Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Environmental Technology & Innovation 35 (2024) 103687

Contents lists available at ScienceDirect

Environmental Technology & Innovation


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

Overlooked effects of chlorides and bicarbonates on the intensity


of peroxydisulfate activation in Fe(II)/citric acid-S2O2−
8 process

Radek Škarohlíd a, Doris Kraljič a, Jan Suchan a, Karel Kühnl a, Barbora Hanzlová a,
Pavlína Těšínská a, Marina Kholomyeva a, Marek Martinec a, Alena Michalcová b,
Lenka McGachy a, *, 1
a
Department of Environmental Chemistry, University of Chemistry and Technology Prague, Technická 5, Prague 16628, Czech Republic
b
Department of Metals and Corrosion Engineering, University of Chemistry and Technology Prague, Technická 5, Prague 16628, Czech Republic

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

Keywords: The influence of chlorides (Cl− ) and bicarbonates (HCO−3 ) on the Fe(II)/citric acid peroxydisulfate
AOPs based (Fe(II)/CA-S2O2−8 ) process has been commonly attributed to the formation of secondary
Chlorides reactive species, such as Cl• and CO•3, and their different reactivity with the target pollutant
Bicarbonates
compared to primary reactive species (e.g., SO•− 4 ). However, this conclusion has been entirely
Perchloroethylene
Ferrihydrite
based only on analyzing target pollutants removal without the context of S2O2− 8 consumption and

Nanoparticles stabilization dissolved Fe concentration throughout the process. To address this knowledge gap, we conducted
a series of batch experiments to investigate the influence of Cl− (2 mmol•L-1) and HCO−3
(1 mmol•L-1) on the intensity of S2O2− 2−
8 activation within the Fe(II)/CA-S2O8 process with tet­
rachloroethene (PCE) as the target pollutant. Throughout the experiments, PCE removal effi­
ciency, S2O2−
8 consumption, pH, redox potential, and dissolved Fe were analyzed. Equilibrium
hydrochemical modelling (Visual MINTEQ 4) and high-resolution transmission electron micro­
scopy (TEM) were employed to analyze and interpret the data obtained. We found that both Cl−
and HCO−3 significantly affected the intensity of S2O2− 8 activation. While Cl influenced the

steady-state dissolved Fe(II) concentration, HCO3 primarily increased solution pH, which led to

the formation of CA-stabilized ferrihydrite nanoparticles. This resulted in the inaccessibility of


dissolved Fe(III) to the Fe(III)/Fe(II) regeneration cycle, consequently suppressing S2O2− 8 acti­
vation intensity and PCE removal. Overall, the findings in this study have deepened the funda­
mental knowledge of how Cl− and HCO−3 influence the Fe(II)/CA-S2O2− 8 process, which may help
improve the design and operation of S2O2− 8 -based advanced oxidation processes using the Fe(II)/
CA activation method.

1. Introduction

Peroxydisulfate anion (S2O2− 8 ) is a relatively strong oxidizing agent (Wang and Wang, 2022). However, direct two-electron
oxidation of most organic compounds in water is generally slow at laboratory temperature (Behrman, 2004). To enhance the
oxidation process, various S2O2−
8 activation methods, including physical (e.g., thermal) and chemical (e.g., homogenous/heterogenous

* Corresponding author.
E-mail address: lenka.mc.gachy@vscht.cz (L. McGachy).
1
ORCID: 0000-0002-2251-1798

https://doi.org/10.1016/j.eti.2024.103687
Received 25 January 2024; Received in revised form 7 May 2024; Accepted 21 May 2024
Available online 22 May 2024
2352-1864/© 2024 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

activators/catalysts) activations or their combinations, are employed (Liu et al., 2023; Oh and Lim, 2019; Wen et al., 2023; Xie et al.,
2023; Zhao et al., 2017). Based on the activation method used and reaction conditions (e.g., pH, temperature), this step typically
initiates a chain of radical reactions involving predominantly sulfate radical (SO•− 4 ) and/or hydroxyl radical ( OH).

Despite all the recent advances in heterogeneous catalysis, homogeneous activation using Fe(II) ions (Fe(II)-S2O2− 8 ) remains a
popular choice (Faggiano et al., 2023; Yao et al., 2023; Yoon et al., 2022) within practitioners due to the low cost and easy availability
of iron salts, nonexistent problematic interphase mass transfer resistance, and the overall simplicity of the activation process compared
to heterogeneous catalysts (Karim et al., 2021).
The Fe(II)-S2O2− 8 process consists primary in the formation of SO4 (Eq. (1)), (Fordham and Williams, 1951)) and ferryl-ion species
•−

(Fe(IV)) (Eq. (2)), (Wang et al., 2018)), secondary in forming •OH via Eqs. (3)-(4) (Herrmann et al., 1995) and other reactive species,
such as superoxide (O•− 2 ) (Eq. (5) (Tang et al., 2023)), peroxydisulfate radical (S2O8 ) (Eqs. (6)-(9) (Wang et al., 2019; Zhu et al.,
•−

2018b, 2018a)) and singlet oxygen (1O2) (Eqs. (10) and (11) (Ding et al., 2021; Jiang et al., 2020)). The resulting composition of the
reactive species mix is then dependent on many factors, for example, the S2O2− 8 /Fe(II) initial molar ratio, pH, nature of contaminants,
and presence of aqueous matrix constituents (Dong et al., 2020; Hou et al., 2023; Rao et al., 2022). Although the efficiency of the Fe
(II)-S2O2−
8 process for contaminants removal has been demonstrated in a variety of studies (Cai et al., 2021; Faggiano et al., 2023; Lu
et al., 2023), it still faces some challenges: (i) limited Fe(II) regeneration capability (Eq. (9)) leading to significant suppression of Eqs.
(1) and (2) and, thus, halting the formation of reactive species; (ii) high concentrations of Fe(II) are required to fully activate S2O2− 8 ,
which in turn leads to scavenging of reactive species by excessive Fe(II) via Eqs. (12)–(14) (Buxton et al., 1988; Pestovsky and Bakac,
2006); (iii) when pH > 3, Fe(III) precipitates in the form of oxyhydroxides, which leads to decreased concentrations of total dissolved
iron (Gao et al., 2020; Karim et al., 2021).

Fe2+ + S2 O2−
8 →Fe
3+
+ SO•−4 + SO2−
4 k = 91M− 1 ⋅s− 1
(1)

Fe2+ + S2 O2− IV 2+
8 + H2 O→Fe O + 2SO2−
4 +H
+
k = 20M− 1 ⋅s− 1
(2)

SO•−4 + H2 O→HSO−4 + • OH k < 60M− 1 ⋅s− 1


(3)

SO•−4 + OH− →SO2−


4 + OH

k = 1.4⋅107 M− 1 ⋅s− 1
(4)

O2 + e− →O•−2 (5)

SO•−4 +S2 O2− 2−


8 →S2 O8 + SO4
•−
k = 6.3⋅105 M− 1 ⋅s− 1
(6)


OH+S2 O2−
8 →S2 O8 + OH
•− −
k = 1.4⋅107 M− 1 · s− 1
(7)

FeIV O2+ +S2 O2− + •− 2−


8 + H →S2 O8 + SO4 + OH

k = 104 M− 1 ⋅s− 1
(pH = 1) (8)

Fe3+ +S2 O2−


8 →S2 O8 + Fe
•− 2+
k = 6.6⋅10− 2 M− 1 ⋅s− 1
(9)

O•−2 + • OH→OH− + 1 O2 (10)

2O•−2 + 2H+ →H2 O2 + O12 (11)

SO•−4 + Fe2+ →Fe3+ + SO2−


4 k = 4.6⋅109 M− 1 ⋅s− 1
(12)


OH + Fe2+ →Fe3+ + OH− k = 2.8⋅108 M− 1 ⋅s− 1
(13)

FeIV O2+ + Fe2+ + H2 O→Fe3+ + 2OH− k = 3.6⋅104 M− 1 ⋅s− 1


(pH = 1) (14)

To overcome Fe(II)-S2O2− 8 limitations, iron-chelators such as polycarboxylates (e.g., citric acid (CA), tartaric acid, oxalic acid (Chen
et al., 2023; Han et al., 2015)), have been successfully implemented. Due to its biodegradability, environmental friendliness, and
ability to promote Fe(III)/Fe(II) cycle regeneration through radical chain reactions, CA has been widely investigated as a chelating
agent in the Fe(II)-S2O2− 2−
8 process (hereafter denoted as Fe(II)/CA-S2O8 ) (Table S1 of the Supporting Information (SI)). CA promotes Fe
(III)/Fe(II) cycle regeneration through radical chain reactions (Eqs. (15)-(21)) with the crucial role of carboxyl radicals (•CO−2 ), alkyl
radicals (•R), and O•−
2 (Wang et al., 2019; Zhang and Zhou, 2019), as well as significantly alters the nature of formed primary reactive
species from Fe(IV) to SO•−4 in the activation process (Wang et al., 2019).

