Chem. Eng. J. 213 (2012) 142149.
[186] K. Selvam, M. Muruganandham, I. Muthuvel, M. Swaminathan, The influence of
inorganic oxidants and metal ions on semiconductor sensitized photodegradation of 4-
fluorophenol, Chem. Eng. J. 128 (2007) 5157.
[187] L. Ravichandran, K. Selvam, M. Swaminathan, Effect of oxidants and metal ions on
photodefluoridation of pentafluorobenzoic acid with ZnO, Sep. Purif. Technol. 56
(2007) 192198.
[188] C.H. Yu, C.H. Wu, T.H. Ho, P.K. Andy Hong, Decolorization of C.I. Reactive Black 5
in UV/TiO2, UV/oxidant and UV/TiO2/oxidant systems: A comparative study, Chem.
Eng. J. 158 (2010) 578583.
[189] H. Eskandarloo, A. Badiei, M.A. Behnajady, Optimization of UV/inorganic oxidants
system efficiency for photooxidative removal of an azo textile dye, Desalin. Water
Treat. 55 (2015) 210226.
[190] A. Ghauch, A.M. Tuqan, N. Kibbi, S. Geryes, Methylene blue discoloration by heated
persulfate in aqueous solution, Chem. Eng. J. 213 (2012) 259271.
[191] N.S. Shah, X. He, H.M. Khan, J.A. Khan, K.E. OShea, D.L. Boccelli, D.D. Dionysiou,
Efficient removal of endosulfan from aqueous solution by UV-C/peroxides: A
comparative study, J. Hazard. Mater. 263 (2013) 584592.
[192] C. Tan, N. Gao, Y. Deng, Y. Zhang, M. Sui, J. Deng, S. Zhou, Degradation of
antipyrine by UV, UV/H2O2 and UV/PS, J. Hazard. Mater. 260 (2013) 1008-1016.
[193] T. Olmez-Hanci, C. Imren, I. Kabdash, O. Tnay, I. Arslan-Alaton, Application of the
UV-C photo-assisted peroxymonosulfate oxidation for the mineralization of dimethyl
phthalate in aqueous solutions, Photochem. Photobiol. Sci. 10 (2011) 343349. \
[194] A.R. Zarei, H. Rezaeivahidian, A.R. Soleymani, Mineralization of unsymmetrical
dimethylhydrazine (UDMH) via persulfate activated by zero valent iron nano particles:
modeling, optimization and cost estimation, Desalin. Water Treat. 57 (2016) 16119
16128.
[195] A. Kambhu, S. Comfort, C. Chokejaroenrat, C. Sakulthaew, Developing slow-release
persulfate candles to treat BTEX contaminated groundwater, Chemosphere. 89 (2012)
656664.
[196] Y.T. Lin, C. Liang, J.H. Chen, Feasibility study of ultraviolet activated persulfate
oxidation of phenol, Chemosphere. 82 (2011) 11681172.
[197] J. Chen, Y. Qian, H. Liu, T. Huang, Oxidative degradation of diclofenac by thermally
activated persulfate: implication for ISCO., Environ. Sci. Pollut. Res. Int. 23 (2016)
38243833.

68
[198] P. Manoj, K.P. Prasanthkumar, V.M. Manoj, U.K. Aravind, T.K. Manojkumar, C.T.
Aravindakumar, Oxidation of substituted triazines by sulfate radical anion (SO4-) in
aqueous medium: a laser flash photolysis and steady state radiolysis study, J. Phys. Org.
Chem. 20 (2007) 122129.
[199] J.A. Khan, X. He, N.S. Shah, H.M. Khan, E. Hapeshi, D. Fatta-Kassinos, D.D.
Dionysiou, Kinetic and mechanism investigation on the photochemical degradation of
atrazine with activated H2O2, S2O82- and HSO5-, Chem. Eng. J. 252 (2014) 393403.
[200] M.E.D.G. Azenha, H.D. Burrows, M. Canle L., R. Coimbra, M.I. Fernndez, M. V
Garca, M.A. Peiteado, J.A. Santaballa, Kinetic and mechanistic aspects of the direct
photodegradation of atrazine, atraton, ametryn and 2-hydroxyatrazine by 254 nm light
in aqueous solution, J. Phys. Org. Chem. 16 (2003) 498503.
