Skip to main content
Ultrasonics Sonochemistry logoLink to Ultrasonics Sonochemistry
. 2021 Sep 10;78:105750. doi: 10.1016/j.ultsonch.2021.105750

Ultrasonic role to activate persulfate/chlorite with foamed zero-valent-iron: Sonochemical applications and induced mechanisms

Qihui Xu a, Hong Zhang b, Haoran Leng a, Hong You a,b,, Yuhong Jia b, Shutao Wang a
PMCID: PMC8455865  PMID: 34544014

Highlights

  • The novel foamed-ZVI activating persulfate/chlorite by ultrasonic was applied.

  • The real-time and online continuous detections tracked radicals transformation.

  • The EPR spectra with different systems in different mixtures were detected.

  • The free or non-free radicals processes induced by US were explored.

  • The dynamics on cavitation and the US-field simulation were presented.

Keywords: Persulfate/chlorite composite oxidants, Ultrasonic role, Polishing zero-valent iron foam, Determining active species

Abstract

The novel system, consisting of composite oxidants (persulfate/chlorite, S2O82−/ClO2) and stationary phase activator (zero-valent-iron foam, Fe0f) driven by ultrasonic (US) field, was applied to treat the triphenylmethane derivative effectively even at low temperature (≈ 289 K). By comparisons of sub-systems, the US roles to S2O82−, ClO2, and Fe0f were seriatim analyzed. US made the reaction order of multi-component system tend to within 1 (leading to de-order reaction), and widened pH activating range of the Fe0f by sonicate-polishing during the process of ClO2 co-activating S2O82−. US and Fe0f were affected by fluid eddy on activating S2O82−/ClO2. The Fe0f had slight effect on the temperature of US bubble-water interface but the addition of ClO2 lowered it. The partitioning capacity of the above US reactive zone increased during the reaction. US and ClO2 could enrich the kinds of degradation intermediates. The contributions of free radicals (ClOx-based radicals, sulfate radicals (SO4), and hydroxyl radicals (•OH)) and non-free radicals (ClO2, and O = FeIV/V from ionic Fe under “-O-O-” of S2O82− and cyclic adjustment reaction of ClO2) processes by sonochemical induction were equally important by corresponding detection means. Especially, real-time and online high-resolution mass spectrum by self-developing further confirmed the chain transfers of different free radicals due to US role. The findings expanded the application of sono-persulfate-based systems and improved understanding on activation mechanism.

1. Introduction

A series of trace triphenylmethane derivatives detected in aquatic environment with low concentration (<μM level) as emerging contaminants have drawn the attention [1], therein no lack of triphenylmethane crystal-violet (tpmCV). Besides as the dye, the tpmCV (Fig. S1a), could be used not only in clinical treatment, but also as antimicrobial and antiparasitic agents in aquaculture to treat and prevent fungal and protozoal infections [2]. However, the United States, the European Union, China, and others have not approved the use of tpmCV in aquaculture due to it with the possible carcinogen, mutagen and teratogen effects [1], [2], [3]. Even so, the use of the prohibited and persistent tpmCV would be common occurance, leading to its frequent detection in aquatic environment. Specially, the bio-concentration of organisms made tpmCV convert into leuco crystal-violet (LtpmCV, Fig. S1b) (more toxic and more difficult to be metabolized) [1], [3], which might be seen as the persistent pollutants.

H2O))))OH+H (1)

The alternative US-based advanced oxidation processes (AOPs), as the promising and friendly treatment technologies with physicochemical processes of cavitation (gas-phase reaction zone effective temperature, ≈ 5200 K, and high pressure, in the range of hundreds of bars), pyrolysis (breaking chemical bonds, and yielding hydroxyl radicals (Eq. (1)) [4]), supercritical water oxidation (depth and temperature of the liquid-phase reaction zone, ≈ 0.2 μm and ≈ 1900 K, respectively), and violent turbulence [5], [6]; meanwhile, especially as the sono-catalysis technologies to activate oxidizing agents to produce active species to form combined system [5], [6], [7], have attracted particular concern for environmental remediation. Herein, as a kind of oxidizing agent, persulfate (involving peroxymonosulfate (HSO5, PMS) or peroxydisulfate (S2O82−, PDS)) with direct oxidizing or producing SO4 (E0 (SO4/SO42−) = 2.60 – 3.10 VNHE) by activation to improve the oxidizing capacity [8] via physicochemical means (e.g. alkali, Eq. (2); heat, Eq. (3) [9]; photo-; electro-; microwave; and US, Eq. (4)), metal-free catalysis [10], metal-based materials (e.g. zero-valence metal, Eq. (5)) [11], and other methods (e.g. ozone, superoxide, and plasma), could be used to treat environmental contaminants as well.

2S2O82-+2H2OOH-SO4-+O2-+3SO42-+4H+ (2)
S2O82-heat2SO4-k=1.0×10-7s-1(298K)or5.7×10-5s-1(343K) (3)
S2O82-hv/))))2SO4- (4)
Mn++S2O82-Mn+1++SO42-+SO4- (5)

Therefore, further works for sono-persulfate based AOPs were worth exploring to extend acoustic catalysis and persulfate activation bilaterally, seeming to become prospective treatment methods for the elimination of refractory organic contaminants above.