SO•−4 + R − COOH→SO2−
4 + O2 C/ R + intermediates
− •
(15)


OH + R − COOH→OH− + O−2 C/• R + intermediates (16)

FeIV O2+ + R − COOH→Fe3+ + O−2 C/• R + intermediates (17)

2
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

O−2 C + O2 →CO2 + O•−2 k = 4.6⋅109 M− 1 ⋅s− 1


(18)


R + O2 →ROO• →Rox + O•−2 (19)

Fe3+ − CA + O•−2 →Fe2+ − CA + O2 (20)

Fe3+ − CA + • O2− C→Fe2+ − CA + CO2 (21)

Fe(II)/CA-S2O2−
8 was demonstrated to be capable of effectively degrading a wide range of pollutants (Table S1). However, despite
all the knowledge accumulated within Fe(II)/CA-S2O2− 2−
8 , there is a lack of a deeper understanding of S2O8 activation chemistry in the
context of scavengers influencing the process. To our best knowledge, only McGachy et al. (2021), Škarohlíd et al. (2020) and Gao et al.
(2021) (Table S1) analyzed S2O2− 8 consumption in the presence of HCO3 and Cl . However, McGachy et al. (2021) and Škarohlíd et al.
− −

(2020) conducted their experiments solely in artificial groundwater (c(HCO−3 ) = 1 mmol•L-1; c(Cl− ) = 2 mmol•L-1), not in demin­
eralized water, limiting the reliable interpretation scavengers influence on S2O2− 8 chemistry through quantitative comparison.
Additionally, in the case of Gao et al. (2021), they separately tested the influence of 100 mmol•L-1 Cl− , SO2− 4 , NO3 and H2PO4 on
− −

S2O2−8 consumption, but, surprisingly, not HCO3 , which has been widely recognized as the scavenger with one of the most inhibitory

effects in various S2O2− 2−


8 activation scenarios (Matzek and Carter, 2016; Wacławek et al., 2017) and also in Fe(II)/CA-S2O8 studies
(Dong et al., 2022; Wu et al., 2014; Zeng et al., 2021) (Table S1). Furthermore, although Gao et al. (2021) provide important insights,
their discussion of S2O2−8 chemistry in the presence of scavengers lacks a comprehensive insight into the processes. For instance, the
application of equilibrium hydrochemical modelling, which was beneficially implemented in other S2O2− 8 - based advanced oxidation
processes studies (Chen et al., 2023; Wang et al., 2019; Xu et al., 2019; Zhang et al., 2021).
In this study, S2O2− 2−
8 activation chemistry in Fe(II)/CA-S2O8 in the presence of selected scavengers, HCO3 and Cl , was investi­
− −

gated by using a series of batch experiments with tetrachloroethene (PCE) as a model contaminant. We systematically monitored
crucial parameters including PCE removal efficiency, S2O2− 8 consumption, reactive stoichiometry efficiency, dissolved total Fe and Fe
(II) concentrations, pH, and standard redox potential (ORPH). Furthermore, to interpret the results and some ambiguities, we used the
equilibrium hydrochemical modelling tool Visual MINTEQ 4 and other supporting techniques, such as high-resolution transmission
electron microscopy (TEM). This multifaceted approach allowed a more thorough understanding of the intricate S2O2− 8 activation
chemistry in Fe(II)/CA-S2O2−8 in the presence of HCO3 and Cl .
− −

2. Materials and methods

2.1. Chemicals

A list of all chemicals used is in Text S1 of SI.

2.2. Experimental setup

2.2.1. Batch experiments


All batch experiments were conducted in 500 mL round glass reactors capped with bimetallic caps with aluminum layered silicone
septa (Fig. S1). Preparation of experimental systems with controlled initial molar concentrations for Fe(II)/CA-S2O2− 8 process is
described in Text S2. The initial molar ratio Fe(II)/CA = 2/3 was chosen on the basis of our previous work (Škarohlíd et al., 2020).
Target Cl− (2 mmol•L-1) and HCO−3 (1 mmol•L-1) in the experimental systems were based on concentrations consistent with artificial
groundwater conditions (McGachy et al., 2021; Škarohlíd et al., 2020). Immediately after preparation and each sampling event, the
reactors were placed into a thermostat (FOC 120I, VELP Scientifica) in the dark at a temperature of 23 ± 1 ◦ C. All reactors were
constantly agitated using a shaker (Digital Orbital Shaker, Heathrow Scientific: 141 rpm).
Individual reaction batch systems are identified using the convention: i) all systems contained PCE, S2O2−
8 and Fe(II) (except control
systems), and the abbreviations w/ and w/o standing for words "with" and "without", respectively. For the other components in the
systems, ii) capital lambda (∧) is used to abbreviate word “and”. Components presence details are summarized in Table S2. The initial
concentrations of each component are provided in the corresponding figure caption, and results, conducted in triplicate, are presented
as mean values with standard deviations (SD).

2.3. Probes experiments

To estimate the steady state concentrations of primary (SO•− 4 ) and secondary radicals ( OH, Cl , Cl2 , and CO3) in the selected
• • • − •

reaction systems (w/(CA), w/(CA ∧ Cl-), and w/(CA ∧ HCO-3), chemical multiprobe tests were utilized (Dong et al., 2020; Hong et al.,
2024; Lei et al., 2023). For a detailed description, refer to Text S4 in SI.

2.4. Analytical methods

PCE was analyzed using a DANI Master VH gas chromatograph equipped with an electron capture detector (ECD), a capillary
column Rtx-VMS (Restek Corporation, U.S.), and a Master SHS static headspace auto-sampler. Peaks were identified by comparing

3
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

their retention times, and mass concentrations were then quantified relative to the response factors of standards. PCE samples were
prepared by withdrawing 2.5 mL by Hamilton syringe from reaction systems into 10 mL crimp-cap vials with silicone and PTFE septa.
Conditions of the analysis are provided in SI (Table S3).
S2O2−
8 concentration was determined using the LUMEX CAPEL-205 capillary electrophoresis system (CE) with fused silica capillary
(ID = 75 µm, L = 55 cm) and UV detection system. CE samples were prepared by extracting 200 µL by Hamilton syringe from reaction
systems and, subsequently, diluted with pre-cooled (4 ◦ C) internal standard stock solution (SS-IS) containing molybdates (c(MoO2− 4 )=
50 mg⋅L-1) to achieve a final volume of 1 mL placed in 2 mL plastic vials. CE samples were then stirred on a shaker for 10 s and
subsequently analyzed. Prior to CE analysis, the capillary was conditioned in consecutive order by 0.25 N NaOH and electrolyte, each
for 10 min, and by 5 min blank analysis. Prior to each run, the capillary was flushed with 0.25 N NaOH for 20 s and with electrolyte for
80 s followed by 1 min blank analysis and finally by electrolyte for 2 min. The parameters of analysis were as follows: voltage − 30 kV;
electrolyte Agilent Inorganic Ions Buffer pH 7.7; capillary temperature 20 ◦ C; hydrodynamic injection 10 mbar for 5 s; inverted UV
detection at wavelength 245 nm; duration 5 min.
The concentrations of chemical probes were analyzed by high-performance liquid chromatography (HPLC) with a diode array
detector (DAD) (Agilent 1260 Infinity). A C18 column (Poroshell 120 EC-C18, 4.6 × 150 mm, 4 μm, Agilent) was used to separate the
probes. The injection volume was set to 100 μL. The samples were eluted with a mixture of acetonitrile and water acidified with formic
acid (0.1%). The column temperature was set at 40◦ C. The gradient elution table is provided in SI (Table S4). All the probes were
determined using the maximum UV absorption (Table S5).
Additionally, for each sampling point, another 6 mL was taken from reaction systems for the pH and ORPH measurement conducted
by the Greisinger GMH 3530 digital pH/mV/thermometer. At sampling points in t = 0 min, pH and ORPH values were determined right
before the addition of S2O2−8 and Fe(II)-CA stock solutions.
The concentrations of dissolved Fe and Fe(II) were determined using the ferrozine/thioglycolate method (Lovibond Iron T M220)
by measuring the absorbance at 560 nm on a Lovibond MultiDirect Photometer System. The sample was prepared by removing
1–10 mL of the reaction systems as needed by dilution (measuring range 0.02–1 mg•L-1) into Lovibond 24 mm diameter cuvettes and
topped up as needed to the line (10 mL). Subsequently, Iron LR and Iron II LR tablets were added to determine total dissolved Fe or

Fig. 1. The influence of Cl− and HCO−3 on Fe(II)/CA-S2O2− 2−


8 process: (a) PCE concentration, (b) S2O8 concentration, (c) pH, (d) ORPH. Experimental
conditions: [S2O2− ]
8 0 /[PCE] 0 = 15/1, [Fe(II)] 0/[CA] 0 = 2/3, [S O 2−
2 8 0 ] /[Fe(II)]0 = 10/1, [PCE] -1 2− -1
0 = 0.06 mmol•L , [S2O8 ]0 = 0.9 mmol•L , [Fe
(II)]0 = 0.09 mmol•L-1, [SO2− -1 -1 -1
4 ]0 = 0.09 mmol•L , [CA]0 = 0.135 mmol•L (if present), [Cl ]0 = 2 mmol•L (if present), [HCO3 ]0 = 1 mmol•L
− − -1

(if present).

