You are on page 1of 19

Chemical Engineering Journal 330 (2017) 44–62

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Review

Chemistry of persulfates in water and wastewater treatment: A review MARK


a,⁎ b,c,d e a
Stanisław Wacławek , Holger V. Lutze , Klaudiusz Grübel , Vinod V.T. Padil ,
Miroslav Černíka, Dionysios.D. Dionysiouf,⁎
a
Centre for Nanomaterials, Advanced Technologies and Innovation, Technical University of Liberec, Studentská 1402/2, 461 17 Liberec 1, Czech Republic
b
University Duisburg-Essen, Faculty of Chemistry, Instrumental Analytical Chemistry, Universitaetsstraβe 5, D-45141 Essen, Germany
c
IWW Water Centre, Moritzstr. 26, D-45476 Mülheim an der Ruhr, Germany
d
Centre for Water and Environmental Research (ZWU) Universitätsstr. 5, D-45141 Essen, Germany
e
Institute of Environmental Protection and Engineering, Department of Environmental Microbiology and Biotechnology, University of Bielsko-Biala, Willowa 2, 43-309
Bielsko-Biała, Poland
f
Environmental Engineering and Science Program, University of Cincinnati, 705 Engineering Research Center, Cincinnati, OH 45221-0012, USA

G RA P H I C A L AB S T R A C T

A R T I C L E I N F O A B S T R A C T

Keywords: Persulfate decontamination technologies either utilizing radical driven processes or direct electron transfer are
Persulfates very powerful tools for the treatment of a broad range of impurities, including halogenated olefins, BTEXs
Peroxydisulfate (benzene, toluene, ethylbenzene and xylenes), perfluorinated chemicals, phenols, pharmaceuticals, inorganics
Peroxymonosulfate and pesticides. Furthermore, the reactivity of persulfates is extremely dependent on the related activation
Remediation
techniques and the composition of the treated water matrix. Direct reactions of peroxydisulfate (PDS) or per-
Oxidation
Sulfate radical
oxymonosulfate (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.

1. Introduction 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
Over the last few decades, ubiquitous contamination with various pathogens [3]. In many countries with a low human development index
inorganic and organic substances has caused serious problems all over (HDI), particularly in Africa and Asia, surface- and groundwater


Corresponding authors.
E-mail addresses: stanislaw.waclawek@tul.cz (S. Wacławek), dionysios.d.dionysiou@uc.edu (D.D. Dionysiou).

http://dx.doi.org/10.1016/j.cej.2017.07.132
Received 5 June 2017; Received in revised form 18 July 2017; Accepted 19 July 2017
Available online 21 July 2017
1385-8947/ © 2017 Elsevier B.V. All rights reserved.
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

contamination can be a direct cause of water scarcity. It is important, Table 1


therefore, to develop cheap and efficient water and wastewater treat- Various properties of common oxidants used for in situ remediation.
ment methods [4].
Oxidant OeO bond- Solubility Average Price Price
There are many different water and wastewater decontamination dissociation in water at estimated [USD [USD
methods, which can be universally divided into biological [5], physical energy [kJ 25 °C [g lifetime in kg−1]2 mol−1]
[6], and chemical methods [7]. mol−1] L−1] groundwater4
Oxidants used for environmental purposes can be divided into ones
H2O2 213 soluble hours to days 1.5 0.05
that possess a peroxide bond and ones that do not. Oxidizing agents O3 364 0.1 <1 h 2.3 0.11
such as ozone and hypochlorite have long been employed for water and Potassium – 0.7–33 unstable in 2000 396
wastewater treatment [8,9]. Oxidants possessing the OeO bond (known ferrate water at
as peroxide or the peroxo group), typically form free radicals which in pH ≠ 9
PDS5 926 730 >5 months 0.74 0.18
turn may result in pollutant degradation.
PMS 3776 298 hours to days 2.21 1.36
Peroxydisulfate (PDS) and peroxymonosulfate (PMS) are gaining
increasing attention in water and wastewater treatment [10–12] The remaining data was taken from [16,19].
1
(Fig. 1). Many studies have shown that they are capable of degrading PMS is herein considered as the Oxone® salt (2KHSO5·KHSO4·K2SO4).
2
highly toxic and persistent pollutants, e.g. polychlorinated biphenyls 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#.
(PCBs) [13,14], and are relatively cheap in comparison to other oxi-
Vu2SzOZxBqF; [21].
dants [15] (Table 1). 3
[22].
This paper reviews the chemical and physical properties of persul- 4
Depends on water hardness, transition metal concentration, dissolved organic carbon
fates in addition to their determination and activation methods. (DOC) concentration, and many other factors.
5
Moreover, the direct oxidation reactions of persulfates as well as oxi- Sodium PDS.
6
[23].
dation with generated radicals are described.

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 elec-
trolysis 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 am-
monium PDS can cause residual ammonia and secondary contamina-
tion. 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 °C) [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 Fig. 2. (a) Sodium PDS and (b) Oxone triple salt molecular structure (*potassium PMS
applications. The bond dissociation energy and bond length of [O3SO- (KHSO5) - active part of Oxone; bond lengths were taken from [28]).
OSO3]2− were respectively determined to be 140 kJ mol−1 and 1.497 Å
[15,16,19].
potential applicability [25–27]. In comparison to PDS, PMS has a
Peroxymonosulfate (PMS; monopersulfate) has its origin from the
shorter bond length (1.46 Å), which translates to a higher bond dis-
Caro’s acid (H2SO5) also known as persulfuric, peroxymonosulfuric, and
sociation energy (377 kJ mol−1). In other words, in theory, PMS re-
peroxysulfuric acid [24]. It possesses reactive oxygen closest to the
quires more energy to produce radicals in the homolytic cleavage of the
hydrogen atom that carries the active power (Fig. 2b). Since the oxygen
peroxide bond.
is readily reactive, KHSO5 is usually found in a more stable form of a
The cost of Oxone is the highest of all of the conventionally used
white triple salt, 2KHSO5·KHSO4·K2SO4 (potassium hydrogen PMS).
oxidants in ISCO, although it is still much cheaper than ferrates [21].
Commercially, the product goes under the trade-name of Oxone® and
Persulfates are very stable in the solid state and remain stable for
has a high solubility in water, is safe to handle, hence it has a good

Fig. 1. Annual number of a) publications concerning PDS or PMS and b) citations. Source: Web of knowledge (data as of July 2017).

45
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

several months in pure water. One of the advantages of PDS in com- long term application.
parison to the other commonly used ISCO reagents (hydrogen peroxide A detailed summary of the activation methods of PDS and PMS is
or ozone) is its much higher stability i.e. smaller transportation issues; presented in Tables 2 and 3, respectively.
elevated concentrations of PDS can be pumped into the contaminated The application of PMS and PDS in oxidation treatment belongs to
area, and it could be moved by the density driven dispersion through advanced oxidation processes (AOPs) because of the free radicals pre-
the treated soil [12]. sence [67].
Yen et al. [14] determined that PDS anions can persist in the soil AOPs can be either homogeneous or heterogeneous. Whilst homo-
system for over five months. PMS does not show such a high persistence genous AOPs are the most efficacious at an acidic pH, heterogeneous
as PDS, although when being stored in a dry and cold place, it loses only advanced oxidation processes could function in a wide pH spectrum
around 1% activity/month due to the oxygen and heat release (http:// [68,69]. Despite the fact that heterogeneous AOPs exhibit numerous
www2.dupont.com/Oxone/en_US/index.html). Under the influence of benefits in comparison to homogenous ones, the point that the catalyst
elevated temperatures (>300 °C) it decomposes to SO2 and SO3. PMS is creation is not easy nor economical should also be considered [60]
also not stable in water with higher pH values (at pH 9 the stability (however there are some exceptions of this rule [51]).
reaches minimum), whereby concentration of its protonated form
(HSO5−) equals the concentration of the unprotonated one (SO52−) 3.1. Homogeneous activation process
[29]. More described information on the stability of PMS at various pH
values can be found in Bouchard et al. [29]. In the case of ISCO, activation should be slow to allow the long-term
Peroxydisulfate and peroxymonosulfate are among the strongest formation of radicals and the treatment of heterogeneously distributed
oxidizing agents that are applied in environmental remediation. The contaminants. For homogenous radical based processes, one of the most
standard oxidation–reduction potential (ORP) for the reduction of PDS common methods is to use chelated transition metals, which sig-
anion (Eq. (1)) equals to 2.01 V, hence it is higher than that of PMS i.e. nificantly decreases the metal concentration required for the activation
1.4 V (Eq. (2)) [30,31]. [70,71]. Rastogi et al. [72] assessed the efficacy of several chelating
agents i.e. ethylenediaminedisuccinate, citrate, and pyrophosphate on
S2O82 − + 2H+ + 2e− → 2HSO−4 (1) ferrous iron activation of PMS and PDS under neutral pH conditions.
HSO− + − − They found that PMS was the all-embracing oxidizing agent catalysed
5 + 2H + 2e → HSO4 + H2O (2)
by all ferrous iron-chelates, whereas citrate coupled with ferrous iron
In case persulfates are used for pollutant degradation, the radicals was the best chelate/iron configuration.
formed upon cleavage of the peroxide bond are the most important. The
radical formation can be initiated by different means, i.e., photo- 3.2. Heterogeneous activation processes with metal catalysts
chemical or thermal cleavage of the peroxide bond or chemical re-
duction. PDS typically decays in a radical pathways often resulting in On the other hand, heterogeneous processes with material (often in
sulfate radicals. However, PMS may also decompose into sulfate radical nano-scale) synthesized beforehand, have drawn much attention re-
and hydroxyl radical [32]. cently due to their excellent performance. In addition, as mentioned
above, cobalt has been found to be the very best homogenous activator
3. Activation mechanism for PMS. Therefore, heterogeneous catalysts often also contain cobalt
(especially the ones that are used for peroxymonosulfate activation).
The key in PDS and PMS-based oxidation is formation of highly Co3O4 was used for instance by Chen et al. [73] and Muhammad et al.
reactive species, which themselves have a potential to degrade pollu- [74] as a catalyst for PMS to degrade Acid Orange 7 (AO7) and phenol,
tants. This can be achieved mainly by thermal [33–36], photolytic [37], respectively (see also [75–78]). Interestingly, Zhang et al. [79] syn-
sonolytic [38], radiolytic activation [39], as well as by the reactions of thesized nanoscale Co3O4 and applied it in heterogeneous activation of
PDS and PMS with iron oxide magnetic composites [40,41] including as PDS. The highest degradation rate of Orange G (as a model compound)
well in situ formed iron hydroxides [42] and quinones [43]. Further- was at pH ∼ 7, where the catalyst dissolution is very low. The authors
more, activation of persulfates by hydroxide ions [44,45] in ozonation also confirmed that the hydroxyl and sulfate radicals were the main
[46] or in presence of phenol [47] was described. Moreover, sulfate oxidants.
radicals could be also generated by chemical reduction of PMS and PDS One of the major problems of the activation of persulfates with
using low valent transition metals as reductant [48,49] including also Co3O4 is an excessive quantity of cobalt leaching. To improve catalytic
bimetallic or trimetallic iron-based systems as well as iron scrap capacity and decrease leaching, cobalt can be attached to several sup-
[50,51]. Photo and radiodissociation or thermal activation of PMS ports (most often metal oxide) i.e. Al2O3, TiO2, SiO2, MgO or activated
precedes the evolution of both SO4%− and %OH [49,52]. The formation carbon/carbon aerogel or graphene. For example, Yang et al. [78]
of SO4%− as the main oxidizing species can be performed only via confirmed that the TiO2 support increased the content of –OH groups on
transition-metal catalysis (although the interchange of radicals in water the Co/TiO2 catalysts, enhancing the generation of the Co-OH complex,
solutions can be also notable and will be further discussed in Section that is believed to be a crucial stage in heterogeneous activation of PMS
5.3). In the group of the transition element ions (Fe2+, Co2+, Ni2+, (Eq. (3)).
Cu2+, Ru3+, Ag+), divalent cobalt was determined to be the most ef- In another study, Yao et al. [80] used bimetallic oxides CoFe2O4 and
ficient catalyst for PMS according to Anipsitakis and Dionysiou [53] CoFe2O4− on graphene to catalyse PMS and thus abate phenol. Su et al.
and Fernandez et al. [54]. As for PDS, similarly to the Fenton reaction, [81] successfully synthesized heterogeneous CoxFe3−xO4 material and
Fe2+ is the most commonly used metal for the homogeneous catalysis. discovered that the iron-cobalt interactions are vital for effective het-
The type and dose of the transition metal catalysis is of significant erogeneous catalysis of PMS. Lin et al. [82] demonstrated an alternative
importance, since when it is applied in excess, scavenging of sulfate path to synthesize a magnetic nano-Co-graphene composite from car-
radicals may become a problem [55], this problem however can be bonizing of a cobalt-based metal organic frameworks and graphene
solved by adopting the use of solid iron particles where the release of oxide, and used it to activate PMS. In order to measure the long-run
iron species responsible for PDS activation occurred smoothly without catalytic activity, a 50-cycle decolorization of Acid yellow 17 was
the risk of sulfate radical quenching as reported by Naim and Ghauch performed and the efficacy of degradation remained at 97.6%, dis-
[51]. Ayoub and Ghauch [50] demonstrated as well that PDS activation playing its stability and efficient catalytic activity. In their further
can be better sustained in solution especially while using bimetallic and study, a nanocomposite was prepared by the carbonization (one-step) of
trimetallic iron-based systems making the process more efficient toward a Co-based zeolitic imidazolate [83]. Higher cobalt/carbon

46
Table 2
PDS activation methods.

Method Mechanism Predominant radical species Comments Reference


S. Wacławek et al.

Heat Homolysis of peroxide bond Sulfate radical/Hydroxyl radical (in Because of the low bond-dissociation energy, often low temperature [35,36,56]
presence of chloride / at elevated increase can effectively cleave OeO bond
pH)
UV radiation Homolysis of peroxide bond Sulfate radical Often used λ = 254 nm, Quantum yield 1.4 (depends on dissolved [57–59]
oxygen concentration)
Homogenous: Transition metals One electron transfer Sulfate radical Often requires low pH [48]
Heterogeneous: Transition metals One electron transfer Sulfate radical Preparation of catalyst is not economical [50,60]
(mono, bi, and trimetallic
systems)
Chelated transition metals One electron transfer Sulfate radical Stabilizes ferrous iron at neutral pH, widely used in situ [61]
Alkaline pH Base-catalysed hydrolysis of PDS to hydroperoxide, which later initiates Sulfate radical/Hydroxyl radical/ Often pH > 11 [44]
radical formation Superoxide radical
Electrolysis Ferrous iron generated in the chemical and electrochemical corrosion of Sulfate radical/Hydroxyl radical Additional Fe3+ reduction on the cathode. Yuan et al. [62] claim that OH [62]
iron radical contribution in this system is more significant
Nanocarbons Peroxide bond of PDS is debilitated on the defective edges and Hydroxyl radical Reduced mesoporous carbon, carbon nanotubes, and graphene oxide, [15,63]
oxygengroups (from which the carbonyl group was found as the most displayed great catalyticproperties, in contrary to nanodiamonds,
active one) of carbocatalysts, than persulfate oxidizes adsorbed water or fullerenes and graphitic carbon nitride
hydroxide ions
Other organics One electron transfer Sulfate radical Low molecular weight, anionic organic compounds [43,47]
Radiolysis Reaction with the solvated electron Sulfate radical Electron beam irradiation of aqueous solution [39]