[201] B. Neppolian, E. Celik, H. Choi, Photochemical Oxidation of Arsenic(III) to Arsenic(V)
using Peroxydisulfate Ions as an Oxidizing Agent, Environ. Sci. Technol. 42 (2008)
61796184.
[202] Z.H. Diao, X.R. Xu, H. Chen, D. Jiang, Y.X. Yang, L.J. Kong, Y.X. Sun, Y.X. Hu,
Q.W. Hao, L. Liu, Simultaneous removal of Cr(VI) and phenol by persulfate activated
with bentonite-supported nanoscale zero-valent iron: Reactivity and mechanism, J.
Hazard. Mater. 316 (2016) 186193.
[203] Z.H. Diao, X.R. Xu, D. Jiang, L.J. Kong, Y.X. Sun, Y.X. Hu, Q.W. Hao, H. Chen,
Bentonite-supported nanoscale zero-valent iron/persulfate system for the simultaneous
removal of Cr(VI) and phenol from aqueous solutions, Chem. Eng. J. 302 (2016) 213
222.
[204] J. Gerritse, V. Renard, J. Visser, J.C. Gottschal, Complete degradation of
tetrachloroethene by combining anaerobic dechlorinating and aerobic methanotrophic
enrichment cultures, Appl. Microbiol. Biotechnol. 43 (1995) 920928.
[205] A. Hirvonen, T. Tuhkanen, P. Kalliokoski, Treatment of TCE- and PCE-contaminated
groundwater using UV/H2O2 and O3/H2O2 oxidation processes, Water Sci. Technol. 33
(1996) 6773.
[206] N. Yan, F. Liu, Q. Xue, M.L. Brusseau, Y. Liu, J. Wang, Degradation of trichloroethene
by siderite-catalyzed hydrogen peroxide and persulfate: Investigation of reaction
mechanisms and degradation products, Chem. Eng. J. 274 (2015) 6168.
[207] M. Xu, H. Du, X. Gu, S. Lu, Z. Qiu, Q. Sui, Generation and intensity of active oxygen
species in thermally activated persulfate systems for the degradation of
trichloroethylene, RSC Adv. 4 (2014) 4051140517.
[208] B.M. Sharma, G.K. Bharat, S. Tayal, L. Nizzetto, P. Cupr, T. Larssen, Environment and
human exposure to persistent organic pollutants (POPs) in India: a systematic review of
recent and historical data., Environ. Int. 66 (2014) 4864.
[209] W. Qin, G. Fang, Y. Wang, T. Wu, C. Zhu, D. Zhou, Efficient transformation of DDT
by peroxymonosulfate activated with cobalt in aqueous systems: Kinetics, products, and
reactive species identification, Chemosphere. 148 (2016) 6876.
[210] C. Zhu, G. Fang, D.D. Dionysiou, C. Liu, J. Gao, W. Qin, D. Zhou, Efficient
transformation of DDTs with persulfate activation by zero-valent iron nanoparticles: a
mechanistic study, J. Hazard. Mater. 316 (2016) 232-241.
[211] A. Tauber, C. von Sonntag, products and kinetics of the OH-radical-induced
dealkylation of atrazine, Acta Hydrochim. Hydrobiol. 28 (2000) 1523.
[212] J.L. Acero, K. Stemmler, U. Von Gunten, Degradation kinetics of atrazine and its
degradation products with ozone and OH radicals: A predictive tool for drinking water
treatment, Environ. Sci. Technol. 34 (2000) 591597.

69
[213] M.J. Shipitalo, L.B. Owens, Atrazine, deethylatrazine, and deisopropylatrazine in
surface runoff from conservation tilled watersheds., Environ. Sci. Technol. 37 (2003)
944950.
[214] J. De Laat, N. Chramosta, M. Dor, H. Suty, M. Pouillot, Constantes cintiques de
raction des radicaux hydroxyles sur quelques sous produits doxydation de latrazine
par O3 ou par O3/H2O2, Environ. Technol. 15 (1994) 419428.
[215] S. Khan, X. He, J.A. Khan, H.M. Khan, D.L. Boccelli, D.D. Dionysiou, Kinetics and
mechanism of sulfate radical- and hydroxyl radical-induced degradation of highly
chlorinated pesticide lindane in UV/peroxymonosulfate system, Chem. Eng. J. 318
(2016) 135-142.