4
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

dissolved Fe(II), respectively. The internal calibration of the Lovibond Multidirect system was used for quantification.
Ferrihydrite nanoparticles separation from an aqueous matrix for morphological and elemental analysis was achieved via ultra­
filtration using a Sterlitech (USA) CF042P Cell Assembly with a Synder Flat Sheet Membrane XT, PES, UF (1 kDa), active surface
42 cm2. The membrane underwent a pressure-free conditioning process, starting with a ten-minute flush with clean drinking water,
followed by a ten-minute flush with NaOH solution at pH = 10.5. Subsequently, the membrane was rinsed with potable water until it
reached a neutral pH and then with artificial groundwater for 10 minutes. After conditioning, the membrane was first flushed with a
500 mL water sample without pressure, followed by a 500 mL flush at 6 bar working pressure, driven by a Verder VGS096.17 pump
operating at 80% capacity. The concentration factor was 5, which corresponds to an 80% concentration of the input water. The
membrane unit had no cooling capability, and the inlet water/concentrate temperature increased from approximately 20 ◦ C to 26.2 ◦ C
during the separation process.
TEM were carried out using a Jeol 2200 FS (Jeol, Japan) field emission gun equipped with an in-column energy filter and a TVIPS
camera controlled by EM-Menu software. Energy dispersive spectroscopy (EDS) by Oxford Instruments was used for chemical analysis.
To enhance the contrast of Fe nanoparticles in EELS analysis, a high annular angular dark field (HAADF) detector with a spot size of
1 nm was used in scanning TEM. A small portion of the sample obtained by ultrafiltration was suspended in isopropanol and deposited
on a 300 mesh copper lacey carbon grid (Agar Scientific, Essex, UK).

3. Results and discussions

3.1. Influence of Cl− and HCO−3 on Fe(II)/CA-S2O2−


8 process

We conducted a total of four main experiments varying in the presence of Cl− , HCO−3 (systems denoted as w/(CA ∧ Cl− ), w/(CA ∧
HCO−3 ), and w/(CA ∧ Cl− ∧ HCO−3 )), along with two additional experiments, one with the absence of CA (w/o(CA)) and the second in
borate buffer (w/(CA ∧ BB). All six experiments were conducted with fixed initial molar ratios of S2O2− 2−
8 /PCE = 15/1, S2O8 /Fe(II) =
10/1 and Fe(II)/CA = 2/3 (except for the system w/o(CA) where CA was absent) based on the experiments regarding the influence of
initial S2O2- 2-
8 /Fe(II) molar ratios on Fe(II)/CA-S2O8 process (Text S3).

3.1.1. Influence of Cl− on Fe(II)/CA-S2O2− 8 process


By comparing w/(CA ∧ Cl− ) with w/(CA) in terms of overall removal efficiency (Fig. 1a), it is evident that the kinetics of w/(CA ∧
Cl− ) is significantly slower. After 55 minutes, 88% of PCE had been removed in w/(CA), whereas only 44% in w/(CA ∧ Cl− ). However,
at the end of the experiment (155 min), the difference is less distinctive, as 99% of PCE had been removed in w/(CA), whereas 94% in
w/(CA ∧ Cl− ). From the perspective of pH and ORPH (Figs. 1c and 1d), both systems can be considered identical, as only slight dif­
ferences were observed in the course of the experiments with pH decreasing from an initial value of ≈6 to a final value of ≈3 and ORPH
increasing from value of ≈395 mV to ≈495 mV in the first 5 min and further increasing to ≈540 mV at the end of the experiment.
Wu et al. (2014) and Zeng et al. (2021) also observed negative effects of Cl− (1–100 mmol•L-1) within Fe(II)/CA-S2O2− 8 on tri­
chloroethene (TCE) and naphthalene removal, respectively, using initial molar ratios of S2O2− 2−
8 /TCE = 15/1, S2O8 /Fe(II) = 7.5/1, Fe
(II)/CA = 2/1 and S2O2− 2−
8 /naphthalene = 15/1, S2O8 /Fe(II) = 3/1, Fe(II)/CA = 5/1, respectively. However, using Fe(II)/CA-S2O8 ,
2−

Dong et al. (2022) identified no effect of Cl− (1–50 mmol•L-1) on fluoranthene removal, and Gao et al. (2021) even found a slightly
positive effect of Cl− (100 mmol•L-1) for pyrene removal. Different initial molar ratios in Fe(II)/CA-S2O2− 8 were used in these studies, i.
e., S2O2− 2− 2−
8 /fluoranthene = 60/1, S2O8 /Fe(II) = 3/1 and Fe(II)/CA = 4/1 (Dong et al., 2022), and S2O8 /pyrene = 2000/1, S2O8 /Fe
2−

(II) = 1/1 and Fe(II)/CA = 1/1 (Gao et al., 2021), suggesting that in particular lower S2O2− 8 /Fe(II) initial molar ratios, which imply
higher Fe(II) concentrations relative to S2O2− 8 , led to suppression of the negative effect of Cl .

Both negative and positive effects of Cl− on the removal efficiency in Fe(II)/CA-S2O2− 8 process have been commonly attributed to
formation of Cl•, Cl•− 2 and [ClOH]
•−
radicals via Eq. (22)–(26) and, consequently, their potential lower/higher reactivities with the
target contaminant compared to the reactivity SO•− 4 and OH (Dong et al., 2022; Gao et al., 2021; Wu et al., 2014).

However, in our experiments, Cl− also significantly altered S2O2− 8 consumption (Fig. 1b). Compared to w/(CA), where only 46% of
S2O2− 2−
8 remained after 155 min, in w/(CA ∧ Cl ) 68% of S2O8 was still present in the system at the end of the experiment.

SO•−4 + Cl− →SO2−


4 + Cl

k = 2.5⋅108 M− 1 ⋅s− 1
(22)

Cl• + Cl− ↔ Cl•−2 k→ = 2.0⋅1010 M− 1 ⋅s− 1


k← = 1.1⋅105 ⋅s− 1
(23)

Cl•−2 + H2 O→[ClOH]•− + H+ + Cl− k[H2 O] < 100s− 1


(24)

[ClOH]•− + H+ ↔ Cl• + H2 O k→ = 2.1⋅1010 M− 1 ⋅s− 11


k← = 1.3⋅103 s− 1
(25)

Cl− + • OH ↔ [ClOH]•− k→ = 4.3 ⋅ 109 M− 1 ⋅ s− 1


k← = 6.1 ⋅ 109 ⋅ s− 1
(26)

Despite the fact that both w/(CA) and w/(CA ∧ Cl ) were able to sustain dissolved Fe(II) throughout the experiments in comparison

with other systems (Fig. 2a), dissolved Fe(II) concentration dropped in the first 5 min to a steady-state value of ≈5% of the initial
concentration in w/(CA), while in w/(CA ∧ Cl− ) dissolved Fe(II) concentration dropped further to the lower stable value of ≈2% of the
initial concentration. The concentrations of total dissolved Fe were almost identical in both w/(CA) and w/(CA ∧ Cl-) (Fig. 2b). From

5
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

all Fe(II)/CA-S2O2− 2−
8 studies (Table S1), only Gao et al. (2021) analyzed S2O8 consumption throughout their experiments with Cl , but

2−
unfortunately not dissolved Fe(II) concentration, and they found no noteworthy distinction in S2O8 consumption between systems
with and without Cl− (100 mmol•L-1). However, as mentioned earlier, Gao et al. (2021) used a completely different setting of initial
molar ratios in the Fe(II)/CA-S2O2− 8 process compared to this studym making a direct comparison difficult.
Therefore, it might be suggested that the lower stable dissolved Fe(II) concentration in w/(CA ∧ Cl− ) led primarily to less intense
S2O2− 8 activation, which probably resulted secondarily along with Eq. (22)–(26) in qualitatively and quantitatively different mix of
reactive species compared to w/(CA) that could attack PCE. And this was indeed observed by implementing probe tests in systems w/
(CA) and w/(CA ∧ Cl− ), for details see Text S4. While the estimated steady state radical concentrations (marked as [radical]SS) in the
-13
post-initial stage (t ≥5 min) were [SO•− 4 ]SS = 2.1⋅10 mol•L-1 and [•OH]SS = 2.4⋅10-14 mol•L-1 in w(CA) (Table S7), system w/(CA ∧
Cl ) contained [SO4 ]SS = 1.1⋅10 mol•L and [ OH]SS = 3.0⋅10-14 mol•L-1 including [Cl•−
− •− -13 -1 •
2 ]SS = 3.0⋅10
-13
and [Cl•]SS = 6.4⋅10-16
mol•L-1 (Table S7). A shift in the radical mix, although it was not reflected in the values of ORPH compared to w/(CA), entails an
inherent change in reactivity towards the contaminant of concern, in our case, PCE.
It is well documented that in system w/(CA ∧ Cl− ) prevailing Cl•− 2 has different reactivity about one to two orders of magnitude
lower than SO•− 4 , OH, and Cl towards trace organic compounds (Duan et al., 2022; Lei et al., 2019; H. Zhang et al., 2024) and about a
• •

few orders of magnitude lower than Cl• towards simple alcohols, aldehydes, ketones, and organic acids (Buxton et al., 2000). In the
case of PCE, the reported second-order rate constant for •OH (k(PCE, •OH) = 2.6 ⋅ 109 M-1⋅s-1 (Mertens and von Sonntag, 1995)) is one
order magnitude higher than the rate constant for Cl• (k(PCE, Cl•) = 2.8 ⋅ 108 M-1⋅s-1 (Mertens and Sonntag, 1994)), while the
second-order rate constants for SO•− 4 and Cl2 have not been reported.
•−