47
Table 3
PMS activation methods.

Method Mechanism Predominant radical species Comments Reference

Heat Homolysis of peroxide bond Sulfate radical/ Hydroxyl Higher temperatures are required to split the OeO bond, due to a higher bond-dissociation energy in [49]
radical comparison to S2O82−
UV radiation Homolysis of peroxide bond Sulfate radical/ Hydroxyl Often used λ = 254 nm, Quantum yield 0.52 [49,52]
radical
Homogenous: Transition metals One electron transfer Sulfate radical Often requires a low pH to have the metals in a desirable oxidation state [49]
Heterogeneous: Transition One electron transfer Sulfate radical Preparation of catalyst is not economical [60]
metals
Alkaline pH Base-catalysed hydrolysis of PMS to hydrogen Superoxide radical [45]
peroxide
Electrolysis One electron transfer from electrochemically/ Sulfate radical The contaminant degradation rate by an oxidant assisted with electrolysis was observed in the following [64]
chemically produced Fe2+ order: PMS > PDS > H2O2
Nanocarbons One electron transfer Sulfate radical Graphene demonstrated better catalytic performance than few other allotropes of carbon, like: activated [65]
carbon (AC), graphite, graphene oxide, and carbon nanotubes
Organics One electron transfer Sulfate radical/ Hydroxyl Polyimide as an electron donor [66]
radical
Ozone Formation of an – O3SO5− adduct that decomposes Sulfate radical/ Hydroxyl – [46,52]
into radicals radical
Radiolysis Homolysis of peroxide bond Sulfate radical – [49]
Chemical Engineering Journal 330 (2017) 44–62
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

nanocomposite loading, PMS dose, acidic conditions and temperature solution to a minimum and the repeated experiments did not show any
greatly ameliorated the degradation of caffeine. Following the change efficiency loss [51]. Moreover, their research on the reaction stoichio-
of the reaction rate after addition of different radical scavengers (i.e., metric efficiencies (RSE) allows to better understand and compare the
methanol (MeOH) and tert-Butyl alcohol (TBA)), it was shown that the reaction mechanism of different activation methods and its correlation
process involves rather sulfate radicals than hydroxyl radicals. Shi et al. to the mineralization extent of the organic contaminant [95].
[84] demonstrated Co3O4/expanded graphite, developed in a sol-
vothermal synthesis, as an extremely effective heterogeneous catalyst of 3.3. Heterogeneous activation with metal-free catalysts
PMS. They concluded that the abatement of Orange II in water is due to
SO4%−, and 100% removal can be reached after 8 min. Ding et al. [85] Recently, catalytic materials with no metal content are gaining
used cobalt and bismuth salts and sodium hydroxide (as a precipitation popularity due to the many advantages, i.e. lack of secondary pollution,
agent) to create the Co3O4-Bi2O3 nanocomposite as a heterogeneous chemical stability, and that generally they are considered as en-
catalyst for PMS. The nanocomposite demonstrated substantial catalytic vironmentally compatible [66]. It was confirmed that the functional
performance towards PMS for the removal of organic contaminants. groups of carbonaceous materials that possess oxygen (especially the
Moreover, leaching of cobalt was decreased to 43 µg L−1, which is an carbonyl group) can effectively contribute to the activation of persul-
improvement when compared to that of Co3O4 (158 µg L−1) under fates [96]. From the metal-free heterogeneous catalyst, Sun et al. [65]
equivalent testing conditions. Yang et al. [76] claimed that the het- found that the reduced graphene oxide (rGO) can serve as an effective
erogeneous oxide CoFe2O4 was capable to steadily catalyse PMS for the heterogeneous catalyst to activate PMS in order to form radicals. Zhang
abatement of 2,4-dichlorophenol at pH 7. Furthermore, in this system, et al. [97] used granular activated carbon (AC) as a “green” catalytic
leaching of cobalt was reduced to a very low extent. However, such a material for PMS to degrade Acid Orange 7 in an aqueous solution.
reduction in cobalt leaching requires the pH of the solution to be close Also, Saputra et al. [98] reported that AC powder can be an en-
to 7.0, which is problematic in the applicability for treatment of water vironmentally friendly catalyst for the efficient activation of PMS that
or wastewater due to substantial acidification during activation [86]. later exhibited excellent potential for phenol degradation mediated by
Hence, it is still an important challenge to develop an efficacious cobalt- sulfate radicals.
carrying heterogeneous catalyst for PMS without causing a secondary Lee et al. [63] discovered that carbon nanotubes could catalyse
contamination. One way to overcome this problem was proposed by persulfates to form species that are more reactive. Similarly, Duan et al.
Rhadfi et al. [87], who determined that the partial replacement of co- [15] evaluated the ability of various nanocarbons to activate PDS for
balt in Co3O4 with Mn can be a strategy for decreasing the quantity of degradation of phenolic compounds, dyes, and intermediates formed
the harmful Co element. Mn is more abundant in nature, more en- during their treatment process. As PDS activators, single wall carbon
vironmentally friendly, and is twenty times cheaper than cobalt [88]. nanotubes, rGO, and mesoporous carbon displayed great catalytic
Wang et al. [89] provided an example of the PMS catalytic mechanism performances, while nanodiamonds, fullerenes, and graphitic carbon
using an iron oxide/C core shell magnetic nanosphere supported by nitride showed lower efficiencies. Furthermore, the carbo-catalysts
manganese oxide nanoparticle. They found that SO4%− was the primary manifested much higher activity towards PDS activation in comparison
radical for phenol degradation. The same authors also attempted to use to the universally applied AC and metal oxides, i.e.: Fe3O4, Co3O4,
3D-hierarchically structured MnO2 to activate PMS [90]. The catalytic MnO2, and CuO.
mechanism of PMS activation was determined therein by electron In a recent study, Andrew Lin and Zhang [99] showed that not only
paramagnetic resonance (EPR (also known as electron spin resonance, carbons can act as metal-free peroxymonosulfate activators. They have
ESR)) spectra and showed that hydroxyl and sulfate radicals are si- demonstrated that the use of orthorhombic α-sulfur as a metal-free
multaneously produced in the activation processes, however SO4%− photo-catalyst for PMS (under irradiation with visible light) is efficient,
played a more crucial role in oxidation of carbolic acid. Similarly, Li promising, and friendly for the environment.
et al. [91] synthesized three trivalent Mn (oxyhydr)oxides, namely Considering the fact that large amounts of solid waste that could be
bixbyite, hausmannite and manganite, from which only manganite used for the activation of persulfates are produced from various in-
showed good catalytic activity for PDS activation. Analogous to Wang dustries, there is no need to synthesize expensive catalysis. Fly ashes
et al. [90], it was also shown that both SO4%− and %OH are present derived from coal, biomass or oil combustion are major contributors to
during the treatment, although the sulfate radicals were selected as the the production of solid waste and could be used to prepare (Co)-based
predominant oxidative species in phenol oxidation. catalysts for PMS activation [74]. It was found that the fly ash does not
Similarly to the homogeneous systems, iron-based compounds are adsorb phenol, however it consists of cobalt oxide, which can be used
frequently used for persulfate activation due to their lower toxicity for the activation of PMS. In other studies, steel waste powder [100]
compared to cobalt. Tan et al. [92] reported a nano-Fe3O4 catalyst in and the iron scrap from car rotary disc [51] were applied as activators
PMS activation process, whose stability decreased significantly from the for PDS. The iron scrap should be cleaned under dry process where a
first to the third run. In another study, authors of a recent paper con- layer of 3 µm is removed. This layer contains excellent iron alloys
cerning heterogeneous catalysis [60], fabricated CuFe2O4-Fe2O3 for the formed during the friction process between the pad and the disc. This
activation of peroxymonosulfate and subsequent degradation of bi- material is widely available and not expensive.
sphenol A [93]. The performance of CuFe2O4-Fe2O3 was compared with Electron donor catalysis has various benefits but can also, as was
the alternative heterogeneous catalytic materials and shown to be su- stated before, become a new source of contamination. In addition, it is
perior in comparison to CuFe2O4, CoFe2O4, CuBi2O4, CuAl2O4, Fe2O3, often much less efficient, e.g. only one mole of radical can be obtained
and MnFe2O4 (whereas the catalyst order indicates their activity to- from one mole of oxidant, in contrast to UV or heat activation. A de-
wards PMS). Moreover, it is possible to reuse CuFe2O4-Fe2O3 catalyst tailed study of the activation of oxidants with OeO bond by UV and
(at least several times) without significant performance loss. In addi- elevated temperature was performed by Yang et al. [19]. They found
tion, Zhang et al. [94] observed that the property of nano-Fe3O4 to that the order for the decontamination efficacy of peroxides activated
activated PMS can be recycled by hydroxylamine. Furthermore, the by heat is: peroxydisulfate > peroxymonosulfate > hydrogen peroxide
catalytic ability of Fe3O4 coupled with hydroxylamine soared gradually and of UV (254 nm) activated peroxides: peroxydisulfate > hydrogen
in the ten consecutive runs. peroxide > peroxymonosulfate (conditions: experiments were done in
Ghauch’s group by working on bi and trimetallic systems has sig- double distilled water; concentration of AO7 (contaminant of concern)
nificantly contributed to the heterogeneous catalyst in the field of −20 mg L−1; the molar ratio of hydrogen peroxide/AO7 - 10/1). Sur-
peroxydisulfate [50]. They have determined inter alia that the iron prisingly, they also determined that relatively common (in ground,
scrap of special properties can reduce the release of metal ions in surface, and wastewaters) anions (i.e. HCO3−, HPO42−, Cl−, and

48
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

CO32−) can initiate formation of more reactive species from PMS. It has can serve as an additional catalyst for PDS due to continuous generation
to be mentioned, that the efficiency of the above processes are very of ferrous iron by the reduction and transformation of ferric iron on the
dependent on the water matrix. In natural matrices the presence of surface of goethite. Whereas Jaafarzadeh et al. [104] concluded that
chloride and HCO3−/CO32− can greatly decrease the efficiency of only with the simultaneous use of electrolysis and UV, PMS (and
sulfate radical based processes [101] (however, it depends on the probe electro-generated H2O2) can effectively decolorize acid brown 14.
being treated; see below, section Matrix effects). Moreover, the great efficacy of electro-activation of PDS was shown in
Wacławek et al. [55]. This study has shown that pseudo first-order
3.4. Alkaline activation reaction rate constant of lindane removal was (at the optimal condi-
tions) 0.04 min−1, which is higher than the one observed by Cao et al.
Another persulfate activation technique often used in situ is alkaline [106] −0.0083 min−1 (in the PDS/Fe2+ system) but lower than that
activation involving an increase in pH (often > 11) by dosing sodium presented by Khan et al. [107] −0.124 min−1 (in the PMS/Fe2+/UV
hydroxide (NaOH) or potassium hydroxide (KOH) solutions [12]. On system).
the other hand, Cassidy et al. [102] concluded that in situ stabilization
amendments, including fields containing Ca(OH)2 and/or CaO, can ef- 3.6. Other activation methods
fectively 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). Recent work reported by Cong et al. [108], focusing on the si-
Activated PDS decreased concentrations of BTEX compounds and PAHs. multaneous use of PMS and ozone for degradation of 4-chlorobenzoic
The proposed mechanism of alkaline activation of PDS [44] relies on acid, indicated that PMS may act in a similar way to H2O2 in the pro-
the hydrolysis of PDS to hydrogen peroxide anion (HO2−) and motion of %OH production in ozonation. This theory was confirmed by
subsequent reduction of PDS by this anion with the production of Yang et al. [46], who demonstrated that the reaction between PMS and
sulfate and superoxide radicals (O2%−). Very recently, Qi et al. [45] ozone is chiefly responsible for promoting ozone disappearance with a
proved a similar activation tendency for PMS, which can take place determined second-order rate constant of 2 × 104 M−1 s−1. Both sul-
at high pH with base-catalysed hydrolysis of PMS to hydrogen fate and hydroxyl radicals were present in this system, which was
peroxide. It was discovered that O2%− was the predominant radical confirmed by chemical probes and their yields were found to be
species in this system, but the role of singlet oxygen was also found to 0.43 ± 0.1 and 0.45 ± 0.1 per mol of ozone, for SO4%− and %OH, re-
be significant. spectively. The first step of the reaction is assigned to formation of an
adduct (–O3SOO– + O3 → –O3SO5−), which can further decompose
3.5. Electro-activation into more reactive radicals (SO5%− and O3%−). The following reaction
of SO5%− with ozone is assumed to produce SO4%−, while O3%− can
Several studies have evaluated the “electro-activation” of persul- convert to %OH [8].
fates. Yuan et al. [62], investigated the activation of PDS using ferrous
iron produced in an electrolytic system. They have determined that PDS 4. Determination methods
is mainly decomposed by Fe2+ created in two steps – (1) from the
corrosion of Fe0 (chemical and electro-chemical) and (2) from the Numerous determination methods for PDS and PMS can be found in
cathodic reduction of Fe3+. They also found %OH to be the predominant the literature. These methods vary in their execution time and sensi-
radical in their system. Recently, Govindan et al. [64] proved the effi- tivity of detection (see Tables 4 and 5).
cacy of electrochemical activation for persulfates. Electrochemically Tables 4 and 5 present a summary of PDS and PMS determination
activated PMS was the best oxidant (from PDS, PMS and hydrogen methods, respectively.
peroxide) for pentachlorophenol decontamination. Long and Zhang The first methods used for the determination of PDS and PMS were
[103] stated that the electro-activation of PDS can be effective for de- most probably iodometric titrations [16]. Their spectrophotometric
gradation of methylbenzene (toluene) from a surfactant flushing solu- alternatives possess much higher sensitivity and do not require as much
tion. The results indicated that in the reduction process of ferric iron on reagent and time [111,120]. Determination of persulfates with the use
the cathode, there could be produced ferrous iron that could further of dyes has gained popularity lately, although the method has been
activate PDS. This is in agreement with the research work (mentioned known for a long time [115]. Nevertheless, these methods possess very
before in this subsection) of Yuan et al. [62]. Also, simultaneous use of high sensitivity and short measurement time. More information on the
electrolysis and UV process/goethite can enhance the radical formation determination of persulfates with organic dyes can be found in Ding
from PMS and PDS (Jaafarzadeh et al. [104]; Lin et al. [105]). Ac- et al. [124] and Zhang et al. [125]. In addition, the fast and accurate
cording to Lin et al. [105], during the electrolysis, goethite (α-FeOOH) analysis of both PDS and PMS can be performed with the use of ion

Table 4
PDS determination methods.

Method Substance used Time of measurement Limit of quantification Reference

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


Titration Ce4+/Fe2+ <20 min >10−4 M [110]
Spectrophotometry KI/HCO3− 15 min 7 × 10−2 M [111]
Spectrophotometry H2SO4/NH4SCN ∼40 min 2.1 × 10−4 M [112]
Spectrophotometry Methylene blue under microwave radiation 1 min 3 × 10−6 M [113]
Spectrophotometry N,N-diethyl-p-phenylenediamine 10 min 10−5 M [114]
Spectrophotometry Alcian blue 120 min 8.8 × 10−8 M [115]
Amperometry Electrodeposited poly-brilliant cresyl blue on carbon nanotube – 10−5 M [116]
Amperometry Electrode (glassy carbon) with a nanocomposite containing ruthenium oxide nanoparticles – 10−5 M [117]
and thionine
Amperometry Electrode (glassy carbon) with a nanocomposite containing ruthenium oxide nanoparticles – 10−5 M [117]
and celestine blue
Voltammetry Prussian blue - modified platinum disc electrode – 5 × 10−5 M [118]
Ion chromatography Mobile phase: 50 mM KOH; flow rate: 1 mL/min 18 min 5.2 × 10−8 M [119]

49
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

Table 5
PMS determination methods.

Method Substance used Time of Limit of quantification Reference


measurement

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


Spectrophotometry KI/HCO3− 5 min 1.2 × 10−5 M [120]
Spectrophotometry Co2+/methyl orange 1 min 5 × 10−7 M [121]
Spectrophotometry N,N-diethyl-p-phenylenediamine 10 min 10−5 M [114]
Ion chromatography Stationary phase: Super-Sep anion-exchange column; Mobile phase: 2 mM phthalic acid with 11 min 4.5 × 10−5 M [122]
5% (v/v) acetonitrile content, pH 3.0; flow rate 1.5 mL/min; cond. 290 μS/cm
HPLC Stationary phase: Varian Polaris 3 C-18 A column; Mobile phase: methanol/phosphoric acid – ∼5 × 10−7 M [123]
buffer; flow rate 1.0 mL/min; UV at 260 nm

chromatography and HPLC [119,122,123]. 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
5. Persulfate decontamination technologies the electron transfer towards the PMS.
In addition, PMS may oxidize As(III) without an external activation
5.1. Direct oxidation as suggested by Wang et al. [134]. They stated that the PMS can
completely oxidize As(III) to As(V) within 24 h, which was possible
Although reactions of non-catalysed persulfates occur at rates that even after addition of radical scavengers in high concentrations (1.6 M
are often slow, several studies reported their direct oxidation processes. of methanol).
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) 5.2. Basic reactions in water matrices
acts as the nucleophile, whereas in the Boyland-Sims reaction, the nu-
cleophile is a neutral aromatic amine. In both of these processes, there Electron transfer reactions of oxygen yield reactive oxygen radicals
is no radical involvement (apart from the side reactions). [135]. Also, oxygen’s role in SO4%− systems was determined to be
In addition, during catalysed persulfate oxidation, persulfates can significant and was described carefully in a recent paper by Xu et al.
react with generated radicals in several ways. [136] and Ghauch et al. [42]. The step-by-step reduction of an oxygen
In Fig. 3 the second-order reaction rate constants of persulfate re- molecule to water followed by the formation of radicals during per-
actions with several species in an aqueous phase, can be observed. sulfate decomposition is shown in Fig. 4.
Further example of PDS oxidation not relying on the radical for- Superoxide disproportion may result in the formation of H2O2
mation was presented by Zhang et al. [21]. According to them, PDS which in turn can react to generate %OH (e.g., by chemical reduction).
weakly interacts with the surface of copper oxide (under mild condi- Furthermore, %OH can also be generated together with SO4%− in re-
tions) from which it can easily react with the contaminant of concern actions of PDS or PMS with transition metals [49] or in the photolysis of
(2,4-dichlorophenol). The reaction between PDS and the surface of CuO PMS [137] and SO4%− can also form %OH in the reaction with OH−
was assumed therein to be the rate-limiting step. (k = 1.4 × 107 M−1 s−1, Herrmann et al. [57], and reaction 4). Con-
Furthermore, Lei et al. [133] have shown that not only PDS can effi- sidering the concentration of OH− at normal conditions of water and
ciently oxidize target pollutants without the free radicals involvement. wastewater treatment (e.g., pH 6–8) the observed reaction rate constant
They have discovered that PMS can directly react with cationic pigments of the reaction of SO4%− with OH− is in the range of 10−1 to
without catalyst and in a broad pH range (2–12). Furthermore, they 102 M−1 s−1. However, reactions of SO4%− with other matrix compo-
concluded that Cl− anions improved the degradation efficacy of the target nents such as chloride or dissolved organic compounds are likely more
pollutants. Radical quenching experiments and ESR studies revealed that important at that pH-values (for relevant reaction rate constants see

Fig. 3. Comparison of the second-order reaction rate constants of persul-


fates 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].

50
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

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 information see [127,141]. For a more detailed ex-
planation of radical identification, e.g. with the use of radical sca-
vengers, see [71,138–140].