At present, the zero-valent-iron (ZVI, Fe0)-based materials in disperse mediums by modified techniques (pre-magnetized [12], sulfidated [13], loaded/embedded [14], micro/nano [15]) or in stationary phase ZVI (e.g. foamed-ZVI (Fe0f) consisting of three-dimensional porous form (Fig. S1c) with properties [16], [17] of permeability, controlled pore size, foveolate structure, shape stability, and machinability) as the activated intermediary to connect the US and persulfate, have obtained great attention. Herein, the Fe0f for oxidant activation was rarely reported, however, functional materials based on Fe0f have drawn important attention, such as, electrode materials for hydrogen evolution reaction or oxygen evolution reaction (electrocatalytic water splitting) [17], which also had greater potential for environmental remediation. However, the surface passivation of ZVI has been a fatal problem affecting its activation to persulfate. Nevertheless, it’s just that introducing US could polish the above ZVI in activation process, and ease the burden of passivation to provide the fresh ZVI.

In addition, in order to promote corrosion of ZVI and regulate the activation cycle (e.g. ionic Fe cycle) under the action of US, the chlorite (ClO2, Cl(III)) could be introduced into the aforementioned process and combined with persulfate to form composite oxidants. Commonly, ClO2 with a higher quantum yield at 254 nm, 1.0–1.53 mol⋅Einstein-1 [18], [19], comparing with H2O2 (0.5 mol⋅Einstein-1) and persulfate (0.7 mol⋅Einstein-1) [20], could be activated by physical fields including UV-light stimuli ([18], Eq. (6)), heat [21], and US, or chemical agents including acid (Eq. (7)) and PDS [22] to produce chlorine dioxide (ClO2).

3ClO2-+H2Ohv2ClO2+Cl-+2OH-+0.5O2 (6)
5ClO2-+4H+Cl-+4ClO2+2H2O (7)

The ensuing surface passivation of ZVI (Cl role) may impede its further activation, however, introducing US, as a “bridge”, could not only activate composite oxidants, persulfate/chlorite (OSC), to generate SO4 and ClO2 by co-catalysis (Eq. (8)) [23], but also polish the Fe0f to avoid aforementioned Fe0f passivation and to activate OSC deeply (Eqs. (9)-(13)).

S2O82-+ClO2-))))SO4-+SO42-+ClO2 (8)
Fef0+S2O82-FeII+2SO42- (9)
FeII+S2O82-Fe(aq&s)III+SO4-+SO42- (10)
FeII+ClO2Fe(aq&s)III+ClO2- (11)
4FeII+ClO2-+2H2O4Fe(aq&s)III+Cl-+4OH- (12)
2Fe(aq)III+polishedFef03FeII (13)

In this study, the novel treatment systems, namely, the combined system (US/Fe0f-OSC) based on co-catalysis involving acoustic catalysis and multiple activation of composite oxidants, and sub-systems, were established. The commonly apparent effects of US may have been mentioned, however, because of that US has always been seemed as just an assistant method in environmental remediation, the researches on its applied potential (e.g. couple function as a “bridge” of systems above by series connection) and feasible mechanism of US role (e.g. US role to transformations of generated active species) may be neglected.

This study was therefore designed to (1) evaluate applications of the combined- and sub- systems based on US from comparisons of performances, affecting parameters, and pollutant degradation paths; (2) reveal the detailed sonochemical roles in apparent remediation by comparisons of coupling systems; and (3) explore the roles of US-induction in remediation mechanism by comparisons of free radicals and non-free radicals processes.

2. Experiment

2.1. Operating

The operating processes including chemicals, self-developed US/catalytic/activated reactors (UCARs), etc. were presented in Section S1.

2.2. Analytical methods

The intermediates of tpmCV and methyl phenyl sulfoxide (PMSO) were detected by gas chromatograph-mass spectrometer (GC/MS, Trace 1300 ISQ-QD, Thermo Fisher). The temperature programming (GC/MS) was: column temperature 323 K, hold time 3 min; 10 K·min−1 up to 573 K, hold time 10 min. Before the above analysis, 50 mL subsample at a certain treating time (5, 15, 25, 35, and 45 min) was extracted with 15 mL dichloromethane, respectively. Afterward, the extracted solution by mixing was dehydrated by anhydrous sodium sulfate and concentrated to 2 mL. The ClO2 and tpmCV concentration were detected by UV–vis spectrum (TU-1810S, PGENERAL, China) at 360 nm [19] and 584 nm, respectively. The concentrations of “Fe” were determined by phenanthroline spectrophotometry at 510 nm (HJ/T 345–2007). The accurate acoustic-intensity entering the system was measured by the calorimetry method [24], presenting in Section S2 (Eq. S1). The compositions and micrographs of the Fe0f, before and after treatment, were detected and examined by X-ray diffraction (XRD, HAOYUAN Instrument, DX-2700) and scanning electron microscopy (SEM, Zeiss, Germany), respectively. Before XRD and SEM, the Fe0f samples taken out from reaction solution needed dehydration immediately by adding into absolute ethyl alcohol, then drying out with filter papers, finally vacuum drying. The anions of S2O82−/ClO2 were detected by ion chromatography (IC, ThermoScientific, Aquion).

2.3. Spin trapping and EPR measurements

The US filed was applied in solution consisting of NaClO2 (1–20 mM) and K2S2O8 (ClO2: S2O82− = 1:5 in molar ratio) to conduct sonochemical experiments, then Fe0f was added as needed. DMPO with equal proportion of OSC was utilized to trap the generated active species above. In addition, dimethyl sulfoxide (DMSO, as needed) was added into water to form mixture for comparison.