However, the question remains: what mechanism causes the relatively lower stable dissolved Fe(II) concentration in w/(CA ∧ Cl− )
compared to w/(CA)? We hypothesize that it may be due to one or a combination of the following two reasons: (i) similarly to SO•− 4 ,

OH, both Cl•/Cl•−2 may act as quenchers of dissolved Fe(II) (Eq. (27)–Eq. (28)) (Thornton and Laurence, 1973; Truong et al., 2004). Cl

with a slightly higher reaction rate than that of SO•−4 (Eq. (12)), and a few orders of magnitude higher than that of OH (Eq. (13)), and

Cl•−
2 with a few orders of magnitude lower than that of SO4 and OH (ii) Cl / Cl2 might be unable to contribute to CO2 / R formation
•− • • •− • − •

within the Fe(III)/Fe(II) regeneration cycle at a comparable intensity to SO•- 4 , •


OH(Eqs. (15)–(16)).
To clarify the relative significance of these two potential mechanisms, further comprehensive investigation combined with the
development of complex kinetic model is needed. But considering [Cl•− 2 ]SS is of the same order as [SO4 ]SS in w/(CA ∧ Cl ), and,
•− −

generally, lower reactivity of Cl•−


2 towards organic compounds and dissolved Fe(II), as discussed earlier, we are more inclined towards
the second option (ii).

Cl• + Fe2+ →Cl− + Fe3+ k = 5.9⋅109 M− 1 ⋅s− 1


(27)

Cl•−2 + Fe2+ →Cl− + Fe3+ k = 1.4⋅107 M− 1 ⋅s− 1


(28)

In summary, we suggest that the effect of Cl on the



Fe(II)/CA-S2O2−
8 process lies
not only in the altered reactivity of the generated
Cl• and Cl•−
2 towards the target pollutant compared to SO•−
4 and OH, as described

elsewhere, but also in the ability of Cl• and Cl•−
2
radicals to reduce the steady-state dissolved Fe(II) concentration and thus the intensity of S2O2−
8 activation/consumption, resulting in a
lower rate of target pollutant removal.

3.1.2. Influence of HCO−3 on Fe(II)/CA-S2O2−8 process


The presence of 1 mmol•L-1 HCO−3 led to a significant decrease in the removal efficiency in both w/(CA ∧ HCO−3 ) and w/(CA ∧ Cl− ∧
HCO−3 ) compared to w/(CA) after 155 min, specifically 13% in the case of w/(CA ∧ HCO−3 ) and 7% in w/(CA ∧ Cl− ∧ HCO−3 ) (Fig. 1a).
In terms of pH (Fig. 1c), w/(CA ∧ HCO−3 ) and w/(CA ∧ Cl− ∧ HCO−3 ) exhibited a decrease from an initial higher value of 8.4 (caused by
HCO−3 presence) compared to 6.1 in w/(CA) to values in the range of 6.0–6.5 in the first minutes due to the addition of stock solutions

Fig. 2. The influence of Cl− and HCO−3 on Fe(II)/CA-S2O2− 8 process: (a) concentration of dissolved Fe(II) (ferrozine method), (b) concentration of
total dissolved Fe (ferrozine method). Experimental conditions: [S2O2− 2−
8 ]0/[PCE]0 = 15/1, [Fe(II)]0/[CA]0 = 2/3, [S2O8 ]0/[Fe(II)]0 = 10/1, [PCE]0
= 0.06 mmol•L-1, [S2O2− -1 -1 2− -1 -1
8 ]0 = 0.9 mmol•L , [Fe(II)]0 = 0.09 mmol•L , [SO4 ]0 = 0.09 mmol•L , [CA]0 = 0.135 mmol•L (if present), [Cl ]0 =

2 mmol•L-1 (if present), [HCO−3 ]0 = 1 mmol•L-1 (if present).

6
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

of Fe(II)-CA and S2O2−8 and remained at the same level throughout the experiment. Thus, the stable pH level from 5 min onwards was
substantially different from w/(CA) and w/(CA ∧ Cl− ).
Wu et al. (2014) and Zeng et al. (2021) also observed a significant decrease in TCE and naphthalene removal from nearly
100–93.5%, 16.6%, 3.2% and 53.3%, 22.2%, 9.2% when adding 1 mmol•L-1, 10 mmol•L-1 and 100 mmol•L-1 to Fe(II)/CA-S2O2− 8
system accompanied by a change in initial/final pH in the range 5.9/3.0–8.1/8.1. They argued that the decrease was associated with
the formation of less reactive CO•−
3 radicals (Eqs. (29)-(33)) (Wu et al., 2014; Zeng et al., 2021) and with an increase of the initial pH
values by the presence of HCO−3 consequently suppressing SO•− 4 and OH generation (Wu et al., 2014). Their reasoning behind the

suppression of SO4 and OH generation assumes a higher SO4 redox potential in acidic solutions and a higher intensity of •OH
•− • •−

formation at lower pH. Dong et al. (2022) also identified a significant negative effect of HCO−3 presence on fluoranthene removal. After
adding 1 mmol•L-1, 10 mmol•L-1 and 50 mmol•L-1 of HCO−3 to Fe(II)/CA-S2O2− 8 system, fluoranthene removal decreased from 96.3%
to 11.7%, 3.1% and 0%, and initial/final pH values changed from 4.69/3.98–6.47/7.91, 8.30/8.42, and 8.41/8.37, respectively.
Similarly, to Wu et al. (2014) and Zeng et al. (2021), Dong et al. (2022) argued that the transformation of SO•−4 and OH into CO3 (Eqs.
• •

(29) and (31) (Bühl et al., 2015; Buxton et al., 1988; Huie and Clifton, 1990; Padmaja et al., 1993)) was the main reason for the
negative effect of HCO−3 observed and the effect of pH was attributed to HCO−3 acting as a buffer, preventing significant changes in pH,
thus altering fluoranthene removal, without deeper discussion or reasoning.

SO•−4 + HCO−3 →SO2−


4 + CO3 + H
• +
k = 2.8⋅106 M− 1 ⋅s− 1
(29)

SO•−4 + CO2− 2−
3 →SO4 + CO3
•−
k = 4.1⋅106 M− 1 ⋅s− 1
(30)


OH + HCO−3 →CO•−3 + H2 O k = 8.5⋅106 M− 1 ⋅s− 1
(31)


OH + CO2−
3 →CO3 + OH
•− −
k = 3.9⋅108 M− 1 ⋅s− 1
(32)

HCO•3 ↔ H+ + CO•−3 pKa ≈ − 2.0 (33)

In above-mentioned studies, S2O2−8 consumption was not determined throughout the experiments with HCO3 . However, as shown

- 2−
in Fig. 1b, the presence of HCO3 in w/(CA ∧ HCO3 ) and w/(CA ∧ Cl ∧ HCO3 ) significantly affected S2O8 consumption in our ex­
− − −

periments. Compared to w/(CA), where only 46% of S2O2− 8 remained after 155 min, in w/(CA ∧ HCO3 ) and w/(CA ∧ Cl ∧ HCO3 ) 96%
− − −
2− 2−
and 95% of S2O8 , respectively, were still present in the system. Considering the uncertainty in S2O8 measurement, a ≈5% loss in
S2O2− 2−
8 could be approximately expected from the initial conversion of dissolved Fe(II) to Fe(III) by the reaction with S2O8 (Fig. 2a).
2−
Furthermore, very similar S2O8 consumption (4%) was also observed in (w/(CA ∧ BB) where the initial pH was adjusted using borate
buffer to roughly match the initial pH of w/(CA ∧ HCO−3 ) and w/(CA ∧ Cl− ∧ HCO−3 ). The indications mentioned above, together with
the fact that w/(CA ∧ HCO−3 ), w/(CA ∧ Cl− ∧ HCO−3 ) and (w/(CA ∧ BB) did not establish a stable concentration of dissolved Fe(II) in
contrast to w/(CA) and w/(CA ∧ Cl ) (Fig. 2a), show that there must be a mechanism linked to pH inhibiting Fe(III)/Fe(II) cyclic

regeneration and consequently S2O2− 8 consumption/activation.


To gain a deeper understanding of this mechanism, we used the equilibrium hydrochemical model implemented in Visual MINTEQ
4 to approximately simulate the distribution of Fe(III) species in all systems in the first minutes of experiments after the expected initial
conversion of Fe(II) to Fe(III). However, for the equilibrium model to be meaningful for this purpose, the modeled systems must at least
approximately satisfy the assumption of instantaneous equilibration. Zhu et al. (2016) showed that due to the very fast dynamics of

Fig. 3. Fe(III) species distribution based on equilibrium hydrochemical model in individual systems. Only Fe(III) species with molar fraction
>0.1% showed.