5.3. Matrix effects: Scavenging and secondary oxidants

In persulfate based oxidation processes, pollutants are mainly de-


graded by reactions of SO4%− which are formed by the different acti-
vation 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
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
In order to define how fast the reaction proceeds, second order rate
of the PMS peroxide bond). (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)
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
Table 6
Second-order reaction rate constants of hydroxyl and sulfate radicals with common an- SO4%− with common anions and natural organic matter (NOM) present
ions and NOM. in real matrices.
Scavenging is mainly driven by dissolved organic matter (DOM),
Rate constants(M−1 s−1)
Radical Compound Reference
HCO3− and CO32− and in some cases nitrite and reduced metal species
%
OH −
Cl 3.0–4.3 × 10 9
[142,143] such as Fe2+. DOM and reduced metal species mainly consume oxida-
SO4%− Cl− 1.3–6.6 × 108 [144–146] tion capacity because the products formed in these reactions are not
%
OH Br− 1.9 × 109 [147] reactive towards pollutants. For the reaction of the sulfate radical and
SO4%− Br− 3.5 × 109 [148]
%
hydroxyl radical with humic acids the kinetic rate constant was mea-
OH HCO3− n × 107 [149]
SO4%− HCO3− 2.6–9.1 × 106 [109,150]
sured to be 6.8 × 103 L mg C−1 s−1 and 1.4 × 104 L mg C−1 s−1 (mg
%
OH CO32− 4 × 108 [149] C = mg carbon), respectively [152]. On the contrary, Luo et al. [32]
SO4%− CO32− 4.1 × 106 [151] observed that the oxidation capacity of the hydroxyl radical and sulfate
%
OH Humic acid 1.4 × 104* [152] radical for humic acid degradation (in the presence of Cl− and Br−)
SO4%− Humic acid 6.8 × 103* [152]
was nearly in the same order. This can be explained by the below dis-
*(mg of C L−1)−1 s−1. cussed conversion of SO4%− into %OH in presence of chloride. Hence in
both systems the same oxidant prevails, i.e., the hydroxyl radical.
Table 6). Studies based on radical scavenging analyses paired with the Especially Br− had a negative impact on the efficiency of these pro-
use of the ESR technique were performed to identify the dominant ra- cesses, which is in agreement with the research work of Yang et al.
dical species involved in persulfate-based oxidation [56,89,90]. [153] (Although, in several studies a concentration of 1 mM of Br−
Several compounds can be used for scavenging of SO4%− and %OH. and/or Cl− seemed to enhance the removal of some compounds
For choosing the correct scavenging agent, one has to consider reaction [51,59]).
kinetics and product formation. It has to be provided that only reactions Electron transfer reactions of SO4%− with phenolic moieties of DOM
under study are relevant and that products which could interfere the result in radical cations which may rearranged in presence of water into
reactions under study or analytical procedures are not formed. carbon centred radicals in analogy to the reaction of benzenes with
TBA (khydroxyl radical = 6.0 × 108 M−1 s−1, ksulfate radical = 4.0 SO4%− [154]. Carbon centred radicals react rapidly with oxygen
× 105 M−1 s−1) and ethanol (khydroxyl radical = 1.2 × 109 M−1 s−1, (≈2 × 109 M−1 s−1 [155]). The ensuing peroxyl radicals eliminate
ksulfate radical = 1.6 × 107 M−1 s−1) are the most often used chemical O2%− or HO2% or decompose in bimolecular reactions [155]. Both
probes to evaluate the relative contribution of SO4%− and %OH because pathways often result in carbonyls [155]. Halide ions also react with
of their different reaction rate constants [138]. Also with the ESR SO4%−. The reactivity of bromide towards SO4%− is high
technique, the radicals can be determined qualitatively (rough esti- (k = 3.5 × 109 M−1 s−1 [148]) and in pure water bromate is formed as
mation of the intensity of radicals in the spectra). The spin trap com- a stable product [156], which is undesired due to its carcinogenic po-
pound, 5,5- dimethyl-1-pyrroline N-oxide (DMPO) reacts with radical tential (EU and US-EPA drinking water standard 10 µg L−1). The pre-
intermediates in oxidation systems and forms a unique radical adduct sence of organic matter, however, suppresses formation of bromate,
that is stable, relative to the radicals of interest. indicating that no bromate is formed in field water matrices [156]. Fang
Indeed with these techniques, it was shown that in most cases and Shang [157] came to a similar conclusion. Liu et al. [158] in-
SO4%− prevails in acidic media and %OH predominates in alkaline vestigated the transformation of Br− in a PMS (activated by cobalt)
media (e.g. tested at the pH of 5 and 11 by Fang et al. [139,140] (at the oxidation process with the presence of phenol as a model compound
pH of 9 both radicals were in similar quantity [140]). Similar conclu- imitating NOM. It was determined that bromide can be efficiently
sions were obtained by Liang and Su, [138] by using various probes: converted to free bromine and bromine radicals. These reactive species
TBA, phenol, and nitrobenzene). caused bromination of phenol and the formation of brominated disin-
Reactions of SO4%− are very different from reactions of %OH. SO4%− fection products including bromoform and bromoacetic acids and bro-
are prone to react by electron transfer reactions while %OH favours H- mophenols as intermediates. Brominated disinfection by-products were
abstraction and addition reactions. This can explain that reaction rate also degraded by surplus SO4%−. Free bromine was also created when
cobalt was not present, indicating that Br− might be directly oxidized

51
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

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]
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
DOC 2.5 × 104 L mg−1 s−1 [166]
HCO3− 8.5 × 106 M−1 s−1 [149]

by PMS. In another study these conclusions were confirmed and it was potential to degrade pollutants (depending on their chemical structure
also found that the brominated intermediates cannot be degraded in the and affinity with carbonate radicals). Table 7 compiles rate constants of
absence of SO4%− [159]. Similar work was performed by Lu et al. CO3%− with several organic compounds. The reaction rates are in the
[160], who determined that reactive bromine species formed in the range of 106–109 M−1 s−1 and thus, on average slower compared to
reaction with sulfate radicals can react with NOM and might form SO4%− or %OH, which often react at second order rate constants in the
brominated products, including brominated disinfection by-products. In range of 108–1010 M−1 s−1 [149,162]. However, this may be com-
addition, bromoacetic acid and bromoform were formed in the presence pensated by the small reaction rate of CO3%− with the scavengers in the
of humic acids. water matrix, as will be explained below.
Iodide also reacts fast with SO4%−, which may form iodate in ana- The efficiency of pollutant degradation is depending on the
logy to bromide. scavenging rate of the water matrix and the kinetics of the radical re-
The products of the reaction of SO4%− with chloride can result in action with the pollutant. With the kinetics of scavenger reactions and
chlorate formation in an H+ catalysed reaction [101]. However, at pollutant reactions at hand one can calculate the apparent degradation
typical conditions of water treatment (pH > 7) no chlorate formation rate (ḱ) in a given water sample according to Katsoyiannis et al. [167].
was observed in UV/S2O82− [101]. The primary step in the reaction of Fig. 5 shows ḱ for the compounds of Table 7 for reactions of CO3%−
SO4%− with chloride is an electron transfer yielding chlorine (Cl%) assuming a water matrix with 1 mg L–1 DOC and 1 mM HCO3−. ḱ of
atoms and sulfate. Cl% forms a H2OCl% complex in water [146,161]. This CO3%− is shown for the two reaction rate constants with DOC (Table 7).
complex can deprotonate and the resulting HOCl%− is in equilibrium As will be explained below, HCO3− presumably does not contribute in
with %OH [101,146,161]. At pH 7 and in presence of 1 mM chloride, the scavenging of CO3%−. Fig. 5 also shows ḱ for the same water matrix in
reaction of SO4%− with chloride yields %OH with a nearly 100% yield SO4%− and OH% reactions (dashed lines) for a defined reaction kinetics
[101]. In case of chloride concentrations in the lower mM range most of towards trace compounds (dashed lines in Fig. 5).
SO4%− are scavenged by reactions with chloride resulting in a con- The herbicide propanil (compound 10, Table 7) reacts with CO3%−
ventional %OH based AOP [101]. CO32− or HCO3− react fast with Cl% at a reaction rate constant of 1.4 × 107 M−1 s−1, arriving at
and dichloride (Cl2%−) and thus, interrupt the %OH forming reactions ḱ = 0.25 s−1, in case CO3%− reacts with DOC at a rate of 40 L
[101]. Since typical natural waters contain chloride and HCO3−, the mg−1 s−1. The same ḱ is achieved in %OH based processes for pollutants
%
OH yield is likely much below the SO4%− yield [101]. However, oxi- with a much higher reaction rate constant (k ≥ 9 × 109 M−1 s−1) and
dation of HCO3− and CO32− gives rise to CO3%−, which also exhibits a in SO4%− based reactions with k ≥ 1 × 109 M−1 s−1 (dashed lines in

52
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

Fig. 5. Calculated apparent first order kinetics (ḱ) of pollu-


tants degradation by carbonate-, sulfate- and hydroxyl radi-
cals in presence of 1 mg L−1 DOC and 1 mM HCO3−.
Calculations are based on reaction rate constants shown in
Table 7. Circles: ḱ(carbonate radicals) assuming: k
(CO3%− + DOC) = 40 L mg−1 s−1, squares ḱ(carbonate ra-
dicals): assuming k(CO3%− + DOC) = 40 L mg−1 s−1, dashed
lines ḱ(sulfate radicals or hydroxyl radicals, as indicated in the
figure) for a given reaction rate constant towards a pollutant,
as indicated in the figure.

Fig. 5). This is due to fact, that %OH and SO4%− have a high reactivity reactivity of alkenes with halogen radicals. This statement was further
towards DOC (k(SO4%−) = 6.8 × 103 L mg−1 s−1 [152], and k(OH%) confirmed in a study of Liu et al. [169] who found that in the presence
= 2.5 × 104 L mg−1 s−1 [163] and towards HCO3− (k(SO4%−) of chloride there could be a different mechanisms of oxidation than that
= 2.8–9.6 × 106 M−1 s−1 [165], and k(OH%) ≈ 1 × 107 M−1 s−1 of the SO4%−, and carbon isotope fractionation of trichloroethene (TCE)
[149]. The reaction of CO3%− with organic matter is considerably lower was used to prove this statement.
(see above). Furthermore, HCO3− or CO32− is not important for pol- In another work, Xie et al. [159] focused on the generation of
lutant degradation, because it presumably yields CO3%−. Hence, the chlorinated intermediates in a sulfate radical system, and they found
overall scavenger rate of CO3%− is considerably lower compared to that the formation of carbonaceous disinfection by-products, i.e. ha-
SO4%−. In case the CO3%− reacts with DOC with a rate constant of loacetic acid and chloroform, only increased a little, but the generation
280 L mg−1 s−1 the efficiency of CO3%− for pollutant degradation of nitrogenous disinfection intermediates, i.e. haloacetonitriles and
drops down and CO3%− has to react with rate constants > 108 M−1 s−1 trichloronitromethane, slightly decreased. On the contrary, Lu et al.
to yield a faster degradation kinetics compared to %OH. Compound 18 [170] found that after treatment of surface water with 0.1 μM PDS for
(vanilinate) represents such a compound, assuming that %OH does not 48 h, caused increase in concentration of compounds such as: chloro-
react faster than 9 × 109 M−1 s−1 with vanilinate. In case of SO4%− form, trichloroacetic acid, and dichloroacetic acid by 16%, 37%, and
which react slower with DOC, CO3%− has to react even faster with 52%, respectively. The above results indicate that before the applica-
́
pollutants to achieve comparable k. Pollutants with electron rich moi- tion of persulfates in the field, bench scale studies should be performed
eties such as anilines reveal a fast reactivity towards CO3%−. It is in detail due to the fact that the formation of dangerous by-products
conceivable that the degradation of electron rich compounds is en- depends highly on the treated water matrix properties.
hanced by HCO3−/CO32− in SO4%− or %OH based processes. This ap- To sum up, at typical conditions of drinking water treatment (i.e.,
plies for compounds 15 to 32 (Table 7, phenols and anilines). Most pH 6–9) presence of SO4%− scavenged by chloride are largely trans-
pesticides listed in Table 7, however, reveal reaction rate constants formed into %OH. Reactive chlorine species are prone to be scavenged
of < 107 M−1 s−1. Their degradation is probably mitigated in presence by carbonate or bicarbonate, resulting in CO3%−. The fraction of radi-
of HCO3−/CO32− in a SO4%− or %OH based process, as observed for cals turned into %OH or CO3%− is controlled by kinetics and con-
atrazine degradation in river water [101] (cf. further discussion centration of the corresponding reaction partners (further details on
[101,164]). that reaction system see above and Lutze et al. [101]). CO3%− may
It was also reported that reactive chlorine species, formed in the become important for pollutant degradation due to their higher se-
reaction of SO4%− with Cl–, may also react with other matrix compo- lectivity compared to SO4%− and %OH, provided that CO3%− readily
nent and thus, form chlorinated products [144]. For example, Fang react with such pollutants (k ≈ 107 M−1 s−1 or faster) (see also Fig. 5
et al. [140] reported that the total concentration of radicals was greatly and Table 7).
increased in the presence of chloride ion but the degradation efficiency
of PCBs in this system decreased. Yang et al. [153] also investigated the
5.4. Post-treatment toxicity assessment
conversion of %OH and SO4%− to the halogen radicals. Halogens re-
duced the abatement efficacy of cyclohexanecarboxylic acid and ben-
Although persulfates treatment has many benefits, there are also
zoic acid (in the presence of seawater), which were chosen as the target
several downsides that have to be taken into consideration, i.e. con-
pollutants. It was also concluded therein that the activated PDS was
tamination with sulfate salts or even worse newly created hazardous
more affected by Cl− than the activated hydrogen peroxide system
compounds due to e.g. additional chlorination and/or bromination (as
probably due to the fact that Cl− has a higher reactivity with sulfate
discussed in Section 5.3).
radicals than the hydroxyl radicals at pH of 7. Indeed %OH is hardly
In addition, many researchers study toxicity in clean systems, al-
effected by Cl− at typical pH values of water treatment (6 −8) since the
though in actual matrix conditions, the background DOC and other
reaction of %OH with Cl− has a fast back reaction. The oxidation of Cl−
constituents may additionally generate toxicity.
by %OH is acid catalysed and only becomes important in %OH reactions
For instance, in our recent study concerning the decontamination of
at pH < 3 [168]. The degradation efficiency of cyclohexanecarboxylic
groundwater polluted with various chlorinated olefins (Wacławek and
acid was not altered by the halogens, probably due to the high
Kudlek, unpublished data), many new substances (with a molecular

