You are on page 1of 26

journal of environmental sciences 147 (2025) 688–713

Available online at www.sciencedirect.com

www.elsevier.com/locate/jes

A review on reactive oxygen species-induced


mechanism pathways of pharmaceutical waste
degradation: Acetaminophen as a drug waste
model

Saba Humayun 1,2, Maan Hayyan 3,∗, Yatimah Alias 1,2,∗


1 Department of Chemistry, Faculty of Science, University of Malaya, Kuala Lumpur 50603, Malaysia
2 Universityof Malaya Centre for Ionic Liquids, University of Malaya, Kuala Lumpur 50603, Malaysia
3 Chemical Engineering Program, Faculty of Engineering and Technology, Muscat University, Muscat P.C.130, Oman

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

Article history: Innately designed to induce physiological changes, pharmaceuticals are foreknowingly haz-
Received 4 August 2023 ardous to the ecosystem. Advanced oxidation processes (AOPs) are recognized as a set of
Revised 22 November 2023 contemporary and highly efficient methods being used as a contrivance for the removal of
Accepted 22 November 2023 pharmaceutical residues. Since reactive oxygen species (ROS) are formed in these processes
Available online 4 December 2023 to interact and contribute directly toward the oxidation of target contaminant(s), a profound
insight regarding the mechanisms of ROS leading to the degradation of pharmaceuticals
Keywords: is fundamentally significant. The conceptualization of some specific reaction mechanisms
Wastewater treatment allows the design of an effective and safe degradation process that can empirically reduce
Ionic liquid the environmental impact of the micropollutants. This review mainly deliberates the mech-
Hydroxyl radical anistic reaction pathways for ROS-mediated degradation of pharmaceuticals often leading
Superoxide to complete mineralization, with a focus on acetaminophen as a drug waste model.
Pharmaceutical contaminant © 2024 The Research Center for Eco-Environmental Sciences, Chinese Academy of
Advanced oxidation process Sciences. Published by Elsevier B.V.
Deep eutectic solvent
Hydrogen peroxide

nential growth in the last two decades. According to an es-


Introduction timation (Council, 2018), in the United States approximately
one-third of the number of prescription items (around four
The degree of upsurge in the manufacture and diversifi- billion) turn into annual waste. The evolution and advance-
cation of pharmaceuticals outpaced one of the most early ment of the pharmaceutical industry and modern medicine
and eminent drivers of global environmental shifts, such have ensued undisputed benefits in enhancing the well-being
as the rise in the concentration of atmospheric CO2 , loss of the masses by increasing the health span and providing
of biodiversity, nutrient contamination, habitat damage, etc. remedies to many fatal ailments.
(Bernhardt et al., 2017). Pharmaceuticals in the environment A global threat to the ecosystem and human health is
as a transdisciplinary field of research have manifested expo- posed by pollution caused by pharmaceutical waste. In the


Corresponding authors.
E-mails: maan_hayyan@yahoo.com (M. Hayyan), yatimah70@um.edu.my (Y. Alias).

https://doi.org/10.1016/j.jes.2023.11.021
1001-0742/© 2024 The Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences. Published by Elsevier B.V.
journal of environmental sciences 147 (2025) 688–713 689

1990s, the detection of estrogen in sewage discharge was quired from 1052 locations in 104 countries appearing from all
found to be a cause of alteration of sex functions of a continents, analyzing 61 different APIs. This elaborate work
species of fish, which prompted curiosity in apprehending reported that of the four APIs, acetaminophen/paracetamol
how drug substances make their way to the environment was the most frequently encountered, and the only over-the-
(Larsson, 2014). It was reported in a study that about 85 per- counter analgesic that has been detected across all conti-
cent of male smallmouth bass collected from North-eastern nents (the other three being stimulant or lifestyle compounds
U.S. had eggs growing in male testes, suggesting some sort were caffeine, nicotine, and cotinine). Moreover, 14 APIs
of hormonal disarray (Iwanowicz et al., 2016). This led to the were also additionally detected as contaminants in all con-
evaluation of the latent threats posed to the quality of water tinents except Antarctica. These compounds along with their
from endocrine-disrupting compounds (EDCs) (Bhandari et al., pharmaceutical class and chemical structure are enlisted in
2015; Blazer et al., 2007), as contemporary research specifies Appendix A Table S1.
that exposure to pharmaceuticals through drinking water can Oxidation processes taking place in ecological compart-
also affect the human reproductive system, particularly in ments are one of the many natural ways to revive nutrients
males (Schreiber et al., 2011; Westerling, 2013). and purify water by removing contaminants. Among other
A potential environmental hazard has been posed by the treatments, the oxidation processes in engineered systems
pharmaceutical waste originating from an estimated 10% are used as well to remove chemical pollutants and for wa-
of such products according to the German Environment ter purification (Connors et al., 1986; Johnson, 1988). Oxidation
Agency (UBA) (Küster and Adler, 2014). Since a distinct in- reactions signify an important degradation pathway for phar-
crease in the apprehension regarding the repercussions aris- maceuticals, and organic substances in general. Advanced ox-
ing from pharmaceuticals polluting our environment is evi- idation processes (AOPs) are currently celebrated to assert an
dent (Donnachie et al., 2016), it is also found that an enor- ever-increasing implementation for green remediation of wa-
mous number of pharmaceuticals have not been gauged for ter polluted with recalcitrant compounds, which are not easy
their occurrence, toxicity, or fate in the environment, there- to remove using conventional techniques. AOPs are surmised
fore making it even more challenging to generalize the harm as viable technologies contributing to the degradation of phar-
they might contribute to. Many researchers have put emphasis maceutical contaminants through the in-situ formation of re-
on the dearth of knowledge concerning the effects of low-level active oxygen species (ROS) acting as strong oxidants (He et al.,
exposure of pharmaceutical mixtures in the environment. 2011; Yang et al., 2013). ROS concern free radicals, e.g., hy-
Pharmaceuticals that target the endocrine system could droxyl radical (● OH), hydroperoxyl radical (HO2 ● ), superoxide
epitomize the excessive environmental influence and risk as radical anion (O2 ●− ), sulfate radical (SO4 ●− ), as well as non-
demonstrated in a recent investigation (Gunnarsson et al., radical species (Wang et al., 2011), for e.g., singlet oxygen (1 O2 )
2019). The release of hormones such as synthetic oestro- and hydrogen peroxide (H2 O2 ).
gen ethinylestradiol (EE2) and progestin levonorgestrel (LNG), The degradation of pharmaceutical waste by ROS can
might also have inadvertent adverse effects on humans by involve a complex interplay of several mechanisms, such as
way of exposure through drinking water as well as food. oxidation, reduction, radical formation, catalytic or enzymatic
Oestrogens are known to induce increased risk of breast degradation, and photochemical reactions (Ao et al., 2023;
cancer in women (Moore et al., 2016) and prostate cancer Asif et al., 2023; Chen et al., 2022, 2023; Constantin et al., 2022;
in men (Nelles et al., 2011). Pregnant women, fetuses and Malefane et al., 2023; Flores et al., 2023; Gonzaga et al., 2023;
neonates are exceptionally threatened by a continuous sub- Honarmandrad et al., 2023; Jiang et al., 2021; Jin et al., 2023;
jection to endocrine disruptors even if the concentrations are Mohapatra et al., 2023; Papac et al., 2023; Pérez-Lucas et al.,
low. Several potential diseases associated to endocrine dis- 2023; Piccirillo et al., 2021; Qi et al., 2022; Rapti et al., 2023;
rupting chemicals (EDCs) include those correlated to such Sarker and Ahn, 2023; Scaria and Nidheesh, 2023; Singh et al.,
systems as reproductive and endocrine (infertility, early pu- 2023; Usman et al., 2023; Valadez-Renteria et al., 2023;
berty, diabetes, breast or prostate cancer), immune and au- Wang et al., 2023; Xie et al., 2023; You et al., 2023; Yu et al.,
toimmune, cardiopulmonary (heart disease or asthma), and 2023; Zare et al., 2023; Zilberman et al., 2023). The specific
nervous systems (Parkinson’s disease, Alzheimer’s disease, mechanism involved depends on the type of ROS used or
and ADHD) (Bergmanet al., 2013). Moreover, antimicrobial re- generated as well as on the chemical structure of the pharma-
sistance (AMR) is well-recognized as a phenomenon that is ceutical being degraded. The underlying significance of these
posing severe risks to global health and livelihoods. Freshwa- reaction mechanisms by ROS is indicated as these critically
ter ecosystems turn out to be an ideal site for the acquire- determine the efficiency and selectivity of the degradation
ment and outspread of AMR owing to the continual pollution process. By explicitly understanding the comprising mech-
caused by antimicrobial and antibacterial compounds ema- anisms pertaining to the degradation process, it is possible
nating from anthropogenic activities. to optimize the conditions of the reaction, particularly to
Acetaminophen (ACT) has been selected as a drug waste attain the preferred level of degradation while minimizing
model for this study because it is one of the most acclaimed undesirable side reactions to eventually increase the efficacy
APIs known for its ubiquitousness as a pharmaceutical con- and selectivity of the procedures.
taminant. Published in 2022, a global study on API pollu- While process efficiency and procedure selectivity are sig-
tion in 258 of the world’s rivers represented the environmen- nificant considerations in designing a chemical process to de-
tal influence of 471.4 million people across 137 geographic grade contaminants, the development and optimization of a
regions (Wilkinson et al., 2022), in which samples were ac- specific method is required. In that context as an example,
690 journal of environmental sciences 147 (2025) 688–713

some approaches and strategies that can be effectively imple- that could avoid the formation of harmful intermediates or
mented to help attain these goals are specified as follows: by-products leading to safety hazards, hence evading any
risk to the ecosystem and concerns to human health dur-
(a) Suitable catalysts or promoters that enhance the de- ing or after the process.
sired degradation reactions and suppress undesired
pathways can be selected. Moreover, catalyst loading This perspective reiterates the state-of-the-art ROS-
and compatibility with the reactants and reaction con- mediated reaction mechanisms implicated in the degradation
ditions should also be optimized. of pharmaceutical waste which seems pivotal while designing
(b) Understanding the kinetics of the reactions involved in a degradation strategy to remove a specific target pollutant
the degradation process is useful. This includes distin- because it largely enables the optimization of selectivity,
guishing the rate equations, identifying reaction mech- efficiency, by-product formation, and safety. Hence an insight
anisms, and the influence of temperature/pressure. into these mechanistic routes makes it possible to scheme an
(c) Selectivity for the desired reaction pathways increases effective and safe degradation process that can precisely re-
by choosing optimum reaction conditions and reagents, duce the environmental impact of pollutants. While primarily
such as adjusting pH, using specific solvents/media, or emphasizing the implication of understanding such reac-
altering the redox potential. Optimization of the reac- tions, we draw the inference by outlining valuable frontiers
tion conditions (pH, catalyst and reactant concentra- for imminent research and opportunities to make the most
tion, temperature, etc.) to favor or control the desired from these underlying chemical insights for pharmaceutical
reaction pathway can aid in terminating the side reac- removal processes being developed in the present day.
tions. Furthermore, selectivity can be achieved by em-
ploying reaction engineering techniques like staged re-
actions, or multi-step processes, etc.
(d) Utilizing effective separation techniques to isolate or re-
1. ROS implicated in pharmaceutical
cycle the catalysts, medium or unreacted starting mate-
degradation
rials or products can also minimize waste and improve
Once generated, the ROS reacts rapidly, most often close to
process efficiency.
diffusion-controlled rates (Buxton et al., 1988). Some more fre-
quently concerned ROS for reaction mechanisms with phar-
The implication of various types of ROS-mediated mecha-
maceutical waste include ● OH, O3 , O2 ●− , SO4 ●− , 1 O2, and some
nisms in the degradation of pharmaceutical waste is deemed
organic radicals.
imperative for several reasons, which essentially revolve
around aiming the following assistive gains:
1.1. Hydroxyl radicals (● OH)
• Optimization of the degradation process: Discernment of
different types of mechanisms by which ROS can degrade Hydroxyl radicals (● OH) as a frequently detected ROS engaged
pharmaceutical waste allows optimizing the degradation in oxidative degradation reactions, are one of the strongest
process in a way so as to achieve maximum efficiency and oxidizers since these possess a very high electronic poten-
selectivity. Particularly, this understanding can help tailor tial. ● OH have a very short half-life and are hence highly re-
the degradation process subject to the kind of ROS used active and very non-selective species that are able to interact
and on the specific type of waste. directly with dissolved solids, including most organic and in-
• Development of new ROS-based treatment systems: Re- organic pollutants. Reactions with ● OH in water can be initi-
viewing different types of mechanisms pertaining to ROS- ated by means of certain types of substances known as acti-
based degradation of pharmaceutical waste can also pro- vators. This ● OH activation is in fact a highly intricate process
vide insight into the development of new treatments. For occurring through various mechanisms.
example, a new ROS-based treatment could be designed Selected pharmaceutical contaminants in drinking water
that specifically targets a particular mechanism, therefore and their second-order rate constants for oxidation via ● OH
leading to a more efficient degradation of the waste. are indicated in Table 1. It is evident from the magnitude of
• Prediction of transformation products and by-products: these rate constants that the oxidation of many pharmaceuti-
The different types of mechanisms involved in ROS- cal pollutants with ● OH is quite rapid. Moreover, the reactivity
mediated degradation of pharmaceutical waste can also of ● OH is also very high with the background constituents in-
impact the types and amounts of by-products that would herently present in natural waters, which results in short life-
be more likely generated during the degradation process. times of ● OH (in the order of microseconds). Therefore, in an
The expected transformation products, their nature and AOP system, the ● OH steady-state concentration during treat-
toxicity, as well as the extent of mineralization, are con- ment of natural waters characteristically remains very low i.e.,
ceivably more predictable after comprehensively study- 10−12 –10−15 mol/L (Elovitz et al., 2008), causing the species to
ing the mechanisms. This would assist in optimizing the not increase over time due to accumulation.
degradation process so as to minimize the generation of Based on mechanisms, the reactions with radicals are in
more by-products with higher toxicity. general classified into (a) addition reaction, (b) hydrogen ab-
• Safety considerations: The in-depth analysis of ROS-based straction, (c) electron transfer, with addition reaction as the
process mechanisms can also aid in optimizing the process most preferred pathway, if possible.
journal of environmental sciences 147 (2025) 688–713 691

Table 1 – Kinetics of the ● OH oxidation with selected pharmaceutical contaminants present in water.