The electron paramagnetic resonance (EPR) spectra were obtained by a Bruker A200 spectrometer. General instrument settings were as follows unless otherwise noted: center field 3490 G, modulation frequency 100 kHz, modulation amplitude 1 G, receiver gain 1.42 × 104, time constant 20.5 ms, conversion time 40 ms, sweep time 41 s, frequency 9.82 GHz, and power 19.47 mW. The hyperfine splitting constants were measured by using Bruker Win-EPR SimFonia (Version 1.2).

2.4. Real-time and online continuous detection

In Scheme 1, the real-time and online continuous (ROC) detections of reaction mixtures were realized with a self-developed pneumatic nebulization (PN)-droplet spray ionization (DSI) (electrospray ionization (ESI) with CH3OH as an assist as required) coupled to LTQ and LTQ-Orbitrap high resolution mass spectrum (HRMS) (Thermo Scientific) in positive or negative mode with a resolution of 60 000. Online MS/MS analysis of the captured intermediates could be performed during the reaction process. Data were acquired in full scan in the m/z range 50–300 and in MS/MS with a collision-induced dissociation energy of 25. The instrument was calibrated routinely with the provided calibration solution by Thermo Scientific to ensure mass accuracy within ± 3 ppm.

Scheme 1.

Scheme 1

Schematic diagram of ROC detection device.

The self-developed PN-DSI as Scheme 1 shown: the sealed glass device (7 mL) was equipped with mixture (5 mL) of S2O82− at 5 × 10-6 M, ClO2 at 1 × 10-6 M, and DMPO at 1 × 10-4 M under US generator; then the liquid along the capillary tube entered into the site of triple valve by N2 inflating; and the charged samples from the above liquid, suffering the action of PN (its description in Section S3) by another N2 flow, were transmitted into HRMS.

2.5. Simulation on US reactor

The Acoustic-Piezoelectric Interaction of COMSOL Multiphysics, including pressure acoustics Eq. (14), was applied to simulate heterogeneous reaction process and visualize the action of US-field in the solution. The detailed simulation process were described in Section S4 (Eqs. S2-S25). The UCAR2 was simulated as shown in Fig. 1a-c and Fig. S2a-g.

-1ρcpt-qd-keq2ptρc=Qm (14)

Fig. 1.

Fig. 1

Simulation of acoustic pressure distribution and piezoelectric mechanics deformation at 28 kHz with Fe0f: (a) sound pressure level, (b) vertical slice, (c) horizontal slice; and comparisons of OSC (6 mM)-based systems and PDS (6 mM)-based systems: (d) removal efficiencies, (e) fitting kinetics (molar ratioS2O82−/ClO2 5:1, f 28 kHz, I 30 W⋅L-1, pH without adjustment (initial pH 6.5, near-neutral), Fe0f 0.358 mM, tpmCV 26.47 μM, 0 r⋅min−1, and initial T 289 K).

3. Results and discussion

3.1. Comparisons of US-based systems

At low temperature (≈ initial 289 K), low pollutant concentration (10-6 M, μM level), and near-neutral pH (without adjustment), the novel US/Fe0f-OSC presented good performance on removing tpmCV. OSC, consisting of S2O82− and ClO2 with molar ratio 5:1, was sourced from our previous study. From Fig. 1d, the treating efficiencies of tpmCV by US/OSC (6 mM)-based systems were greater than those by US/PDS (6 mM)-based systems, due to the degradation of tpmCV in US/PDS-based systems presenting weak implementation on activation in later stage (e.g. after 15 min).

The exponential analysis of apparent dynamics could be utilized to further understand the removal process of tpmCV (details in Section S5, Eqs. S26-S28). By fitting concentration functions of the systems mentioned above, the optimal fitting orders (non-integer, R2Adj > 0.999, Fig. 1e) of OSC, Fe0f-OSC, US-OSC, and US/Fe0f-OSC systems were 1.35, 1.10, 0.90, and 0.75, respectively.

In order to further measure the synergistic effect among the above-mentioned systems (Fig. 1e) conveniently, the S and fi (details in Section S6, Eqs. S29-S33) were calculated to evaluate interactions of each OSC-based system with pseudo-first-order kinetics (R2Adj > 0.990) as below (kobs values (10-3, min−1) of OSC, Fe0f-OSC, US-OSC, and US/Fe0f-OSC systems: 46.82, 51.56, 57.22, and 85.64, respectively).

S=kobsnkobsi (15)
fi=kobsnkobsn-1 (16)

where S is the synergistic effect of the multicomponent system, fi is the action of single i system, and kobs represents pseudo-first-order kinetic constant unless otherwise indicated.

The S was 1.83 (>1), and the fUS and fFe0f in US/Fe0f-OSC (1.66, 1.50) > those in US-OSC (1.22, 1.10), which indicated that: i) US/Fe0f-OSC had well synergistic effect; ii) The action of US was slightly better than that of Fe0f; iii) Compared with dual systems, both action of US and Fe0f enlarged in ternary systems (further confirming their synergistic effect).

In addition, the adding of US (or ClO2) may enrich the varieties and amounts of active substances in the reaction system, which greatly weakened the influence of pollutant concentration, leading to de-order reaction, but Fe0f alone seemed to play not obvious role on the above.