53
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

mass > 250 g mol−1) were observed after UV/persulfate treatment be used for the removal of inorganic pollutants in water. They in-
that were unnoticed after e.g. UV/hydrogen peroxide treatment. vestigated the UV/PDS oxidation of As(III) to the less harmful As(V). It
Therefore, during bench-scale testing performed before field applica- was established that humic acid had no effect on the reaction rate (si-
tion, toxicity tests (that are relevant to the treated site) should be milar results to Wang et al. [134], although therein 1.6 M methanol was
performed in order to avoid any environmental side effects. used as a scavenger), even at 20 mg L−1. Yet, the continual addition of
However, several authors reported that the toxicity of the matrix nitrogen considerably minimized the rate of the reaction (by 20%),
after persulfates treatment is significantly lower. One example was attributing this to the role of dissolved oxygen in the reaction (at high
provided by Zhang et al. [171] that evaluated toxicity with Vibrio qin- concentrations of PDS this phenomenon was not observed, presumably
ghaiensis sp. Q67 test after the PDS treatment of carbamazepine and due to formation of oxygen in the reaction of peroxydisulfate with
concluded that the acute toxicity has decreased together with the re- SO4%−; [58]).
moval of the pharmaceutical (the inhibitory effect of the solution after Surprisingly, according to Diao et al. [202,203] SO4%− coupled
treatment decreased to 65% within 60 min). Also Temiz et al. [172] with nano zero-valent iron (supported on bentonite, and used as a ra-
evaluated the toxicological safety (using two different bioassays) of the dical initiator), can be effective for the simultaneous abatement of Cr
zero-valent iron/PDS oxidation of Triton X-45 (TX-45). Vibrio fischeri (VI) and phenol from water. The reaction mechanism according to them
and Pseudokirchneriella subcapitata bioassays were used therein and the involved the removal of Cr(VI) mainly by reduction with nano zero-
toxicity profiles of the treated matrices significantly decreased from an valent iron and phenol removal mainly by the SO4%− generated from
original value of 66% relative inhibition to 21% (Vibrio fischeri) and the PDS.
from 16% relative inhibition to non-toxic values (Pseudokirchneriella Chlorinated olefins are ubiquitous contaminants, and although they
subcapitata). Vibrio fischeri has shown higher sensitivity to TX-45 and its can be degraded with many biological [204] and less invasive chemical
degradation products than the microalgae Pseudokirchneriella sub- treatments (i.e.·H2O2) [205], persulfates are often used for their de-
capitata. gradation in situ. Recently, Yan et al. [206] combined siderite-catalysed
Olmez-Hanci et al. [173] went even further and used three toxicity H2O2 with PDS and effectively used it for the remediation of tri-
tests (Vibrio fischeri, Daphnia magna, and Pseudokirchneriella subcapitata) chloroethene contamination from groundwater. It was claimed therein,
and the Yeast Estrogen Screen bioassay to assess the possible estrogenic that in the absence of PDS (only catalysed peroxide), most of the hy-
and toxic properties of nonionic surfactant octylphenol ethoxylate and drogen peroxide was reduced within the first hour of the test, resulting
its oxidation products. In the case of Vibrio fischeri and Daphnia magna in non-efficient use of OH-radicals. After the addition of PDS, the de-
tests the inhibitory effect of nonionic surfactant octylphenol ethoxylate composition rate of H2O2 was mitigated due to a more sustainable re-
dropped considerably after the application of PMS. However, treatment lease of OH-radicals. Furthermore, the heat given by the decomposition
with PMS/UV-C generated oxidation products that possessed highly reaction of H2O2 activated the PDS, and the generated sulfate radicals
toxic effect towards Pseudokirchneriella subcapitata (opposite observa- were claimed to be the main oxidative species. In addition, di-
tions to Temiz et al. [172], concerning toxicity of TX-45 and its oxi- chloroacetic acid has been detected as an intermediate. However, it has
dation by-products). 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
5.5. Decontamination of water and wastewater with free radicals generated PDS. Xu et al. [207] also studied the abatement of trichloroethene
in persulfate systems (TCE) but in a thermally activated peroxydisulfate system. Their results
showed that TCE can be completely removed from the solution within
As shown in Section 3, numerous methods of persulfate activation 9 min at 50 °C with an initial trichloroethene concentration of 0.15 mM
can be applied in water and wastewater treatment technology in order and a PDS dose of 0.3 M, as a consequence of the active oxygen species
to achieve the expected results. The decontamination of water (Section formation (SO4%−, %OH). Moreover, Zhao et al. [48] studied the si-
5.5.1) and wastewater (Section 5.5.2) with the radicals generated from multaneous decontamination of 1,4-dioxane, the inherent associate of
persulfates is described in detail below. TCE (frequently used as a solvent stabilizer for TCE), with heat- and
Fe2+ PDS activation. Analysis of carbon balance revealed that 96% and
5.5.1. Water 93% of the organic carbon was removed after the 1,4-dioxane abate-
SO4%− are very reactive and able to degrade very persistent organic ment with and without activation (addition of Fe2+), respectively.
compounds. Perfluorooctanoic acid (PFOA) e.g., that is unreactive to- Another commonly found and very toxic group of contaminants is
wards %OH, can be degraded in the SO4%− system according to many pesticides, which contribute to nine out of the twelve most hazardous
[174–179]. Recently, Qian et al. [180] determined the degradation and assiduous organic compounds defined by the Stockholm
kinetics of PFOA in a UV/PDS system and proposed its degradation Convention on Persistent Organic Pollutants (POPs) [208]. A very re-
mechanism, which relies on a sequential loss of CF2 units from it and its cent study by Qin et al. [209] presented 1,1,1-trichloro-2,2-bis(p-
intermediates [179,180]. However, the degradation kinetics of the re- chlorophenyl) ethane (DDT) removal with Co2+ catalysed PMS. It was
action SO4%− with perflurocarbonic acid are very slow (approximately found that DDT was efficiently decomposed within several hours, pro-
104 M−1 s−1). Since the reaction of SO4%− with other matrix compo- portionally to PMS/Co2+ concentrations. The decontamination rates of
nents is much faster, a degradation of perfluorinated compounds is not DDT were determined with pseudo-first-order reaction rate equations in
feasible in SO4%− based oxidation. several temperatures, which enabled calculation of the activation en-
SO4%− can readily oxidize other organic pollutants such as 2,4-di- ergy (72 kJ mol−1). Several by-products of the reaction were de-
chlorophenol [75], 2-chlorobiphenyl [181], aniline [24], bisphenol A termined including: 4-chlorobenzoic acid, benzylalcohol, di-
[182], calcon [183], Acid Orange 7 [19], hexachlorocyclohexanes chlorobenzophenone, and the possible degradation pathway of DDT
[55,184], Ponceau S [185], 4-fluorophenol [186], pentafluorobenzoic was suggested on the foundation of the detected by-products. Zhu et al.
acid [187], C.I. Reactive Black 5 [188], C.I. Basic Red 46 [189], me- [210] have also followed the DDT degradation (and the reaction in-
thylene blue [190], endosulfan [191], antipyrine [192], naproxen [95], termediates) but using PDS (activated by nanoscale zero-valent iron) as
chloramphenicol [59], ranitidine [51], sulfamethoxazole [42,50], bi- a source of radicals. The degradation pathway was very similar to that
soprolol [36], ibuprofen [35], dimethyl phthalate [193] or di- of Qin et al. [209] (in the PMS/cobalt system), although it should be
methylhydrazine [194] and chlorotriazine pesticides. Table 8 includes noted that instead of 4-chlorobenzoic acid, its dechlorinated version has
pseudo first-order rate constants of various contaminants with gener- been found. In addition, ESR results showed simultaneous involvement
ated from persulfates radicals. of sulfate radicals and hydroxyl radicals in the degradation process.
Moreover, Neppolian et al. [201] proved that these radicals can also Several authors examined atrazine degradation with persulfates.

54
Table 8
Decontamination of common water pollutants with radicals generated in persulfate systems.

Group of compounds Model contaminant Main oxidative species Comments Pseudo first-order reaction rate constant Reference
S. Wacławek et al.

Chlorinated olefins Perchloroethene Possibly SO4%− (not Trace quantities of hexachloroethane as an intermediate. 7.6 × 10−3 min−1 (activation temperature of 50 °C) [144]
identified)
−3 −1
Trichloroethene Possibly SO4%− (not Trace quantities of hexachloroethane as an intermediate. 4.7 × 10 min (activation temperature of 50 °C) [144]
identified)
−3 −1
cis-dichloroethene Possibly SO4%− (not Isomerization between trans and cis-DCE presumably after 1.6 × 10 min (activation temperature of 50 °C) [144]
identified) formation of a single bonded intermediate.
trans- dichloroethene Possibly SO4%− (not Isomerization between trans and cis-DCE presumably after 3 × 10−3 min−1 (activation temperature of 50 °C) [144]
identified) formation of a single bonded intermediate.

BTEXs Benzene Possibly SO4%− (not Of all the BTEX compounds studied, benzene was most 9.5 × 102 day−1 (activation temperature of 20 °C; Oxidant/BTEX molar [195]
identified) resistant to PDS oxidation. ratio 100/1)
Toluene Possibly SO4%− (not – 23.2 × 102 day−1 (activation temperature of 20 °C; Oxidant/BTEX molar [195]
identified) ratio 100/1)
Ethylbenzene Possibly SO4%− (not – 14.5 × 102 day−1 (activation temperature of 20 °C; Oxidant/BTEX molar [195]
identified) ratio 100/1)
Xylene Possibly SO4%− (not – 21.9 × 102 day−1 (activation temperature of 20 °C; Oxidant/BTEX molar [195]
identified) ratio 100/1)

Phenols Phenol Possibly SO4%− and/or Greater total organic carbon (TOC) removal was observed 0.14–0.16 min−1 (UV activation parameters: λ = 254 nm; Oxidant/ [196]
%
OH (not identified) at elevated pH (11). Phenol molar ratio 168/1)
Bisphenol A SO4%− (scavenging tests Bisphenol A abatement with sulfate radicals was found to 0.025 min−1 (UV activation parameters: λ = 254 nm; 40 W power; [182]
were performed) proceed via one electron transfer reaction mechanism. Io = 1.26 μE s−1; Oxidant/Bisphenol A molar ratio 3/1)

Pharmaceuticals Diclofenac SO4%− (scavenging tests Degradation involved one electron transfer, 0.032 h−1 (Activation temperature of 50 °C; Oxidant/Diclofenac molar [197]
were performed) hydroxylation, decarboxylation. ratio 10/1)
Carbamazepine SO4%− (scavenging tests Electron-transfer between sulfate radicals and 0.087 min−1 (UV activation parameters: λ = 254 nm; 9 W power; [171]

55
were performed) carbamazepine was found to be the major mechanism. Oxidant/Carbamazepine molar ratio 10/1)
Bisoprolol SO4%− and %OH The formation of hydroxylated products through 8.5 × 10−2 min−1 (activation temperature of 60 °C; Oxidant/Bisoprolol [36]
(scavenging tests were hydroxylation was proposed. molar ratio 20/1)
performed)
Chloramphenicol SO4%− and %OH – 29 × 10−2 min−1 (UV activation parameters: λ = 254 nm; Oxidant/ [59]
(scavenging tests were Chloramphenicol molar ratio 80/1)
performed)

Pesticides Lindane Possibly SO4%− and/or Approx. 96.4% of chloride ion was released after the 10−3 sec−1 (UV activation parameters: λ = 254 nm; 10 µM = Fe2+; [107]
%
OH (not identified) treatment, which was consistent with the TOC analysis. Oxidant/Lindane molar ratio 73/1)
Triazine pesticides (atrazine, Possibly SO4%− and/or – 1.7 × 10−3 sec−1 (UV activation parameters: λ = 254 nm; Oxidant/ [32,152,198–200]
%
tert-butylazine, propazine) OH Atrazine molar ratio 50/1) Manoj et al. [198] has determined bimolecular
rate constants of the SO4%− reaction with triazines in the range of:
4.6 × 107–3 × 109 M−1 s−1, whereas Lutze et al. [152]:
2.2–3.5 × 109 M−1 s–1; Khan et al. [199]: 2.59 × 109 M−1 s–1 and
2.25 × 109 M−1 s–1 for the reaction with SO4%− and %OH, respectively;
Endosulfan SO4%− and %OH Degradation of endosulfan was began at the endosulfans 4.14 × 10−4 sec−1 (UV activation parameters: λ = 254 nm; Oxidant/ [191]
(identified) S ] O group. Endosulfan molar ratio 10/1)

Other Methylene blue Possibly SO4%− and/or The first step is an electron transfer from a benzene rings 14.8 × 10−2 min−1 (activation temperature of 60 °C; Oxidant/Methylene [190]
%
OH of MB to SO4%−. After 20 min of reaction, complete blue molar ratio 640/1)
disappearance of all intermediates was noticed.
Acid orange 7 Possibly SO4%− (not Reaction rates differed between UV/PMS and UV/PDS, 0.175 min−1 (UV activation parameters: λ = 254 nm; Oxidant/Acid [19]
identified) possibly due to the production of the hydroxyl radical orange 7 M ratio 10/1)
during the photolysis of PMS and different quantum
yields.
PFOA SO4%− and S2O8%− SO4%− accepts an electron from the carboxylate group in a 0.18 h−1 (UV activation parameters: λ = 254 nm; Oxidant/PFOA molar [180]
primary step. ratio 33/1)
Chemical Engineering Journal 330 (2017) 44–62
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

One group was Luo et al. [32] who tested degradation of atrazine with groundwater. The results implied that the degradation of diclofenac
three oxidants – H2O2, PMS and PDS (UV activated, 254 nm). The could be fitted well to a pseudo 1st-order kinetic model, and that the
matrix effects, i.e. water hardness, Cl−, and NOM, were evaluated on rate constants were larger at higher temperatures. Activation energy
these three AOPs. It was determined that the concentrations of sulfate was also calculated and equalled to 158 kJ mol−1. The presence of a
radicals and hydroxyl radicals decrease with an increase of carbonate/ small dose of chloride (0–10 mM) enhanced the abatement of diclo-
bicarbonate concentrations. A detailed description on the reactions of fenac, whereas larger chloride addition (>10 mM) had opposite effect.
%
OH and SO4%− with Cl− in presence of HCO3− is described in Section HCO3− demonstrated an insignificant effect on diclofenac elimination,
5.3, see also Lutze et al. [152]. Main transformation products of atra- while NOM, e.g., humic acids, slightly inhibited diclofenac removal.
zine in reactions with %OH and SO4%− are desethyl-atrazine and desi- The fast oxidation of diclofenac was further observed in a groundwater
sopropyl atrazine [152,211,212] which are similarly toxic as the parent sample from contaminated site. In addition, radical quenching tests
compound [213]. The second order rate constant of different chloro- revealed that SO4%− were the leading reactive species for diclofenac
triazine pesticides with SO4%− was determined to be in the range of oxidation. Ibuprofen is also considered as an emerging contaminant and
1–5 × 109 M−1 s−1. The primary dealkylation products still react fast its degradation in the activated PDS system was carried out by Ghauch
0.8–2 × 109 M−1 s−1. However, in case no alkyl group is attached at et al. [35]. The by-products were not detected throughout the treatment
the chlorotriazine ring (i.e., the chlorotriazine diamine) the reaction process. It was concluded that this method could be an adequate ap-
becomes very slow for both %OH (k < 107 M−1 s−1) [214] and SO4%− proach for specific treatment of small volumes of hot spot wastewater
(k 1.5 × 108 M−1 s−1) [152], hence desethyldesisopropyl atrazine can containing this contaminant (for example hospital effluents). PDS was
be considered to survive oxidative treatment (cf. mechanistic aspects of also tested on hospital effluent spiked with naproxen. This study clearly
chlorotriazine pesticides degradation [152,211,212]. Research work of demonstrated that thermally activated peroxydisulfate is a valid and
Wacławek et al. [184] was one of the first evaluating the degradation efficient method that can be used for the removal of dissolved phar-
efficiency of hexachlorocyclohexane isomers by PMS. Cao et al. [106] maceuticals in water and sewage water [95].
and Khan et al. [107] provided more detailed study focusing on the SO4%− treatment also proved to be effective for highly con-
oxidation of one HCH isomer - lindane (γ-hexachlorocyclohexane) by taminated mature landfill leachate. Li et al. [219] used ferrous iron
Fe2+ activated PDS and PMS, respectively. These studies revealed that loaded AC as a heterogeneous PDS catalyst for its pretreatment. The
oxidation of HCH with activated persulfates is not only beneficial for effects of the iron/PDS dose and initial pH on the abatement of the
the complete removal of the parent compounds but also for obtaining organic pollution in the landfill leachate were determined. It was shown
complete mineralization. Trichlorophenol was the main by-product that the chemical oxygen demand (COD) degradation rate exceeded
detected in all of these studies (although there is no agreement on the 87.8% when simultaneous conditions were applied i.e. Fe2+ dose of
exact isomer generated). In addition, in a further study, Khan et al. 127 mg L−1, PDS concentration of 0.5 M and initial pH of 3.0.
[215] determined the second-order rate constant of lindane with SO4%−
(1.3 × 109 M−1 s−1). Kuśmierek et al. [216] investigated the de- 5.5.2. Wastewater and sludge
gradation of 2,4-dichlorophenol and 2,4-dichlorophenoxyacetic acid by Several studies can be found in the literature that describe the use of
ammonium PDS activated in several ways. 2,4-dichlorophenol de- persulfates in wastewater treatment technologies. Their use is focused
graded faster and more efficiently in an alkaline environment on exploiting their oxidative potential (removal of contaminants) and
(pH = 9.0), whereas 2,4-dichlorophenoxyacetic degraded faster in an improving the properties of sludge, i.e. dewaterability. Kronholm and
acidic environment (pH = 3.0). They have also examined the sy- Riekkola [220] tried to answer the question whether potassium per-
nergistic activation of PDS with heat and ferrous iron to improve the oxydisulfate is a good choice for wastewater oxidation below the cri-
oxidation of 2,4-dichlorophenol and 2,4-dichlorophenoxyacetic acid. tical temperature of water. The efficiency of phenol, 1-naphthol, and
They were able to select the optimal degradation conditions (molar 2,3-dichlorophenol oxidation in high-temperature (75–340 °C) pres-
ratio between PDS/Fe2+ - 1:2, temp. - 50 °C), where complete removal surized (25–45 MPa) wastewater was investigated in an aqueous en-
of contaminants was obtained after approximately 45 and 60 min, re- vironment. The removal percentages of phenol were good even at
spectively. 115 °C. Nonetheless, it has to be noted that although this radical in-
As was mentioned in the introduction, the rapid emergence of re- itiation method is not economically feasible and there were reported
sistant bacteria worldwide is probably due to the overuse of medica- many other methods for treatment of phenolic wastewater [221,222],
tions that can later become a contamination of concern. Several authors the study of Kronholm and Riekkola [220] has shown new alternative
have focussed on the remediation of pharmaceuticals and since per- way for wastewater treatment. Recently, the removal of COD from
sulfates are one of the newest ISCO reagents used, there are also several petrochemical wastewater and from real high-strength industrial was-
new studies describing their reactivity towards pharmaceutical drugs. tewater was studied by Babaei and Ghanbari [223] and Kattel et al.
Monteagudo et al. [217] investigated ISCO of a carbamazepine solution [224], respectively. In both studies, the persulfates proved to be a vi-
by PDS (at the same time) activated by UV, heat, Fe2+ ions, and H2O2. able alternative to the conventionally used oxidants (i.e., percarbonate,
Zhang et al. [171] determined the main by-products generated during hydrogen peroxide).
the oxidation process, including 10,11-epoxy-carbamazepine, acridine- Generally, oxidative wastewater treatment options are considered
9-carbaldehyde, acridine, and other compounds with smaller molecular as more efficient than the conventional ones; however, one of the
weight. downsides of them is above-mentioned high cost of the chemicals and
Trimethoprim and sulfamethoxazole in expired sulfamethoxazole energy involved. This drawback could be partially overcome by the
tablets were the subject of research conducted by Liu et al. [218], who combination of various techniques (e.g. membrane ones), which could
studied their degradation by catalysed PDS treatment. Zero valent-iron not only reduce the amount of oxidant needed but also improve the
showed much better catalytic properties than alkaline activation, which second process, e.g. by decrease membrane fouling [225].
was completed after 0.5 h, while full mineralization was achieved after Fagier et al. [226] presented the efficiency of coagulation-floccu-
2 h. Also Ayoub and Ghauch [50] applied activated PDS for sulfa- lation pretreatment coupled with a SO4%− oxidation process in the
methoxazole degradation and determined that the metallic iron-based removal and mineralization of organic matter of sugarcane vinasse.
particles (heterogeneous systems) are more efficient than conventional Ferric chloride (15 g L−1), a standard coagulation agent in wastewater
Fe2+ fed systems (homogeneous systems) for the removal of this treatment plants (WWTP), was used and achieved a 70% TOC removal.
pharmaceutical. The pretreated vinasse subjected to a PDS/PMS oxidation process (ac-
Chen et al. [197] examined the performance of thermally activated tivated by Fe2+) showed the highest TOC removal efficiency at pH 7.
PDS on the degradation of diclofenac in both water and polluted Under the selected optimum conditions, approximately 70 and 49%