Pharmaceuticals Reaction rate constant (k) Reference


(mol/(L . sec))

Carbamazepine 8.8 × 109 (Huber et al., 2003)


Diclofenac 7.5 × 109 (Huber et al., 2003)
Sulfamethoxazole 5.5 × 109 (Huber et al., 2003)
17α-Ehinylestradiol 9.8 × 109 (Huber et al., 2003)
Bezafibrate 7.4 × 109 (Huber et al., 2005)
Clofibric acid 4.7 × 109 (Huber et al., 2005)
Diazepam 7.2 × 109 (Huber et al., 2005)
Ibuprofen 7.4 × 109 (Huber et al., 2005)
Iopromide 3.3 × 109 (Huber et al., 2005)
Naproxen 9.6 × 109 (Huber et al., 2005)
Tetracycline 6.3 × 109 (Jeong et al., 2010)
Chlortetracycline 5.2 × 109 (Jeong et al., 2010)
Atorvastatin (lipitor) 1.2 × 1010 (Razavi et al., 2011)

1.1.1. Addition reactions according to Eq. (7). Therefore, a peroxyl radical being part
The ● OH readily undergoes reaction with carbon-carbon of several unimolecular reactions appears to be in a much
(Eq. (1)), carbon-nitrogen, and sulfur-oxygen double bonds (ex- more complex chemistry. These reactions include the elim-
cept SO4 2− , Eq. (2)), including reactions with the transition ination reaction of HO2 ● /O2 ●− couple via double bond for-
metal ions (Eq. (3)) via addition reaction, resulting in the for- mation, as well as carbon-carbon double bond addition or
mation of an adduct followed by its subsequent decomposi- oxygen-transfer reactions (Sonntag, 2006b).
tion. The electrophilic character of carbon atoms causes their
electron-rich positions to get preferably attacked. For instance, HRH + ● OH → HR● + H2 O (4)
C5 is more prone to the attack than C6 , when these two car-
bons are in a double bond (Sonntag, 2006), because of the more HR● + O2 → HROO● (5)
electron-rich zone localized at position C5 .
HROO● + HRH → ROOH + HR● (6)

(not at normal wastewater conditions)

HROO● → HO2 ● + R (7)


(1)
It is evident from the aforementioned reactions that the oc-
R2 S=O + ● OH → R2 S(O● )OH → RS(O)OH + R● (2) currence of oxygen has a considerable influence on radical re-
action systems. In the absence or insufficiency of O2 , the radi-
Tl2+ + ● OH → HOTl2+ → Tl2+ + H2 O (3) cal combination reaction (Eq. (8)) is prevalent and there is mi-
nor significance of the reaction represented by (Eq. (5)), specifi-
1.1.2. Hydrogen abstraction reactions cally in comparison with other reactions. Alternatively, a large
Since the HO–H bond has a higher dissociation energy than quantity of molecular O2 increases the relative fraction of the
that of a carbon-bound hydrogen (C–H) bond, removing a H- bimolecular recombination (von Sonntag and Schuchmann,
atom from an organic compound (HRH) is more likely (Eq. (4)), 1991). The reaction of most carbon-centered radicals with O2
hence resulting in the production of carbon-centered radical (Eq. (5)) is fast and irreversible, whereas other organic com-
(● R) (Sonntag et al., 1997). It is suggested by many authors that pounds such as thiyl and hexandienyl radicals merely react
the interaction of this carbon-centered radical with molecu- reversibly with molecular oxygen, while the hetero-centered
lar oxygen (Eq. (5)) initiates a chain reaction that generates radicals like phenoxyl do not tend to react with oxygen at all
peroxyl radicals. This peroxyl radical then successively re- (Sonntag, 2006).
acts with other target organics (Eq. (6)) ideally leading to a The existence of scavengers or inhibitors determines the
final conversion into CO2 , water, and inorganic salts. On the rate of reaction cycles. CO3 2− in several of its dissociated con-
other hand, as hydrogen abstraction is typically utilized to figurations is one of the familiar scavengers; CO3 ●− formed
describe the oxidation of target substances, the recognition by reactions of CO3 2− and HCO3 − are generally recognized as
of peroxyl radicals as weak oxidants (low reduction poten- poor reactants (see Section 2.2.2).
tial) is also noteworthy. In the case of strong electron-donor
species, these peroxyl radicals might also perform as one- 1.1.3. Electron transfer reactions
electron oxidants. Consequently, at standard AOP conditions Electron transfer reactions tend to compete with the addi-
during water treatment, these radicals instead of undergo- tion reaction; despite the electron transfer reaction being
ing an auto-oxidation chain reaction prefer the elimination thermodynamically favorable it has been observed less of-
of O2 ●− (Sonntag, 2006; von Sonntag and Schuchmann, 1991), ten, while addition reaction pathways are generally the pre-
692 journal of environmental sciences 147 (2025) 688–713

ferred ones (Sonntag, 2006). However, the direct electron trans-


fers are favored in such reactions as those including halo-
genated phenolate ions, since ortho- and para-positions for
the favored ● OH addition are obstructed by the substituents,
in which case electron transfer could turn out to be domi-
nant (Sonntag, 2006). Furthermore, the electron transfer yield
contributing to the overall reaction is found to increase with
the involved halogen size and predominates for bromine
and iodine with 73% and 97%, respectively (Akhlaq and
Clemens, 1987). Likewise, reactions with thiol and thiolate also
preferentially undergo electron transfers.

1.1.4. Radical recombination reactions


Radical recombination reactions terminate the chain reac-
tions through the production of a combined molecule, while
larger molecules tend to decompose according to a reaction
represented in Eq. (8), as an instance.

HROO● + HROO● → ROH + ROH + O2 (8)

1.1.5. Competition kinetics


The ● OH however, not only reacts with target contami-
nants, but also with water matrix constituents (e.g., NOM,
CO3 2− /HCO3 − , and inorganic species) and other radicals (e.g.,
HO2 ● ) collectively which act as “radical scavengers”, hence
initiating a complex flow of oxidative reactions. The impact
Scheme 1 – Pathways for various interactions of ozone with
of radical scavengers present in the water matrices is there-
organic molecules (von Gunten, 2003). (a) Formation of
fore deemed as one of the major complications encountered
ozonide (A) in dipolar cycloaddition reactions. (b)
in the application of AOPs, because there is much likelihood
Disintegration of ozonide. (c) Reaction between phenol and
of scavengers instigating an inhibition of the reaction leading
O3 .
to degradation of the target pollutant. The predominant in-
fluence of scavengers depends on all the involved moieties in
compounds existent in the type of water under investigation.
Carbonate in its various dissociation forms is the key scav-
formed during ozonation were persistent (Almomani et al.,
enger, typically prevalent in natural and wastewaters. Other
2016). This outcome also points out the need to estimate the
commonly occurring scavengers include natural organic mat-
toxicity following an ozonation process. Complete abatement
ters (NOMs) also referred to as humic substances, while amine
of investigated pharmaceuticals has been achieved using low
and phenol are among the more frequent moieties arising
doses of O3 but with incomplete mineralization rates.
from NOM (Buffle et al., 2006). Appendix A Table S2 lists the re-
During the O3 oxidation process, degradation by means of
action rate constants of some typical chemical species found
both the direct O3 molecule and indirect ● OH reactions can
in water that ● OH can react with.
occur effectively. One of these routes will predominate based
Factors that majorly influence the system of radical reac-
on several factors, for example, pH, temperature, and chemi-
tions involve the inhibiting (scavenging) or promoting species,
cal composition of the investigated water medium. O3 can ox-
which in turn depends on the composition of water or
idize organic compounds selectively and partially. In the case
wastewater (De Laat and Le, 2005). In addition to the reactions,
of direct reactions by O3 , the mechanisms of oxidation by O3
it is also required to take into consideration the fluctuations
and ozonide (O3 ●− ) are determined. It is generally known to
in pH during oxidation along with the pH-dependent disso-
undergo three categories of reactions in solution form con-
ciation equilibriums. Likewise, in the case of heterogeneous
taining organic contaminants in which O3 either behaves as
catalysis for instance, the pH value affects the surface charge
an electrophilic or nucleophilic agent or as a 1,3-dipole.
and consequently the adsorption/desorption kinetics as well.
Apart from pH, several parameters (Pignatello et al., 2006) that
might be able to frequently inhibit the removal of target con-
taminants involve carbonate/bicarbonate, dissolved oxygen, 1.2.1. Cycloaddition (Criegee mechanism)
nitrate ion (NO3 − ), NOM, hydroxyl ion (OH− ), Fe2+ , primary and Saturated compounds are capable of undergoing a 1–3 dipo-
secondary alcohols, phosphate ion (PO3 4− ) and specific chem- lar cycloaddition with O3 molecule because of its structure
icals, such as pBA, tert–butyl alcohol, etc. which is dipolar in nature. This process leads to the formation
of a species called ‘ozonide’ (Scheme 1a) (von Gunten, 2003). In
1.2. Ozone (O3 ) and ozonide radicals (O3 ●− ) water or other protic solutions, the primary ozonide disinte-
grates into an aldehyde, a ketone, or a zwitter ion (Scheme 1b),
Various reports have documented low mineralization despite which is eventually further broken down into H2 O2 and car-
the high removal rates of pharmaceuticals, as the byproducts boxyl compounds.
journal of environmental sciences 147 (2025) 688–713 693

1.2.2. Electrophilic reactions


Electrophilic reactions essentially take place in solution forms
containing organic compounds with high electronic density
such as solutions with elevated levels of aromatic compounds
(von Gunten, 2003). Moreover, aromatic substances substi-
tuted by electron-donating groups (e.g., OH and NH2 ) possess
higher electronic density in ortho and para positions where
O3 can react more actively. For instance, hydroxyl groups in
phenol react more rapidly with O3 (Scheme 1c).

1.2.3. Nucleophilic reactions


Nucleophilic reactions generally occur in the presence of or-
ganic compounds having unavailability of electrons, in par-
ticular with compounds having electron-withdrawing groups,
e.g., carboxyl and nitro groups. However, the rate of reaction
for such groups is significantly slower.
Therefore, to summarize, the reaction mechanisms in-
Scheme 2 – Schematic representation of probable
volving direct oxidation of organic pollutants are relatively
movements for transfer of protons (green) and electrons
selective, during which the compounds containing double
(blue/red) between superoxide ion (O2 ●− /HO2 ● ) and the
(or triple) bonds, activated aromatic rings or amines are
organic compound (XH/X− ). The constant red and blue
oxidized by O3 . Furthermore, with ionized and dissociated
arrows relate to the oxidation and reduction, respectively.
molecules O3 reacts more rapidly than with the neutral or
The black arrows indicate transfer of hydrogen, viz.
non-dissociated type of organics (von Gunten, 2003).
combined transfer of electrons and protons. The
significance of hydrogen transfer is linked with the relative
1.3. Superoxide anion radical (O2 ●− )
BDE (bond dissociation energy) of X–H. Adapted with
permission from Nolte and Peijnenburg (2018). Copyright
O2 ●− can initiate the mineralization of organic pollutants
2018 Elsevier.
by hydrogen abstraction as a primary step often resulting in
the generation of some carbon-based radicals (Baum, 1984;
Sawyer and Valentine, 1981). Peroxy intermediates are gener-
ated after the combination of oxygen with these carbon-based by speciation of the organic substance as a function of pH
radicals. A consecutive breakdown of peroxy-intermediates (Nolte and Peijnenburg, 2018). The O2 ●− reacts with the hy-
results in the production of degradation compounds drogen donors (such as, (poly)phenols) through the hydrogen
(Frimer et al., 1986). Degradation of various pollutants abstraction process (top left pointer in Scheme 2). Although it
like dyes and bisphenol A involves O2 ●− as a major contribut- is noted yet again that the HO2 ● may also perform as an H-
ing species as reported in numerous studies (Fu et al., 2015; donor species (bottom-right dashed pointer in Scheme 2), ow-
Wang et al., 2017b). Moreover, O2 ●− is even capable of breaking ing to the frail bonding between oxygen and hydrogen (O–H;
the carbon-chlorine bonds (Li et al., 2011), thereby forming 49 kcal/mol) (Fuller, 2008). This event is estimated to take place
dechlorinated products. essentially at a pH that is considered acidic. At this pH (< 4.9)
Since hydrogen abstraction is recognized as a predomi- it might also exhibit a competition with the one-electron oxi-
nant route for O2 ●− reaction in the case of (poly)phenols such dation.
as hydroquinone (Nakarada and Petković, 2018), concurrently,
during the electron transfer by HO2 ● may also result in the 1.4. Sulfate radicals (SO4 ●− )
same product as with H-abstraction by O2 ●− (Nolte and Peij-
nenburg, 2018), presented in Eq. (9). SO4 ●− -based oxidative degradation of pharmaceuticals has
been receiving more and more consideration in wastewater
treatment technology as SO4 ●− has a high redox potential
(Eberson, 1982). The reduction potential (standard) of SO4 ●−
(E0 = 2.5 to 3.1 V) is much comparable to that of ● OH (E0 = 1.9
to 2.7 V), which allows SO4 ●− to perform as a potent oxidant
for contaminant degradation (Zhang et al., 2015). Despite the
similar oxidation potential for both SO4 ●− and ● OH (Oh et al.,
(9) 2016), the SO4 ●− -driven oxidations are much more selective
(Lee et al., 2020). In order to bring this higher selectivity of
The reaction of O2 ●− with an extensive range of pollu- SO4 ●− into perspective, the rate of reaction between humic
tants or organic compounds takes place by means of hydro- acids and SO4 ●− can be related to humic acids and ● OH, which
gen transfer, reduction, and one-electron oxidation. The in- is approximately 103 L mg/C and 104 L mg/C, respectively
vestigations involving O2 ●− and organic species are intrin- (Lutze et al., 2015b). While SO4 ●− and ● OH have analogous re-
sically complicated because of the competing oxidative and dox potentials, it is of note to specify that the reaction rate of
reductive routes, which may become even more challenging organic compounds with SO4 ●− is usually lower, around 106 to
694 journal of environmental sciences 147 (2025) 688–713

107 mol/(L. sec) relative to that with ● OH (Buxton et al., 1988; Kearns, 1971). 1 O2 tends to function as an electron acceptor
Lee et al., 2020; Lutze et al., 2015b; Sonntag et al., 1997), with to form O2 ●− with phenolic molecules comprising electron-
reaction rates between 108 and 109 mol/(L. sec). donating groups. Concurrently, such compounds may also
Nevertheless, SO4 ●− is suggestively more stable than ● OH, generate corresponding phenolic radicals eventually yielding
and largely undergoes reaction with organic substances via ring-opening and also quinone-like products, as a result of re-
electron transfer mechanism (Neta et al., 1977). It has also arrangement or further combination with oxygen (Scully and
been established that compared to ● OH, the decay of SO4 ●− Hoigné, 1987). In general, 1 O2 exhibited higher reactivity to-
by the non-target compounds present in water (e.g., NOM) is wards compounds comprising electron-rich sites, e.g., unsat-
quite lower. This infers that AOPs based on SO4 ●− as the re- urated sulfide and amino groups (Kim et al., 2012; Min and
active species may possibly be more effective than ● OH-based Boff, 2002; Schweitzer and Schmidt, 2003; Xiao et al., 2021;
processes, particularly in the treatment of real water matri- Zhou et al., 2015).
ces (Ghauch and Tuqan, 2012; Lutze et al., 2015a). Therefore, The rates of reaction of 1 O2 with many organic compounds
for such applications, remediation through in-situ oxidation are lower (about 104 to 107 mol/(L. sec) (Lee et al., 2020) as com-
using activated PS (persulfate) could be favored over ● OH or pared to ● OH and SO4 ●− , on account of the low redox potential,
peroxide-based treatment processes (Waldemer et al., 2007; i.e., 0.65 V (Rayaroth et al., 2023), indicating that 1 O2 is a weak
Yan et al., 2013), because the more stable persulfate anion oxidant relative to other radical species. More specifically, a
might be transferred beyond, to the sub-surface prior to be probe compound for 1 O2 such as furfuryl alcohol has a range
able to get activated for oxidation of the target pollutants. of reaction rate reaching approximately 108 mol/(L. sec), which
SO4 ●− may undergo a reaction with organic pollutants is too high to detect the rapid removal of an organic contami-
through electron transfer, hydrogen abstraction, and ad- nant by 1 O2 implicated in real wastewater application.
dition to the double bond (Neta et al., 1977). Owing to Although several reports have studied 1 O2 as a major ROS
the electrophilic nature of SO4 ●− , electron-donating groups for degradation of organic compounds (Cheng et al., 2017;
demonstrate rates of reaction with SO4 ●− which are way Luo et al., 2019; Zhu et al., 2020), on account of its lower re-
more rapid when compared to electron-withdrawing groups dox potential, mineralization of these pollutants via 1 O2 does
(Tsitonaki et al., 2010). Persulfate (PS) directly undergoes reac- not seem to be a much prudent approach. Hence such infer-
tion with some organic compounds, generating SO4 ●− which ences are more of a dubious nature and need to be reaffirmed
can propagate secondary reactions (Eq. (10)) or might form or- via further experimentation as additional evidence. Moreover,
ganic radicals (Eq. (11)) (Huang et al., 2002; Tanner and Os- while exhibiting a rate constant of ∼105 mol/(L. sec), water can
man, 1987), thus progressively decomposing the target con- also act as a 1 O2 scavenger (Cheng et al., 2017). In most cases
taminant (He et al., 2014; Wang and Liang, 2014). The overall where the medium is water, its concentration is naturally
rates of pollutant degradation are reliant on an interrelated much higher as compared to the organic compounds. A major
sequence of the SO4 ●− chain involved in propagation and ter- contribution by 1 O2 is supposedly provided for the degrada-
mination reactions. However, unlike H2 O2 , PS also efficiently tion of waste chemicals present in water compartments; from
oxidizes some organics directly, without the contribution of the viewpoint of competitive kinetics, the reaction rate of 1 O2
reactive radical species (Ahn et al., 2019; Duan et al., 2018a, with organics should be at least higher than 108 mol/(L. sec) in
2018b; Yang et al., 2018b; Zhou et al., 2015; Zhu et al., 2019). that instance. The second-order rate constants for reactions of
different contaminants with 1 O2 are summarized in Appendix
S2 O8 2− + R → SO4 2− + SO4 ●− + R∗ (10) A Table S3.
Recent findings (Hu et al., 2021) reported a 1 O2 -dominated
S2 O8 2− + R → 2SO4 ●− + R● (11)
activation of PMS for catalytic degradation of tetracycline
Furthermore, SO4 ●− is reported to be highly sensitive to- through a non-radical oxidation production pathway. A com-
ward substituent groups of the aromatic molecules (Luo et al., prehensive mechanism for the generation of 1 O2 in the
2017), however, the nature of substituent groups determines process has been described which was supported through
the route of the reaction. For example, following the initial quenching experiments, as well as EPR studies. The produc-
step of oxidation via SO4 ●− , the aromatic radical cations gen- tion of other reactive species, such as ● OH, SO4 . ●− and O2 ●− ac-
erated could rearrange themselves in a way so as to undergo companied the activation process of PMS by biochar but were
oxidation of the side chain while substituent moieties may observed to be reduced in the subsequent reactions. The H2 O2
act as electron donors (Caregnato et al., 2008). Contrary to this generated during activation of PMS reacted with ● OH to pro-
when the substituents act as electron acceptors, dehalogena- duce HO2 ● , which further decomposed to yield O2 ●− (Eqs. (12)-
tion of the halo-aromatic radical cations is a preferred path- (14)) (Reshetnyak et al., 2003; Yan et al., 2017). The recombina-
way (Madhavan et al., 1978). tion of O2 ●− , along with the reaction of ● OH with O2 ●− succes-
sively resulted in the formation of 1 O2 as depicted in Eqs. (15)-
1.5. Singlet oxygen (1 O2 ) (18) (Duan et al., 2020, 2018a; Qi et al., 2020).