3.2. Effects of factors

3.2.1. Effect of pH with/without US

The pH values greatly affected the systems of Fe0f-OSC and US/Fe0f-OSC on removing tpmCV (Fig. 2a). Generally, the acidic condition was beneficial for degradation. After introducing US, organic pollutants could evaporate into the cavitation bubbles easily, and then be pyrolyzed directly in them under acid condition. Since pH decreased from 11 to 2, the kobs (10-3, min−1) increased from 41.37 to 345.80 (from 38.54 to 132.73) of US/Fe0f-OSC (Fe0f-OSC). However, due to US role, the pH application rage of system widened (specifically, the kobs of Fe0f-OSC at pH 2 ≈ that of US/Fe0f-OSC at pH 4; the kobs of Fe0f-OSC at pH 6.8 ≈ that of US/Fe0f-OSC at pH 10.1) as shown in embedded Fig. 2a. Meanwhile, when pH presented alkaline, the reaction rate both in Fe0f-OSC and US/Fe0f-OSC seemed to be accelerated, but it would decelerate. The greater the alkaline, the more obvious the above-mentioned phenomenon, which could be ascribed to ClO2 (alkali activation to form ClO• being different from acidic reaction to generate ClO2) [20], [22]. The S2O82− was dominant, thus the alkali role to ClO2 could not sustain over the whole reaction process. Due to US polishing (comparing the Fig. 3) and ClO2 co-catalysis, the systems actually kept certain ability of treatment, and the passivation of Fe0f obtained remission.

Fig. 2.

Fig. 2

Effects of (a) acid-base (comparisons of US/Fe0f-OSC and Fe0f-OSC), (b) fluid eddy (ibid. (a)), (c) temperature, and (d) ultrasonic intensity & frequency (basic conditions referring to Fig. 1).

Fig. 3.

Fig. 3

Comparisons of Fe0f with SEM (1 μm, 2 μm, and 20 μm) and XRD (10-90°): (a) original sample without treating, (b) with US-OSC 30 min, and (c) & (d) with OSC but without US, 15 min & 30 min (molar ratioS2O82−/ClO2 5:1, f 28 kHz, I 30 W⋅L-1, pH without adjustment, 0 r⋅min−1, and initial T 289 K).

3.2.2. Effect of fluid eddy with/without US

The fluid eddy was realized by mechanical mixing. The liquid height [4] was determined based on the wavelength (λ) of the applied frequency (f), namely, λ = cs/f where cs was obtained as 1469.60 m⋅s−1 with the modified temperature equation (Eq. S34). The site of mechanical mixing was placed at 1/2λ (distance from bottom) [5]. In order to describe the inherent attribute of fluid by the action of mechanical mixing, herein, the angular momentum (L), moment of inertia (Iv), and rotational kinetic energy (Erot) were introduced to describe the role of fluid eddy with following Eqs. S35-S37 (Section S7).

By comparison of Fe0f-OSC and US/Fe0f-OSC systems in Fig. 2b, the fluid eddy both affected the removal of tpmCV varying in three stages (increased → decreased → restrained), which was partly consistent with the study of [5]. The difference was that the introduction of OSC (especially, ClO2) and Fe0f led to the appearance of restrained stage. Enlarging the fluid eddy made S2O82− and ClO2 react preferentially to consume themselves by Eq. (17) (sonic catalysis weakening) and accelerate Fe0f surface oxidation (OSC and O2) and even passivation.

S2O82-+2ClO2-2SO42-+2ClO2 (17)

Due to the action of US, the peak-shift (a right-ward shift) occurred in terms of variation of kobs. The enhancement in the sonochemical oxidation was attributed mainly to the direct disturbance of the ultrasound transmission and the resulting change in the cavitation-active zone.

3.2.3. Effect of initial temperature with US

The temperature affected the performance of US/Fe0f-OSC on removing tpmCV (kobs, (10-3, min−1) increasing from 85.64 to 309.47 when T increasing from 289 K to 318 K in Fig. 2c). The reaction thermodynamics of tpmCV removal was described by Eyring equation (Eqs. (18), (19)) and Arrhenius equation (Eq. (20)) to obtain the experimentally derived free energy of activation (ΔG), fitted enthalpy (ΔH), entropy (ΔS) of activation, and activation energy (Ea), respectively.

kobs=kbThe-(ΔG/RT) (18)
ΔG=ΔH-T×ΔS (19)
kobs=Ae-(Ea/RT) (20)

where R is the gas constant (8.314 J·mol−1·K−1); kb is the Boltzmann constant (1.38 × 10-23 J·K−1); h is the Planck’s constant (6.626 × 10-34 J·s); and the above equations were extended in Section S8 (Eqs. S38 and S39).

As shown in Fig. 2c, the liner relation in ln(kobs/T) vs. 1/T (R2Adj = 0.984) and ln(kobs) vs. 1/T (R2Adj = 0.985). Therefore, the ΔH, ΔS, Ea, and A were calculated as 31.117 kJ·mol−1, −156.956 J·mol−1·K−1, 33.227 kJ·mol−1, and 91.858 × 103, respectively. Therefore, for US/Fe0f-OSC system, the relationships (ΔG vs. T and kobs vs. T) were expressed as ΔG= 31.117 + 0.157 T (kJ·mol−1) and kobs = 9.186 × 104exp(–33.227 × 103/RT) (min−1).

3.2.4. Effect of US intensity and frequency

The effects of ultrasonic intensity and frequency (Fig. 2d) were discussed in Section S9.

3.3. Comparative analysis of degradation intermediates with/without US

In terms of US/Fe0f-OSC, mono-(S1-S13), di-(D1-D9), and tri-(T1-T4) phenylic compounds (Table S1) were detected. Comparing with double-systems Fe0f-OSC and US-OSC, the US/Fe0f-OSC seemed to make the intermediates more diversified (Fig. S3) [25], especially, to produce more Cl-adducts (because of ClO2) [22] and mono-benzene compounds (Fig. 4a) [26]. The detailed analysis was shown in Section S10.