[216] K. Kumierek, L. Dbek, A. witkowski, A comparative study on oxidative
degradation of 2,4-dichlorophenol and 2,4-dichlorophenoxyacetic acid by ammonium
persulfate, Desalin. Water Treat. 57 (2016) 10981106.
[217] J.M. Monteagudo, A. Durn, R. Gonzlez, A.J. Expsito, In situ chemical oxidation of
carbamazepine solutions using persulfate simultaneously activated by heat energy, UV
light, Fe2+ ions, and H2O2, Appl. Catal. B Environ. 176177 (2015) 120129.
[218] X. Liu, X. Zhang, K. Shao, C. Lin, C. Li, F. Ge, Y. Dong, Fe0-activated persulfate-
assisted mechanochemical destruction of expired compound sulfamethoxazole tablets,
RSC Adv. 6 (2016) 2093820948.
[219] Z. Li, Q. Yang, Y. Zhong, X. Li, L. Zhou, X. Li, G. Zeng, Granular activated carbon
supported iron as a heterogeneous persulfate catalyst for the pretreatment of mature
landfill leachate, RSC Adv. 6 (2016) 987994.
[220] J. Kronholm, M.L. Riekkola, Potassium persulfate as oxidant in pressurized hot water,
Environ. Sci. Technol. 33 (1999) 20952099.
[221] S.A.A. Nakhli, Biological removal of phenol from saline wastewater using a moving
bed biofilm reactor containing acclimated mixed consortia, Springerplus. 3 (2014).
[222] A. Rubalcaba, M.E. Surez-Ojeda, F. Stber, A. Fortuny, C. Bengoa, I. Metcalfe, J.
Font, J. Carrera, A. Fabregat, Phenol wastewater remediation: advanced oxidation
processes coupled to a biological treatment., Water Sci. Technol. 55 (2007) 2217.
[223] A.A. Babaei, F. Ghanbari, COD removal from petrochemical wastewater by
UV/hydrogen peroxide, UV/persulfate and UV/percarbonate: biodegradability
improvement and cost evaluation, J. Water Reuse Desalin. 7 (2016).
[224] E. Kattel, N. Dulova, M. Viisimaa, T. Tenno, M. Trapido, Treatment of high-strength
wastewater by Fe2+-activated persulphate and hydrogen peroxide, Environ. Technol. 37
(2016) 352-359.
[225] S.O. Ganiyu, E.D. van Hullebusch, M. Cretin, G. Esposito, M.A. Oturan, Coupling of
membrane filtration and advanced oxidation processes for removal of pharmaceutical
residues: A critical review, Sep. Purif. Technol. 156 (2015) 891914.
[226] M.A. Fagier, E.A. Ali, K.S. Tay, M.R.B. Abas, Mineralization of organic matter from
vinasse using physicochemical treatment coupled with Fe2+-activated persulfate and
peroxymonosulfate oxidation, Int. J. Environ. Sci. Technol. 13 (2016) 11891194.
[227] J. Rodrguez-Chueca, C. Amor, T. Silva, D.D. Dionysiou, G. Li Puma, M.S. Lucas, J.A.
Peres, Treatment of winery wastewater by sulphate radicals: HSO5 /transition
metal/UV-A LEDs, Chem. Eng. J. 310 (2017) 473483.
[228] R. Ocampo-Prez, M. Snchez-Polo, J. Rivera-Utrilla, R. Leyva-Ramos, Degradation of
antineoplastic cytarabine in aqueous phase by advanced oxidation processes based on
ultraviolet radiation, Chem. Eng. J. 165 (2010) 581588.
[229] H.-Y. Shu, M.-C. Chang, S.-W. Huang, Decolorization and mineralization of azo dye
Acid Blue 113 by the UV/Oxone process and optimization of operating parameters,
Desalin. Water Treat. 57 (2016) 79517962.

70
[230] G. Zhen, X. Lu, B. Wang, Y. Zhao, X. Chai, D. Niu, A. Zhao, Y. Li, Y. Song, X. Cao,
Synergetic pretreatment of waste activated sludge by Fe(II)-activated persulfate
oxidation under mild temperature for enhanced dewaterability, Bioresour. Technol. 124
(2012) 2936.