1O is known to be an extremely selective oxidant capa- HSO5 − + H2 O → HSO4 − + H2 O2 (12)


2
ble of oxidizing unsaturated organic substances either via
H2 O2 + ● OH → HO2 ● + H2 O (13)
electron abstraction or an electrophilic attack (Kearns, 1971;
Thomas and Foote, 1978). For example, reactions of 1 O2 with HO2 ● → O2 ●− + H+ (14)
alkenes have been reported to form various degradation prod-
ucts depending on the structure of alkenes (Frimer, 1979; ● OH + O2 ●− → 1 O2 + OH− (15)
journal of environmental sciences 147 (2025) 688–713 695

2O2 ●. − + 2OH+ → 1 O2 + H2 O2 (16) Fe2+ ). A wide range of organic contaminants can be oxidized
using SO4 ●− due to its high reactivity. 1 O2 , on the other hand,
2O2 ●− + 2H2 O → H2 O2 + 1 O2 + 2OH− (17) is typically formed through the reaction of molecular oxygen
(O2 ) with a photosensitizer, such as organic matter or certain
HO2 ● + O2 ●− → 1 O2 + HO2 − (18) transition metals, upon exposure to light (Dong et al., 2023;
Gao et al., 2020; Jin et al., 2023b; Xie et al., 2022; Yang et al.,
Another prominent production pathway for 1 O2 during 2019).
Fenton-like reactions is specifically during PS activation by Reaction conditions: The conditions under which PS activa-
metal ions. Although their contribution differs in each system, tion takes place can influence the relative contribution or sig-
the 1 O2 is either generated by direct reaction of oxidants with nificance of SO4 ●− and 1 O2 (Ghanbari and Moradi, 2017). For
the heterogeneous catalyst or by the interconversion of differ- example, several catalysts or pH conditions may favor the
ent ROS, particularly the O2 ●− . This has been abundantly re- abundant generation of one ROS over the other.
ported in processes involving metal ion-based heterogeneous Complex reaction pathways: In some cases, multiple ROS can
catalysts, such as zerovalent iron-based materials and tran- be produced concurrently and their relative role in degrada-
sition metal-doped catalysts (Hayat et al., 2021; Huang et al., tion may shift during the process as reaction intermediates
2022; Tan et al., 2021a). For example, in the instance of Mn- are being formed and consumed (Wang et al., 2021).
doped g-C3 N4 , Mn(III) as the primary form of Mn functions Contaminant reactivity: The choice between SO4 ●− and 1 O2
to activate PMS to produce O2 ●− and Mn(II). The reaction of as the dominant ROS might also depend on the specific con-
Mn(II) with PMS can result in the regeneration of Mn(III). Both taminants present in the medium. Some contaminants are
reactions are followed by the creation of a medium copious more susceptible to oxidation by SO4 ●− , while others may be
with O2 ●− , which predominantly serves as the precursor of more reactive in the presence of 1 O2 , resulting in their con-
1 O (Eqs. (19)-(25)) (Fan et al., 2019).
2 sumption while decreasing the concentration during degra-
dation of target pollutant.
Mn(III) + HSO5 − + H2 O → Mn(II) + 3H+ + O2 ●− + SO4 2− (19)
Detection methods: Identifying and quantifying the particu-
lar ROS involved in a given system can be challenging, and
Mn(II) + 2HSO5 − → O2 ●− + Mn(III) + 2HSO4 − (20)
various analytical techniques might be utilized for monitor-
ing and detection purposes. Also, some techniques are better
2O2 ●− + 2H2 O → 1 O2 + H2 O2 + 2OH− (21)
suited to detect one ROS over the other which can potentially
S2 O8 2− + 2OH− → 2SO4 2− + HO2 − + H+ (22) lead to variations in the dominant ROS that are reported.

S2 O8 2− + HO2 − → SO4 ●− + 2SO4 2− + O2 ●− + H+ (23) 1.6. Hydroperoxyl radicals (HO2 ● )

S2 O8 2− + 4OH− → 2SO4 2− + 1 O2 + 2H2 O (24) HO2 ● are also formed in AOPs (in addition to the production
of aforementioned reactive radicals). Although, owing to the
HSO5 − + SO5 − → HSO4 − + SO4 2− + 1 O2 (25)
lower oxidizing potential of this ROS, there barely exists any
data regarding the degradation of pharmaceuticals by HO2 ● (or
Therefore, this non-radical pathway involving 1 O2 for the
even organic pollutants in general (Wang and Wang, 2020c).
degradation of contaminants by AOPs is largely relevant
Nonetheless, in the course of indirect (● OH-based) reac-
(Rayaroth et al., 2023). The highly selective 1 O2 is produced in
tions of the ozonation process, HO2 ● are formed during the
the medium through various pathways, the most significant
terminal phase which may play a role in initiating the radical
of which involves activated persulfates, modified photocata-
reactions all over again. Consequently, a chain reaction is pre-
lysts, and photosensitizers.
served by the supposed promoters, which can be specified as
It is also observed through literature (Gao et al., 2018;
substances (including organic molecules) that transform ● OH
Xiao et al., 2021) that in many systems based on the activa-
to O2 ●− .
tion of persulfate (PMS or PDS), 1 O2 is emphasized as the dom-
Some properties of the reactive oxygen species (ROS) piv-
inant ROS contributing toward degradation rather than the
otally involved in AOPs have been abridged in Table 2, since the
SO4 ●− . In effect, the predominant ROS in a persulfate-based
degradation efficiency of AOPs tends to vary with the proper-
system might not be the same for all contaminants or under
ties of the ROS contributive in a process.
all conditions. Both the SO4 ●− and 1 O2 can contribute to pol-
lutant degradation, and the relative significance of each ROS
can vary. The prevalence of SO4 ●− vs. 1 O2 as the primary re-
active species in a system based on the activation of PS de- 2. The case of acetaminophen (ACT)
pends on various factors, including reaction conditions, reac- degradation
tant concentrations, and the nature of the contaminants being
degraded (Shao et al., 2023; Xia et al., 2020). The contribution of Also renowned as paracetamol, acetaminophen (ACT) is
SO4 ●− or 1 O2 toward pollutant degradation in a given system known for its pervasive detection as pharmaceutical waste
seems to be reliant upon some considerations, for example: residues in drinking water (Gusseme et al., 2011), wastewater
Reaction mechanisms: The formation of SO4 ●− occurs when (Chen et al., 2021), surface water (Yang et al., 2020), groundwa-
persulfate ions (S2 O8 2− ) are activated via different methods, ter (Vulliet and Cren-Olivé, 2011) and landfill leachate (Yu et al.,
such as heat, UV radiation, or the presence of catalysts (e.g., 2020). For the most part, this antipyretic and analgesic drug
696 journal of environmental sciences 147 (2025) 688–713

Table 2 – Properties of ROS.

ROS Reduction Half-life in water Mechanism/pathway Ref.


potential

OH 2.8 V 10−9 sec (Morris et al., 2022) H-abstraction, one-electron oxidation, (Buxton et al., 1988)
and aromatic substitution reaction
(non-selective reactions)
SO4 ●− 2.5 – 3.1 V 10−6 sec One-electron oxidation (Neta et al., 1988)
(Uthirakrishnan et al., 2020)
O2 ●− 0.95 V 10−6 sec (Morris et al., 2022) Disproportionation, one-electron transfer, (Hayyan et al., 2016)
deprotonation, and nucleophilic
substitution
1
O2 0.65 V 10−6 sec (Morris et al., 2022) Oxidation of the electron-rich sites (Min and Boff, 2002)

enters the water matrices by excretion of humans and an- treatment process. As an instance, the ACT molecules default
imals, a major part of which is known to reach the water to ● OH attack since ● OH are generated in profusion which is
bodies following incomplete elimination by the sewage treat- responsible for the degradation of ACT, as highlighted by a re-
ment plants (Langford and Thomas, 2009; Phillips et al., 2010; cent report on its ultrasonic treatment (Gao et al., 2022). The
Wu et al., 2012). Pharmaceutical contamination brings po- tendency of ● OH as a non-selective oxidant to undergo reac-
tential risks to human health and ecological niches essen- tion with organic compounds is depicted through three key
tially because of bioaccumulation, strong persistence, and mechanisms, viz., electron transfer, hydrogen abstraction, and
poor biodegradability (Akhtar et al., 2016; Papageorgiou et al., addition of hydroxyl group (hydroxylation) to unsaturated car-
2016). The impending adverse impacts of ACT in small con- bons.
centrations include inhibited cell proliferation, as well as re- Initiation of the different degradation routes shown in
productive damage (Van et al., 2020). Scheme 3a occurs by parallel attacks of the oxidant radical
The concentrations of ACT detected in effluents of Euro- either on the C(2)-position or C(4)-position of ACT. In these
pean STP rose to 6 μg/L (Ternes, 1998), about 10 μg/L in natural transformations, the first pathway corresponds to the ACT
water bodies in the USA (Kolpin et al., 2002), and a concen- aromatic ring hydroxylation giving rise to the emergence of
tration of more than 65 μg/L was determined in Tyne River in its hydroxylated product (TP1) (El Najjar et al., 2014). This was
the UK (Roberts and Thomas, 2006). The maximum concen- subsequently followed by cleavage of the aromatic ring of TP1
trations of ACT in China (Jiulong), Portugal (Arade), Spain (Bil- resulting in the generation of the ring-opening compound;
bao), and USA (New York Bay) have been reported as 13 ng/L, TP2 (Moctezuma et al., 2012).
88 ng/L, 440 ng/L, and 162 ng/L, respectively (Letsinger et al., An initiating ● OH attack on the para-position of
2019). Owing to the simplicity of chemical structure and am- phenol-formed hydroquinone during a second pathway,
ple investigation on the degradation of ACT via various ROS, it with consequent oxidation of hydroquinone yielding TP3
was chosen to thoroughly explore the in-depth mechanisms (Moctezuma et al., 2012). Whereas, as a third pathway,
of ACT oxidation with various oxygen-based radicals. the acetyl-amino group underwent ● OH attack preceding
The structure of the parent compound, acetaminophen the appearance of TP4, which was further oxidized to TP5
(ACT), has two vulnerable locations; (a) the acetamido group (Wang et al., 2019).
(–NH–CO–CH3 ) and (b) an aromatic ring. The reactive oxida-
tive species (● OH, O2 ●− , SO4 ●− , 1 O2 ) and the free electrons can
2.1.2. Attack of ● OH on aromatic and acetamido group of ACT
readily attack such susceptible sites in a molecule. Therefore,
A study recently reported on catalytic degradation of ACT
the various ACT degradation mechanisms can recurrently and
through potassium peroxydisulfate (PDS) activation via N-
most likely follow the ROS attacking the acetamido moiety or
doped CNT (carbon nanotube) modified surfaces elucidates
the aromatic ring.
a mechanism pathway that might pursue the attack by ● OH
The ideal consequence of any ROS-mediated degradation
on ACT aromatic ring, as well as on the acetamido group
for pharmaceuticals is considered to be the total mineraliza-
(Govindan et al., 2022). This attack could result in the produc-
tion of target contaminants, as represented in Eq. (26):
tion of N-(3,4-dihydroxylphenol)-acetamide (m/z = 167) and
hydroquinone (m/z = 110) (Li et al., 2020). Additionally, the
ROS + API → CO2 + H2 O (26)
hydroquinone underwent oxidation through ● OH and subse-
quently formed 1,4-benzoquinone (m/z = 108) and acetamide
2.1. The ● OH attack on ACT (m/z = 60).
Contrariwise, the incursion by ● OH can also propagate di-
2.1.1. Hydroxylation of the ACT aromatic ring mediated by hydroxylation, hence generating a dehydroxylated ACT prod-

OH uct (m/z = 167), which possibly could undergo additional
The nature or type of the contributing ROS has a strong role conversion to malonic acid (m/z = 104)) and 1-amino-1,3,4,-
in determining the degradation pathway during an oxidative trihydroxybutan-2-one (m/z = 135) as illustrated in Scheme 3b.
journal of environmental sciences 147 (2025) 688–713 697

(a) Degradation route followed by ACT under the attack of OH.

(b) Attack of OH on aromatic and acetamido group of ACT. (c) Electrophilic adduct reaction of ACT with ROS.

Scheme 3 – ● OH-mediated reactions with ACT (a) Degradation route followed by ACT under the attack of ● OH. Adapted with
permission from Gao et al. (2022). (b) Attack of ● OH on aromatic and acetamido group of ACT. Adapted with permission from
Govindan et al. (2022). (c) Electrophilic adduct reaction of ACT with ROS. Adapted with permission from Govindan et al.
(2022). Copyright 2022 Elsevier.