Fig. 4.

Fig. 4

(a) Comparisons of intermediates detected by GC/MS under different systems and possible degradation pathway of tpmCV on US/Fe0f-OSC (S: single benzene ring, D: double benzene rings, T: three benzene rings, Z: straight chain compounds, SMC: small molecule compounds (organic acid, CO2, H2O, etc.), the detailed intermediates with tR and m/z seeing in Table S1); (b) GC/MS detection of PMSO and PMSO2 (initial PMSO 6 mM) (the detailed intermediates with tR and m/z seeing in Table 1); and (c) the transformations of PMSO in US/Fe0f-OSC and Fe0f-OSC.

3.4. Assessing US bubble/water interface

The kobs could be measured at the macro-level based on the total UCAR solution. However, in order to obtain concrete rate constants for catalyactivation according to effectively reactive volume (kint at bubble–liquid interface zone of cavitation nucleus where the reaction sites for catalysis and activation of US (or Fe0f) to OSC (or PDS) mainly located), the mass balance expression, as Wei et al. [27] mentioned, could be used to make a conversion from kobs to kint:

kint=VtotalVintkobs (21)

where kint was the rate constant at the phase interface, Vtotal was the volume of wastewater in treatment, and Vint was the volume of interfacial shell surrounding cavitation bubbles. Here, assuming the interface thickness was 10% of the radius of the cavitating bubble: an empirical value of Vtotal/Vint was obtained from the calculation of [27].

The effective mean temperature of the interfacial region was estimated by kint and Arrhenius equation. The Arrhenius parameters (lnA = 36.6 and Ea = 134 kJ·mol−1) were referred to the study of thermal dissociation of PS [27], [28]. Therefore, at optimal degradation order, the corresponding temperatures of US/Fe0f-OSC, US-OSC, US/Fe0f-PDS, and US-PDS were calculated to be 531 K, 535 K, 737 K, and 744 K, respectively, which illustrated that i) Fe0f had slight effect on the interface temperature; ii) ClO2 lowered the interface temperature. Although the interfacial zone would have a steep temperature gradient between the hot cavitation bubble and the ambient temperature of the bulk solution [27], the calculated temperatures above could be considered as the mean effective temperatures of the interfacial region for OSC (or PDS) dissociation (activated by Fe0f under the US roles), but < 1900 K found in the liquid region surrounding the collapsing cavitation bubbles in alkane solvents at 5 Torr vapor pressure under argon [29], [30]. As Misik et al. [31] described, the contaminants with different structures may also influence the interface temperature at sonochemical regions.

In addition, the calculated Ea of US/Fe0f-OSC was 33.227 kJ·mol−1 > 10–13 kJ·mol−1, indicating that degradation process was dominated by an intrinsic chemical reaction rather than a diffusion process. According to study of Cui et al. [32], the modified Freundlich Isotherm (Eq. (22)) could be applied to describe the relationship between the amounts of molecules partitioning (adsorption) at the interface of cavitation bubble and the initial concentration at equilibrium in US-based systems.

logrt=1krlogctpmCV,0+log(Kf) (22)

where rt (µM·min−1) was the decomposed rate of tpmCV (Eq. S40), kr was called as the reactivity constant, and Kf was the adsorption capacity. According to the calculating of rt, the Section S11 was presented as the general form.

Therefore, according to US/Fe0f-OSC, the log(r0) vs. log(ctpmCV,0) had a good liner relation (R2Adj = 0.996, Fig. 5a), and the kr,0, indicating the adsorption intensity, was 0.823, while the Kf,0, as [32] described, represented the capacity of tpmCV molecules adsorbing or partitioning to the reactive zone (gas–liquid interface) and was 0.0442. Similarly, the kr,end was 0.728, while the Kf,end was 0.0659 (R2Adj = 0.987). By comparison, the adsorption or partitioning capacity of tpmCV to the reactive zone was increased as the degradation progressed.

Fig. 5.

Fig. 5

(a) Freundlich degradation isotherms of tpmCV in Fe0f-OSC solution media under US irradiation (f 28 kHz, I 30 W⋅L-1), (b) inhibition of different systems via screening agents (6 M MA and 1 M TBA), (c) variation of various Fe concentrations in US/Fe0f-OSC, and (d) ClO2 detection of US-OSC by UV–vis spectrum (T 303 K).

3.5. Discussion of degradation mechanism induced by US

3.5.1. Contribution of main active species with/without US

The reaction rate of tert butyl alcohol (TBA) vs. •OH (3.8–7.6 × 108 M−1·s−1) was much higher than that of TBA vs. SO4 (4.0–9.1 × 105 M−1·s−1) [33], while the reaction rate of methyl alcohol (MA) vs. •OH (9.7 × 108 M−1·s−1) was close to that of MA vs. SO4 (2.5 × 107 M−1·s−1) [34]. The results of inhibition reactions of different systems with different scavengers were shown in Fig. 5b and Fig. S4. According to Eq. (23), the relative degrees of dependence in •OH or SO4 could be obtained.

δi=Δkobsikobs(none) (23)

where δi represented the shielding effects of different systems with different scavengers, Δkobs, i represented the difference values of different scavengers, and i represented MA or TBA.