7
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

reaching equilibrium of dissolved and solid Fe(III) species, the equilibrium model was able to fit the experimental data remarkably.
Owing to the very low concentrations of dissolved Fe(II) compared to dissolved Fe(III) in all systems in the first 5 min (Fig. 2a),
dissolved Fe(II) was omitted from the model inputs for simplicity. S2O2− 8 concentrations were then approximately considered as the
initial S2O2− 2−
8 concentrations minus S2O8 consumed for the stoichiometric conversions of all dissolved Fe(II) to Fe(III). SO4 con­
2−

centrations were based on the amount of FeSO4⋅7 H2O added to the system as an activator summed with the stoichiometric amount
produced by the decomposition of S2O2− 2−
8 for the conversion of dissolved Fe(II) to Fe(III). S2O8 was added as a component to the
default Visual MINTEQ 4 database assuming its nonreactivity with other components (lack of hydrochemical data). Due to the
considered time frame of minutes, only ferrihydrite was considered as a Fe(III) solid specie (Schwertmann et al., 2004; Stefánsson,
2007; Zhu et al., 2016). The measured values in each system at 5 min were used as model input pH. Ionic strength was automatically
calculated by Visual MINTEQ 4 and Davies activity correction was used with the temperature set to 23 ◦ C. The inputs to the Visual
MINTEQ 4 model are summarized in Table S8.
Model results in the form of Fe(III) species distribution are shown in Fig. 3. Charge differences were all <10% in all modeled
systems. It is evident that only in w/(CA) and w/(CA ∧ Cl-) all Fe(III) was available in dissolved form, predominantly (≈98%) as [Fe(III)
Cit], with the remainder as [Fe(III)HCit]+, according to the model. On the other hand, in systems w/o(CA), w/(CA ∧ HCO-3), w/(CA ∧
Cl- ∧ HCO-3) and (w/(CA ∧ BB) the model outputs indicate a majority to absolute Fe(III) distribution of 86.54%, 98.83%, 100% and
100%, respectively, in the form of ferrihydrite. Comparing the model Fe(III) species distribution as a function of pH of systems w/o
(CA) and w/(CA) (Fig. 4), it can be seen that the presence of CA in w/(CA) shifts the pH limit of ferrihydrite formation by about 2 pH
points higher from 2.8 (w/o(CA)) to 4.6 w/(CA). In the extreme case of an initial molar ratio of Fe(II)/CA = 1/10, the theoretical pH
limit of ferrihydrite formation can be shifted to a pH value of ≈5.5 (Fig. S5). However, excess CA could potentially significantly
compete with the target contaminant by interacting with the reactive species, thus decreasing removal efficiency.
Based on the results of the model, the potential formation of ferrihydrite particles in systems w/(CA ∧ HCO−3 ), w/(CA ∧ Cl- ∧
HCO−3 ), and (w/(CA ∧ BB) would have explained the unavailability of dissolved Fe(III) for Fe(III)/Fe(II) cyclic regeneration (Eqs. (20)-
(21)) and consequently stopped S2O2− 8 activation by dissolved Fe(II), thereby significantly suppressing the generation of reactive
species. This suppression effect was indeed observed in the probe tests of the system w/(CA ∧ HCO−3 ) (Text S4) by stagnation of the
decay of the used probe compounds in the post-initiation stage (t ≥ 5 min), see Fig. S4c and Table S7, and it was further reflected in the
ORPH values of w/(CA ∧ HCO−3 ), see Fig. 1d, which were completely divergent compared to w/(CA) and w/(CA ∧ Cl− ), reaching lower
values up to 250 mV at the end of the experiment.
It has been well documented that the presence of Fe-chelating agents such as polycarboxylates (e.g., citrate, oxalate) might in­
fluence the structural and surface properties of Fe(III)-(hydro)oxides, including their formation, solubility, and suspension stability
(Cornell and Schwertmann, 2006; Liu and Huang, 2003, 1999). Moreover, Mikutta et al. (2010) found that after centrifugation of a
suspension of ferrihydrite formed in a solution (pH = 6.5) with Fe(III) and CA concentrations of 200 mmol•L-1 and 100 mmol•L-1,

Fig. 4. Fe(III) species distribution as a function of pH based on equilibrium hydrochemical model in systems w/o(CA) (a) and w/(CA) (b). Only Fe
(III) species with molar fraction >0.1% showed.

8
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

respectively, the vast majority of Fe (98%) was contained in the supernatant, which was incidentally filtered through a 25-nm
membrane filter. In the supernatant, ferrihydrite nanoparticles were identified to represent 65% of the Fe(III) species compared to
35% of the Fe(III)-CA complex. These ferrihydrite nanoparticles were electrostatically stabilized by adsorbed citrate, preventing the
formation of visible, sedimentable macroscopic precipitates (Hofmann and Liang, 2007). Mikutta et al. (2010) further found that the
5-k Da ultrafiltrate of supernatant contained only 5% ferrihydrite in contrast to 95% of Fe(III)-CA complex. Based on these findings, we
decided to determine if w/(CA ∧ Cl− ∧ HCO−3 ) also contained ferrihydrite nanoparticles by TEM analysis of the filter cake from 1-k Da
ultrafiltration of specially prepared 1000 mL of w/(CA ∧ Cl− ∧ HCO−3 ) after 30 min reaction time.
The high-resolution TEM images clearly reveal that the filter cake contained nanoparticles with a size less than 20 nm that
aggregated into larger clusters (Fig. 5). Elemental analysis performed by EDS confirmed that the nanoparticles were composed mostly
of Fe and O. Additionally, the EDS analysis identified the presence of Cu and C derived from the measurement grid, accompanied by
ubiquitous dust particles, as indicated by traces of Si and Ca (Fig. S6–S7). To gain further insight into the nanoparticles structure, the
HRTEM image analyses performed by CrystBox analysis tool proved presence of nanocrystalline ferrihydrite (Fig. 5b). Selected single
crystals are labelled by red circles for better visibility. The FFT of bold-labelled particle is show in the inset of the image.
Interestingly, despite the formation of ferrihydrite nanoparticles, we observed only a slight decrease in total dissolved Fe con­
centration in w/(CA ∧ HCO−3 ), w/(CA ∧ Cl− ∧ HCO−3 ) and w/(CA ∧ BB) compared to systems w/(CA) and w/(CA ∧ Cl− ) after 120 min
(Fig. 2b). Moreover, this reduction was not as significant as that observed in the case of w/o(CA). Because the ferrozine method detects
Fe particles <50 nm within the indicator of total dissolved Fe (Pullin and Cabaniss, 2001) it can be assumed that in w/(CA ∧ HCO−3 ),
w/(CA ∧ Cl− ∧ HCO−3 ) and w/(CA ∧ BB), the ferrihydrite nanoparticles likely dissolved during the measurement due to reduced pH (≈
4) and excess ferrozine acting as a chelating agent (Cornell and Schwertmann, 2006).
To further investigate the effect of pH on the Fe(II)/CA-S2O2−8 process, we prepared systems of the same composition as w/(CA) but
with varied initial pH values by titrating with 0.1 mol•L-1 NaOH (Fig. S8). As can be seen from Fig. S8, the system with an initial pH of
≈11 exhibits similar behavior to systems w/(CA ∧ HCO−3 ), w/(CA ∧ Cl− ∧ HCO−3 ), and w/(CA ∧ BB) in all monitored parameters (PCE
removal efficiency, S2O2− 8 consumption, and dissolved Fe(II)) except the concentration of total dissolved Fe, which was affected by the
high value of the final pH ≈10, possibly preventing the dissolution of CA-stabilized ferrihydrite nanoparticles by the exposition to
ferrozine method, as discussed earlier. The common feature of the mentioned systems is that the final pH is greater than the pH
threshold of ≈4.6 for ferrihydrite formation in w/(CA) (Fig. 4b). On the other hand, the remaining systems that did not exceed the pH
threshold of ≈4.6 exhibited similar behavior to the w/(CA) system. A comparable trend can also be found in other studies investigating
the effect of initial/final pH on the Fe(II)/CA-S2O2− 8 process (Dong et al., 2022; Wang et al., 2019; Xu et al., 2022; Zeng et al., 2021; R.
Zhang et al. 2024), where final pH values exceeding the pH threshold of ferrihydrite formation led to a significant decrease in
contaminant removal efficiency. The pH threshold in these studies was in the range of 3–5 based on the chemical composition of Fe
(II)/CA-S2O2− 8 , especially the initial Fe(II) and CA molar concentrations and the Fe(II)/CA molar ratio. However, this fact is left mostly
without deeper discussion, lacking the context in the form of S2O2− 8 consumption, dissolved Fe(II), and the absence of hydrochemical
modeling.
In summary, the major effect of HCO−3 on Fe(II)/CA-S2O2− 8 was in altering solution pH, leading to the formation of CA-stabilized
ferrihydrite nanoparticles, not in the formation of less reactive CO•− 3 radicals as commonly reported. This primarily resulted in the
inaccessibility of dissolved Fe(III) for the Fe(III)/Fe(II) regeneration cycle (Eqs. (20)–(21)), consequently resulting in the absence of
available Fe(II) for S2O2− 2−
8 activation, hindering PCE removal. In the absence of available dissolved Fe(II) for S2O8 activation, the main
2−
process contributing to the initiation of S2O8 decomposition is thermolysis, producing two SO4 (Eq. (34)), initiating a chain of
•−

reactions. Although S2O2− 8 can be converted to SO4 by thermolysis Eq. (34), the rate of radical production is too slow, at weakly basic
•−

Fig. 5. TEM images of samples separated from a 1-k Da ultrafiltration cake a) cluster overview and b) detailed HRTEM with red labelled particles,
indexed zone axe and interplanar distance and indexed FFT in the inset.