56
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

TOC removal was achieved for PMS/Fe2+ and PDS/Fe2+, respectively. 89.8 cm3 g−1 to 30.6 cm3 g−1 were also observed. A large issue for
Also, Rodríguez-Chueca et al. [227] investigated winery wastewater introduction of the presented method in WWTP, concerns heat activa-
treatment using PMS coupled with a transition metal and UV light. High tion of persulfates. However, the temperature threshold sufficient for
COD and TOC removal efficacy (79% and 64%, respectively) was ob- PDS to rapidly form radicals (50–90 °C) can be reached using e.g. the
served under optimal conditions after three hours of treatment. heat generated in the fermentation process, steam or hot air injection
In another study, wastewater containing cytosine arabinoside (ara- [237].
C) was treated with PDS and H2O2 activated by UV radiation [228]. It Probably the first reference to the use of PMS for chemical disin-
was found that addition of oxidants considerably increased the ara-C tegration of waste activated sludge was made in our recent papers
removal effectiveness, and the TOC content in the wastewater declined [238,239]. Similarly to an earlier study [236], it was concluded that
with longer oxidation period but the toxicity increased, surprisingly, heat application (50, 70 and 90 °C) for PMS activation causes an in-
mainly in UV/H2O2 system. Shu et al. [229] investigated the UV/PMS crease in the soluble COD value and protein concentration in the su-
degradation of Acid Blue 113 containing wastewater. They observed pernatant and positively influences the SVI, which decreased from 89.8
that there was no correlation between initial pH value and dye removal to 17.2 mL g−1. Also, Niu et al. [240] and Liu et al. [241] observed
efficiency but UV light intensity significantly affected the efficiency of positive effects of WAS oxidation with PMS. Sludge disintegration was
TOC removal. characterised by a change in disintegration degree (DD), sludge particle
Heterogeneous activation of PMS could also be used for the re- size, and the properties of extracellular polymeric substances.
mediation of organic contaminants in wastewater according to [27]. Although thermally activated persulfates are efficacious for the
They concluded that a CoMn2O4 catalyst was efficient for the hetero- disintegration and improvement of sludge sedimentation properties
geneous activation of PMS and environmentally friendly. However, it [238], Zhen et al. [242] observed a possible inhibitory effect on
showed almost no catalytic activity to PDS and H2O2. They observed anaerobic digestion. On the contrary, Sun et al. [243] found that PDS
that Rhodamine B degradation in wastewater was enhanced with an disintegration had a positive influence on the biogas yield. Therefore,
increase in reaction temperature (15–55 °C) and inhibited with an in- the composition of the sludge and the type of fermentation could be
crease in fulvic acid concentration (0–0.08 g L−1). crucial for assessing the benefits of persulfates for WAS disintegration.
To date, a very limited number of papers have been published Sludge treatment with persulfates can also be focused on anaero-
concerning the use of persulfates for sludge disintegration. There are bically digested sludge as reported in a recent study [244] and in a
many investigations in the matter of sludge disintegration by PDS and recent article published in Nature: Scientific Reports [245]. In addition
only a few focusing on the disintegration of activated sludge by PMS. to the enhancement of dewaterability, good efficiency of toluene re-
The methods often used for determining the degree of waste activated moval from anaerobically digested sludge could be observed after the
sludge (WAS) disintegration include the measurement of soluble che- treatment with persulfates catalysed to form radicals with elevated
mical oxygen demand (SCOD) and the sludge volume index (SVI). temperatures from meso- or thermophilic digestion [244].
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 6. Conclusions
the settleability of the sludge, which can be measured in a 1000 mL
measuring cylinder after 30 min of sedimentation and expressed for a Despite persulfates being efficacious substances for the remediation
known initial sludge concentration. One of the first references to per- of water, wastewater and sludge media, the decay of pollutants is ex-
sulfates being used for sludge disintegration can be found in a study by tremely dependent on activation techniques and the composition of the
[230], who observed that ferrous iron activated PDS has a positive ef- treated water matrix.
fect on enhancing sludge dewaterability with an 88.8% capillary suc- Non-activated persulfates react at rates that are often considered
tion time (CST) reduction within 1 min. The purpose of a CST test is to slow, but fast reactions between the persulfates and free radicals gen-
characterize the sludge dewaterability rapidly and easily. The time the erated from them can increase the significance of these processes.
filtrate requires to travel a fixed distance in the filter paper is referred to Although non-catalysed persulfate reactions possess advantages, i.e.
as the capillary suction time [231]. Similar results to Zhen et al. [232] lower cost, no secondary contamination due to the catalyst load and
were obtained by Shi et al. [233], whereby the highest specific re- higher stability in the subsurface, drawbacks i.e. often much slower
sistance to filtration (SRF; which is another often used method for de- reaction rates with contaminants and formation of stable disinfection
waterability assessment) and CST reduction efficiencies of 88.5 and by-products, which reduce natural attenuation, favour the use of acti-
91.5%, respectively, were acquired after PDS/Fe2+ oxidation. Electro- vated persulfate reactions.
activated PDS, has also been applied for sludge treatment [234] and it Activation can be achieved by versatile means such as heat, UV
was concluded that the process can be potentially applied to deal with radiation, radiolysis or by chemical methods, allowing to establish
wastewater from toluene nitration processes. In addition, it was de- SO4%−-based oxidation in very different fields such as remediation,
termined that 2,4-dinitrotoluene in wastewater under electro-activated wastewater and drinking water treatment.
PDS oxidation can mainly be treated by virtue of SO4%− descended Chemical activation of persulfates can be considered either a
from the reduction of PDS anions. Also, Zhen et al. [235] performed homogenous or a heterogeneous reaction. Both of these radical initia-
electrolysis/PDS/Fe2+ oxidation to improve sludge dewaterability by tion types have benefits and drawbacks, although especially the high
disrupting the protective barrier and cracking the entrapped cells, cost, chemical stability (leaching of catalyst constituents), long pre-
which resulted in releasing the water inside extracellular polymeric paration time, and chemical stability of heterogeneous catalysts in some
substances and cells. Zhen et al. [232] found that a combination of PDS cases still largely limits their use in water treatment. In the subsurface
and thermal processes (at a mild temperature) is efficient in enhancing the water quality parameters are of great importance. The pH value and
the dewaterability of sludge. They concluded that when the tempera- concentrations of halogens/natural organic matter were found to be
ture is increased to 80 °C in the presence of PDS, the flocs of waste especially crucial. In most of the systems pH of ∼ 3 was found to be
activated sludge were drastically changed and that this pretreatment beneficial for rapid decontamination (although typically the pH of
resulted in the disruption of sludge flocs by degrading extracellular natural waters is not that low). However, elevated pH values can en-
polymeric substances. These results were confirmed in a recent study hance the reaction rates due to the additional activation (proven for
focusing on the disintegration of sludge with heat activated PDS [236]. both persulfates) and the faster reaction of radicals with deprotonated
It was observed that organic matter and polymer transfer from the solid compounds. Natural organic matter and halogens are of relevance due
phase to the liquid phase occurred. An increase in SCOD, (almost a 15- to their radical scavenging and them being a precursor to organic dis-
fold increase over the WAS value) and a decrease in the SVI from infection by-products. On the contrary, activation can be caused

57
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

incidentally by specific site conditions (e.g. large transition metal [13] G. Fan, L. Cang, G. Fang, W. Qin, L. Ge, D. Zhou, Electrokinetic delivery of per-
sulfate to remediate PCBs polluted soils: effect of injection spot, Chemosphere 117
content or elevated pH). (2014) 410–418.
In the catalysed persulfate systems the SO4%− is often considered to [14] C.H. Yen, K.F. Chen, C.M. Kao, S.H. Liang, T.Y. Chen, Application of persulfate to
be the main oxidative species. However, depending on reaction con- remediate petroleum hydrocarbon-contaminated soil: feasibility and comparison
with common oxidants, J. Hazard. Mater. 186 (2011) 2097–2102.
ditions other reactive species such as hydroxyl radical and superoxide [15] X. Duan, H. Sun, J. Kang, Y. Wang, S. Indrawirawan, S. Wang, Insights into het-
radical were determined by many authors to play an important role, erogeneous catalysis of persulfate activation on dimensional-structured nano-
too. In case SO4%− is converted to other reactive species, sometimes the carbons, ACS Catal. 5 (2015) 4629–4636.
[16] I.M. Kolthoff, I.K. Miller, The Chemistry of Persulfate. I. The kinetics and me-
unique features of SO4%− cannot be exploited. This is important to chanism of the decomposition of the persulfate ion in aqueous medium 1, J. Am.
know in case very persistent compounds such as chlorotriazine diamine Chem. Soc. 73 (1951) 3055–3059.
or polyfluorinated compounds (e.g. PFOA) have to be degraded, which [17] E.J. Behrman, D.H. Dean, Sodium peroxydisulfate is a stable and cheap substitute
for ammonium peroxydisulfate (persulfate) in polyacrylamide gel electrophoresis,
are inert towards the conventional degradation methods.
J. Chromatogr. B. Biomed. Sci. Appl. 723 (1999) 325–326.
In the present review analytical methods for quantification of per- [18] S.G. Huling, B. Pivetz, In-situ chemical oxidation–engineering issue. EPA/600/R-
sulfates were presented and assessed. Spectrophotometric methods are 06/072, (2007).
among the most commonly used for persulfate determination; however, [19] S. Yang, P. Wang, X. Yang, L. Shan, W. Zhang, X. Shao, R. Niu, Degradation ef-
ficiencies of azo dye Acid Orange 7 by the interaction of heat, UV and anions with
it should be noted that the liquid chromatographic methods are among common oxidants: persulfate, peroxymonosulfate and hydrogen peroxide, J.
the most reliable with a low quantification limit. Hazard. Mater. 179 (2010) 552–558.
Data on decontamination kinetics using activated persulfates show [20] D.A. Reckhow, B. Langlais, D.R. Brink, AWWA Research Foundation., Compagnie
générale des eaux (Paris, France), Ozone in water treatment : application and
that such reactions are regarded as extremely fast in comparison to engineering : cooperative research report / American Water Works Association
biological or chemical reductive treatment (However, it should be Research Foundation, Compagnie generale des eaux ; edited by Bruno Langlais,
noted, that some compounds might be degraded only by the chemical 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. Croué, Efficient perox-
reduction). The pseudo-first order kinetic model (experiments con- ydisulfate activation process not relying on sulfate radical generation for water
ducted with an excess of oxidant) is mostly used. pollutant degradation, Environ. Sci. Technol. 48 (2014) 5868–5875.
Persulfate decontamination technologies either with oxidation via [22] A.G. Bailie, K. Bouzek, P. Lukášek, I. Roušar, A.A. Wragg, Solubility of potassium
ferrate in 12 m alkaline solutions between 20°C and 60°C, J. Chem. Technol.
direct electron transfer or free radical driven processes were found to be
Biotechnol. 66 (1996) 35–40.
very powerful tools for the remediation of a wide range of con- [23] S.W. Benson, Thermochemistry and kinetics of sulfur-containing molecules and
taminants, including chlorinated olefins, BTEXs, phenols, pharmaceu- radicals, Chem. Rev. 78 (1978) 23–35.
[24] H. Hussain, I.R. Green, I. Ahmed, Journey describing applications of oxone in
ticals, inorganics and pesticides. However, several disadvantages, i.e.
synthetic chemistry, Chem. Rev. 113 (2013) 3329–3371.
pH changes, salinity of the soil and creation of hazardous decontami- [25] P.R. Shukla, S. Wang, H. Sun, H.M. Ang, M. Tadé, Activated carbon supported
nation by-products should be seriously taken into account before their cobalt catalysts for advanced oxidation of organic contaminants in aqueous so-
application. lution, Appl. Catal. B Environ. 100 (2010) 529–534.
[26] G. Lente, J. Kalmár, Z. Baranyai, A. Kun, I. Kék, D. Bajusz, M. Takács, L. Veres,
I. Fábián, One - versus two - electron oxidation with peroxomonosulfate ion: re-
Acknowledgements actions with iron(II), vanadium(IV), halide ions, and photoreaction with cerium
(III), Inorg. Chem. 48 (2009) 1763–1773.
[27] Y. Yao, Y. Cai, G. Wu, F. Wei, X. Li, H. Chen, S. Wang, Sulfate radicals induced
The work was supported by the project LO1201, the financial sup- from peroxymonosulfate by cobalt manganese oxides (CoxMn3−xO4) for Fenton-
port of the Ministry of Education, Youth and Sports in the framework of Like reaction in water, J. Hazard. Mater. 296 (2015) 128–137.
the targeted support of the “National Programme for Sustainability I” [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
and the OPR & DI project “Centre for Nanomaterials, Advanced in-situ dimethyldioxirane: mechanism and reactivity of the oxidants,
Technologies and Innovation - CZ.1.05/2.1.00/01.0005”. The authors Holzforschung 52 (1998) 499–505.
also acknowledge the assistance provided by the Research [30] P. Bajpai, Chapter Seven – Peroxyacids bleaching, in: Environ. Benign Approaches
Pulp Bleach, Gulf Professional Publishing, 2012 167–188.
Infrastructure NanoEnviCz, supported by the Ministry of Education, [31] P.A. Block, R.A. Brown, D. Robinson, Novel activation technologies for sodium
Youth and Sports of the Czech Republic under Project No. LM2015073. persulfate In Situ chemical oxidation. in: Proceedings, Fourth International
Conference on Remediation of Chlorinated and Recalcitrant Compounds,
Monterey, CA, USA. May 24-27. Monterey, (2004) 2 A–05.
References
[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-,
[1] A. Gosset, Y. Ferro, C. Durrieu, Methods for evaluating the pollution impact of UV/S2O82-, Water Res. 80 (2015) 99–108.
urban wet weather discharges on biocenosis: a review, Water Res. 89 (2016) [33] Y. Ji, W. Xie, Y. Fan, Y. Shi, D. Kong, J. Lu, Degradation of trimethoprim by
330–354. thermo-activated persulfate oxidation: reaction kinetics and transformation me-
[2] P.J. Crutzen, S. Waclawek, Atmospheric chemistry and climate in the anthro- chanisms, Chem. Eng. J. 286 (2016) 16–24.
pocene / Chemia atmosferyczna i klimat w antropocenie, Chemistry-Didactics- [34] R.L. Johnson, P.G. Tratnyek, R.O.B. Johnson, Persulfate persistence under thermal
Ecology-Metrology 19 (2014) 9–28. activation conditions, Environ. Sci. Technol. 42 (2008) 9350–9356.
[3] J.L. Martinez, Environmental pollution by antibiotics and by antibiotic resistance [35] A. Ghauch, A.M. Tuqan, N. Kibbi, Ibuprofen removal by heated persulfate in
determinants, Environ. Pollut. 157 (2009) 2893–2902. aqueous solution: a kinetics study, Chem. Eng. J. 197 (2012) 483–492.
[4] T.A. Larsen, S. Hoffmann, C. Lüthi, B. Truffer, M. Maurer, Emerging solutions to [36] A. Ghauch, A.M. Tuqan, Oxidation of bisoprolol in heated persulfate/H2O sys-
the water challenges of an urbanizing world, Science 352 (2016) 928–933. tems: kinetics and products, Chem. Eng. J. 183 (2012) 162–171.
[5] T.M. Phillips, A.G. Seech, H. Lee, J.T. Trevors, Biodegradation of hexa- [37] R. Zhang, P. Sun, T.H. Boyer, L. Zhao, C.H. Huang, Degradation of pharmaceu-
chlorocyclohexane (HCH) by microorganisms, Biodegradation 16 (2005) 363–392. ticals and metabolite in synthetic human urine by UV, UV/H2O2, and UV/PDS,
[6] D. Bass, N. Hastings, R. Brown, Performance of air sparging systems: a review of Environ. Sci. Technol. 49 (2015) 3056–3066.
case studies, J. Hazard. Mater. 72 (2000) 101–119. [38] W.S. Chen, Y.C. Su, Removal of dinitrotoluenes in wastewater by sono-activated
[7] D.W. Elliott, H.-L. Lien, W.-X. Zhang, Degradation of Lindane by zero-valent iron persulfate, Ultrason. Sonochem. 19 (2012) 921–927.
nanoparticles, J. Environ. Eng. 135 (2009) 317–324. [39] J. Criquet, N. Karpel, Vel Leitner, Electron beam irradiation of aqueous solution of
[8] C. von Sonntag, U. von Gunten, Chemistry of Ozone in Water and Wastewater persulfate ions, Chem. Eng. J. 169 (2011) 258–262.
Treatment, IWA Pub., London; New York, 2012. [40] Y. Lei, C.-S. Chen, Y.-J. Tu, Y.-H. Huang, H. Zhang, Heterogeneous Degradation of
[9] F.R. Spellman, Choosing Disinfection Alternatives for Water/Wastewater Organic Pollutants by Persulfate Activated by CuO-Fe3O4: mechanism, Stability,
Treatment Plants, CRC Press, 1999 https://books.google.com/books?id=Fnk- and Effects of pH and Bicarbonate Ions, Environ. Sci. Technol. 49 (2015)
FkablroC & pgis=1. 6838–6845.
[10] K. Kaur, M. Crimi, Release of chromium from soils with persulfate chemical oxi- [41] Y. Ding, L. Zhu, N. Wang, H. Tang, Sulfate radicals induced degradation of tet-
dation, Ground Water 52 (2014) 748–755. rabromobisphenol A with nanoscaled magnetic CuFe2O4 as a heterogeneous cat-
[11] X. Xu, S. Li, Q. Hao, J. Liu, Y. Yu, H. Li, Activation of persulfate and its en- alyst of peroxymonosulfate, Appl. Catal. B Environ. 129 (2013) 153–162.
vironmental application, Int. J. Environ. Bioenergy. 1 (2012) 60–81. [42] A. Ghauch, G. Ayoub, S. Naim, Degradation of sulfamethoxazole by persulfate
[12] R.L. Siegrist, M. Crimi, T.J. Simpkin, In Situ Chemical Oxidation for Groundwater assisted micrometric Fe0 in aqueous solution, Chem. Eng. J. 228 (2013)
Remediation. Chapter 2: fundamentals of ISCO using hydrogen peroxide, Springer; 1168–1181.
2011 edition (March 22, 2011). [43] G. Fang, J. Gao, D.D. Dionysiou, C. Liu, D. Zhou, Activation of persulfate by