2.1.3. Electrophilic adduct reaction tion potential. These dehydroxylated and deacetylated prod-
An electrophilic adduct reaction has also been reported to ucts were further oxidized to produce simpler molecules like
have occurred either between ACT and ● OH or through 2-hydroxypropanoic acid (m/z = 90) along with acetamide
electron transfer utilizing other ROS, i.e., SO4 ●− , O2 ●− , 1 O2 (m/z = 59).
(Govindan et al., 2022). This mechanistic pathway (Scheme 3c)
resulted in the generation of dehydroxylated byproducts, such 2.1.4. Oligomerization of ACT mediated by ● OH
as N-(3,4,5-trihydroxylphenyl)acetamide (m/z = 167), along Another investigation (Tong et al., 2021) reports the reac-
with the elimination of CH3 CO- group hence forming 4- tion of ACT with saturated ferric (Fe3+ ) clay particles under
aminophenol (m/z = 109) via SO4 ●− having greater oxida- UV–visible irradiation which was observed to be 1.7 times
698 journal of environmental sciences 147 (2025) 688–713

Scheme 4 – ACT photo-transformation through ● OH attack. Adapted with permission from Tong et al. (2021). Copyright 2021
Elsevier.

greater than the ACT removal in the dark. This could be as- termediates was likely to take place, along with the dona-
cribed to ROS formation in the presence of light. The study tion of an electron to form cationic radicals with consequent
involved the elucidation of six products via mass spectra oligomerization. Successively, the adjacent hydroxyl groups of
of which three (TP4, TP5, and TP6) were specifically gen- the dimer underwent an intermolecular cyclization reaction.
erated under light conditions, most likely after undergo- Moreover, the formation of TP6 was prompted by oxidation
ing ● OH-based reactions, as represented in Scheme 4. N-(5– in the para-position of the amido in, followed by the abstrac-
hydroxyl-1,6-dioxohex-3-en-2-yl)acetamide (C8 H11 NO4 ) and tion of an electron by ● OH from the hydroxyl group of ACT to
dimer of ACT or 2,4-hexadienal (C14 H13 NO3 ) were referred produce phenoxyl radicals (Deng et al., 2010). This was subse-
as TP4 (m/z = 186.031) and TP5 (m/z = 246.242) respec- quently followed by the transformation process similar to the
tively, while 9-nitrodibenzofuran-2,3,6,7-tetraone (C12 H3 NO7 ), ACT degradation pathway via single electron transfer (SET) as
the compound yielded via cross-linking of the phenol-based illustrated in Scheme 5a. TP1 was yielded as a result of ● OH
radicals was referred as TP6 (m/z = 274.274) according to the attack or hydrolysis leading to the elimination of acetamide
MS analysis (Zhang et al., 2018). Induced by an initial at- from the dimer (Li et al., 2017). Furthermore, oxidation by ● OH
tack by ● OH on the benzene ring, followed by subsequent de- in the para-position of P1 formed TP3 (Zhang et al., 2008).
protonation and formation of the double bond between car- Also, TP1 might as well be produced by the cross-coupling
bon and oxygen, eventually leading to the breakage of the between ACT and hydroquinone, while further oxidation of
carbon-carbon bond resulting in the production of TP4. More- TP1 could subsequently lead to TP5 formation. The photo-
over, the attack of ● OH on ACT could also lead to the gen- transformation of ACT included two reaction pathways, in-
eration of hydroquinone and p-hydroxyaniline (Liang et al., cluding both single electron transfer and the attacking ● OH
2016; Zhang et al., 2008). Further hydroxylation of these in- pathway (as proposed in Scheme 4).
journal of environmental sciences 147 (2025) 688–713 699

Scheme 5 – Electron transfer processes for ACT (a) Single electron transfer pathway for ACT transformation. Adapted with
permission from Tong et al. (2021) Copyright 2021 Elsevier.. (b) Electron-transfer process in ACT. Adapted with permission
from Govindan et al. (2022). Copyright 2022 Elsevier.

2.2. The electron-transfer process the predominant pathway in the absence of light (Tong et al.,
2021).
2.2.1. ACT dimerization via single electron transfer (SET)
The degradation of ACT through single electron transfer in- 2.2.2. Electron transfer from the acetamide group of ACT
volves the formation of three dimers of ACT, referred to as Another degradation pathway has been reported very recently
TP1, TP2, and TP3 in Scheme 5a, with molecular formulae (Govindan et al., 2022) as depicted in Scheme 5b. It represents
C14 H13 NO4 , C16 H16 N2 O4 and C14 H11 NO4 respectively. Compa- an electron-transfer process occurring in the acetamide group
rable compounds have been documented on phenol oligomers which could produce an N-radical cation (m/z = 151). This N-
forming via the electron transfer process (Gu et al., 2008, radical cation intermediate species could likely be stabilized
2011; Polubesova et al., 2010; Qin et al., 2015). The ACT un- through the neighboring phenyl ring and the carbonyl group.
derwent a transformation into cationic radicals after the loss Consequently, the intermediate reacted with the oxidative
of one electron from the hydroxyl functional group. Carbon- species to yield a peroxide-based product (m/z = 141), namely
centered radicals were produced as an outcome of the elec- 4-(hydroperoxyamino)phenol, and the dehydration product
tron (unpaired) in phenoxyl radical being delocalized via res- (m/z = 123) called 4-nitrosophenol. Such transformation prod-
onance in the adjacent conjugated benzene ring positions ucts have been identified and experiential in prior studies
(Lu and Huang, 2009). These carbon-centered radicals would (Li et al., 2017). Owing to this electron-transfer process taking
react among one another to produce carbon-carbon or carbon- place in ACT, the formation of phenoxyl radical (m/z = 150)
oxygen bonds at several positions as cross-connecting dimers was observed as well, which could be oxidized further to pro-
observed as TP2, which successively yielded TP1 and TP3 after duce a transformation product of ACT, i.e., N-(3–hydroxy–4-
hydrolysis and oxidation (Li et al., 2017; Zhang et al., 2017b). oxocyclohexa-1,5–dien–1-yl)acetamide with m/z of 168. Hence
This electron transfer process has been stated in reports as it was evidenced that the electron transfer in the acetamido
700 journal of environmental sciences 147 (2025) 688–713

radical addition elimination reaction (Anipsitakis et al., 2006;


Table 3 – The reduction potentials (E0 ) of ● OH and organic
Li et al., 2020b).
radicals. Adapted with permission from (Li et al., 2021).
Copyright 2021 Elsevier. By means of SET reaction of CH3 COO● and ACT, it is possi-
ble to directly generate the cationic radical of ACT (ACT●+ ), to
Radical species Reaction equation E0 (V)
which HO− might further add to produce the adduct of HO (IM-
2). These HO-adducts could subsequently undertake H-atom
● ●
OH OH + e− → − OH 1.90a shift, oxygen addition, and eventually the exit of HO2 , hence

OH + e− + H+ → H2 O 2.73a
yielding the hydroxylated product (TP1) of ACT. Such ACT-
CH3 COO● CH3 COO● + e− → CH3 COO− 2.12
CH3 COO● + e− + H+ → CH3 COOH 2.25
based hydroxylated products have been identified in recent re-
CH3 O● CH3 O● + e− → CH3 O− 0.65 ports on AOPs with peracetic acid as oxidants (Ghanbari et al.,
CH3 O● + e− + H+ → CH3 OH 2.08 2021).
● ●
CH3 CH3 + e− → CH3 − − 0.68 A reaction between CH3 COO-adduct and molecular oxy-

CH3 + e− + H+ → CH4 1.94 gen in a parallel route toward the HO-adduct was predicted,
CH3 COOO● CH3 COOO● + e− → CH3 COOO− 1.22
which finally resulted in the formation of an acetoxylated
CH3 COOO● + e− + H+ → CH3 COOOH 1.64
product (TP2) of ACT. The anticipated formation pathway of
CH3 OO● CH3 OO● + e− → CH3 OO− 0.55
CH3 OO● + e− + H+ → CH3 OOH 1.26 TP2 seemed to be more promising, while the acetoxylated
CH3 C● O CH3 C● O + e− → CH3 CO− − 0.77 products of aromatics have not been particularly detected dur-
CH3 C● O + e− + H+ → CH3 COH 1.39 ing AOPs based on PAA (Ghanbari et al., 2021; Kim et al., 2019;
a
Wang et al., 2020c). Benzoate esters were previously reported
The E0 values for the reference compound (Wardman, 1989).
to catalytically hydrolyze via base to form benzoates (Xu et al.,
2021, 2019). As observed in Scheme 6, TP2 was likely to un-
group of ACT is also capable of inducing degradation of the dergo base-catalyzed hydrolysis.
drug compound. The initiated attack of HO− on TP2 formed IM-4 followed
by the withdrawal of CH3 COOH from IM-4. All the pathways
proposed for CH3 COO● -mediated hydroxylation of ACT were
2.3. Hydroxylation mechanisms of act mediated by observed to be thermodynamically possible. A viable hydroly-
CH3 COO● sis may lead to no accumulation of acetyl compounds.
In this observed scenario, the ACT reaction rate constants
Further investigations (Li et al., 2021) disclosed that among with organic radicals and ● OH were found to be linearly
the studied organic radicals, CH3 COO● is the most potent related to the reduction potential values of these radicals
as an oxidizing species. According to some other studies on (Li et al., 2021). Enlisted in Table S4 are the second-order re-
UV/peracetic acid (PAA) processes (Zhang and Huang, 2020), action rate constants (kACT ) calculated for ACT with the or-
CH3 COOO● was the most abundantly generated organic rad- ganic radicals and ● OH at 298 K in the aqueous solution. The
ical but it possesses a lower reduction potential. With ref- order of kACT was reported as CH3 COO● > ● OH > CH3 O● >
erence to ● OH, the E0 (one-electron reduction potentials) of CH3 COOO● >> ● CH3 > CH3 C● O > CH3 OO● . Therefore, it can be
several organic radicals produced in the PAA-based system deduced that among these radicals, CH3 COO● demonstrates
were investigated and enlisted in Table 3 for comparisons. the maximum reactivity for ACT, followed by ● OH, CH3 O● and
At pH 7, the E0 values calculated for ● OH were 2.73 V, and CH3 COOO● .
2.12 V and 2.08 V for CH3 COOO● and CH3 O● respectively, while The degradation of ACT via ● OH was initiated via numer-
the E0 values for all other organic radicals turned out to be ous reaction paths, with 58.78% of branching ratios (, %) for
< 2 V. RAF (radical adduct formation), 17.55% for HAA (H-atom ab-
The ACT degradation reactions initiated by the organic straction), and 23.67% for SET (single electron transfer). It can
radicals have been assumed to comprise the radical adduct be further interpreted from the values reported in Table 3, that
formation (RAF) channels on both the benzene ring and ac- the reaction rate constant of CH3 COO● with ACT (5.44 × 1010 )
etamide group, H-atom abstraction (HAA) channels from the is even higher than that with ● OH. ACT degradation initi-
benzene ring, the -OH group, and the acetamide group, and ated by CH3 COO● appeared to be dominated by multiple re-
single electron transfer (SET) channels. Moreover, in order to action channels, involving RAF channels on the benzene ring
abstract an H-atom from ACT, as well as for addition to oc- as well as the acetamide group, HAA channels of the OH group
cur in the ACT unsaturated bonds, such sites of organic rad- and the acetamide group, and the SET channel. The prevalent
icals are speculated which possess maximum unpaired elec- pathways for CH3 COOO● and CH3 O● with ACT were HAA reac-
tron spin density. tions alone, whereas the analyzed reaction rate constants for
The proposed CH3 COO● -mediated hydroxylation route of ACT with ● CH3 , CH3 C● O, and CH3 OO● could be overlooked for
ACT is illustrated in Scheme 6. The successive investigated re- being too low.
actions of RAF and SET channels revealed that CH3 COO● pos-
sibly will react with ACT and hence lead to yield an adduct of 2.4. Degradation mechanism of ACT mediated by 1 O2
acetoxyl (CH3 COO-adduct, IM-1) via the RAF channel. It was
further found that CH3 COO-adduct underwent a unimolecu- More recent works (Lin et al., 2021) revealed that peroxymono-
lar reaction hence eliminating CH3 COO− . This phenomenon sulfate (PMS) was exploited as an oxidant for the efficient
of the addition of CH3 COO● to a molecule proceeding to the degradation of ACT in hydrolyzed urine (HU). It was found that
elimination of CH3 COO− is recognized as SRAE, i.e., sequential 1 O and ● OH both occurred in the HU/PMS system, however,
2
journal of environmental sciences 147 (2025) 688–713 701

Scheme 6 – Proposed CH3 COO● -mediated hydroxylation mechanisms of ACT. Adapted with permission from Li et al. (2021).
Copyright 2021 Elsevier.

Scheme 7 – Proposed photocatalytic 1 O2 -mediated degradation process of ACT. Adapted with permission from Lin et al.
(2021). Copyright 2021 Elsevier.

1O was the predominant ROS contributing toward the ACT as well as photo-induced holes as vital contributors in the re-
2
degradation. moval process.
While reacting with ACT, 1 O2 was able to fragment the A photocatalytic pathway for ACT degradation mediated by
amide bone between the C and N in the benzene ring, be- O2 ●− was suggested as depicted in Scheme 8. Assisted by Pd-
sides possibly resulting in the appearance of three major BiVO4 photocatalyst under the irradiation of visible light, the
transformation products, namely 4-nitrophenol (TP1), 1,2,4- study implicated two reaction routes. In the initial course, the
benzentriol (TP2), and benzaldehyde (TP3) (Le et al., 2017) as decomposition of the amide group of ACT takes place as an-
shown in Scheme 7. In addition, ACT and its byproducts un- ticipated by the outcomes of DFT (density functional theory)
derwent decomposition into oxalic acid (TP4) and maleic acid calculation data of ACT optimization (Wang and Bian, 2020). In
(TP5) since breakage of the benzene ring occurred (Mu et al., addition, the monophenyl aminophenol derivative (TP2) was
2020). found to be produced subsequent to the elimination of the
COCH3 group from ACT. Eventually, the benzene ring of TP2
2.5. Degradation mechanism of ACT mediated by O2 ●− underwent the attack by reactive species resulting in its open-
ing, resulting in the production of the maleic acid intermedi-
Wang and Bian (2020) prepared and utilized a photocatalytic ate (TP1). In the presence of the employed photocatalyst, such
system comprising small, highly dispersed Pd nanoparticles acids with minimal molecular weights were further mineral-
in Pd-BiVO4 for the degradation of ACT to afford simpler ized to CO2 and H2 O.
molecules leading to complete mineralization. Under visible The radical activation results in a second process which
light irradiation, Pd-BiVO4 displayed remarkable activity for comprises ACT dimerization via an amide-methyl func-
ACT achieving 100% removal in one hour. During the degra- tional group to produce the dimer (TP6). The link composed
dation process, ACT mineralization percentage stretched to of amide and carbonyl species of the dimer TP6 under-
about 40% as depicted by values of TOC removal. This sys- went cleavage via light-induced ROS to form TP5 (N,N-bis(4-
tem for ACT degradation involves O2 ●− (superoxide radicals) hydroxyphenyl)oxalamide) after loss of the ethene group.
702 journal of environmental sciences 147 (2025) 688–713

Scheme 8 – Proposed photocatalytic O2 ●− -mediated degradation process of ACT. Adapted with permission from Wang and
Bian (2020). Copyright 2020 Elsevier.