9
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

pH and temperature of 23◦ C, to be useful for PCE removal (McGachy and Sedlak, 2024).

S2 O2−
8 →2SO4 (34)
Δ•−

3.2. Influence of increasing Fe(II) and S2O2− 2−


8 concentrations on Fe(II)/CA-S2O8 process in the presence of Cl and HCO3
− −

Due to the strong inhibitory effect of the presence of HCO−3 and Cl− on Fe(II)/CA-S2O2− 2−
8 at the initial molar ratios of S2O8 /PCE =
15/1, S2O2−8 /Fe(II) = 10/1 and Fe(II)/CA = 2/3, it was convenient to investigate the effect of Cl −
and HCO −
3 on Fe(II)/CA-S 2−
2O8 at
elevated Fe(II) concentrations, specifically quadrupled in this case. The system with initial molar ratios of S2O8 /PCE = 15/1, S2O2−
2−
8 /
Fe(II) = 2.5/1 was able to remove 81% of PCE after 130 min compared to only 2% in the system with initial molar ratios of S2O2− 8 /PCE
= 15/1, S2O2-8 /Fe(II) = 10/1 (Fig. S9a). The addition of higher concentration of Fe(II) and hence CA (Fe(II)/CA = 2/3 fixed in all
experiments) led to a pH decrease to values in the interval 2.5–3 (Fig. S9c), which prevented ferrihydrite formation and resulted in
continuous S2O2− 2− 2-
8 activation (Fig. S9b) and higher ORPH values compared to system S2O8 /PCE = 15/1, S2O8 /Fe(II) = 10/1. It should
be noted that pH decrease induced the transformation of HCO−3 to H2CO3* (Fig. S10), thus, eliminating the potential scavenging effect
of HCO−3 on Fe(II)/CA-S2O2− 8 , see Eqs. (28) and (30). On the other hand, Cl was still actively present, and its negative effect on S2O8
− 2−

activation (mechanisms described in previous text) can be seen by comparing systems S2O2− 8 /PCE = 15/1, S O
2 8
2−
/Fe(II) = 2.5/1 with/
without the presence of Cl− (Fig. S11).
Similarly, increasing initial molar ratios S2O2− 2−
8 /PCE to 60/1 (S2O8 /Fe(II) = 10/1, Fe(II)/CA = 2/3 fixed) resulted in overcoming
the negative effects of Cl− and HCO−3 by achieving sufficient target pollutant removal, this strategy was, in principle, also successful in
other studies that examined Fe(II)/CA-S2O2− 8 efficiency in real waters containing Cl and HCO3 (Dong et al., 2022; Zeng et al., 2021;
− −

Zhou et al., 2021).

4. Implications for Advanced Oxidation Processes (AOPs) design

To our best knowledge, this study is the first to critically investigate the influence of Cl− and HCO−3 presence on the intensity of
S2O2− 2−
8 activation within the Fe(II)/CA- S2O8 process.
In addition to the previously reported potential effect on the occurrence of individual radical species in the radical mix, the
presence of Cl− significantly impacts the steady-state dissolved Fe(II) concentration within the Fe(II)/CA- S2O2−8 process, leading to a
diminished intensity of S2O2−8 activation and consequently a lower contaminant removal rate. Similarly, the presence of HCO3 pri­

2−
marily elevates solution pH, resulting in the formation of CA-stabilized ferrihydrite nanoparticles, which hinder S2O8 activation and
suppress contaminant removal. The CA presence can shift the pH limit of ferrihydrite formation only by about 2 pH points, from ≈2.8
to ≈4.6 in experimental systems. Therefore, at higher pH and in buffering environments the formation of CA-stabilized ferrihydrite
nanoparticles can significantly impact the process efficacy. To overcome this challenge, it would be necessary to use relatively higher
initial molar ratios S2O2− 2-
8 /contaminant with simultaneously fixed S2O8 /Fe(II) and Fe(II)/CA, or employ a higher Fe/CA ratio, to
counteract the buffering effect of HCO3 by lowering the pH value. However, while increasing S2O2−

8 concentration relative to the
target pollutant can mitigate the negative effects of both Cl− and HCO−3 and achieve sufficient removal efficiency, it may lead to an
increase in SO2−
4 levels, compromising water quality. Therefore, a careful assessment of scavenger effects and secondary measures,
such as increasing S2O2− 2−
8 concentration, is crucial before implementing Fe(II)/CA activation within S2O8 based AOPs. The conven­
tional bench treatability tests, commonly utilized before full-scale AOPs implementation, should thus incorporate a comprehensive
analysis of water matrix constituents alongside equilibrium hydrochemical modelling when utilizing Fe(II)/CA-S2O2− 8 .
These insights not only contribute new information to the existing understanding of AOPs mechanisms but also provide valuable
guidance for the practical design and implementation of Fe(II)/CA-S2O2− 8 systems in water treatment applications. Future research
should explore the effects of other water matrix constituents, such as NO−3 , SO2− 2−
4 , and humic acids, on S2O8 activation intensity within
the Fe(II)/CA-S2O2−8 process, further advancing our understanding and optimization of AOPs for contaminated water treatment.

CRediT authorship contribution statement

Lenka McGachy (Honetschlägerová): Writing – review & editing, Supervision, Investigation, Conceptualization. Marek Mar­
tinec: Methodology, Formal analysis. Alena Michalcova: Formal analysis. Radek Skarohlid: Writing – original draft, Investigation,
Conceptualization. Doris Kraljic: Formal analysis. Jan Suchan: Formal analysis. Pavlina Tesinska: Formal analysis. Marina Kho­
lomyeva: Formal analysis. Karel Kuhnl: Formal analysis. Barbora Hanzlova: Formal analysis.

Declaration of Competing Interest

The authors declare the following financial interests/personal relationships which may be considered as potential competing in­
terests:Lenka McGachy reports financial support was provided by Czech Science Foundation. If there are other authors, they declare
that they have no known competing financial interests or personal relationships that could have appeared to influence the work re­
ported in this paper

10
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

Data availability

Data will be made available on request.

Acknowledgment

This research was financially supported by Institutional Program of University of Chemistry and Technology Prague (IP
2019–2020). Additional support was provided by the Czech Science Foundation (Grant 23–05901S). Graphical abstract was created
with BioRender.com.

Appendix A. Supporting information

Supplementary data associated with this article can be found in the online version at doi:10.1016/j.eti.2024.103687.