By comparison of δ (MA 6 M, TBA 1 M, relatively enough):

i) δMA-(Fe0f-OSC) (0.692) ≈ δMA-(OSC) (0.691): both generated •OH and SO4 of OSC and Fe0f-OSC could be controlled by 6 M MA.

ii) δTBA-(Fe0f-OSC) (0.283) ≪ δTBA-(OSC) (0.528) but kobs-(Fe0f-OSC) > kobs-(OSC): some other active species (maybe FeIV/V) existed.

iii) δMA-(US-OSC) (0.805) > δMA-(OSC) and δTBA-(US-OSC) (0.594) > δTBA-(OSC): the introduction of US could promote the system to generate more both of •OH and SO4.

iiii) δMA-(US/Fe0f-OSC) (0.515) < δMA-(all the others) but δTBA-(US/Fe0f-OSC) (0.643) > δTBA-(all the others): the •OH and SO4 were also the main active species of systems, but the introduction of US tended to promote the active species to transform to more •OH.

3.5.2. Variation of common Fe and detection of ClO2 with/without US

The variations of the dissolved Fe (FeII and FeIII), suspending Fe & dissolved Fe, and Fe weight loss (suspending Fe, dissolved Fe, and sedimentary Fe) in US/Fe0f-OSC were shown at Fig. 5c. The dissolved Fe seemed to exist a dynamic (fluctuant) equilibrium, indicating a ferrikinetics utilization process. By comparisons of Fe0f morphology changes among Fe0f-OSC and US/Fe0f-OSC (SEM, Fig. 3), and Fe0f-PDS and US/Fe0f-PDS (previous study), US also had the polishing role in US/Fe0f-OSC, but more serious Fe-corrosion in OSC indicated that ferrikinetics may be not sourced from Fe-self circulation (different from PDS), which be mainly attributed to the ClO2 role. ClO2/generated-ClO2 took part in the reaction of Fe, enriching the transformation ways, due to US polishing role, to not only ease Fe0f surface passivation but also accelerate Fe0f surface corrosion (Fig. 3b). But interestingly, without US polishing, the Fe corrosion and formation process of FexOy on Fe0f surface could be observed under enhanced role of OSC (Fig. 3a smooth → Fig. 3c rough → Fig. 3d “flower”-like). Meanwhile, the relative intensities of characteristic peaks (three) of Fe0f were also changed under OSC.

In addition, the ClO2 was detected by UV–Vis spectrum (ClO2 different from other chlorines in water, existing absorption peak around 360 nm) [20], [21]. From Fig. 5d, the produced ClO2 presented throughout the reaction process (gradually increasing) on US-OSC by the reaction of Eq. (17). In addition, the anions of S2O82−/ClO2 after reaction consisted of ClO2, Cl, SO42−, and a few ClO3 (Fig. S5) by IC detection.

3.5.3. ROC-HRMS and EPR detection with/without US

3.5.3.1. ROC-HRMS detection

The spin trapping of DMPO•-SO4 was the predominant O-centered adduct for SO4, and the adduct could be protonated and decomposed over time to DMPO•–OH [35].

According to US-OSC system (at negative mode), the peak at m/z 209 was detected by LTQ and LTQ-Orbitrap, respectively. The MS/MS results of m/z 209 in US-OSC system was totally different to those in background, and the peak (m/z 209) in background was weak and sporadic but in US-OSC system was stable by parallel experiments, thus indicating that the above peak in background belonged to signal noise but in US-OSC system was a new peak. By MS/MS results, the peak at m/z 209 was much close to DMPO•-SO4 (Interestingly, the peak at m/z 209 being observed in both positive and negative modes with similar but different MS/MS results). Especially, the fragment “m/z 97” (failing to capture MS/MS/MS result) in m/z 209 in negative mode implied an unstable whole structure just like “-OSO3(H)” [36], [37], [38]. In addition, the unstable peaks at m/z 209(+1, +2) with Δ m/z (peaks at m/z 209(+1, +2) and 112(+1)) around 97(-1, +1), would appear alternately in positive mode (inferring as Scheme 2b).

Scheme 2.

Scheme 2

Transformations of DMPO-R, -R[35], [40], [41], [42].

PN-HRMS analysis also confirmed the presence of a peak at m/z 130 (only in positive modes) by MS/MS results, which could be explained with DMPO•–OH after abstraction of •H from another molecule in the sample mixture (observed mass was shifted at + 1 Da relative to the mass of DMPO•–OH [39]) (inferring as Scheme 2a). Previous studies associated m/z 130 with an adduct between the DMPO and •OH [40], [41], [42]m/z 18 between m/z 130 and 112 representing de –OH and -H), which was also observed in EPR in our experiments.

As continuous reaction and detection, the peak at m/z 209 gradually weakened even disappeared, but the peak at m/z 130 enhanced (compared with Fig. 6b & 6d, the peak could be clearly captured in LTQ-Orbitrap over time). Meanwhile, the ESI with the adding MA (compared with Fig. 6c & 6d) attempted to enhance the effect of peaks appearance but resulted in that peaks at m/z 130 and 209 were weaken, and other peaks remained unchanged or indeed enhanced. Herein, two points (mutual supporting) could be further obtained: i) MA indeed could quench both of SO4 and •OH; and ii) peaks at m/z 130 and 209 most likely corresponded to DMPO•–OH and DMPO•-SO4, respectively.

Fig. 6.