58
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

quinones: free radical reactions and implication for the degradation of PCBs, naturally occurring chelating agents on Fe(II) mediated advanced oxidation of
Environ. Sci. Technol. 47 (2013) 4605–4611. chlorophenols, Water Res. 43 (2009) 684–694.
[44] O.S. Furman, A.L. Teel, R.J. Watts, Mechanism of base activation of persulfate, [73] X. Chen, J. Chen, X. Qiao, D. Wang, X. Cai, Performance of nano-Co3O4/perox-
Environ. Sci. Technol. 44 (2010) 6423–6428. ymonosulfate system: kinetics and mechanism study using Acid Orange 7 as a
[45] C. Qi, X. Liu, J. Ma, C. Lin, X. Li, H. Zhang, Activation of peroxymonosulfate by model compound, Appl. Catal. B Environ. 80 (2008) 116–121.
base: implications for the degradation of organic pollutants, Chemosphere 151 [74] S. Muhammad, E. Saputra, H. Sun, J. de C. Izidoro, D.A. Fungaro, H.M. Ang,
(2016) 280–288. M.O. Tadé, S. Wang, Coal fly ash supported Co3O4 catalysts for phenol degradation
[46] Y. Yang, J. Jiang, X. Lu, J. Ma, Y. Liu, Production of sulfate radical and hydroxyl using peroxymonosulfate, RSC Adv. 2 (2012) 5645–5652.
radical by reaction of ozone with peroxymonosulfate: a novel advanced oxidation [75] G.P. Anipsitakis, E. Stathatos, D.D. Dionysiou, Heterogeneous activation of Oxone
process, Environ. Sci. Technol. 49 (2015) 7330–7339. using Co3O4, J. Phys. Chem. B. 109 (2005) 13052–13055.
[47] M. Ahmad, A.L. Teel, R.J. Watts, Mechanism of persulfate activation by phenols, [76] Q. Yang, H. Choi, S.R. Al-Abed, D.D. Dionysiou, Iron-cobalt mixed oxide nano-
Environ. Sci. Technol. 47 (2013) 5864–5871. catalysts: heterogeneous peroxymonosulfate activation, cobalt leaching, and fer-
[48] L. Zhao, H. Hou, A. Fujii, M. Hosomi, F. Li, Degradation of 1,4-dioxane in water romagnetic properties for environmental applications, Appl. Catal. B Environ. 88
with heat- and Fe2+-activated persulfate oxidation, Environ. Sci. Pollut. Res. 21 (2009) 462–469.
(2014) 7457–7465. [77] Q. Yang, H. Choi, Y. Chen, D.D. Dionysiou, Heterogeneous activation of perox-
[49] G.P. Anipsitakis, D.D. Dionysiou, Transition metal/UV-based advanced oxidation ymonosulfate by supported cobalt catalysts for the degradation of 2,4-di-
technologies for water decontamination, Appl. Catal. B Environ. 54 (2004) chlorophenol in water: the effect of support, cobalt precursor, and UV radiation,
155–163. Appl. Catal. B Environ. 77 (2008) 300–307.
[50] G. Ayoub, A. Ghauch, Assessment of bimetallic and trimetallic iron-based systems [78] Q. Yang, H. Choi, D.D. Dionysiou, Nanocrystalline cobalt oxide immobilized on
for persulfate activation: application to sulfamethoxazole degradation, Chem. Eng. titanium dioxide nanoparticles for the heterogeneous activation of perox-
J. 256 (2014) 280–292. ymonosulfate, Appl. Catal. B Environ. 74 (2007) 170–178.
[51] S. Naim, A. Ghauch, Ranitidine abatement in chemically activated persulfate [79] J. Zhang, M. Chen, L. Zhu, Activation of persulfate by Co3O4 nanoparticles for
systems: assessment of industrial iron waste for sustainable applications, Chem. orange G degradation, RSC Adv. 6 (2016) 758–768.
Eng. J. 288 (2016) 276–288. [80] Y. Yao, Z. Yang, D. Zhang, W. Peng, H. Sun, S. Wang, Magnetic CoFe2O4–graphene
[52] Y.H. Guan, J. Ma, X.C. Li, J.Y. Fang, L.W. Chen, Influence of pH on the formation hybrids: facile synthesis, characterization, and catalytic properties, Ind. Eng.
of sulfate and hydroxyl radicals in the UV/Peroxymonosulfate system, Environ. Chem. Res. 51 (2012) 6044–6051.
Sci. Technol. 45 (2011) 9308–9314. [81] S. Su, W. Guo, Y. Leng, C. Yi, Z. Ma, Heterogeneous activation of Oxone by
[53] G.P. Anipsitakis, D.D. Dionysiou, Degradation of organic contaminants in water CoxFe3-xO4 nanocatalysts for degradation of rhodamine B, J. Hazard. Mater.
with sulfate radicals generated by the conjunction of peroxymonosulfate with 244–245 (2013) 736–742.
cobalt, Environ. Sci. Technol. 37 (2003) 4790–4797. [82] K.-Y.A. Lin, F.-K. Hsu, W.-D. Lee, Magnetic cobalt–graphene nanocomposite de-
[54] J. Fernandez, P. Maruthamuthu, A. Renken, J. Kiwi, Bleaching and photobleaching rived from self-assembly of MOFs with graphene oxide as an activator for perox-
of Orange II within seconds by the oxone/Co2+ reagent in Fenton-like processes, ymonosulfate, J. Mater. Chem. A. 3 (2015) 9480–9490.
Appl. Catal. B Environ. 49 (2004) 207–215. [83] K.-Y.A. Lin, B.-C. Chen, Efficient elimination of caffeine from water using Oxone
[55] S. Wacławek, V. Antoš, P. Hrabák, M. Černík, D. Elliott, Remediation of hexa- activated by a magnetic and recyclable cobalt/carbon nanocomposite derived
chlorocyclohexanes by electrochemically activated persulfates, Environ. Sci. from ZIF-67, Dalt. Trans. 45 (2016) 3541–3551.
Pollut. Res. 23 (2016) 765–773. [84] P.H. Shi, S.B. Zhu, H.G. Zheng, D.X. Li, S.H. Xu, Supported Co3O4 on expanded
[56] D. Zhao, X. Liao, X. Yan, S.G. Huling, T. Chai, H. Tao, Effect and mechanism of graphite as a catalyst for the degradation of Orange II in water using sulfate ra-
persulfate activated by different methods for PAHs removal in soil, J. Hazard. dicals, Desalin. Water Treat. 52 (2014) 3384–3391.
Mater. 254–255 (2013) 228–235. [85] Y. Ding, L. Zhu, A. Huang, X. Zhao, X. Zhang, H. Tang, A heterogeneous
[57] H. Herrmann, A. Reese, R. Zellner, Time-resolved UV/VIS diode array absorption Co3O4–Bi2O3 composite catalyst for oxidative degradation of organic pollutants in
spectroscopy of SOx- (x=3, 4, 5) radical anions in aqueous solution, J. Mol. Struct. the presence of peroxymonosulfate, Catal. Sci. Technol. 2 (2012) 1977–1984.
348 (1995) 183–186. [86] S. Sarkar, M.S. Adhikari, M. Banerjee, R.S. Konar, Thermal decomposition of po-
[58] G. Mark, M.N. Schuchmann, H.-P. Schuchmann, C. von Sonntag, The photolysis of tassium persulfate in aqueous solution at 50 °C in an inert atmosphere of nitrogen
potassium peroxodisulphate in aqueous solution in the presence of tert-butanol: a in the presence of acrylonitrile monomer, J. Appl. Polym. Sci. 35 (1988)
simple actinometer for 254 nm radiation, J. Photochem. Photobiol. A Chem. 55 1441–1458.
(1990) 157–168. [87] T. Rhadfi, J.Y. Piquemal, L. Sicard, F. Herbst, E. Briot, M. Benedetti, A. Atlamsani,
[59] A. Ghauch, A. Baalbaki, M. Amasha, R. El Asmar, O. Tantawi, Contribution of Polyol-made Mn3O4 nanocrystals as efficient Fenton-like catalysts, Appl. Catal. A
persulfate in UV-254 nm activated systems for complete degradation of chlor- Gen. 386 (2010) 132–139.
amphenicol antibiotic in water, Chem. Eng. J. 317 (2017) 1012–1025. [88] X. Zhai, W. Yang, M. Li, G. Lv, J. Liu, X. Zhang, Noncovalent hybrid of CoMn2O4
[60] W.-D. Oh, Z. Dong, T.-T. Lim, Generation of sulfate radical through heterogeneous spinel nanocrystals and poly (diallyldimethylammonium chloride) functionalized
catalysis for organic contaminants removal: current development, challenges and carbon nanotubes as efficient electrocatalysts for oxygen reduction reaction,
prospects, Appl. Catal. B Environ. 194 (2016) 169–201. Carbon N. Y. 65 (2013) 277–286.
[61] A. Rastogi, S.R. Al-Abed, D.D. Dionysiou, Sulfate radical-based ferrous-perox- [89] Y. Wang, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, Synthesis of magnetic core/shell
ymonosulfate oxidative system for PCBs degradation in aqueous and sediment carbon nanosphere supported manganese catalysts for oxidation of organics in
systems, Appl. Catal. B Environ. 85 (2009) 171–179. water by peroxymonosulfate, J. Colloid Interface Sci. 433 (2014) 68–75.
[62] S. Yuan, P. Liao, A.N. Alshawabkeh, Electrolytic manipulation of persulfate re- [90] Y. Wang, H. Sun, H.M. Ang, M.O. Tadé, S. Wang, 3D-hierarchically structured
activity by iron electrodes for trichloroethylene degradation in groundwater, MnO2 for catalytic oxidation of phenol solutions by activation of perox-
Environ. Sci. Technol. 48 (2014) 656–663. ymonosulfate: structure dependence and mechanism, Appl. Catal. B Environ. 164
[63] H. Lee, H.J. Lee, J. Jeong, J. Lee, N.B. Park, C. Lee, Activation of persulfates by (2015) 159–167.
carbon nanotubes: oxidation of organic compounds by nonradical mechanism, [91] Y. Li, L.D. Liu, L. Liu, Y. Liu, H.W. Zhang, X. Han, Efficient oxidation of phenol by
Chem. Eng. J. 266 (2015) 28–33. persulfate using manganite as a catalyst, J. Mol. Catal. A Chem. 411 (2016)
[64] K. Govindan, M. Raja, M. Noel, E.J. James, Degradation of pentachlorophenol by 264–271.
hydroxyl radicals and sulfate radicals using electrochemical activation of perox- [92] C. Tan, N. Gao, Y. Deng, J. Deng, S. Zhou, J. Li, X. Xin, Radical induced de-
omonosulfate, peroxodisulfate and hydrogen peroxide, J. Hazard. Mater. 272 gradation of acetaminophen with Fe3O4 magnetic nanoparticles as heterogeneous
(2014) 42–51. activator of peroxymonosulfate, J. Hazard. Mater. 276 (2014) 452–460.
[65] H. Sun, S. Liu, G. Zhou, H.M. Ang, M.O. Tadé, S. Wang, Reduced graphene oxide [93] W.-D. Oh, Z. Dong, Z.-T. Hu, T.-T. Lim, A novel quasi-cubic CuFe2O4-Fe2O3 cat-
for catalytic oxidation of aqueous organic pollutants, ACS Appl. Mater. Interfaces. alyst prepared at low temperature for enhanced oxidation of bisphenol A via
4 (2012) 5466–5471. peroxymonosulfate activation, J. Mater. Chem. A. 3 (2015) 22208–22217.
[66] Y. Tao, M. Wei, D. Xia, A. Xu, X. Li, Polyimides as metal-free catalysts for organic [94] J. Zhang, M. Chen, L. Zhu, Activation of peroxymonosulfate by iron-based cata-
dye degradation in the presence peroxymonosulfate under visible light irradiation, lysts for orange G degradation: role of hydroxylamine, RSC Adv. 6 (2016)
RSC Adv. 5 (2015) 98231–98240. 47562–47569.
[67] Y. Deng, C.M. Ezyske, Sulfate radical-advanced oxidation process (SR-AOP) for [95] A. Ghauch, A.M. Tuqan, N. Kibbi, Naproxen abatement by thermally activated
simultaneous removal of refractory organic contaminants and ammonia in landfill persulfate in aqueous systems, Chem. Eng. J. 279 (2015) 861–873.
leachate, Water Res. 45 (2011) 6189–6194. [96] Y. Wang, Z. Ao, H. Sun, X. Duan, S. Wang, Activation of peroxymonosulfate by
[68] S. Ahmed, M.G. Rasul, W.N. Martens, R. Brown, M.A. Hashib, Heterogeneous carbonaceous oxygen groups: experimental and density functional theory calcu-
photocatalytic degradation of phenols in wastewater: a review on current status lations, Appl. Catal. B Environ. 198 (2016) 295–302.
and developments, Desalination 261 (2010) 3–18. [97] J. Zhang, X. Shao, C. Shi, S. Yang, Decolorization of Acid Orange 7 with perox-
[69] H. Zhang, H. Fu, D. Zhang, Degradation of C.I. Acid Orange 7 by ultrasound en- ymonosulfate oxidation catalyzed by granular activated carbon, Chem. Eng. J. 232
hanced heterogeneous Fenton-like process, J. Hazard. Mater. 172 (2009) 654–660. (2013) 259–265.
[70] A. Tsitonaki, B. Petri, M. Crimi, H. Mosbaek, R.L. Siegrist, P.L. Bjerg, In situ [98] E. Saputra, S. Muhammad, H. Sun, S. Wang, Activated carbons as green and ef-
chemical oxidation of contaminated soil and groundwater using persulfate: a re- fective catalysts for generation of reactive radicals in degradation of aqueous
view, Crit Rev Env Sci Tec. 40 (2010) 55–91. phenol, RSC Adv. 3 (2013) 21905–21910.
[71] B.-T. Zhang, Y. Zhang, Y. Teng, M. Fan, Sulfate radical and its application in de- [99] K.-Y. Andrew Lin, Z.-Y. Zhang, α-Sulfur as a metal-free catalyst to activate per-
contamination technologies, Crit. Rev. Environ. Sci. Technol. 45 (2015) oxymonosulfate under visible light irradiation for decolorization, RSC Adv. 6
1756–1800. (2016) 15027–15034.
[72] A. Rastogi, S.R. Al-Abed, D.D. Dionysiou, Effect of inorganic, synthetic and [100] S.-Y. Oh, S.-G. Kang, Degradation of 2,4-dinitrotoluene by persulfate and steel