References

Behrman, E.J., 2004. The Persulfate Oxidation of Phenols and Arylamines (The Elbs and the Boyland–Sims Oxidations). Organic Reactions. John Wiley & Sons, Ltd,
pp. 421–511. https://doi.org/10.1002/0471264180.or035.02.
Bühl, M., DaBell, P., Manley, D.W., McCaughan, R.P., Walton, J.C., 2015. Bicarbonate and alkyl carbonate radicals: structural integrity and reactions with lipid
components. J. Am. Chem. Soc. 137, 16153–16162. https://doi.org/10.1021/jacs.5b10693.
Buxton, G.V., Bydder, M., Salmon, G.A., Williams, J.E., 2000. The reactivity of chlorine atoms in aqueous solution. Part III. The reactions of Cl• with solutes. Phys.
Chem. Chem. Phys. 2, 237–245. https://doi.org/10.1039/A907133D.
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 Aqueous Solution. J. Phys. Chem. Ref. Data 17, 513–886. https://doi.org/10.1063/1.555805.
Cai, S., Hu, X., Lu, D., Zhang, L., Jiang, C., Cai, T., 2021. Ferrous-activated persulfate oxidation of triclosan in soil and groundwater: The roles of natural mineral and
organic matter. Sci. Total Environ. 762, 143092 https://doi.org/10.1016/j.scitotenv.2020.143092.
Chen, L., Dong, X., Feng, R., Li, W., Ding, D., Cai, T., Jiang, C., 2023. Oxalic acid enhanced ferrous/persulfate process for the degradation of triclosan in soil:
Efficiency, mechanism and a column study. Chem. Eng. J., 144961 https://doi.org/10.1016/j.cej.2023.144961.
Cornell, R.M., Schwertmann, U., 2006. The iron oxides: structure, properties, reactions, occurrences and uses, 2., compl. rev. and extended ed., repr. ed. Wiley-VCH,
Weinheim.
Ding, Y., Wang, X., Fu, L., Peng, X., Pan, C., Mao, Q., Wang, C., Yan, J., 2021. Nonradicals induced degradation of organic pollutants by peroxydisulfate (PDS) and
peroxymonosulfate (PMS): Recent advances and perspective. Sci. Total Environ. 765, 142794 https://doi.org/10.1016/j.scitotenv.2020.142794.
Dong, H., Li, Y., Wang, S., Liu, W., Zhou, G., Xie, Y., Guan, X., 2020. Both Fe(IV) and radicals are active oxidants in the Fe(II)/peroxydisulfate process. Environ. Sci.
Technol. Lett. 7, 219–224. https://doi.org/10.1021/acs.estlett.0c00025.
Dong, J., Sheng, X., Liu, Y., Wang, P., Lu, Z., Sui, Q., Lyu, S., 2022. Insights into the enhanced fluoranthene degradation in citric acid coupled Fe(II)-activated sodium
persulfate system. Water Supply 22, 4822–4838. https://doi.org/10.2166/ws.2022.190.
Duan, X., Niu, X., Gao, J., Wacławek, S., Tang, L., Dionysiou, D.D., 2022. Comparison of sulfate radical with other reactive species. Curr. Opin. Chem. Eng. 38, 100867
https://doi.org/10.1016/j.coche.2022.100867.
Faggiano, A., Ricciardi, M., Motta, O., Fiorentino, A., Proto, A., 2023. Greywater treatment for reuse: Effect of combined foam fractionation and persulfate-iron based
fenton process in the bacterial removal and degradation of organic matter and surfactants. J. Clean. Prod. 415, 137792 https://doi.org/10.1016/j.
jclepro.2023.137792.
Fordham, J.W.L., Williams, H.L., 1951. The Persulfate-Iron(II) Initiator System for Free Radical Polymerizations1. J. Am. Chem. Soc. 73, 4855–4859. https://doi.org/
10.1021/ja01154a114.
Gao, Y., Champagne, P., Blair, D., He, O., Song, T., 2020. Activated persulfate by iron-based materials used for refractory organics degradation: a review. Water Sci.
Technol. 81, 853–875. https://doi.org/10.2166/wst.2020.190.
Gao, Y., Yang, F., Jian, H., Zhen, K., Zhang, P., Tang, X., Fu, Z., Xu, W., Wang, C., Sun, H., 2021. Pyrene degradation in an aqueous system using ferrous citrate
complex activated persulfate over a wide pH range. J. Environ. Chem. Eng. 9, 106733 https://doi.org/10.1016/j.jece.2021.106733.
Han, D., Wan, J., Ma, Y., Wang, Y., Li, Y., Li, D., Guan, Z., 2015. New insights into the role of organic chelating agents in Fe(II) activated persulfate processes. Chem.
Eng. J. 269, 425–433. https://doi.org/10.1016/j.cej.2015.01.106.
Herrmann, H., Reese, A., Zellner, R., 1995. Time-resolved UV/VIS diode array absorption spectroscopy of SOx− (x=3, 4, 5) radical anions in aqueous solution. J. Mol.
Struct., Mol. Spectrosc. Mol. Struct. 1994 348, 183–186. https://doi.org/10.1016/0022-2860(95)08619-7.
Hofmann, A., Liang, L., 2007. Mobilization of colloidal ferrihydrite particles in porous media—An inner-sphere complexation approach. Geochim. Cosmochim. Acta
71, 5847–5861. https://doi.org/10.1016/j.gca.2007.06.050.
Hong, W., Zou, J., Zhao, M., Yan, S., Song, W., 2024. Development of a Five-Chemical-Probe Method to Determine Multiple Radicals Simultaneously in Hydroxyl and
Sulfate Radical-Mediated Advanced Oxidation Processes. Environ. Sci. Technol. 58, 5616–5626. https://doi.org/10.1021/acs.est.4c00669.
Hou, K., Shen, P., Wang, Z., Pi, Z., Chen, F., Li, X., Dong, H., Yang, Q., 2023. Revisiting the contribution of FeIVO2+ in Fe(II)/peroxydisulfate system. Chin. Chem.
Lett. 34, 107555 https://doi.org/10.1016/j.cclet.2022.05.069.
Huie, R.E., Clifton, C.L., 1990. Temperature dependence of the rate constants for reactions of the sulfate radical, SO4-, with anions. J. Phys. Chem. 94, 8561–8567.
https://doi.org/10.1021/j100386a015.
Jiang, J., Zhang, D., Zhang, H., Yu, K., Li, N., Zheng, G., 2020. Degradation mechanism study of fluoroquinolones in UV/Fe2+/peroxydisulfate by on-line mass
spectrometry. Chemosphere 239, 124737. https://doi.org/10.1016/j.chemosphere.2019.124737.
Karim, A.V., Jiao, Y.L., Zhou, M.H., Nidheesh, P.V., 2021. Iron-based persulfate activation process for environmental decontamination in water and soil. Chemosphere
265, 20. https://doi.org/10.1016/j.chemosphere.2020.129057.
Lei, Y., Cheng, S., Luo, N., Yang, X., An, T., 2019. Rate constants and mechanisms of the reactions of Cl• and Cl2•– with trace organic contaminants. Environ. Sci.
Technol. 53, 11170–11182. https://doi.org/10.1021/acs.est.9b02462.
Lei, Y., Yu, Y., Lei, X., Liang, X., Cheng, S., Ouyang, G., Yang, X., 2023. Assessing the use of probes and quenchers for understanding the reactive species in advanced
oxidation processes. Environ. Sci. Technol. 57, 5433–5444. https://doi.org/10.1021/acs.est.2c09338.
Liu, B., Huang, B., Wang, Z., Tang, L., Ji, C., Zhao, C., Feng, L., Feng, Y., 2023. Homogeneous/heterogeneous metal-catalyzed persulfate oxidation technology for
organic pollutants elimination: a review. J. Environ. Chem. Eng. 11, 109586 https://doi.org/10.1016/j.jece.2023.109586.
Liu, C., Huang, P.M., 2003. Kinetics of lead adsorption by iron oxides formed under the influence of citrate. Geochim. Cosmochim. Acta, Adv. Oxide Sulfide Miner.
Surf. Chem. 67, 1045–1054. https://doi.org/10.1016/S0016-7037(02)01036-0.