Fig. 6

ROC detection by MS or MS/MS (molar ratioS2O82−/ClO2 5:1, pH without adjustment): (a) OSC, positive mode, LTQ; (b) US-OSC, positive mode, LTQ& LTQ-Orbitrap (long-time); (c) US-OSC, negative mode, LTQ& LTQ-Orbitrap; and (d) US-OSC, positive mode, LTQ& LTQ-Orbitrap (note: i. operating parameters in Section 2.4; ii. MS/MS results in embedded figure; iii. here presenting comprehensive MS results of continuous period not representing one time point, and continuous period ranging from 3 to 15 min).

According to OSC system (Fig. 6a), however, the peak (m/z 209) was clearly detected but the peak (m/z 130) was hardly detected, suggesting that the main free radicals of OSC were SO4. Actually, because of the Scheme 3 [35], the difficulty of free radical identification was increased [43]. However, by PN-HRMS with ROC detecting, the detection (consumption) was along with the reaction (from “without US originally” to “after introducing US”) to gain the variations of peaks at m/z 130 and 209 in real time, illustrating that i) •OH mainly originated from reaction of US-OSC, not Scheme 3; ii) after introducing US, the type of dominant free radicals tended to be •OH in the system and the transformation may be mainly initiated from SO4 (being consistent with Section 3.5.1-iiii); and iii) the PN-HRMS with ROC detecting (comparing with traditional detecting means, liquid chromatograph/mass spectrometer (LC/MS) or EPR with intermittent sample testing one by one) at a certain degree could avoid the potential misjudgement due to Scheme 3.

Scheme 3.

Scheme 3

Intramolecular cleavage decomposition of O-centered sulfate radical adducts [35].

But most importantly, the ROC-HRMS seemed to prove the radical chain reaction (chain transfer SO4 to •OH) directly in the reaction process.

3.5.3.2. EPR signals

The EPR signals of different systems (ClO2, US-ClO2, OSC with different molar ratio, US/PDS, US-OSC, Fe0f-PDS, Fe0f-OSC, US/Fe0f-PDS, and US/Fe0f-OSC) were detected to conduct comparative analyses. Commonly, introducing US would not change the peak shapes of EPR signals (e.g. ClO2 vs. US-ClO2, OSC vs. US-OSC, Fe0f-OSC vs. US/Fe0f-OSC, etc.). Differently, single PDS with DMPO would still have the signals, which was consistent with study of Zamora and Villamena [35], possibly due to -OSO3(H) and –OH.

The DMPO-X1 (♥) signal with a single nitrogen (aN = 16.50 G (16.30–16.60), a triplet 1:1:1) [44] may contribute from the oxidation of DMPO itself, but according to study of Kondo et al. [45], peaks labeled “♥” may be due to DMPO•-H adducts showing a primary triplet (aN = 16.50 G) further split by two secondary hydrogens (⋄, aHβ = 22.40 G). From Fig. 7a, no •OH signal was observed in ClO2 and US-ClO2, however, besides DMPO-X1 signal, the remaining signal DMPO-X3 (×) seemed to be closed to ClO• signal [46] and ClO•+Cl• signal [47]; in addition, according to description of Ozawa et al. [48], combining with the DMPO-X1 and DMPO-X3, it may contribute from oxidation of spin-traps by ClO2•; while according to the study of Nishikawa et al. [49], ClOx (x = 2, 3) radicals led to the above phenomenon. Due to the Cl(III) in ClO2, an intermediate valence [50], it could be explained as ClOx (x = 0–3) radicals contributed to the above-mentioned results.

Fig. 7.

Fig. 7

Comparisons of EPR results under different systems: (a) the step-by-step EPR results of coupled systems, ClO2 & US-ClO2 100 mM, DMPO 100 mM, 5 min; S2O82−/ClO2 50 mM/50 mM, DMPO 100 mM, 3 min; OSC & US-OSC (S2O82−/ClO2 100 mM/20 mM), DMPO 100 mM, 1 min; Fe0f-OSC & US/Fe0f-OSC (S2O82−/ClO2 5 mM/1 mM, Fe0f 0.358 mM), DMPO 6 mM, 30 s. (b) the EPR results of OSC-based systems after adding DMSO, OSC (30 s) & US-OSC (5 min) with DMSO (S2O82−/ClO2 100 mM/20 mM, DMPO 100 mM, DMSO 100 mM); Fe0f-OSC (10 s) & US/Fe0f-OSC (i: 1 min, ii: 2 min) with DMSO (S2O82−/ClO2 5 mM/1 mM, Fe0f 0.358 mM, DMPO 6 mM, DMSO 6 mM). (c) the EPR results of PDS-based systems after adding DMSO, S2O82− 100 mM, DMPO 100 mM, 10 min; Fe0f-S2O82− (5 min) and US/Fe0f- S2O82− (2 min) (S2O82− 6 mM, Fe0f 0.358 mM, DMPO 6 mM, DMSO 6 mM). (♦ DMPO•-SO4, DMPO•–OH, *(*) DMPO•- SO3 & OH, ♥, × & ++ DMPO-Xs, ⋄ DMPO•-H).

In terms of OSC and US-OSC, DMPO-X1 signal was almost disappeared, the DMPO•-SO4 (♦) signal (aN = 13.78 G, aHβ = 9.97 G, aHγ1 = 1.47 G, aHγ2 = 0.98 G) (O-centered adducts) and DMPO•–OH () signal (aN = aH = 14.91 G, a quartet 1:2:2:1) were clearly obtained. Specially, in terms of relative peak intensity in same spectrum, the relative DMPO•-SO4 signals in OSC (S2O82−/ClO2 5:1) or after introducing US were more highlighted than that of OSC (1:1) or single PDS, which illustrated co-activation between ClO2 and PDS.