59
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

waste powder, Geosystem Eng. 13 (2010) 105–110. aliphatic radicals by the peroxydisulfate anion, J. Chem. Soc. Perkin Trans. 2 Phys.
[101] H.V. Lutze, N. Kerlin, T.C. Schmidt, Sulfate radical-based water treatment in Org. Chem. (1984) 503–509.
presence of chloride: formation of chlorate, inter-conversion of sulfate radicals [130] W. Roebke, M. Renz, A. Henglein, Pulsradiolyse der anionen S2O82− und HSO5−
into hydroxyl radicals and influence of bicarbonate, Water Res. 72 (2015) in Wässriger Lösung, Int. J. Radiat. Phys. Chem. 1 (1969) 39–44.
349–360. [131] P. Maruthamuthu, P. Neta, Radiolytic chain decomposition of peroxomonopho-
[102] D.P. Cassidy, V.J. Srivastava, F.J. Dombrowski, J.W. Lingle, Combining in situ sphoric and peroxomonosulfuric acids, J. Phys. Chem. 81 (1977) 937–940.
chemical oxidation, stabilization, and anaerobic bioremediation in a single ap- [132] B.C. Gilbert, J.K. Stell, Mechanisms of peroxide decomposition. An ESR study of
plication to reduce contaminant mass and leachability in soil, J. Hazard. Mater. the reactions of the peroxomonosulfate anion (HOOSO3-) with titanium(III), iron
297 (2015) 347–355. (II), and α-oxygen substituted radicals, J. Chem. Soc. Perkin Trans. 2 Phys. Org.
[103] A. Long, H. Zhang, Selective oxidative degradation of toluene for the recovery of Chem. (1990) 1281–1288.
surfactant by an electro/Fe2+/persulfate process, Environ. Sci. Pollut. Res. 22 [133] Y. Lei, C.-S. Chen, J. Ai, H. Lin, Y.-H. Huang, H. Zhang, Selective decolorization of
(2015) 11606–11616. cationic dyes by peroxymonosulfate: non-radical mechanism and effect of
[104] N. Jaafarzadeh, F. Ghanbari, M. Moradi, Photo-electro-oxidation assisted perox- chloride, RSC Adv. 6 (2016) 866–871.
ymonosulfate for decolorization of acid brown 14 from aqueous solution, Korean [134] Z. Wang, R.T. Bush, L.A. Sullivan, C. Chen, J. Liu, Selective oxidation of arsenite
J. Chem. Eng. 32 (2015) 458–464. by peroxymonosulfate with high utilization efficiency of oxidant, Environ. Sci.
[105] H. Lin, Y. Li, X. Mao, H. Zhang, Electro-enhanced goethite activation of perox- Technol. 48 (2014) 3978–3985.
ydisulfate for the decolorization of Orange II at neutral pH: efficiency, stability and [135] W. Stumm, J.J. Morgan, Aquatic Chemistry: Chemical Equilibria and Rates in
mechanism, J. Taiwan Inst. Chem. Eng. 65 (2016) 390–398. Natural Waters, John Wiley & Sons Inc, New York, 1995.
[106] J.S. Cao, W.X. Zhang, D.G. Brown, D. Sethi, Oxidation of lindane with Fe(II)-ac- [136] X. Xu, G. Pliego, J.A. Zazo, J.A. Casas, J.J. Rodriguez, Mineralization of naphtenic
tivated sodium persulfate, Environ. Eng. Sci. 25 (2008) 221–228. acids with thermally-activated persulfate: the important role of oxygen, J. Hazard.
[107] S. Khan, X. He, H.M. Khan, D. Boccelli, D.D. Dionysiou, Efficient degradation of Mater. 318 (2016) 355–362.
lindane in aqueous solution by iron (II) and/or UV activated peroxymonosulfate, [137] H. Herrmann, On the photolysis of simple anions and neutral molecules as sources
J. Photochem. Photobiol. A Chem. 316 (2016) 37–43. of O-/OH, SOx- and Cl in aqueous solution, Phys. Chem. Chem. Phys. 9 (2007)
[108] J. Cong, G. Wen, T. Huang, L. Deng, J. Ma, Study on enhanced ozonation de- 3935–3964.
gradation of para-chlorobenzoic acid by peroxymonosulfate in aqueous solution, [138] C. Liang, H.-W. Su, Identification of Sulfate and Hydroxyl Radicals in Thermally
Chem. Eng. J. 264 (2015) 399–403. Activated Persulfate, Ind. Eng. Chem. Res. 48 (2009) 5558–5562.
[109] L. Dogliotti, E. Hayon, Flash photolysis of persulfate ions in aqueous solutions. [139] G.D. Fang, D.D. Dionysiou, D.M. Zhou, Y. Wang, X.D. Zhu, J.X. Fan, L. Cang,
Study of the sulfate and ozonide radical anions, J. Phys. Chem. 71 (1967) Y.J. Wang, Transformation of polychlorinated biphenyls by persulfate at ambient
2511–2516. temperature, Chemosphere 90 (2013) 1573–1580.
[110] P.E.A. Boudeville, Simultaneous determination of hydrogen peroxide, perox- [140] G.D. Fang, D.D. Dionysiou, Y. Wang, S.R. Al-Abed, D.M. Zhou, Sulfate radical-
ymonosulfuric acid, and peroxydisulfuric acid by thermometric titrimetry, Anal. based degradation of polychlorinated biphenyls: effects of chloride ion and reac-
Chem. 55 (1983) 612–615. tion kinetics, J. Hazard. Mater. 227–228 (2012) 394–401.
[111] C. Liang, C.F. Huang, N. Mohanty, R.M. Kurakalva, A rapid spectrophotometric [141] P. Neta, V. Madhavan, H. Zemel, R.W. Fessenden, Rate constants and mechanism
determination of persulfate anion in ISCO, Chemosphere 73 (2008) 1540–1543. of reaction of sulfate radical anion with aromatic compounds, J. Am. Chem. Soc.
[112] K.-C. Huang, R.A. Couttenye, G.E. Hoag, Kinetics of heat-assisted persulfate oxi- 99 (1977) 163–164.
dation of methyl tert-butyl ether (MTBE), Chemosphere 49 (2002) 413–420. [142] A.E. Grigo’rev, I.E. Makarov, A.K. Pikaev, Formation of Cl2- in the bulk solution
[113] L. Zhao, S. Yang, L. Wang, C. Shi, M. Huo, Y. Li, Rapid and simple spectro- during the radiolysis of concentrated aqueous solutions of chlorides, Khimiya
photometric determination of persulfate in water by microwave assisted deco- Vysokikh Ehnergij 18 (1987) 123–126.
lorization of Methylene Blue, J. Environ. Sci. (China) 31 (2015) 235–239. [143] G.G. Jayson, B.J. Parsons, A.J. Swallow, Some simple, highly reactive, inorganic
[114] S. Gokulakrishnan, A. Mohammed, H. Prakash, Determination of persulphates chlorine derivatives in aqueous solution. Their formation using pulses of radiation
using N, N-diethyl-p-phenylenediamine as colorimetric reagent: oxidative col- and their role in the mechanism of the Fricke dosimeter, J. Chem. Soc Faraday
oration and degradation of the reagent without bactericidal effect in water, Chem. Trans. 1 (69) (1973) 1597–1607.
Eng. J. 286 (2016) 223–231. [144] R.H. Waldemer, P.G. Tratnyek, R.L. Johnson, J.T. Nurmi, Oxidation of chlorinated
[115] E. Villegas, Y. Pomeranz, J.A. Shellenberger, Colorimetric determination of per- ethenes by heat-activated persulfate: kinetics and products, Environ. Sci. Technol.
sulfate with alcian blue, Anal. Chim. Acta. 29 (1963) 145–148. 41 (2007) 1010–1015.
[116] K.C. Lin, J.Y. Huang, S.M. Chen, Poly(brilliant cresyl blue) electrodeposited on [145] K.-J. Kim, W.H. Hamill, Direct and indirect effects in pulse irradiated concentrated
multi-walled carbon nanotubes modified electrode and its application for persul- aqueous solutions of chloride and sulfate ions, J. Phys. Chem. 80 (1976)
fate determination, Int. J. Electrochem. Sci. 7 (2012) 9161–9173. 2320–2325.
[117] M. Roushani, E. Karami, Electrochemical detection of persulfate at the modified [146] W.J. McElroy, A laser photolysis study of the reaction of sulfate(1-) with chloride
glassy carbon electrode with nanocomposite containing nano-ruthenium oxide/ and the subsequent decay of chlorine(1-) in aqueous solution, J. Phys. Chem. 94
thionine and nano-ruthenium oxide/celestine blue, Electroanalysis 26 (2014) (1990) 2435–2441.
1761–1772. [147] D. Zehavi, J. Rabani, Oxidation of aqueous bromide ions by hydroxyl radicals.
[118] M.F. De Oliveira, R.J. Mortimer, N.R. Stradiotto, Voltammetric determination of Pulse radiolytic investigation, J. Phys. Chem. 76 (1972) 312–319.
persulfate anions using an electrode modified with a Prussian blue film, [148] J.L. Redpath, R.L. Willson, Chain reactions and radiosensitization: model enzyme
Microchem. J. 64 (2000) 155–159. studies, Int. J. Radiat. Biol. Relat. Stud. Physics, Chem. Med. 27 (1975) 389–398.
[119] Z. Huang, C. Ni, F. Wang, Z. Zhu, Q. Subhani, M. Wang, Y. Zhu, Simultaneous [149] G.V. Buxton, N.D. Wood, S. Dyster, Ionisation constants of ·OH and HO·2 in aqueous
determination of peroxydisulfate and conventional inorganic anions by ion chro- solution up to 200°C. A pulse radiolysis study, J. Chem. Soc. Faraday Trans. 1
matography with the column-switching technique, J. Sep. Sci. 37 (2014) 198–203. Phys. Chem. Condens. Phases. 84 (1988) 1113–1121.
[120] S. Wacławek, K. Grübel, M. Černík, Simple spectrophotometric determination of [150] R.E. Huie, C.L. Clifton, Temperature dependence of the rate constants for reactions
monopersulfate, Spectrochim, Acta - Part A Mol. Biomol.Spectrosc. 149 (2015) of the sulfate radical, SO4-, with anions, J. Phys. Chem. 94 (1990) 8561–8567.
928–933. [151] S. Padmaja, P. Neta, R.E. Huie, Rate constants for some reactions of inorganic
[121] J. Zou, J. Ma, X. Zhang, P. Xie, Rapid spectrophotometric determination of per- radicals with inorganic ions. Temperature and solvent dependence, Int. J. Chem.
oxymonosulfate in water with cobalt-mediated oxidation decolorization of methyl Kinet. 25 (1993) 445–455.
orange, Chem. Eng. J. 253 (2014) 34–39. [152] H.V. Lutze, S. Bircher, I. Rapp, N. Kerlin, R. Bakkour, M. Geisler, C. Von Sonntag,
[122] S. Ossadnik, G. Schwedt, Comparative study of the determination of perox- T.C. Schmidt, Degradation of chlorotriazine pesticides by sulfate radicals and the
omonosulfate, in the presence of other oxidants, by capillary zone electrophoresis, influence of organic matter, Environ. Sci. Technol. 49 (2015) 1673–1680.
ion chromatography, and photometry, Fresenius. J. Anal. Chem. 371 (2001) [153] Y. Yang, J.J. Pignatello, J. Ma, W.A. Mitch, Comparison of halide impacts on the
420–424. efficiency of contaminant degradation by sulfate and hydroxyl radical-based ad-
[123] T. Zhang, H. Zhu, J.P. Croué, Production of sulfate radical from perox- vanced oxidation processes (AOPs), Environ. Sci. Technol. 48 (2014) 2344–2351.
ymonosulfate induced by a magnetically separable CuFe2O4 spinel in water: effi- [154] R.O.C. Norman, P.M. Storey, P.R. West, Electron spin resonance studies. Part XXV.
ciency, stability, and mechanism, Environ. Sci. Technol. 47 (2013) 2784–2791. Reactions of the sulphate radical anion with organic compounds, J. Chem. Soc. B.
[124] Y. Ding, L. Zhu, J. Yan, Q. Xiang, H. Tang, Spectrophotometric determination of (1970) 1087–1095.
persulfate by oxidative decolorization of azo dyes for wastewater treatment, J. [155] C. von Sonntag, H.-P. Schuchmann, Peroxyl radicals in aqueous solutions, John
Environ. Monit. 13 (2011) 3057–3063. Wiley & Sons Inc, Chiehester, England, 1997.
[125] J.Q. Zhang, J. Ma, J. Zou, H.Z. Chi, Y. Song, Spectrophotometric determination of [156] H.V. Lutze, R. Bakkour, N. Kerlin, C. von Sonntag, T.C. Schmidt, Formation of
peroxymonosulfate anions via oxidative decolorization of dyes induced by cobalt, bromate in sulfate radical based oxidation: mechanistic aspects and suppression
Anal. Methods. 8 (2016) 973–978. bydissolved organic matter, Water Res. 53 (2014) 370–377.
[126] E.J. Behrman, The Persulfate Oxidation of Phenols and Arylamines (The Elbs and [157] J.Y. Fang, C. Shang, Bromate formation from bromide oxidation by the UV/per-
the Boyland–Sims Oxidations), Org. React. (2004). sulfate process, Environ. Sci. Technol. 46 (2012) 8976–8983.
[127] G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical Review of rate [158] K. Liu, J. Lu, Y. Ji, Formation of brominated disinfection by-products and bromate
constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl ra- in cobalt catalyzed peroxymonosulfate oxidation of phenol, Water Res. 84
dicals in aqueous solution, J. Phys. Chem. Ref. Data. 17 (1988) 513–520. (2015) 1–7.
[128] G. Restelli, G. Angeletti (Eds.), Physico-Chemical Behaviour of Atmospheric [159] P. Xie, J. Ma, W. Liu, J. Zou, S. Yue, Impact of UV/persulfate pretreatment on the
Pollutants, Springer, Netherlands, Dordrecht, 1990. formation of disinfection byproducts during subsequent chlorination of natural
[129] M.J. Davies, B.C. Gilbert, R.O.C. Norman, Electron spin resonance. Part 67. organic matter, Chem. Eng. J. 269 (2015) 203–211.
Oxidation of aliphatic sulfides and sulfoxides by the sulfate radical anion and of [160] J. Lu, J. Wu, Y. Ji, D. Kong, Transformation of bromide in thermo activated