This was followed by the transformation of TP5 to TP4 and can take place to yield maleic acid (TP1), which can in turn
TP3 when the carbonyl group was cleaved. TP4 and TP3 were afford small molecules leading to complete mineralization.
then converted to ACT and TP2 (4-(hydroxyamino)phenol), Another study recently explored O2 ●− as an oxidant
respectively, after undergoing the addition of the hydroxyl for degradation of ACT in a binary system consisting
group. By way of further oxidation via the addition of HO- of bis(trifluoromethylsulfonyl)imide-based hydrophobic ionic
groups in the presence of O2 ●− , ring cleavages of ACT and TP2 liquids (ILs) and acetonitrile (AcN) as an aprotic solvent. The
journal of environmental sciences 147 (2025) 688–713 703

Scheme 9 – Suggested pathway for SO4 ●− -dominated degradation process of ACT. Adapted with permission from
Rodríguez-Narvaez et al. (2020). Copyright 2020 Elsevier.

subsequent in situ degradation of ACT occurred via O2 ●− therefore underlining the higher oxidizing power of SO4 ●− in
which was chemically generated by the dissolution of potas- comparison with ● OH.
sium superoxide (KO2 ) (Humayun et al., 2021). O2 ●− was N-(3,4- dihydroxyphenyl) acetamide (TP1), 4-aminophenol
demonstrated as the major ROS in the reaction system con- (TP2), benzene-1,4-diol (TP3) and 3,4-dihydroxybenzene-1-
tributing to 98% removal of the drug under optimum condi- sulfonic acid (TP4) were the primary reaction byproducts af-
tions. This opens the door to explore further the potential ter the treatment via Co3 O4 /PMS process. Scheme 9 explicitly
use of ILs as multi-task agents, namely as media to gener- indicates the pathway for ACT oxidation presenting the by-
ate radical species and as extractive agents for pharmaceu- products generated by the contribution of SO4 ●− . This route
tical wastes. It can also be extended to investigate the use of for ACT degradation via ● OH and SO4 ●− has also been stated in
IL analogues, and deep eutectic solvents, being more biocom- other reports (Zhang et al., 2017b). The study speculated that
patible, less cost of synthesis, and more sustainable, as they either the ● OH oxidizes ACT at carbon-3 of its benzene ring
can prepared from natural metabolites. to form N-(3,4- dihydroxyphenyl) acetamide (TP1) or a substi-
tution reaction would be carried out at carbon-1 of ACT ben-
zene ring giving rise to the formic and amino group, thus pro-
2.6. Degradation mechanism of ACT mediated by SO4 ●− ducing benzene-1,4-diol (TP3). Both of these transformation
products (TP1 and TP3) have been identified in the work of
Degradation based on SO4 ●− has been lately acclaimed as one Rodriguez-Narvaez et al. (2020) as well, suggesting that some
of the most effective ROS used in recent AOPs, primarily owing degradation pathways may implicate ● OH. However, as illus-
to its high oxidizing power (E0 = (SO4 ●− /SO4 2− ) = 2.5 to 3.1 V) trated in Scheme 9, the ACT oxidation by SO4 ●− was initi-
as compared to that of ● OH (E0 = (● OH/H2 O) = 1.9 to 2.7 V) ated by replacing the amide group followed by generating the
(Wang et al., 2017a; Zhang et al., 2019). acetic acid derivatives leading to produce TP2, which was fol-
The investigation of the Co3 O4 /PMS system for ACT re- lowed by the likely substitution of the amine group by ● OH or
moval (Rodríguez-Narvaez et al., 2020) reports a negligible SO4 ●− to generate TP3 or TP4, respectively. TP5 was formed by
concentration of ● OH in the heterogeneous process, which further oxidation of TP4 by ● OH, whereas not any linear or-
was hence predominated by the SO4 ●− -based degradation re- ganic arrangement comprising connected sulfate groups was
actions. By means of the identified by-products and kinetic detected. After several oxidation steps, either TP3 or TP5 even-
data, a mechanistic degradation pathway for ACT was put for- tually yielded carboxylic acids TP6 and TP7.
ward. The outcomes suggested that simpler structures, such Although the study byZhang et al. (2017a) never identified
as malic acid along with succinic acid were generated suc- the by-products produced due to SO4 ●− intervention, they pro-
ceeding the Co3 O4 /PMS process, which could be mineralized posed that SO4 ●− -mediation would be releasing formic acid
suitably by conventional water treatment processes (Le et al., from the amino group pursued by rupturing of the benzene
2017; Pérez-Estrada et al., 2005). Another study (Le et al., 2017) ring hence generating carboxylic acids as well as amines.
also agreed with these results which studied ACT degrada- As the aforementioned mechanisms elucidate the differ-
tion via the electro-Fenton process. It has been found that ence in pathways acquired, intermediates, and transforma-
● OH-based conversions using an electro-Fenton process re-
tion product formation which could arise depending on the
quired 60 mins of reaction time to form several carboxylic ROS employed in the process, Table 4 summarizes the var-
acids (for example, oxalic, maleic, and formic acids). However, ious systems and processes reported to generate these ROS
for the degradation dominated by SO4 ●− as in the case of the along with their relative efficiencies and outcomes through
Co3 O4 /PMS system, merely 5 min of reaction was sufficient each mechanism which was adopted by each ROS after its in-
to yield the carboxylic acids, such as succinic and malic acid, teraction with a specific chemical structure such as ACT.
704
Table 4 – Proposed oxidation routes of ACT for various recent AOPs and the transformation products (TPs) produced.

AOP systems/ Conditions ROS Mechanism of degradation Removal efficiency TP formation Ref.

US irradiation (555 kHz/ OH (i) hydroxylation of the aromatic 50.6% – 86.9% in (i) ACT → m/z 168 → m/z 200 (Gao et al., 2022)
60 W) /ACT 20 μmol/L ring 10 min (ii) ACT → hydroquinone → m/z
(ii) attack of the ● OH on the para 127
position to the phenolic functional (iii) ACT → m/z 110 → m/z 140
group
(iii) attack of ● OH on the
acetyl-amino group

journal of environmental sciences 147 (2025) 688–713


UV/PAA (peracetic CH3 COO● , ● OH and CH3 COO● -mediated hydroxylation – Hydroxylated products (Li et al., 2021)
acid-based) process CH3 COOO● of aromatics with radical adduct (HO-adduct) of ACT and
formation (RAF) reaction as the acetoxylated products
dominant channel (addition as the (CH3 COO-ACT) of ACT
dominant pathway)
Pd-BiVO4 /visible light O2 ●− (i) ACT amide group 100% in 1 hr (40% TOC (i) ACT → monophenyl (Wang and
system decomposition → removal of ACT removal) aminophenol derivative → Bian, 2020)
COCH3 group → benzene ring maleic acid → CO2 and H2 O
opening leading to carboxylic acid (ii) ACT → dimer of
formation ACT → N,N-bis(4-
(ii) ACT dimerization → cleavage hydroxyphenyl)oxalamide → 1,3-
of the carbonyl and amide groups bis(4-
dimer → oxidative addition of hydroxyphenyl)urea → 4,4-́(hydrazine-
hydroxyl groups → ring cleavage 1,2-diyl)diphenol → 4-
of ACT → maleic acid formation (hydroxyamino)phenol/ACT→
leading to small molecules maleic acid → CO2 and H2 O
Co-FeOCl/H2 O2 system, ACT O2 ●− – 87.5% – (Tan et al., 2021)
10 μmol/L, Co-FeOCl 0.2 g/L,
H2 O2 0.5 mmol/L, pH 7
Fe3+ saturated ●
OH (a) in dark: Electron transfer 47.6% ± 1.1% in dark / (a) ACT → cationic (Tong et al., 2021)
montmorillonite clay process → hydrolysis and 78.9% ± 0.5% in light radicals → carbon-centered
(Fe-SMF) (a) in dark and (b) oxidation after 10 hr radicals → cross-coupling dimers
under simulated sunlight, (b) under simulated sunlight: of ACT → dimer of ACT and
Clay 5 g/L, ACT 50 μmol/L, attack of ● OH on the benzene hydroquinone and dimer of ACT
pH 3 ring → deprotonation → hydroxy- and benzoquinone
lation (b) N-(5–hydroxy–1,6-dioxohex-3-
→ oligomerization → electron en-2-yl)
abstraction → transformation as acetamide → 9-nitrodibenzo-
in dark conditions furan-2,3,6,7-
tetraone → hydroquinone and
p-hydroxyaniline → phenoxyl
radicals → dimerization as in (a)

(continued on next page)


Table 4 (continued)

AOP systems/ Conditions ROS Mechanism of degradation Removal efficiency TP formation Ref.

CuMgFe2 O4 /PS system, Electron transfer Non-radical pathway (direct e− 90% in 40 min (59% ACT → dimeric product of ACT (Ali et al., 2021)
temperature 30 °C, ACT (from ACT to PS) transfer route) → oxidation of TOC removal in 60 (N,N-(5,6-dihydroxy-[1,1-
20 mg/L, catalyst loading aminophenol → cleavage of the min) biphenyl]−2,3-diyl)diacetamide)
0.1 g/L and PS 1.5 mmol/L aromatic ring and aminophenol.
Aminophenol → hydroquinone or
to other open-chain compounds,
including N-acetylacetamide,
acetylcarbamic acid and

journal of environmental sciences 147 (2025) 688–713


acetamide, etc. and many
open-chain compounds (acetone,
acetamide, formic and acetic acids,
etc.)
UVA/BiOCl process, ACT HO2 ● /O2 ●− – > 80% TOC removal – (Wang et al., 2020a)
50 μmol/L, catalyst 0.3 g/L, after 180 min
pH 5.4
Modified CNT/PDS SO4 ●− , ● OH, O2 ●− , (i) ● OH attacks the aromatic ring 95% TOC removal after (i) ACT →N-3,4-dihydroxylphenol)- (Govindan et al.,
1
(peroxydisulfate) system, O2 and acetamido group 60 min acetamide and 2022)
ACT 10 mg/L, catalyst (ii) electrophilic adduct reaction hydroquinone → 1–4-
100 mg/L, PDS 1 mmol/L among AAP and ● OH benzoquinone and acetamide
(iii) e− -transfer process to the ACT → dehydroxylated ACT
acetamide group (N-(4-hydroxyphenyl)acetamide)
→1-amino-1,3,4-trihydroxybutan-
2-one and malonic acid
(ii) ACT → dehydroxylated
by-products → acetamide and
2-hydroxypropanoic acid
(iii) ACT → N-radical
cation → peroxide product or
dehydration product/
ACT → phenoxyl radical → N-(3–
hydroxy–4-oxocyclohexa-1,5–dien–
1-yl)acetamide

Ag/ZnO@NiFe2 O4 OH and SO4 ●− Hydroxylation → oxidation of 100% in 15 min (at pH ACT → m/z 168 → (Kohantorabi et al.,
nanorod/UVA/PMS process, aromatic structure and amide 7.0) 1,4-dihydroxybenzene (m/z 110) 2021)
ACT 12 mg/L, catalyst bonds and glycolic acid (m/z 76) /
loading 0.1 g/ L, PMS m/z 110 and m/z
0.2 mmol/L 149 → 1,4-benzoquinone (m/z
108) → 3- hydroxypropanoic acid
(m/z 90) → formic acid (m/z
46) → CO2 and H2 O / nitrate (NO3 − )
(m/z 62)

(continued on next page)

705
706
Table 4 (continued)

AOP systems/ Conditions ROS Mechanism of degradation Removal efficiency TP formation Ref.


PAA/UVC-LED/ Fe(II) OH and RO● –NH– bond disruption, > 95% in 30 min ACT → acetamide, 4-aminophenol (Ghanbari et al.,
system, ACT 20 mg/L, PAA hydroxylation, -OH addition, and N-(3,4- dihydroxyphenyl) 2021)
4 mmol/L, Fe(II) 0.5 mmol/L, aromatic ring opening /ACT acetamide → 4-aminobenzene-
pH 5 polymerization with 1,2-diol → 4-aminophenol and

journal of environmental sciences 147 (2025) 688–713


intermediates/organic radicals hydroquinone
4-aminophenol→ hydroquinone
and 4-nitrophenol→
hydroquinone
4-nitrophenol and
hydroquinone → carboxylic acids
(acetic acid - m/z 60) /
ACT → unknown polymerized
product (m/z = 203)
HU/PMS (hydrolyzed 1
O2 Chlorine substitution and electron 96.1% (at 25 °C) in 60 ACT → 4-nitrophenol, (Lin et al., 2021)
urine/peroxymonosulfate) transfer → benzene ring breakage min 1,2,4-benzentriol and
system, ACT 5 μmol/L, PMS benzaldehyde → oxalic acid,
3 mmol/L, T = 10–50 °C, pH maleic acid
9 ACT → dimer of ACT (m/z = 299)

Co3 O4 /PMS process, OH and SO4 ●− Substitution in the amide group 99% in 30 min ACT → 4- (Rodríguez-
3 mmol/L PMS, 0.2 g/L aminophenol → benzene-1,4-diol Narvaez et al.,
Co3 O4 or 4-hydroxybenzene-1-sulfonic 2020)
acid,
benzene-1,4-diol → malic acid and
succinic acid,
4-hydroxybenzene-1-sulfonic
acid → 3,4-hydroxybenzene-1-
sulfonic acid → malic acid and
succinic acid
Fabricated Mn-g-C3 N4 1
O2 and O2 ●− 100% in 15 min ACT → hydroquinone (Fan et al., 2019)
composited (MnCN) for PMS –
activation (PMS/MnCN
system), ACT 20 mg/L,
0.8 g/L PMS, 0.2 g/L catalyst
0.5-MnCN, initial pH 6.5
journal of environmental sciences 147 (2025) 688–713 707

Numerous types of treatment systems and AOPs can be (iii) Conducting thorough kinetic studies to comprehend
used for the degradation and mineralization of ACT as is ev- the rate and mechanism of reactions, enabling better
ident from the recent literature (Table 4). The preference of control and optimization.
the most appropriate method depends on factors, such as ef- (iv) Applying real-time monitoring and analytical systems
ficiency, cost-effectiveness, and the recyclability of the mate- to maintain optimal conditions and adjust parameters
rials used in the process to avoid any risks of pollution caused as needed.
by the remediation process. (v) Ensuring that the quality of raw materials and purity
The most recommended systems, nevertheless as an opin- of reagents is not compromised as it might lead to by-
ion, for the degradation of ACT can be the AOPs producing ex- product formation. Similarly, precise control over tem-
clusively either O2 ●− or SO4 ●− as the preferred ROS, the rea- perature and/or pressure needs to be maintained to fa-
son being that these are found to counterpoise the reactiv- vor intended reactions.
ity/selectivity phenomenon which is primarily at the heart of (vi) Screening for reaction medium (consisting of solvents
any contaminant degradation process. It is discernible from or solvent mixtures) that are most selective for the de-
the data in Table 2 indicating the relative properties of ROS, sired reactions.
that O2 ●− and SO4 ●− have a medial value of reduction poten- (vii) Employing efficient separation and purification tech-
tial in contrast to the values for other frequently used ROS niques to isolate or recycle the products or catalysts.
which is either considerably high or too low. This makes ● OH (viii) Designing processes with inherent safety features, such
highly reactive while 1 O2 is too selective in nature. Alteration as utilizing safer (less reactive) chemicals, reducing the
in the redox potential is indeed a key factor governing the bal- use of hazardous materials, and putting passive safety
ance such that the ROS is neither too reactive to attack ev- mechanisms into practice.
ery background substance in the medium nor too selective to (ix) Considering the scalability and potential changes in re-
cease to react before the complete mineralization of the target action kinetics while transitioning from lab-scale to pi-
contaminant. Although the choice of the most suitable system lot and industrial-scale processes.
or AOP for ACT degradation would be based on the specific (x) Emphasizing the use of environmentally friendly or
requirements of the application, including the concentration benign reagents to make the processes sustainable.
of the API, the presence of other contaminants, available re- More recyclability would help curtail the waste and by-
sources, and the desired level of removal. Furthermore, pilot products.
studies and a thorough assessment of the local conditions are (xi) Utilizing computational modeling, optimization algo-
often necessary to determine the best treatment approach. rithms, and simulation tools to explore various design
scenarios and identify optimal conditions.
(xii) Conducting a comprehensive risk assessment to iden-
3. Conclusion and perspective tify potential hazards and safety concerns early in the
design process. Performing a life cycle assessment (LCA)
An understanding of the degradation mechanisms is im- to estimate the environmental impact of the process,
portant to the design or optimization of a new ROS-based since such assessments would ensure compliance with
transformation. This Review has presented myriad examples regulations and diminish harm to the ecosystem.
of ROS-mediated reactions with acetaminophen (ACT) that
reflect the most probable pathway for each ROS discussed Although seemingly counterintuitive, studies on reactions
therein. This analytical compilation could help chemists at of pharmaceutical compounds with various ROS have exhib-
all experience levels to predict the degradation products of ited simplistic degradation values for in-lab environments
ACT with nearly any type of reactive species that will be com- which recently has brought to light a greatly enhanced mech-
monly encountered in numerous sorts of advanced oxidative anistic perception. The predominant objective of this study is
treatment processes. As the field of radical chemistry and its to portray the distinctive and inherent reaction mechanisms
applications in environmental remediation continues to ex- at a molecular scale that eventually factor in pharmaceuti-
pand, it is anticipated that this review will serve as a means cal removal via ROS. Since this study relates to the develop-
for providing familiarity with the mechanistic insights into ment of new transformations in pharmaceutical removal, it
drug degradation processes necessary for the optimization of might help to break down the barriers in the way of scheming
the reaction process, selectivity and efficiency, prediction of an apparently unpredictable form of advanced treatment pro-
by-product formation, as well as the safety considerations. cesses. More investigations into enhancing the selectivity in
Some feasible system design strategies to help steer the op- the degradation of pharmaceuticals by making use of varied
timization of reaction performance and improvement of pro- ROS will retrench the costs in addition to widening the scope
cess safety are summarized as follows: of their applications.