[231] G.W. Chen, W.W. Lin, D.J. Lee, Capillary suction time (CST) as a measure of sludge
dewaterability, Water Sci. Technol. 34 (1996) 443448.
[232] G. Zhen, X. Lu, Y. Zhao, X. Chai, D. Niu, Enhanced dewaterability of sewage sludge in
the presence of Fe(II)-activated persulfate oxidation, Bioresour. Technol. 116 (2012)
259265.
[233] Y. Shi, J. Yang, W. Yu, S. Zhang, S. Liang, J. Song, Q. Xu, N. Ye, S. He, C. Yang, J.
Hu, Synergetic conditioning of sewage sludge via Fe2+/persulfate and skeleton builder:
Effect on sludge characteristics and dewaterability, Chem. Eng. J. 270 (2015) 572581.
[234] W.S. Chen, Y.C. Jhou, C.P. Huang, Mineralization of dinitrotoluenes in industrial
wastewater by electro-activated persulfate oxidation, Chem. Eng. J. 252 (2014) 166
172.
[235] G.Y. Zhen, X.Q. Lu, Y.Y. Li, Y.C. Zhao, Innovative combination of electrolysis and
Fe(II)-activated persulfate oxidation for improving the dewaterability of waste activated
sludge, Bioresour. Technol. 136 (2013) 664663.
[236] S. Wacawek, K. Grbel, Z. Chad, M. Dudziak, Impact of peroxydisulphate on
disintegration and sedimentation properties of municipal wastewater activated sludge,
Chem. Pap. 69 (2015) 14731480.
[237] Y. Yang, K. Tsukahara, R. Yang, Z. Zhang, S. Sawayama, Enhancement on
biodegradation and anaerobic digestion efficiency of activated sludge using a dual
irradiation process, Bioresour. Technol. 102 (2011) 1076710771.
[238] S. Wacawek, K. Grbel, M. ernk, The impact of peroxydisulphate and
peroxymonosulphate on disintegration and settleability of activated sludge, Environ.
Technol. 37 (2016) 12961304.
[239] S. Wacawek, K. Grbel, Z. Chd, M. Dudziak, M. ernk, The impact of oxone on
disintegration and dewaterability of waste activated sludge, Water Environ. Res. 88
(2016) 152157.
[240] T. Niu, Z. Zhou, W. Ren, L.-M. Jiang, B. Li, H. Wei, J. Li, L. Wang, Effects of
potassium peroxymonosulfate on disintegration of waste sludge and properties of
extracellular polymeric substances, Int. Biodeterior. Biodegradation. 106 (2016) 170
177.
[241] C. Liu, B. Wu, X. Chen, S. Xie, Waste activated sludge pretreatment by Fe(II)-activated
peroxymonosulfate oxidation under mild temperature, Chem. Pap. (2017) 19.
[242] G. Zhen, X. Lu, J. Niu, L. Su, X. Chai, Y. Zhao, Y.Y. Li, Y. Song, D. Niu, Inhibitory
effects of a shock load of Fe(II)-mediated persulfate oxidation on waste activated sludge
anaerobic digestion, Chem. Eng. J. 233 (2013) 274281.
[243] D. Sun, H. Liang, C. Ma, Enhancement of sewage sludge anaerobic digestibility by
sulfate radical pretreatment, Adv. Mater. Res. (Durnten-Zurich, Switz.). 518523 (2012)
33583362.
[244] S. Wacawek, K. Grbel, P. Dennis, V.T.P. Vinod, M. ernik, A novel approach for
simultaneous improvement of dewaterability, post-digestion liquor properties and
toluene removal from anaerobically digested sludge, Chem. Eng. J. 291 (2016) 192
198.
[245] K. Song, X. Zhou, Y. Liu, Y. Gong, B. Zhou, D. Wang, Q. Wang, Role of oxidants in
enhancing dewaterability of anaerobically digested sludge through Fe (II) activated
oxidation processes: hydrogen peroxide versus persulfate., Sci. Rep. 6 (2016) 24800.

71
 Chemical and physical properties of persulfates were described in detail
 Up-to-date list of persulfates activation and determination methods was created
 Direct oxidation, radical oxidation and matrix effects were discussed
 Recent developments in the use of persulfates in waters and wastewaters were shown

72
73

Anda mungkin juga menyukai