11
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

Liu, C., Huang, P.M., 1999. Atomic force microscopy and surface characteristics of iron oxides formed in citrate solutions. Soil Sci. Soc. Am. J. 63, 65–72. https://doi.
org/10.2136/sssaj1999.03615995006300010011x.
Lu, J., Zhao, H., Zhang, Z., Jin, F., Pang, H., Yang, J., Wang, X., Chen, R., 2023. Control of chemical and microbial pollutants in liquid-solid phases of sewer pipes by Fe
(II)-activated persulfate: Performance and mechanism of simultaneous removal. Chem. Eng. J. 468, 143486 https://doi.org/10.1016/j.cej.2023.143486.
Matzek, L., Carter, K., 2016. Activated persulfate for organic chemical degradation: A review. CHEMOSPHERE 151, 178–188. https://doi.org/10.1016/j.
chemosphere.2016.02.055.
McGachy, L., Sedlak, D.L., 2024. From theory to practice: leveraging chemical principles to improve the performance of peroxydisulfate-based in situ chemical
oxidation of organic contaminants. Environ. Sci. Technol. 58, 17–32. https://doi.org/10.1021/acs.est.3c07409.
McGachy, L., Skarohlid, R., Martinec, M., Roskova, Z., Smrhova, T., Strejcek, M., Uhlik, O., Marek, J., 2021. Effect of chelated iron activated peroxydisulfate oxidation
on perchloroethene-degrading microbial consortium. Chemosphere 266, 128928. https://doi.org/10.1016/j.chemosphere.2020.128928.
Mertens, R., Sonntag, C. von, 1994. Reaction of the OH radical with tetrachloroethene and trichloroacetaldehyde (hydrate) in oxygen-free aqueous solution. J. Chem.
Soc. Perkin Trans. 2, 2181–2185. https://doi.org/10.1039/P29940002181.
Mertens, R., von Sonntag, C., 1995. Photolysis (λ = 354 nm of tetrachloroethene in aqueous solutions. J. Photochem. Photobiol. Chem. 85, 1–9. https://doi.org/
10.1016/1010-6030(94)03903-8.
Mikutta, C., Frommer, J., Voegelin, A., Kaegi, R., Kretzschmar, R., 2010. Effect of citrate on the local Fe coordination in ferrihydrite, arsenate binding, and ternary
arsenate complex formation. Geochim. Cosmochim. Acta 74, 5574–5592. https://doi.org/10.1016/j.gca.2010.06.024.
Oh, W., Lim, T., 2019. Design and application of heterogeneous catalysts as peroxydisulfate activator for organics removal: an overview. Chem. Eng. J. 358, 110–133.
https://doi.org/10.1016/j.cej.2018.09.203.
Padmaja, S., Neta, P., Huie, R.E., 1993. Rate constants for some reactions of inorganic radicals with inorganic ions. Temperature and solvent dependence. Int. J. Chem.
Kinet. 25, 445–455. https://doi.org/10.1002/kin.550250604.
Pestovsky, O., Bakac, A., 2006. Aqueous Ferryl(IV) Ion: kinetics of oxygen atom transfer to substrates and oxo exchange with solvent water. Inorg. Chem. 45,
814–820. https://doi.org/10.1021/ic051868z.
Pullin, M.J., Cabaniss, S.E., 2001. Colorimetric flow-injection analysis of dissolved iron in high DOC waters. Water Res. 35, 363–372. https://doi.org/10.1016/S0043-
1354(00)00259-1.
Rao, D., Dong, H., Niu, M., Wang, X., Qiao, J., Sun, Y., Guan, X., 2022. Mechanistic Insights into the Markedly Decreased Oxidation Capacity of the Fe(II)/S2O82–
Process with Increasing pH. Environ. Sci. Technol. 56, 13131–13141. https://doi.org/10.1021/acs.est.2c04109.
Schwertmann, U., Stanjek, H., Becher, H.-H., 2004. Long-term in vitro transformation of 2-line ferrihydrite to goethite/hematite at 4, 10, 15 and 25◦ C. Clay Min. 39,
433–438. https://doi.org/10.1180/0009855043940145.
Škarohlíd, R., McGachy, L., Martinec, M., Rošková, Z., 2020. Removal of PCE/TCE from groundwater by peroxydisulfate activated with citric acid chelated ferrous
iron at 13 ◦ C. Environ. Technol. Innov. 19, 101004 https://doi.org/10.1016/j.eti.2020.101004.
Stefánsson, A., 2007. Iron(III) hydrolysis and solubility at 25 ◦ C. Environ. Sci. Technol. 41, 6117–6123. https://doi.org/10.1021/es070174h.
Tang, S., Zhu, E., Zhai, Z., Liu, H., Wang, Z., Jiao, T., Zhang, Q., Yuan, D., 2023. Promoted elimination of metronidazole in ferrous ions activated peroxydisulfate
process by gallic acid complexation. Chemosphere 319, 138025. https://doi.org/10.1016/j.chemosphere.2023.138025.
Thornton, A.T., Laurence, G.S., 1973. Kinetics of oxidation of transition-metal ions by halogen radical anions. Part I. The oxidation of iron(II) by dibromide and
dichloride ions generated by flash photolysis. J. Chem. Soc. Dalton Trans. 804–813. https://doi.org/10.1039/DT9730000804.
Truong, G.L., Laat, J.D., Legube, B., 2004. Effects of chloride and sulfate on the rate of oxidation of ferrous ion by H2O2. Water Res. 38, 2384–2394. https://doi.org/
10.1016/j.watres.2004.01.033.
Wacławek, S., Lutze, H.V., Grübel, K., Padil, V.V.T., Černík, M., Dionysiou, D.D., 2017. Chemistry of persulfates in water and wastewater treatment: a review. Chem.
Eng. J. 330, 44–62. https://doi.org/10.1016/j.cej.2017.07.132.
Wang, B.W., Wang, Y., 2022. A comprehensive review on persulfate activation treatment of wastewater. Sci. Total Environ. 831, 24. https://doi.org/10.1016/j.
scitotenv.2022.154906.
Wang, Z., Jiang, J., Pang, S., Zhou, Y., Guan, C., Gao, Y., Li, J., Yang, Y., Qiu, W., Jiang, C., 2018. Is sulfate radical really generated from peroxydisulfate activated by
Iron(II) for environmental decontamination? Environ. Sci. Technol. 52, 11276–11284. https://doi.org/10.1021/acs.est.8b02266.
Wang, Z., Qiu, W., Pang, S., Jiang, J., 2019. Effect of chelators on the production and nature of the reactive intermediates formed in Fe(II) activated peroxydisulfate
and hydrogen peroxide processes. Water Res 164, 114957. https://doi.org/10.1016/j.watres.2019.114957.
Wen, R., Shen, G., Meng, L., 2023. Research progress of metal–organic framework-based material activation of persulfate to degrade organic pollutants in water. RSC
Adv. 13, 24565–24575. https://doi.org/10.1039/D3RA04296K.
Wu, X., Gu, X., Lu, S., Xu, M., Zang, X., Miao, Z., Qiu, Z., Sui, Q., 2014. Degradation of trichloroethylene in aqueous solution by persulfate activated with citric acid
chelated ferrous ion. Chem. Eng. J. 255, 585–592. https://doi.org/10.1016/j.cej.2014.06.085.
Xie, J., Yang, C., Li, X., Wu, S., Lin, Y., 2023. Generation and engineering applications of sulfate radicals in environmental remediation. Chemosphere, 139659.
https://doi.org/10.1016/j.chemosphere.2023.139659.
Xu, Y., Zeng, L., Li, L., Chang, Y.-S., Gong, J., 2019. Enhanced oxidative activity of zero-valent iron by citric acid complexation. Chem. Eng. J. 373, 891–901. https://
doi.org/10.1016/j.cej.2019.05.093.
Xu, Z., Cai, L., Liang, X., Lyu, S., 2022. Insight into trichloroethene removal in alkaline condition with the presence of surfactant based on persulfate system.
J. Environ. Chem. Eng. 10, 108492 https://doi.org/10.1016/j.jece.2022.108492.
Yao, P., Wang, Z., You, A., Hu, G., 2023. Degradation and transformation of sulfapyridine in Fe2+/peroxydisulfate: Kinetics, mechanism and potential products.
J. Water Process Eng. 53, 103808 https://doi.org/10.1016/j.jwpe.2023.103808.
Yoon, S.-E., Kim, C., Hwang, I., 2022. Continuous Fe(II)-dosing scheme for persulfate activation: Performance enhancement mechanisms in a slurry phase reactor.
Chemosphere 308, 136401. https://doi.org/10.1016/j.chemosphere.2022.136401.
Zeng, G., Yang, R., Fu, X., Zhou, Zhengyuan, Xu, Z., Zhou, Zhikang, Qiu, Z., Sui, Q., Lyu, S., 2021. Naphthalene degradation in aqueous solution by Fe(II) activated
persulfate coupled with citric acid. Sep. Purif. Technol. 264, 118441 https://doi.org/10.1016/j.seppur.2021.118441.
Zhang, H., Jiang, M., Su, P., Lv, Q., Zeng, G., An, L., Cao, J., Zhou, Y., Snyder, S.A., Ma, J., Yang, T., 2024. Refinement of kinetic model and understanding the role of
dichloride radical (Cl2•− ) in radical transformation in the UV/NH2Cl process. Water Res, 121440. https://doi.org/10.1016/j.watres.2024.121440.
Zhang, H., Su, X., Sun, B., Xu, Y., Gong, J., 2021. Citrate iron complex induced dramatically enhanced oxidation of atrazine with bimetallic Bi/Fe0: Reactivity,
oxidation and mechanism. Chemosphere 282, 131100. https://doi.org/10.1016/j.chemosphere.2021.131100.
Zhang, R., Fang, X., Liu, Y., Li, M., Zeng, G., Yang, R., Qiu, Y., Lyu, S., 2024. Comparison of naphthalene degradation by hydrogen peroxide, nano-calcium peroxide,
peroxydisulfate, and peroxymonosulfate in the Fe(II)-citric acid catalytic environments. Water Supply, ws2024004. https://doi.org/10.2166/ws.2024.004.
Zhang, Y., Zhou, M., 2019. A critical review of the application of chelating agents to enable Fenton and Fenton-like reactions at high pH values. J. Hazard. Mater. 362,
436–450. https://doi.org/10.1016/j.jhazmat.2018.09.035.
Zhao, Q., Mao, Q., Zhou, Y., Wei, J., Liu, X., Yang, J., Luo, L., Zhang, J., Chen, Hong, Chen, Hongbo, Tang, L., 2017. Metal-free carbon materials-catalyzed sulfate
radical-based advanced oxidation processes: A review on heterogeneous catalysts and applications. Chemosphere 189, 224–238. https://doi.org/10.1016/j.
chemosphere.2017.09.042.
Zhou, Z., Huang, J., Xu, Z., Ali, M., Shan, A., Fu, R., Lyu, S., 2021. Mechanism of contaminants degradation in aqueous solution by persulfate in different Fe(II)-based
synergistic activation environments: Taking chlorinated organic compounds and benzene series as the targets. Sep. Purif. Technol. 273, 118990 https://doi.org/
10.1016/j.seppur.2021.118990.

12
R. Škarohlíd et al. Environmental Technology & Innovation 35 (2024) 103687

Zhu, C., Zhu, F., Dionysiou, D.D., Zhou, D., Fang, G., Gao, J., 2018a. Contribution of alcohol radicals to contaminant degradation in quenching studies of persulfate
activation process. Water Res 139, 66–73. https://doi.org/10.1016/j.watres.2018.03.069.
Zhu, C., Zhu, F., Liu, C., Chen, N., Zhou, D., Fang, G., Gao, J., 2018b. Reductive hexachloroethane degradation by S2O8•– with thermal activation of persulfate under
anaerobic conditions. Environ. Sci. Technol. 52, 8548–8557. https://doi.org/10.1021/acs.est.7b06279.
Zhu, M., Frandsen, C., Wallace, A.F., Legg, B., Khalid, S., Zhang, H., Mørup, S., Banfield, J.F., Waychunas, G.A., 2016. Precipitation pathways for ferrihydrite
formation in acidic solutions. Geochim. Cosmochim. Acta 172, 247–264. https://doi.org/10.1016/j.gca.2015.09.015.

13

You might also like