(i) Implementing process intensification techniques, such


as continuous processing, microreactors, or multiphase
systems, to improve reaction efficiency and reduce by- Declaration of Competing Interest
product formation.
(ii) Designing or selecting sustainable and eco-friendly cat- The authors declare that they have no known competing fi-
alysts that can enhance the desired reaction pathways nancial interests or personal relationships that could have ap-
while minimizing unwanted side reactions in parallel. peared to influence the work reported in this paper.
708 journal of environmental sciences 147 (2025) 688–713

Blazer, V.S., Iwanowicz, L.R., Iwanowicz, D.D., Smith, D.R.,


Acknowledgments Young, J.A., Hedrick, J.D., et al., 2007. Intersex (testicular
oocytes) in smallmouth bass from the Potomac River and
The authors would like to express thanks to the Ministry of selected nearby drainages. J. Aquat. Anim. Health 19,
Higher Education, Research and Innovation − Oman for their 242–253.
support of this research through TRC block funding grant No. Buffle, M.-O., Schumacher, J., Meylan, S., Jekel, M., von Gunten, U.,
2006. Ozonation and advanced oxidation of wastewater: Effect
BFP/RGP/EBR/22/378.
of O3 dose, pH, DOM and HO•-scavengers on ozone
decomposition and HO• generation. Ozone: Sci. Eng. 28,
247–259.
Appendix A Supplementary data Buxton, G.V., Greenstock, C.L., Helman, W.P., Ross, A.B., 1988.
Critical review of rate constants for reactions of hydrated
electrons, hydrogen atoms and hydroxyl radicals (OH/O−) in
Supplementary material associated with this article can be aqueous solution. J. Phys. Chem. Ref. Data 17, 513–886.
found, in the online version, at doi:10.1016/j.jes.2023.11.021. Caregnato, P., Gara, P.M.D., Bosio, G.N., Gonzalez, M.C., Russo, N.,
Michelini, M.D.C., et al., 2008. Theoretical and experimental
investigation on the oxidation of gallic acid by sulfate radical
anions. J. Phys.Chem. A 112, 1188–1194.
References Chen, G., Dong, W., Wang, H., Zhao, Z., Wang, F., Wang, F., et al.,
2022. Carbamazepine degradation by visible-light-driven
Ahn, Y.Y., Bae, H., Kim, H.I., Kim, S.H., Kim, J., Lee, S.G., et al., 2019. photocatalyst Ag3 PO4 /GO: mechanism and pathway. Environ.
Surface-loaded metal nanoparticles for peroxymonosulfate Sci. Ecotechnol. 9, 100143.
activation: efficiency and mechanism reconnaissance. Appl. Chen, H., Lin, T., Wang, P., Zhang, X., Jiang, F., Wang, Y., 2023. Novel
Catal. B: Environ. 241, 561–569. solar/sulfite advanced oxidation process for carbamazepine
Akhlaq, M.S., Clemens, V.S., 1987. Intermolecular H-abstraction of degradation: radical chemistry, transformation pathways,
thiyl radicals from thiols and the intramolecular complexing influence on disinfection byproducts and toxic changes. J.
of the thiyl radical with the thiol group in 1,4-dithiothreitol. A Chem. Eng. 451, 138634.
pulse radiolysis study. Z. Naturforsch. C. 42, 134–140. Chen, L., Fu, W., Tan, Y., Zhang, X., 2021. Emerging organic
Akhtar, J., Amin, N.A.S., Shahzad, K., 2016. A review on removal of contaminants and odorous compounds in secondary effluent
pharmaceuticals from water by adsorption. Desalin. Water wastewater: identification and advanced treatment. J. Hazard.
Treat. 57, 12842–12860. Mater. 408, 124817.
Ali, J., Jiang, W., Shahzad, A., Ifthikar, J., Yang, X., Wu, B., et al., Cheng, X., Guo, H., Zhang, Y., Wu, X., Liu, Y., 2017.
2021. Isolated copper ions and surface hydroxyl groups as a Non-photochemical production of singlet oxygen via
function of non-redox metals to modulate the reactivity and activation of persulfate by carbon nanotubes. Water Res 113,
persulfate activation mechanism of spinel oxides. J. Chem. 80–88.
Eng. 425, 130679. Connors, K.A., Amidon, G.L., Stella, V.J., 1986. In Chemical stability
Almomani, F., Shawaqfah, M., Bhosale, R., Kumar, A., 2016. of pharmaceuticals: A handbook of Pharmacists. John Wiley &
Removal of emerging pharmaceuticals from wastewater by Sons, pp. 82–114.
ozone-based advanced oxidation processes. Environ. Prog. Constantin, M.A., Chiriac, F.L., Gheorghe, S., Constantin, L.A., 2022.
Sustain. 35, 982–995. Degradation of carbamazepine from aqueous solutions via
Anipsitakis, G.P., Dionysiou, D.D., Gonzalez, M.A., 2006. TiO2 -assisted photo catalyze. Toxics 10, 168.
Cobalt-mediated activation of peroxymonosulfate and sulfate Council, P.S., 2018. Global Best Practices For Drug Take-Back
radical attack on phenolic compounds. Implications of Programs - Product Stewardship Institute (PSI). Product
chloride ions. Environ Sci Technol 40, 1000–1007. Stewardship Institute (PSI), Webinar.
Ao, X., Zhang, X., Li, S., Yang, Y., Sun, W., Li, Z., 2023. De Laat, J., Le, T.G., 2005. Kinetics and modeling of the Fe(III)/H2 O2
Comprehensive understanding of fluoroquinolone system in the presence of sulfate in acidic aqueous solutions.
degradation via MPUV/PAA process: radical chemistry, matrix Environ. Sci. Technol. 39, 1811–1818.
effects, degradation pathways, and toxicity. J. Hazard. Mater. Deng, C., Ren, C., Wu, F., Deng, N., Glebov, E.M., Pozdnyakov, I.P.,
445, 130480. et al., 2010. Montmorillonite KSF as catalyst for degradation of
Asif, M.B., Kang, H., Zhang, Z., 2023. Assembling CoAl-layered acetaminophen with heterogeneous Fenton reactions. React.
metal oxide into the gravity-driven catalytic membrane for Kinet. Mech. Catal. 100, 277–288.
Fenton-like catalytic degradation of pharmaceuticals and Dong, C., Yi, Q., He, J., Xing, M., Zhang, J., 2023. Evolution of singlet
personal care products. J. Chem. Eng. 463, 142340. oxygen in peroxymonosulfate activation: a review. EES. Catal.
Baum, R.M., 1984. Superoxide theory of oxygen toxicity is center 1, 103–116.
of heated debate. Chem. Eng. News Archive 62, 20–26. Donnachie, R.L., Johnson, A.C., Sumpter, J.P., 2016. A rational
Bergman, Å., Heindel, J.J., Jobling, S., Kidd, K., Zoeller, T.R., 2013. approach to selecting and ranking some pharmaceuticals of
WHO, UN environment, program, geneva, inter-organization concern for the aquatic environment and their relative
programme for the sound management of chemicals, state of importance compared with other chemicals. Environ. Toxicol.
the science of endocrine disrupting chemicals 2012: summary Chem. 35, 1021–1027.
for decision-makers. Duan, P., Qi, Y., Feng, S., Peng, X., Wang, W., Yue, Y., et al., 2020.
Bernhardt, E.S., Rosi, E.J., Gessner, M.O., 2017. Synthetic chemicals Enhanced degradation of clothianidin in
as agents of global change. Front. Ecol. Environ. 15, 84–90. peroxymonosulfate/catalyst system via core-shell FeMn @
Bhandari, R.K., Deem, S.L., Holliday, D.K., Jandegian, C.M., N-C and phosphate surrounding. Appl. Catal. B: Environ. 267,
Kassotis, C.D., Nagel, S.C., et al., 2015. Effects of the 118717.
environmental estrogenic contaminants bisphenol A and Duan, X., Ao, Z., Zhang, H., Saunders, M., Sun, H., Shao, Z., et al.,
17alpha-ethinyl estradiol on sexual development and adult 2018a. Nanodiamonds in sp2/sp3 configuration for radical to
behaviors in aquatic wildlife species. Gen. Comp. Endocrinol. nonradical oxidation: core-shell layer dependence. Appl.
214, 195–219. Catal. B: Environ. 222, 176–181.
journal of environmental sciences 147 (2025) 688–713 709

Duan, X.G., Sun, H.Q., Shao, Z.P., Wang, S.B., 2018b. Nonradical Fe(III)-Montmorillonite Clay. Environ. Sci. Technol. 42,
reactions in environmental remediation processes: 4758–4763.
uncertainty and challenges. Appl. Catal. B: Environ. 224, Gu, C., Liu, C., Ding, Y., Li, H., Teppen, B.J., Johnston, C.T., et al.,
973–982. 2011. Clay mediated route to natural formation of
Eberson, L., 1982. Electron-transfer Reactions in Organic polychlorodibenzo-p-dioxins. Environ. Sci. Technol. 45,
Chemistry. (Academic Press, Gold, V., Bethell, D., (Eds.)), Adv. in 3445–3451.
Phys. Org. Chem. 18, 79–185. Gunnarsson, L., Snape, J.R., Verbruggen, B., Owen, S.F.,
El Najjar, N.H., Touffet, A., Deborde, M., Journel, R., Leitner, N.K.V., Kristiansson, E., Margiotta-Casaluci, L., et al., 2019.
2014. Kinetics of paracetamol oxidation by ozone and Pharmacology beyond the patient – the environmental risks
hydroxyl radicals, formation of transformation products and of human drugs. Environ. Int. 129, 320–332.
toxicity. Sep. Purif. Technol. 136, 137–143. Gusseme, B.D., Vanhaecke, L., Verstraete, W., Boon, N., 2011.
Elovitz, M.S., Shemer, H., Peller, J.R., Vinodgopal, K., Degradation of acetaminophen by Delftia tsuruhatensis and
Sivaganesan, M., Linden, K.G., 2008. Hydroxyl radical rate Pseudomonas aeruginosa in a membrane bioreactor. Water Res
constants: comparing UV/H2 O2 and pulse radiolysis for 45, 1829–1837.
environmental pollutants. J. Water Supply: Res. Technol.-Aqua Hayat, W., Zhang, Y., Shaobin, H., Huang, R., 2021. Insight into the
57, 391–401. degradation of methomyl in water by peroxymonosulfate. J.
Fan, J., Qin, H., Jiang, S., 2019. Mn-doped g-C3 N4 composite to Environ. Chem. Eng. 9, 105358.
activate peroxymonosulfate for acetaminophen degradation: Hayyan, M., Hashim, M.A., AlNashef, I.M., 2016. Superoxide Ion:
the role of superoxide anion and singlet oxygen. J. Chem. Eng. generation and chemical Implications. Chem. Rev. 116,
359, 723–732. 3029–3085.
Flores, J., Moya, P., Bosca, F., Marin, M.L., 2023. Photoreactivity of He, X., Mezyk, S.P., Michael, I., Fatta-Kassinos, D., Dionysiou, D.D.,
new rose bengal-SiO2 heterogeneous photocatalysts with and 2014. Degradation kinetics and mechanism of β-lactam
without a magnetite core for drug degradation and antibiotics by the activation of H2 O2 and Na2 S2 O8 under
disinfection. Catal. Today 413-415, 113994. UV-254nm irradiation. J. Hazard. Mater. 279, 375–383.
Frimer, A.A., 1979. The reaction of singlet oxygen with olefins: the He, Y., Grieser, F., Ashokkumar, M., 2011. Kinetics and mechanism
question of mechanism. Chem. Rev. 79, 359–387. for the sonophotocatalytic degradation of p-chlorobenzoic
Frimer, A.A., Farkash-Solomon, T., Aljadeff, G., 1986. Mechanism acid. J. Phys. Chem. A 115, 6582–6588.
of the superoxide anion radical (O2 -) mediated oxidation of Honarmandrad, Z., Sun, X., Wang, Z., Naushad, M., Boczkaj, G.,
diarylmethanes. J. Org. Chem. 51, 2093–2098. 2023. Activated persulfate and peroxymonosulfate based
Fu, Y., Huang, T., Zhang, L., Zhu, J., Wang, X., 2015. Ag/g-C3 N4 advanced oxidation processes (AOPs) for antibiotics
catalyst with superior catalytic performance for the degradation – a review. Water Resour. Ind. 29, 100194.
degradation of dyes: a borohydride-generated superoxide Hu, Y., Chen, D., Zhang, R., Ding, Y., Ren, Z., Fu, M., et al., 2021.
radical approach. Nanoscale 7, 13723–13733. Singlet oxygen-dominated activation of peroxymonosulfate
Fuller, T., 2008. Proton Exchange Membrane Fuel Cells. The by passion fruit shell derived biochar for catalytic degradation
Electrochem. Soc., p. 8. of tetracycline through a non-radical oxidation pathway. J.
Gao, L., Guo, Y., Zhan, J., Yu, G., Wang, Y., 2022. Assessment of the Hazard. Mater. 419, 126495.
validity of the quenching method for evaluating the role of Huang, D., Xu, W., Lei, L., Chen, S., Lai, C., Zhou, W., et al., 2022.
reactive species in pollutant abatement during the Promoted generation strategies and corresponding roles of
persulfate-based process. Water Res 221, 118730. singlet oxygen in activation of persulfate by nanoscale
Gao, P., Tian, X., Nie, Y., Yang, C., Zhou, Z., Wang, Y., 2018. zero-valent iron systems. J. Chem. Eng. 449, 137493.
Promoted peroxymonosulfate activation into singlet oxygen Huang, K.-C., Couttenye, R.A., Hoag, G.E., 2002. Kinetics of
over perovskite for ofloxacin degradation by controlling the heat-assisted persulfate oxidation of methyl tert-butyl ether
oxygen defect concentration. J. Chem. Eng. 359, 828–839. (MTBE). Chemosphere 49, 413–420.
Gao, Y., Chen, Z., Zhu, Y., Li, T., Hu, C., 2020. New Insights into the Huber, M.M., Canonica, S., Park, G.Y., Gunten, U.V., 2003. Oxidation
generation of singlet oxygen in the metal-free of pharmaceuticals during ozonation and advanced oxidation
peroxymonosulfate activation process: important role of processes. Environ. Sci. Technol. 37, 1016–1024.
electron-deficient carbon atoms. Environ. Sci. Technol. 54, Huber, M.M., GÖbel, A., Joss, A., Hermann, N., LÖffler, D.,
1232–1241. McArdell, C.S., et al., 2005. Oxidation of pharmaceuticals
Ghanbari, F., Giannakis, S., Lin, K.Y.A., Wu, J., Madihi-Bidgoli, S., during ozonation of municipal wastewater effluents: A pilot
2021. Acetaminophen degradation by a synergistic peracetic study. Environ. Sci. Technol. 39, 4290–4299.
acid/UVC-LED/Fe(II) advanced oxidation process: kinetic Humayun, S., Hayyan, M., Alias, Y., Hayyan, A., 2021. Oxidative
assessment, process feasibility and mechanistic degradation of acetaminophen using superoxide ion
considerations. Chemosphere 263, 128119. generated in ionic liquid/aprotic solvent binary system. Sep.
Ghanbari, F., Moradi, M., 2017. Application of peroxymonosulfate Purif. Technol. 270, 118730.
and its activation methods for degradation of environmental Iwanowicz, L.R., Blazer, V.S., Pinkney, A.E., Guy, C.P., Major, A.M.,
organic pollutants. Review. J. Chem. Eng. 310, 41–62. Munney, K., et al., 2016. Evidence of estrogenic endocrine
Ghauch, A., Tuqan, A.M., 2012. Oxidation of bisoprolol in heated disruption in smallmouth and largemouth bass inhabiting
persulfate/H2 O systems: kinetics and products. J. Chem. Eng. Northeast U.S. national wildlife refuge waters: a
183, 162–171. reconnaissance study. Ecotoxicol. Environ. Saf. 124,
Gonzaga, I.M.D., Almeida, C.V.S., Mascaro, L.H., 2023. A critical 50–59.
review of photo-based advanced oxidation processes to Jeong, J., Song, W., Cooper, W.J., Jung, J., Greaves, J., 2010.
pharmaceutical degradation. Catalysts 13, 221. Degradation of tetracycline antibiotics: mechanisms and
Govindan, K., Kim, D.G., Ko, S.O., 2022. Catalytic oxidation of kinetic studies for advanced oxidation/reduction processes.
acetaminophen through pristine and surface-modified Chemosphere 78, 533–540.
nitrogen-doped carbon-nanotube-catalyzed peroxydisulfate Jiang, Y., Liu, Q., Tan, K.M., Wang, F., Ng, H.Y., 2021. Insights into
activation. J. Environ. Chem. Eng. 10, 108257. mechanisms, kinetics and pathway of continuous visible-light
Gu, C., Li, H., Teppen, B.J., Boyd, S.A., 2008. photodegradation of PPCPs via porous g-C3 N4 with highly
Octachlorodibenzodioxin formation on dispersed Fe(III) active sites. J. Chem. Eng. 423, 130095.
710 journal of environmental sciences 147 (2025) 688–713