60
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

persulfate oxidation processes, Water Res. 78 (2015) 1–8. [189] H. Eskandarloo, A. Badiei, M.A. Behnajady, Optimization of UV/inorganic oxi-
[161] G.V. Buxton, M. Bydder, G. Arthur Salmon, K.J. Radford, C. Yamanaka, Reactivity dants system efficiency for photooxidative removal of an azo textile dye, Desalin.
of chlorine atoms in aqueous solution Part 1 The equilibrium ClMNsbd+Cl-Cl2-, J. Water Treat. 55 (2015) 210–226.
Chem. Soc. Faraday Trans. 94 (1998) 653–657. [190] A. Ghauch, A.M. Tuqan, N. Kibbi, S. Geryes, Methylene blue discoloration by
[162] A.B. Ross, P. Neta, Rate constants for reactions of inorganic radicals in aqueous heated persulfate in aqueous solution, Chem. Eng. J. 213 (2012) 259–271.
solution, U.S. Dept. of Commerce, National Bureau of Standards, U.S. Govt. Print. [191] N.S. Shah, X. He, H.M. Khan, J.A. Khan, K.E. O’Shea, D.L. Boccelli, D.D. Dionysiou,
Off., 1979. https://books.google.pl/books?id=l34-AAAAIAAJ. Efficient removal of endosulfan from aqueous solution by UV-C/peroxides: a
[163] R.A. Larson, R.G. Zepp, Reactivity of the carbonate radical with aniline deriva- comparative study, J. Hazard. Mater. 263 (2013) 584–592.
tives, Environ. Toxicol. Chem. 7 (1988) 265–274. [192] C. Tan, N. Gao, Y. Deng, Y. Zhang, M. Sui, J. Deng, S. Zhou, Degradation of an-
[164] S. Canonica, T. Kohn, M. Mac, F.J. Real, J. Wirz, U. Von Gunten, Photosensitizer tipyrine by UV, UV/H2O2 and UV/PS, J. Hazard. Mater. 260 (2013) 1008–1016.
method to determine rate constants for the reaction of carbonate radical with [193] T. Olmez-Hanci, C. Imren, I. Kabdash, O. Tünay, I. Arslan-Alaton, Application of
organic compounds, Environ. Sci. Technol. 39 (2005) 9182–9188. the UV-C photo-assisted peroxymonosulfate oxidation for the mineralization of
[165] P. Neta, R.E. Huie, A.B. Ross, Rate constants for reactions of inorganic radicals in dimethyl phthalate in aqueous solutions, Photochem. Photobiol. Sci. 10 (2011)
aqueous solution, J. Phys. Chem. Ref. Data. 17 (1988) 1027–1284. 343–349.
[166] D.M. Schwarzenbach, R.P., Gschwend, P.M. and Imboden, Photochemical trans- [194] A.R. Zarei, H. Rezaeivahidian, A.R. Soleymani, Mineralization of unsymmetrical
formation reactions. in: Environ. Org. Chem., John Wiley & Sons, Inc., New York, dimethylhydrazine (UDMH) via persulfate activated by zero valent iron nano
(1993) 436–484. particles: modeling, optimization and cost estimation, Desalin. Water Treat. 57
[167] I.A. Katsoyiannis, S. Canonica, U. von Gunten, Efficiency and energy requirements (2016) 16119–16128.
for the transformation of organic micropollutants by ozone, O3/H2O2 and UV/ [195] A. Kambhu, S. Comfort, C. Chokejaroenrat, C. Sakulthaew, Developing slow-re-
H2O2, Water Res. 45 (2011) 3811–3822. lease persulfate candles to treat BTEX contaminated groundwater, Chemosphere
[168] U. von Gunten, Ozonation of drinking water: part II, Disinfection and by-product 89 (2012) 656–664.
formation in presence of bromide, iodide or chlorine, Water Res. 37 (2003) [196] Y.T. Lin, C. Liang, J.H. Chen, Feasibility study of ultraviolet activated persulfate
1469–1487. oxidation of phenol, Chemosphere 82 (2011) 1168–1172.
[169] Y. Liu, A. Zhou, Y. Gan, X. Li, Variability in carbon isotope fractionation of tri- [197] J. Chen, Y. Qian, H. Liu, T. Huang, Oxidative degradation of diclofenac by ther-
chloroethene during degradation by persulfate activated with zero-valent iron: mally activated persulfate: implication for ISCO, Environ. Sci. Pollut. Res. Int. 23
effects of inorganic anions, Sci. Total Environ. 548–549 (2016) 1–5. (2016) 3824–3833.
[170] J. Lu, W. Dong, Y. Ji, D. Kong, Q. Huang, Natural organic matter exposed to sulfate [198] P. Manoj, K.P. Prasanthkumar, V.M. Manoj, U.K. Aravind, T.K. Manojkumar,
radicals increases its potential to form halogenated disinfection byproducts, C.T. Aravindakumar, Oxidation of substituted triazines by sulfate radical anion
Environ. Sci. Technol. 50 (2016) 5060–5067. (SO4•-) in aqueous medium: a laser flash photolysis and steady state radiolysis
[171] Q. Zhang, J. Chen, C. Dai, Y. Zhang, X. Zhou, Degradation of carbamazepine and study, J. Phys. Org. Chem. 20 (2007) 122–129.
toxicity evaluation using the UV/persulfate process in aqueous solution, J. Chem. [199] J.A. Khan, X. He, N.S. Shah, H.M. Khan, E. Hapeshi, D. Fatta-Kassinos,
Technol. Biotechnol. 90 (2015) 701–708. D.D. Dionysiou, Kinetic and mechanism investigation on the photochemical de-
[172] K. Temiz, T. Olmez-Hanci, I. Arslan-Alaton, Zero-valent iron-activated persulfate gradation of atrazine with activated H2O2, S2O82- and HSO5-, Chem. Eng. J. 252
oxidation of a commercial alkyl phenol polyethoxylate, Environ. Technol. 37 (2014) 393–403.
(2016) 1757–1767. [200] M.E.D.G. Azenha, H.D. Burrows, M. Canle, L., R. Coimbra, M.I. Fernández, M. V
[173] T. Olmez-Hanci, I. Arslan-Alaton, D. Dursun, B. Genc, D.G. Mita, M. Guida, L. Mita, García, M.A. Peiteado, J.A. Santaballa, Kinetic and mechanistic aspects of the
Degradation and toxicity assessment of the nonionic surfactant TritonTM X-45 by direct photodegradation of atrazine, atraton, ametryn and 2-hydroxyatrazine by
the peroxymonosulfate/UV-C process, Photochem. Photobiol. Sci. 14 (2015) 254 nm light in aqueous solution, J. Phys. Org. Chem. 16 (2003) 498–503.
569–575. [201] B. Neppolian, E. Celik, H. Choi, Photochemical Oxidation of Arsenic(III) to Arsenic
[174] H. Hori, A. Yamamoto, E. Hayakawa, S. Taniyasu, N. Yamashita, S. Kutsuna, (V) using Peroxydisulfate Ions as an Oxidizing Agent, Environ. Sci. Technol. 42
H. Kiatagawa, R. Arakawa, Efficient decomposition of environmentally persistent (2008) 6179–6184.
perfluorocarboxylic acids by use of persulfate as a photochemical oxidant, [202] Z.H. Diao, X.R. Xu, H. Chen, D. Jiang, Y.X. Yang, L.J. Kong, Y.X. Sun, Y.X. Hu,
Environ. Sci. Technol. 39 (2005) 2383–2388. Q.W. Hao, L. Liu, Simultaneous removal of Cr(VI) and phenol by persulfate acti-
[175] Y.C. Lee, S.L. Lo, P. Te Chiueh, Y.H. Liou, M.L. Chen, Microwave-hydrothermal vated with bentonite-supported nanoscale zero-valent iron: reactivity and me-
decomposition of perfluorooctanoic acid in water by iron-activated persulfate chanism, J. Hazard. Mater. 316 (2016) 186–193.
oxidation, Water Res. 44 (2010) 886–892. [203] Z.H. Diao, X.R. Xu, D. Jiang, L.J. Kong, Y.X. Sun, Y.X. Hu, Q.W. Hao, H. Chen,
[176] Y. Lee, S. Lo, J. Kuo, C. Hsieh, Decomposition of perfluorooctanoic acid by mi- Bentonite-supported nanoscale zero-valent iron/persulfate system for the si-
crowaveactivated persulfate: effects of temperature, pH, and chloride ions, Front. multaneous removal of Cr(VI) and phenol from aqueous solutions, Chem. Eng. J.
Environ. Sci. Eng. 6 (2011) 17–25. 302 (2016) 213–222.
[177] Y.C. Lee, S.L. Lo, J. Kuo, Y.L. Lin, Persulfate oxidation of perfluorooctanoic acid [204] J. Gerritse, V. Renard, J. Visser, J.C. Gottschal, Complete degradation of tetra-
under the temperatures of 20–40°C, Chem. Eng. J. 198–199 (2012) 27–32. chloroethene by combining anaerobic dechlorinating and aerobic methanotrophic
[178] Y.C. Lee, S.L. Lo, J. Kuo, C.P. Huang, Promoted degradation of perfluorooctanic enrichment cultures, Appl. Microbiol. Biotechnol. 43 (1995) 920–928.
acid by persulfate when adding activated carbon, J. Hazard. Mater. 261 (2013) [205] A. Hirvonen, T. Tuhkanen, P. Kalliokoski, Treatment of TCE- and PCE-con-
463–469. taminated groundwater using UV/H2O2 and O3/H2O2 oxidation processes, Water
[179] C.S. Liu, C.P. Higgins, F. Wang, K. Shih, Effect of temperature on oxidative Sci. Technol. 33 (1996) 67–73.
transformation of perfluorooctanoic acid (PFOA) by persulfate activation in water, [206] N. Yan, F. Liu, Q. Xue, M.L. Brusseau, Y. Liu, J. Wang, Degradation of tri-
Sep. Purif. Technol. 91 (2012) 46–51. chloroethene by siderite-catalyzed hydrogen peroxide and persulfate: investiga-
[180] Y. Qian, X. Guo, Y. Zhang, Y. Peng, P. Sun, C.-H. Huang, J. Niu, X. Zhou, tion of reaction mechanisms and degradation products, Chem. Eng. J. 274 (2015)
J.C. Crittenden, Perfluorooctanoic acid degradation using UV-persulfate process: 61–68.
modeling of the degradation and chlorate formation, Environ. Sci. Technol. 50 [207] M. Xu, H. Du, X. Gu, S. Lu, Z. Qiu, Q. Sui, Generation and intensity of active
(2016) 772–781. oxygen species in thermally activated persulfate systems for the degradation of
[181] Y. Wang, C.S. Hong, Effect of hydrogen peroxide, periodate and persulfate on trichloroethylene, RSC Adv. 4 (2014) 40511–40517.
photocatalysis of 2-chlorobiphenyl in aqueous TiO2 suspensions, Water Res. 33 [208] B.M. Sharma, G.K. Bharat, S. Tayal, L. Nizzetto, P. Cupr, T. Larssen, Environment
(1999) 2031–2036. and human exposure to persistent organic pollutants (POPs) in India: a systematic
[182] J. Sharma, I.M. Mishra, D.D. Dionysiou, V. Kumar, Oxidative removal of bisphenol review of recent and historical data, Environ. Int. 66 (2014) 48–64.
A by UV-C/peroxymonosulfate (PMS): kinetics, influence of co-existing chemicals [209] W. Qin, G. Fang, Y. Wang, T. Wu, C. Zhu, D. Zhou, Efficient transformation of DDT
and degradation pathway, Chem. Eng. J. 276 (2015) 193–204. by peroxymonosulfate activated with cobalt in aqueous systems: kinetics, pro-
[183] M.K. Sahoo, B. Sinha, M. Marbaniang, D.B. Naik, R.N. Sharan, Mineralization of ducts, and reactive species identification, Chemosphere 148 (2016) 68–76.
Calcon by UV/oxidant systems and assessment of biotoxicity of the treated solu- [210] C. Zhu, G. Fang, D.D. Dionysiou, C. Liu, J. Gao, W. Qin, D. Zhou, Efficient trans-
tions by E. coli colony forming unit assay, Chem. Eng. J. 181 (2012) 206–214. formation of DDTs with persulfate activation by zero-valent iron nanoparticles: a
[184] S. Wacławek, V. Antoš, P. Hrabák, M. Černík, Remediation of hexa- mechanistic study, J. Hazard. Mater. 316 (2016) 232–241.
chlorocyclohexanes by cobalt-mediated activation of peroxymonosulfate, Desalin. [211] A. Tauber, C. von Sonntag, products and kinetics of the OH-radical-induced
Water Treat. 57 (2016) 26274–26279. dealkylation of atrazine, Acta Hydrochim. Hydrobiol. 28 (2000) 15–23.
[185] M.K. Sahoo, M. Marbaniang, B. Sinha, D.B. Naik, R.N. Sharan, UVC induced TOC [212] J.L. Acero, K. Stemmler, U. Von Gunten, Degradation kinetics of atrazine and its
removal studies of Ponceau S in the presence of oxidants: evaluation of electrical degradation products with ozone and OH radicals: a predictive tool for drinking
energy efficiency and assessment of biotoxicity of the treated solutions by water treatment, Environ. Sci. Technol. 34 (2000) 591–597.
Escherichia coli colony forming unit assay, Chem. Eng. J. 213 (2012) 142–149. [213] M.J. Shipitalo, L.B. Owens, Atrazine, deethylatrazine, and deisopropylatrazine in
[186] K. Selvam, M. Muruganandham, I. Muthuvel, M. Swaminathan, The influence of surface runoff from conservation tilled watersheds, Environ. Sci. Technol. 37
inorganic oxidants and metal ions on semiconductor sensitized photodegradation (2003) 944–950.
of 4-fluorophenol, Chem. Eng. J. 128 (2007) 51–57. [214] J. De Laat, N. Chramosta, M. Doré, H. Suty, M. Pouillot, Constantes cinétiques de
[187] L. Ravichandran, K. Selvam, M. Swaminathan, Effect of oxidants and metal ions on réaction des radicaux hydroxyles sur quelques sous- produits d’oxydation de l’a-
photodefluoridation of pentafluorobenzoic acid with ZnO, Sep. Purif. Technol. 56 trazine par O3 ou par O3/H2O2, Environ. Technol. 15 (1994) 419–428.
(2007) 192–198. [215] S. Khan, X. He, J.A. Khan, H.M. Khan, D.L. Boccelli, D.D. Dionysiou, Kinetics and
[188] C.H. Yu, C.H. Wu, T.H. Ho, P.K. Andy, Hong, Decolorization of C.I. Reactive Black mechanism of sulfate radical- and hydroxyl radical-induced degradation of highly
5 in UV/TiO2, UV/oxidant and UV/TiO2/oxidant systems: a comparative study, chlorinated pesticide lindane in UV/peroxymonosulfate system, Chem. Eng. J. 318
Chem. Eng. J. 158 (2010) 578–583. (2016) 135–142.

61
S. Wacławek et al. Chemical Engineering Journal 330 (2017) 44–62

[216] K. Kuśmierek, L. Dąbek, A. Świątkowski, A comparative study on oxidative de- [231] G.W. Chen, W.W. Lin, D.J. Lee, Capillary suction time (CST) as a measure of sludge
gradation of 2,4-dichlorophenol and 2,4-dichlorophenoxyacetic acid by ammo- dewaterability, Water Sci. Technol. 34 (1996) 443–448.
nium persulfate, Desalin. Water Treat. 57 (2016) 1098–1106. [232] G. Zhen, X. Lu, Y. Zhao, X. Chai, D. Niu, Enhanced dewaterability of sewage sludge
[217] J.M. Monteagudo, A. Durán, R. González, A.J. Expósito, In situ chemical oxidation in the presence of Fe(II)-activated persulfate oxidation, Bioresour. Technol. 116
of carbamazepine solutions using persulfate simultaneously activated by heat (2012) 259–265.
energy, UV light, Fe2+ ions, and H2O2, Appl. Catal. B Environ. 176–177 (2015) [233] Y. Shi, J. Yang, W. Yu, S. Zhang, S. Liang, J. Song, Q. Xu, N. Ye, S. He, C. Yang,
120–129. J. Hu, Synergetic conditioning of sewage sludge via Fe2+/persulfate and skeleton
[218] X. Liu, X. Zhang, K. Shao, C. Lin, C. Li, F. Ge, Y. Dong, Fe0-activated persulfate- builder: effect on sludge characteristics and dewaterability, Chem. Eng. J. 270
assisted mechanochemical destruction of expired compound sulfamethoxazole (2015) 572–581.
tablets, RSC Adv. 6 (2016) 20938–20948. [234] W.S. Chen, Y.C. Jhou, C.P. Huang, Mineralization of dinitrotoluenes in industrial
[219] Z. Li, Q. Yang, Y. Zhong, X. Li, L. Zhou, X. Li, G. Zeng, Granular activated carbon wastewater by electro-activated persulfate oxidation, Chem. Eng. J. 252 (2014)
supported iron as a heterogeneous persulfate catalyst for the pretreatment of 166–172.
mature landfill leachate, RSC Adv. 6 (2016) 987–994. [235] G.Y. Zhen, X.Q. Lu, Y.Y. Li, Y.C. Zhao, Innovative combination of electrolysis and
[220] J. Kronholm, M.L. Riekkola, Potassium persulfate as oxidant in pressurized hot Fe(II)-activated persulfate oxidation for improving the dewaterability of waste
water, Environ. Sci. Technol. 33 (1999) 2095–2099. activated sludge, Bioresour. Technol. 136 (2013) 664–1663.
[221] S.A.A. Nakhli, Biological removal of phenol from saline wastewater using a [236] S. Wacławek, K. Grübel, Z. Chład, M. Dudziak, Impact of peroxydisulphate on
moving bed biofilm reactor containing acclimated mixed consortia, Springerplus. disintegration and sedimentation properties of municipal wastewater activated
3 (2014). sludge, Chem. Pap. 69 (2015) 1473–1480.
[222] A. Rubalcaba, M.E. Suárez-Ojeda, F. Stüber, A. Fortuny, C. Bengoa, I. Metcalfe, [237] Y. Yang, K. Tsukahara, R. Yang, Z. Zhang, S. Sawayama, Enhancement on biode-
J. Font, J. Carrera, A. Fabregat, Phenol wastewater remediation: advanced oxi- gradation and anaerobic digestion efficiency of activated sludge using a dual ir-
dation processes coupled to a biological treatment, Water Sci. Technol. 55 (2007) radiation process, Bioresour. Technol. 102 (2011) 10767–10771.
221–227. [238] S. Wacławek, K. Grübel, M. Černík, The impact of peroxydisulphate and perox-
[223] A.A. Babaei, F. Ghanbari, COD removal from petrochemical wastewater by UV/ ymonosulphate on disintegration and settleability of activated sludge, Environ.
hydrogen peroxide, UV/persulfate and UV/percarbonate: biodegradability im- Technol. 37 (2016) 1296–1304.
provement and cost evaluation, J. Water Reuse Desalin. 7 (2016). [239] S. Wacławek, K. Grübel, Z. Chłąd, M. Dudziak, M. Černík, The impact of oxone on
[224] E. Kattel, N. Dulova, M. Viisimaa, T. Tenno, M. Trapido, Treatment of high- disintegration and dewaterability of waste activated sludge, Water Environ. Res.
strength wastewater by Fe2+-activated persulphate and hydrogen peroxide, 88 (2016) 152–157.
Environ. Technol. 37 (2016) 352–359. [240] T. Niu, Z. Zhou, W. Ren, L.-M. Jiang, B. Li, H. Wei, J. Li, L. Wang, Effects of
[225] S.O. Ganiyu, E.D. van Hullebusch, M. Cretin, G. Esposito, M.A. Oturan, Coupling of potassium peroxymonosulfate on disintegration of waste sludge and properties of
membrane filtration and advanced oxidation processes for removal of pharma- extracellular polymeric substances, Int. Biodeterior. Biodegradation. 106 (2016)
ceutical residues: a critical review, Sep. Purif. Technol. 156 (2015) 891–914. 170–177.
[226] M.A. Fagier, E.A. Ali, K.S. Tay, M.R.B. Abas, Mineralization of organic matter from [241] C. Liu, B. Wu, X. Chen, S. Xie, Waste activated sludge pretreatment by Fe(II)-
vinasse using physicochemical treatment coupled with Fe2+-activated persulfate activated peroxymonosulfate oxidation under mild temperature, Chem. Pap. 1–9
and peroxymonosulfate oxidation, Int. J. Environ. Sci. Technol. 13 (2016) (2017).
1189–1194. [242] G. Zhen, X. Lu, J. Niu, L. Su, X. Chai, Y. Zhao, Y.Y. Li, Y. Song, D. Niu, Inhibitory
[227] J. Rodríguez-Chueca, C. Amor, T. Silva, D.D. Dionysiou, G. Li Puma, M.S. Lucas, effects of a shock load of Fe(II)-mediated persulfate oxidation on waste activated
J.A. Peres, Treatment of winery wastewater by sulphate radicals: HSO5−/transi- sludge anaerobic digestion, Chem. Eng. J. 233 (2013) 274–281.
tion metal/UV-A LEDs, Chem. Eng. J. 310 (2017) 473–483. [243] D. Sun, H. Liang, C. Ma, Enhancement of sewage sludge anaerobic digestibility by
[228] R. Ocampo-Pérez, M. Sánchez-Polo, J. Rivera-Utrilla, R. Leyva-Ramos, sulfate radical pretreatment, Adv. Mater. Res. (Durnten-Zurich, Switz.) 518–523
Degradation of antineoplastic cytarabine in aqueous phase by advanced oxidation (2012) 3358–3362.
processes based on ultraviolet radiation, Chem. Eng. J. 165 (2010) 581–588. [244] S. Wacławek, K. Grübel, P. Dennis, V.T.P. Vinod, M. Černik, A novel approach for
[229] H.-Y. Shu, M.-C. Chang, S.-W. Huang, Decolorization and mineralization of azo simultaneous improvement of dewaterability, post-digestion liquor properties and
dye Acid Blue 113 by the UV/Oxone process and optimization of operating toluene removal from anaerobically digested sludge, Chem. Eng. J. 291 (2016)
parameters, Desalin. Water Treat. 57 (2016) 7951–7962. 192–198.
[230] G. Zhen, X. Lu, B. Wang, Y. Zhao, X. Chai, D. Niu, A. Zhao, Y. Li, Y. Song, X. Cao, [245] K. Song, X. Zhou, Y. Liu, Y. Gong, B. Zhou, D. Wang, Q. Wang, Role of oxidants in
Synergetic pretreatment of waste activated sludge by Fe(II)-activated persulfate enhancing dewaterability of anaerobically digested sludge through Fe (II) acti-
oxidation under mild temperature for enhanced dewaterability, Bioresour. vated oxidation processes: hydrogen peroxide versus persulfate, Sci. Rep. 6 (2016)
Technol. 124 (2012) 29–36. 24800.

62

You might also like