Jin, H., Li, L., Luo, N., Niu, H., Han, J., Xu, L., et al., 2023a. theoretical calculation and toxicity assessment. J. Hazard.
Biochars-Fex PCu hybrides deriving from solid waste and Mater. 416, 126250.
waste acids for elimination of refractory organic pollutants Li, Y., Niu, J., Yin, L., Wang, W., Bao, Y., Chen, J., et al., 2011.
from pharmaceutical wastewater. J. Chem. Eng. 465, Photocatalytic degradation kinetics and mechanism of
142727 . pentachlorophenol based on Superoxide radicals. J. Environ.
Jin, L., You, S., Ren, N., Liu, Y., 2023b. Selective activation of Sci. 23, 1911–1918.
peroxymonosulfate to singlet oxygen by engineering oxygen Li, Y., Pan, Y., Lian, L., Yan, S., Song, W., Yang, X., 2017.
vacancy defects in Ti3 CNTx MXene for effective removal of Photosensitized degradation of acetaminophen in natural
micropollutants in water. Fundam. Res. 3, 770–776. organic matter solutions: the role of triplet states and oxygen.
Johnson, G.L., 1988. Autoxidation and antioxidants. In: Water Res 109, 266–273.
Swarbrick, J.B. (Ed.), Encyclopedia of pharmaceutical Liang, C., Lan, Z., Zhang, X., Liu, Y., 2016. Mechanism for the
technology. Marcel Dekker, New York, pp. 415–449. primary transformation of acetaminophen in a soil/water
Kearns, D.R., 1971. Physical and chemical properties of singlet system. Water Res 98, 215–224.
molecular oxygen. Chem. Rev. 71, 395–427. Lin, Y., Mo, X., Zhang, Y., Nie, M., Yan, C., Wu, L., 2021. Selective
Kim, H., Kim, W., Mackeyev, Y., Lee, G.S., Kim, H.J., Tachikawa, T., degradation of acetaminophen from hydrolyzed urine by
et al., 2012. Selective oxidative degradation of organic peroxymonosulfate alone: performances and mechanisms.
pollutants by singlet oxygen-mediated photosensitization: tin RSC Adv 11, 40022–40032.
porphyrin versus C-60 aminofullerene systems. Environ. Sci. Lu, J., Huang, Q., 2009. Removal of acetaminophen using
Technol. 46, 9606–9613. enzyme-mediated oxidative coupling processes: II.
Kim, J., Zhang, T., Liu, W., Du, P., Dobson, J.T., Huang, C.H., 2019. Cross-coupling with natural organic matter. Environ. Sci.
Advanced oxidation process with peracetic acid and Fe(II) for Technol. 43, 7068–7073.
contaminant degradation. Environ. Sci. Technol. 53, Luo, R., Li, M., Wang, C., Zhang, M., Nasir Khan, M.A., Sun, X., et al.,
13312–13322. 2019. Singlet oxygen-dominated non-radical oxidation
Kohantorabi, M., Moussavi, G., Mohammadi, S., Oulego, P., process for efficient degradation of bisphenol A under high
Giannakis, S., 2021. Photocatalytic activation of salinity condition. Water Res 148, 416–424.
peroxymonosulfate (PMS) by novel mesoporous Lutze, H.V., Kerlin, N., Schmidt, T., 2015a. Sulfate radical-based
Ag/ZnO@NiFe2 O4 nanorods, inducing radical-mediated water treatment in presence of chloride: formation of
acetaminophen degradation under UVA irradiation. chlorate, inter-conversion of sulfate radicals into hydroxyl
Chemosphere 277, 130271. radicals and influence of bicarbonate. Water Res 72,
Kolpin, D.W., Furlong, E.T., Meyer, M.T., Thurman, E.M., Zaugg, S.D., 349–360.
Barber, L.B., et al., 2002. Pharmaceuticals, hormones, and other Luo, S., Wei, Z., Dionysiou, D.D., Spinney, R., Hu, W.P., Chai, L., et al.,
organic wastewater contaminants in U.S. streams, 1999−2000: 2017. Mechanistic insight into reactivity of sulfate radical with
a national reconnaissance. Environ. Sci. Technol. 36, aromatic contaminants through single-electron transfer
1202–1211. pathway. J. Chem Eng. 327, 1056–1065.
Küster, A., Adler, N., 2014. Pharmaceuticals in the environment: Lutze, H.V., Bircher, S., Rapp, I., Kerlin, N., Bakkour, R., Geisler, M.,
scientific evidence of risks and its regulation. Philos. Trans. R. et al., 2015b. Degradation of chlorotriazine pesticides by
Soc. B: Biol. Sci. 369, 20130587. sulfate radicals and the influence of organic matter. Environ.
Langford, K.H., Thomas, K.V., 2009. Determination of Sci. Technol. 49, 1673–1680.
pharmaceutical compounds in hospital effluents and their Madhavan, V., Levanon, H., Neta, P., 1978. Decarboxylation by
contribution to wastewater treatment works. Environ. Int. 35, SO4 . − radicals. Radiat. Res. 76, 15.
766–770. Malefane, M.E., Mafa, P.J., Nkambule, T.T.I., Managa, M.E.,
Larsson, D.G.J., 2014. Pollution from drug manufacturing: review Kuvarega, A.T., 2023. Modulation of Z-scheme photocatalysts
and perspectives. Philos. Trans. R. Soc. B: Biol. Sci. 369, for pharmaceuticals remediation and pathogen inactivation:
20130571. design devotion, concept examination, and developments. J.
Le, T.X.H., Nguyen, T.V., Yacouba, Z.A., Zoungrana, L., Avril, F., Chem. Eng. 452, 138894.
Nguyen, D.L., et al., 2017. Correlation between degradation Min, D.B., Boff, J.M., 2002. Chemistry and reaction of singlet
pathway and toxicity of acetaminophen and its by-products oxygen in foods. Compr. Rev. Food Sci. Food Saf.
by using the electro-Fenton process in aqueous media. 1, 58–72.
Chemosphere 172, 1–9. Moctezuma, E., Leyva, E., Aguilar, C.A., Luna, R.A., Montalvo, C.,
Lee, J., Gunten, U.V., Kim, J.H., 2020. Persulfate-based advanced 2012. Photocatalytic degradation of paracetamol:
oxidation: critical assessment of opportunities and intermediates and total reaction mechanism. J. Hazard. Mater.
roadblocks. Environ. Sci. Technol. 54, 3064–3081. 243, 130–138.
Letsinger, S., Kay, P., Rodríguez-Mozaz, S., Villagrassa, M., Mohapatra, S., Snow, D., Shea, P., Gálvez-Rodríguez, A., Kumar, M.,
Barceló, D., Rotchell, J.M., 2019. Spatial and temporal Padhye, L.P., et al., 2023. Photodegradation of a mixture of five
occurrence of pharmaceuticals in UK estuaries. Sci.Total pharmaceuticals commonly found in wastewater:
Environ 678, 74–84. experimental and computational analysis. Environ. Res. 216,
Li, B., Ma, X., Deng, J., Li, Q., Chen, W., Li, G., et al., 2020a. 114659.
Comparison of acetaminophen degradation in UV-LED-based Moore, S.C., Matthews, C.E., Ou Shu, X., Yu, K., Gail, M.H., Xu, X.,
advance oxidation processes: reaction kinetics, radicals et al., 2016. Endogenous estrogens, estrogen metabolites, and
contribution, degradation pathways and acute toxicity breast cancer risk in postmenopausal Chinese women. JNCI
assessment. Sci. Total Environ. 723, 137993. 108, djw103.
Li, M., Sun, J., Han, D., Wei, B., Mei, Q., An, Z., et al., 2020b. Morris, J.J., Rose, A.L., Lu, Z., 2022. Reactive oxygen species in the
Theoretical investigation on the contribution of HO, SO4 − and world ocean and their impacts on marine ecosystems. Redox
CO3 − radicals to the degradation of phenacetin in water: Biol 52, 102285.
mechanisms, kinetics, and toxicity evaluation. Ecotoxicol. Mu, R., Ao, Y., Wu, T., Wang, C., Wang, P., 2020. Synthesis of novel
Environ. Saf. 204, 110977. ternary heterogeneous anatase-TiO2 (B) biphase
Li, M., Sun, J., Mei, Q., Wei, B., An, Z., Cao, H., et al., 2021. nanowires/Bi4 O5 I2 composite photocatalysts for the highly
Acetaminophen degradation by hydroxyl and organic radicals efficient degradation of acetaminophen under visible light
in the peracetic acid-based advanced oxidation processes: irradiation. J. Hazard. Mater. 382, 121083.
journal of environmental sciences 147 (2025) 688–713 711

Nakarada, Ð., Petković, M., 2018. Mechanistic insights on how Rapti, I., Kourkouta, T., Malisova, E.M., Albanis, T., Konstantinou, I.,
hydroquinone disarms OH and OOH radicals. Int. J. Quantum 2023. Photocatalytic degradation of inherent pharmaceutical
Chem. 118, e25496. concentration levels in real hospital WWTP effluents using
Nelles, J.L., Hu, W.Y., Prins, G.S., 2011. Estrogen action and prostate g-C3 N4 catalyst on CPC pilot scale reactor. Molecules 28, 1170.
cancer. Expert Rev. Endocrinol. Metab. 6, 437–451. Rayaroth, M.P., Aravind, U.K., Boczkaj, G., Aravindakumar, C.T.,
Neta, P., Huie, R.E., Ross, A.B., 1988. Rate constants for reactions of 2023. Singlet oxygen in the removal of organic pollutants: an
inorganic radicals in aqueous solution (Quarterly report). updated review on the degradation pathways based on mass
JPCRD 17, 1027–1284. spectrometry and DFT calculations. Chemosphere 345, 140203.
Neta, P., Madhavan, V., Zemel, H., Fessenden, R.W., 1977. Rate Razavi, B., Abdelmelek, S.B., Song, W., O’Shea, K.E., Cooper, W.J.,
constants and mechanism of reaction of sulfate radical 2011. Photochemical fate of atorvastatin (lipitor) in simulated
anion with aromatic compounds. J. Am. Chem. Soc. 99, natural waters. Water Res 45, 625–631.
163–164. Reshetnyak, O.V., Koval’chuk, E.P., Skurski, P., Rak, J.,
Nolte, T., Peijnenburg, W., 2018. Use of quantum-chemical Błażejowski, J., 2003. The origin of luminescence
descriptors to analyze reaction rate constants between accompanying electrochemical reduction or chemical
organic chemicals and superoxide/hydroperoxyl (O2 ·− /HO2 · ). decomposition of peroxydisulfates. J. Lumin. 105, 27–34.
Free Radic. Res. 52, 1–411. Roberts, P.H., Thomas, K.V., 2006. The occurrence of selected
Oh, W.-D., Dong, Z., Lim, T., 2016. Generation of sulfate radical pharmaceuticals in wastewater effluent and surface waters of
through heterogeneous catalysis for organic contaminants the lower Tyne catchment. Sci. Total Environ. 356, 143–153.
removal: current development, challenges and prospects. Rodríguez-Narvaez, O.M., Rajapaksha, R.D., Ranasinghe, M.I.,
Appl. Catal. B - Environ. 194, 169–201. Bai, X., Peralta-Hernández, J.M., Bandala, E.R., 2020.
Papac, J., Ballesteros, S.G., Tonkovic, S., Kovacic, M., Tomic, A., Peroxymonosulfate decomposition by homogeneous and
Cvetnić, M., et al., 2023. Degradation of pharmaceutical heterogeneous Co: kinetics and application for the
memantine by photo-based advanced oxidation processes: degradation of acetaminophen. J. Environ. Sci. 93, 30–40.
kinetics, pathways and environmental aspects. J. Environ. Sarker, M.A.R., Ahn, Y.H., 2023. Strategic insight into enhanced
Chem. Eng. 11, 109334. photocatalytic remediation of pharmaceutical contaminants
Papageorgiou, M., Kosma, C., Lambropoulou, D., 2016. Seasonal using spherical CdO nanoparticles in visible light region.
occurrence, removal, mass loading and environmental risk Chemosphere 311, 137040.
assessment of 55 pharmaceuticals and personal care products Sawyer, D.T., Valentine, J.S., 1981. How super is superoxide? Acc.
in a municipal wastewater treatment plant in Central Greece. Chem. Res. 14, 393–400.
Sci. Total Environ. 543, 547–569. Scaria, J., Nidheesh, P.V., 2023. Pre-treatment of real
Pérez-Estrada, L.A., Malato, S., Gernjak, W., Agüera, A., pharmaceutical wastewater by heterogeneous Fenton and
Thurman, E.M., Ferrer, I., et al., 2005. Photo-Fenton degradation persulfate oxidation processes. Environ. Res. 217, 114786.
of diclofenac: identification of main intermediates and Schreiber, R., Gündel, U., Franz, S., Küster, A., Rechenberg, B.,
degradation pathway. Environ. Sci. Technol. 39, 8300–8306. Altenburger, R., 2011. Using the fish plasma model for
Pérez-Lucas, G., Aatik, A.E., Aliste, M., Navarro, G., Fenoll, J., comparative hazard identification for pharmaceuticals in the
Navarro, S., 2023. Removal of contaminants of emerging environment by extrapolation from human therapeutic data.
concern from a wastewater effluent by solar-driven RTP 61, 261–275.
heterogeneous photocatalysis: a case study of Schweitzer, C., Schmidt, R., 2003. Physical mechanisms of
pharmaceuticals. Water, Air, Soil Pollut 234, 55. generation and deactivation of singlet oxygen. Chem. Rev. 103,
Phillips, P.J., Smith, S.G., Kolpin, D.W., Zaugg, S.D., Buxton, H.T., 1685–1757.
Furlong, E.T., et al., 2010. Pharmaceutical formulation facilities Scully, F.E., Hoigné, J., 1987. Rate constants for reactions of singlet
as sources of opioids and other pharmaceuticals to oxygen with phenols and other compounds in water.
wastewater treatment plant effluents. Environ. Sci. Technol. Chemosphere 16, 681–694.
44, 4910–4916. Shao, P., Yin, X., Yu, C., Han, S., Zhao, B., Li, K., et al., 2023.
Piccirillo, G., Aroso, R.T., Rodrigues, F.M.S., Carrilho, R.M.B., Enhanced activation of peroxymonosulfate via sulfate
Pinto, S.M.A., Calvete, M.J.F, et al., 2021. Oxidative degradation radicals and singlet oxygen by SrCox Mn1-x O3 perovskites for
of pharmaceuticals: the role of tetrapyrrole-based catalysts. the degradation of Rhodamine B. Processes 11, 1279.
Catalysts 11, 1335. Singh, S., Sharma, N., Sehrawat, P., Kansal, S.K., 2023.
Pignatello, J. J., Oliveros, E., MacKay, A., 2006. Advanced oxidation Solar-light-driven photocatalytic degradation of
processes for organic contaminant destruction based on the pharmaceutical pollutants utilizing 2D g-C(3)N(4)/BiOCl
Fenton reaction and related chemistry. Crit. Rev. Environ. Sci. composite. Environ. Toxicol. Pharmacol. 99, 104110.
Technol. 36, 1–84. Sonntag, C., 2006. Free-radical-induced DNA Damage and Its
Polubesova, T., Eldad, S., Chefetz, B., 2010. Adsorption and repair: A chemical Perspective, 1 ed. Springer-Verlag, Berlin
oxidative transformation of phenolic acids by Heidelberg.
Fe(III)-montmorillonite. Environ. Sci. Technol. 44, 4203–4209. Sonntag, C.v., Dowideit, P., Fang, X., Mertens, R., Pan, X.,
Qi, Y., Ge, B., Zhang, Y., Jiang, B., Wang, C., Akram, M., et al., 2020. Schuchmann, M.N., et al., 1997. The fate of peroxyl radicals in
Three-dimensional porous graphene-like biochar derived aqueous solution. Water Sci. Technol. 35, 9–15.
from Enteromorpha as a persulfate activator for Tan, C., Sheng, T., Xu, Q., Xu, T., Sun, K., Deng, L., et al., 2021a.
sulfamethoxazole degradation: role of graphitic N and Cobalt doped iron oxychloride as efficient heterogeneous
radicals transformation. J. Hazard. Mater. 399, 123039. Fenton catalyst for degradation of paracetamol and
Qi, Y., Zhou, X., Li, Z., Yin, R., Qin, J., Li, H., et al., 2022. phenacetin. Chemosphere 263, 127989.
Photo-induced holes initiating peroxymonosulfate oxidation Tan, Y., Zheng, S., Di, Y., Li, C., Bian, R., Sun, Z., 2021b. Diatomite
for carbamazepine degradation via singlet oxygen. Catalysts supported nano zero valent iron with 3D network for
12, 1327. peroxymonosulfate activation in efficient degradation of
Qin, C., Troya, D., Shang, C., Hildreth, S., Helm, R., Xia, K., 2015. bisphenol A. J. Mater. Sci. Technol. 95, 57–69.
Surface catalyzed oxidative oligomerization of 17β-Estradiol Tanner, D.D., Osman, S.A.A., 1987. Oxidative decarboxylation. On
by Fe3+ -saturated montmorillonite. Environ. Sci. Technol. 49, the mechanism of the potassium persulfate-promoted
956–964. decarboxylation reaction. J. Org. Chem. 52, 4689–4693.
712 journal of environmental sciences 147 (2025) 688–713

Ternes, T.A., 1998. Occurrence of drugs in German sewage kinetic model and degradation pathways. J. Chem. Eng. 358,
treatment plants and rivers. Water Res 32, 3245–3260. 1091–1100.
Thomas, M.J., Foote, C.S., 1978. Chemistry of Singlet Wang, W., Chen, M., Wang, D., Yan, M., Liu, Z., 2021. Different
Oxygen—Xxvi. Photooxygenation of phenols. Photochem. activation methods in sulfate radical-based oxidation for
Photobiol. 27, 683–693. organic pollutants degradation: catalytic mechanism and
Tong, Y., Wang, X., Sun, Z., Gao, J., 2021. Two transformation toxicity assessment of degradation intermediates. Sci. Total
pathways of Acetaminophen with Fe3+ saturated clay Environ. 772, 145522.
particles in dark or light. Chemosphere 278, 130399. Wang, X., Brigante, M., Dong, W., Wu, Z., Mailhot, G., 2020a.
Tsitonaki, A., Petri, B., Crimi, M., MosbÆK, H., Siegrist, R.L., Degradation of Acetaminophen via UVA-induced advanced
Bjerg, P.L., 2010. In situ chemical oxidation of contaminated oxidation processes (AOPs). Involvement of different radical
soil and groundwater using persulfate: a review. Crit. Rev. species: HO, SO4 − and HO2 /O2 − . Chemosphere 258, 127268.
Environ. Sci. Technol. 40, 55–91. Wang, Y., Liu, M., Zhao, X., 2017a. Insights into heterogeneous
Usman, M., Monfort, O., Gowrisankaran, S., Hameed, B.H., catalytic activation of peroxymonosulfate by Pd/g-C3 N4 : the
Hanna, K., Al-Abri, M., 2023. Dual functional materials capable role of superoxide radical and singlet oxygen. Catal. Commun.
of integrating adsorption and Fenton-based oxidation 102, 85–88.
processes for highly efficient removal of pharmaceutical Wang, Z., Ai, L., Huang, Y., Zhang, J., Li, S., Chen, J., et al., 2017b.
contaminants. JWPE 52, 103566. Degradation of azo dye with activated peroxygens: when
Uthirakrishnan, U., Lu, X., Wang, J., Zhang, Z., Dai, J., Tan, Y., et al., zero-valent iron meets chloride. RSC Adv 7, 30941–30948.
2020. Sulfate radicals-based advanced oxidation technology in Wang, Z., Wang, J., Xiong, B., Bai, F., Wang, S., Wan, Y., et al., 2020b.
various environmental remediation: a state-of-the–art review. Application of cobalt/peracetic acid to degrade
J. Chem. Eng. 402, 126232. sulfamethoxazole at neutral condition: efficiency and
Valadez-Renteria, E., Perez-Gonzalez, R., Gomez-Solis, C., mechanisms. Environ. Sci. Technol. 54, 464–475.
Diaz-Torres, L.A., Encinas, A., Oliva, J., et al., 2023. A novel and Wardman, P., 1989. Reduction potentials of one-electron couples
stretchable carbon-nanotube/Ni@TiO2 :w photocatalytic involving free radicals in aqueous solution. J. Phys. Chem. Ref.
composite for the complete removal of diclofenac drug from Data 18, 1637–1755.
the drinking water. J. Environ. Sci. 126, 575–589. Westerling, K., 2013. Boys will be girls: the life-altering effects of
Van, H.T., Nguyen, L.H., Hoang, T.K., Nguyen, T.T., Tran, T.N.H., PPCPs in drinking water. Water Online 8–10.
Nguyen, T.B.H., et al., 2020. Heterogeneous Fenton oxidation of Wilkinson, J.L., Boxall, A.B.A., Kolpin, D.W., Leung, K.M.Y.,
paracetamol in aqueous solution using iron slag as a catalyst: Lai, R.W.S., Galbán-Malagón, C., et al., 2022. Pharmaceutical
degradation mechanisms and kinetics. Environ. Technol. pollution of the world’s rivers. Environ. Sci. 119, e2113947119.
Innov. 18, 100670. Wu, S., Zhang, L., Chen, J., 2012. Paracetamol in the environment
von Gunten, U., 2003. Ozonation of drinking water: part II. and its degradation by microorganisms. Appl. Microbiol.
Disinfection and by-product formation in presence of Biotechnol. 96, 875–884.
bromide, iodide or chlorine. Water Res 37, 1469–1487. Xia, X., Zhu, F., Li, J., Yang, H., Wei, L., Li, Q., et al., 2020. A review
von Sonntag, C., Schuchmann, H.P., 1991. The elucidation of study on sulfate-radical-based advanced oxidation processes
peroxyl radical reactions in aqueous solution with the help of for domestic/industrial wastewater treatment: degradation,
radiation-chemical methods. Angew. Chem. Int. Ed. Engl. 30, efficiency, and mechanism. Front. Chem. 8, 592056.
1229–1253. Xiao, G., Xu, T., Faheem, M., Xi, Y., Zhou, T., Moryani, H.T., et al.,
Vulliet, E., Cren-Olivé, C., 2011. Screening of pharmaceuticals and 2021. Evolution of singlet oxygen by activating peroxydisulfate
hormones at the regional scale, in surface and groundwaters and peroxymonosulfate: a review. Int. J. Environ. Res. Public
intended to human consumption. Environ. Pollut. 159, Health 18, 3344.
2929–2934. Xie, L., Hao, J., Wu, Y., Xing, S., 2022. Non-radical activation of
Waldemer, R.H., Tratnyek, P.G., Johnson, R.L., Nurmi, J.T., 2007. peroxymonosulfate with oxygen vacancy-rich amorphous
Oxidation of chlorinated ethenes by heat-activated MnOX for removing sulfamethoxazole in water. J. Chem. Eng.
persulfate: kinetics and products. Environ. Sci. Technol. 41, 436, 135260.
1010–1015. Xie, Z.H., He, C.S., He, Y.L., Yang, S.R., Yu, S.Y., Xiong, Z., et al., 2023.
Wang, C.-W., Liang, C., 2014. Oxidative degradation of TMAH Peracetic acid activation via the synergic effect of Co and Fe in
solution with UV persulfate activation. J. Chem. Eng. 254, CoFe-LDH for efficient degradation of pharmaceuticals in
472–478. hospital wastewater. Water Res 232, 119666.
Wang, J., Guo, Y., Liu, B., Jin, X., Liu, L., Xu, R., et al., 2011. Detection Xu, T., Chen, J., Chen, X., Xie, H., Wang, Z., Xia, D., et al., 2021.
and analysis of reactive oxygen species (ROS) generated by Prediction models on pKa and base-catalyzed hydrolysis
nano-sized TiO2 powder under ultrasonic irradiation and kinetics of parabens: experimental and quantum chemical
application in sonocatalytic degradation of organic dyes. studies. Environ. Sci. Technol. 55, 6022–6031.
Ultrason. Sonochem. 18, 177–183. Xu, T., Chen, J., Wang, Z., Tang, W., Xia, D., Fu, Z., et al., 2019.
Wang, J., Wang, S., 2020c. Reactive species in advanced oxidation Development of prediction models on base-catalyzed
processes: formation, identification and reaction mechanism. hydrolysis kinetics of phthalate esters with density functional
J. Chem Eng. 401, 126158. theory calculation. Environ. Sci. Technol. 53, 5828–5837.
Wang, J., Wang, M., Kang, J., Tang, Y., Xu, Z., Dong, Q., et al., 2023. Yan, J., Qian, L., Gao, W., Chen, Y., Ouyang, D., Chen, M., 2017.
Sulfamethoxazole degradation by Ni2+ doped Fe2 O3 on a Enhanced Fenton-like degradation of trichloroethylene by
nickel foam in peroxymonosulfate assisting hydrogen peroxide activated with nanoscale zero valent iron
photoelectrochemical oxidation system: performance, loaded on biochar. Sci. Rep. 7, 43051.
mechanism and degradation pathway. Sep. Purif. Technol. 314, Yan, N., Liu, F., Huang, W., 2013. Interaction of oxidants in siderite
123584. catalyzed hydrogen peroxide and persulfate system using
Wang, L., Bian, Z., 2020. Photocatalytic degradation of trichloroethylene as a target contaminant. J. Chem. Eng. 219,
paracetamol on Pd-BiVO(4) under visible light irradiation. 149–154.
Chemosphere 239, 124815. Yang, L., Zhou, Y., Shi, B., Meng, J., He, B., Yang, H., et al., 2020.
Wang, S., Wu, J., Lu, X., Xu, W., Gong, Q., Ding, J., et al., 2019. Anthropogenic impacts on the contamination of
Removal of acetaminophen in the Fe2+ /persulfate system: pharmaceuticals and personal care products (PPCPs) in the
journal of environmental sciences 147 (2025) 688–713 713

coastal environments of the Yellow and Bohai seas. Environ. methylated quinones: performance and mechanism. RSC Adv
Int. 135, 105306. 9, 27224–27230.
Yang, Q., Ma, Y., Chen, F., Yao, F., Sun, J., Wang, S., et al., 2019. Zhang, T., Huang, C.-H., 2020. Modeling the kinetics of
Recent advances in photo-activated sulfate radical-advanced UV/Peracetic acid advanced oxidation process. Environ. Sci.
oxidation process (SR-AOP) for refractory organic pollutants Technol. 54, 7579–7590.
removal in water. J. Chem. Eng. 378, 122149. Zhang, X., Wu, F., Wu, X., Chen, P., Deng, N., 2008.
Yang, X.J., Xu, X.M., Xu, J., Han, Y.F., 2013. Iron oxychloride (FeOCl): Photodegradation of acetaminophen in TiO2 suspended
an efficient Fenton-like catalyst for producing hydroxyl solution. J. Hazard. Mater. 157, 300–307.
radicals in degradation of organic contaminants. J. Am. Chem. Zhang, Y., Fan, J., Yang, B., Huang, W., Ma, L., 2017a.
Soc. 135, 16058–16061. Copper–catalyzed activation of molecular oxygen for
Yang, Y., Banerjee, G., Brudvig, G.W., Kim, J.-H., Pignatello, J.J., 2018. oxidative destruction of acetaminophen: the mechanism and
Oxidation of organic compounds in water by unactivated superoxide-mediated cycling of copper species. Chemosphere
peroxymonosulfate. Environ. Sci. Technol. 52, 5911–5919. 166, 89–95.
You, Q., Zhang, C., Cao, M., Wang, B., Huang, J., Wang, Y., et al., Zhang, Y., Zhang, Q., Dong, Z., Wu, l., Hong, J., 2018. Degradation of
2023. Defects controlling, elements doping, and crystallinity acetaminophen with ferrous/copperoxide activate persulfate:
improving triple-strategy modified carbon nitride for efficient synergism of iron and copper. Water Res 146, 232–243.
photocatalytic diclofenac degradation and H2 O2 production. Zhang, Y., Zhang, Q., Hong, J., 2017b. Sulfate radical degradation of
Appl. Catal. B: Environ. 321, 121941. acetaminophen by novel iron–copper bimetallic oxidation
Yu, D., Wu, F., He, J., Bai, L., Zheng, Y., Wang, Z., et al., 2023. Tuned catalyzed by persulfate: mechanism and degradation
layered double hydroxide-based catalysts inducing singlet pathways. Appl. Surf. Sci. 422, 443–451.
oxygen evolution: reactive oxygen species evolution Zhu, C., Zhang, Y., Fan, Z., Liu, F., Li, A., 2020. Carbonate-enhanced
mechanism exploration, norfloxacin degradation and catalytic activity and stability of Co(3)O(4) nanowires for
catalysts screen based on machine learning. Appl. Catal. B: (1)O(2)-driven bisphenol A degradation via
Environ. 320, 121880. peroxymonosulfate activation: critical roles of electron and
Yu, X., Sui, Q., Lyu, S., Zhao, W., Liu, J., Cai, Z., et al., 2020. Municipal proton acceptors. J. Hazard. Mater. 393, 122395.
solid waste landfills: an underestimated source of Zhou, Y., Jiang, J., Gao, Y., Ma, J., Pang, S.Y., Li, J., et al., 2015.
pharmaceutical and personal care products in the water Activation of peroxymonosulfate by benzoquinone: a novel
environment. Environ. Sci. Technol. 54, 9757–9768. nonradical oxidation process. Environ. Sci. Technol. 49,
Zare, M., Alfonso-Muniozguren, P., Bussemaker, M.J., Sears, P., 12941–12950.
Serna-Galvis, E.A., Torres-Palma, R.A., et al., 2023. A Zhu, S., Li, X., Kang, J., Duan, X., Wang, S., 2019. Persulfate
fundamental study on the degradation of paracetamol under activation on crystallographic manganese oxides: mechanism
single- and dual-frequency ultrasound. Ultrason. Sonochem. of singlet oxygen evolution for nonradical selective
94, 106320. degradation of aqueous contaminants. Environ. Sci. Technol.
Zhang, B.-T., Zhang, Y., Teng, Y.-G., Fan, M., 2015. Sulfate radical 53, 307–315.
and its application in decontamination technologies. Crit. Rev. Zilberman, A., Gozlan, I., Avisar, D., 2023. Pharmaceutical
Environ. Sci. Technol. 45, 1756–1800. transformation products formed by ozonation - does
Zhang, H., Qiao, L., He, J., Li, N., Zhang, D., Yu, K., et al., 2019. degradation occur? Molecules 28, 1227.
Activating peroxymonosulfate by halogenated and

You might also like