In terms of Fe0f-OSC and US/Fe0f-OSC, the DMPO•-SO4 and DMPO•–OH signals could be still detected but the DMPO-X1 signal was enhanced obviously, which indicated that the introduction of Fe0f accelerated reaction, meanwhile may produce other strongly oxidizing substance besides radicals to result in oxidation of DMPO.

In addition, DMSO as the probe was added the above systems to detect free radicals indirectly [51] due to the reaction of Scheme 4. As the addition of DMSO (Fig. 7b), unfortunately, the •CH3 signal could not be detected; instead, the quartet signal DMPO-X2 (* & *, near to 1:1:1:1, aN = aH = 14.50 G) appeared (Fig. 7b & 7c). The phenomenon was also observed by Zalibera et al. [52] (the mixtures (DMSO/H2O) with the lower dielectric permittivity significantly influenced the hyperfine coupling constants and the observed signals largely deviated from typical DMPO•–OH adduct in aqueous media); but it seemed to be consisted with the study of Jiang et al. [53] (contributing from SO3 [54] which had be explained by Ranguelova et al. [55]). And according to study of Zamora and Villamena [35], the above quartet signal may source from SO3 (S-centered adducts) and •OH with different proportions. In addition, the DMPO•-SO4 signal was disappeared, which may result from the competition reaction of DMSO vs. SO4 [56].

Scheme 4.

Scheme 4

Reactions of DMSO under •OH, SO4, and US in literature reports [56], [58], [59], [60].

For comparison, the EPR of Fe0f-PDS and US/Fe0f-PDS was operated (Fig. 7c). The detected signal of Fe0f-PDS was matched to the study of Zhu et al. [57] (adopting Fe-base activation), belonging to “DMPOX”, but it seemed to be more closed to composite of DMPO-X1 and DMPO-X2 in this study. However, after introducing the US, the signals were complex because of the appearance of DMPO-X4 (++, ?). Meanwhile, the introducing of Fe0f would force and make the EPR signals right-shift and complex (Fe0f-OSC vs. US/Fe0f-OSC, Fe0f-PDS vs. US/Fe0f-PDS).

3.5.4. Exploration of US role to O = FeIV/V

Catalytic iron center formed a highly oxidizing oxo-iron species FeIV/V (namely, O = FeIV and O = FeV) [65] which had the reaction with the sulfoxide in Scheme 5 to generate the corresponding sulfone by oxygen atom transfer step [61], [62]. Formally, O = FeV was best described as an oxo-Fe(IV)-porphyrin radical cation, namely, O = FeIV(Por)+ (O = FeIV+) [65], [66]. The PMSO was used as chemical probe in this study, and the methyl phenyl sulfone (PMSO2) with m/z of 156, 141, 125, etc. in Table 1 was detected by GC/MS (Fig. 4b), which could demonstrate that the existence of FeIV/V in Fe0f-OSC (tR 9.34 min) and US/Fe0f-OSC (tR 9.32 min) systems with larger proportion. Meanwhile, it was also corresponding to the aforementioned δTBA of Fe0f-OSC (namely, TBA could not effectively inhibit the degradation). In terms of Fe0f-OSC (US/Fe0f-OSC) system, the formations of FeIV/V as shown in Scheme 6, on the one hand, sourced from conversion reaction between “-O-O-” (from PDS) and FeII/III (from Fe0) [63], [64], [66], [67]; on the other hand, contributed from adjustment reaction between ClO2 and FeII/III [68], which were operated through a FeII/FeIV cycle [65], [66], [68]. The kinds and relative peak intensities of the intermediates (especially, PMSO2) in US/Fe0f-OSC were greater than those in Fe0f-OSC, implying that US could facilitate the transformation reaction.

Scheme 5.

Scheme 5

Reaction between sulfoxides and Fe(IV) in literature reports [61], [62], [63], [64].

Table 1.

The information of generated products after adding PMSO identified by GC/MS.

Intermediates Retention time (min)
Detected fragments (m/z) m/z
US/Fe0f-OSC Fe0f-OSC
PMSO 8.58 8.66 140 125 97 94 77 65 51 140
PMSO2 9.32 9.34 156 141 125 94 77 65 51 156
P1 10.53 10.55 174 (1 7 6) 158 138 125 97 77 51 174
P2 11.59 11.61 186 (187, 188, 185, 184) 152 125 109 84 (86) 77 65 51 186
P3 14.18 14.21 218 (219, 220) 185 154 140 109 84 65 218
P4 15.16 / 218 152 125 97 77 65 51 218
Scheme 6.

Scheme 6

Reactions of FeII/FeIV cycles [63], [66], [68], [70].

In addition, the detailed information of other formed intermediates P1-P4 was shown in Table 1. PMSO (sulfoxide) would undergo the reaction of “Pummerer rearrangement” to transfer to Ph-thiol (unstable in oxidation system), then further turn into Ph-S-Ph (P2) and Ph-S-S-Ph (P3), meanwhile, P2 would be oxidized into P4 (the detailed process could be seen in Fig. 4c) [69]. In addition, PMSO under the role of ClOx would turn into P1 via substitution.

3.5.5. Possible degradation process with US

By detected results and literatures, the possible mechanism of degradation process in terms of US/Fe0f-OSC was inferred and proposed as follows: at neutral pH, FeIV would be as the important intermediate in Fenton like reaction [71], but FeIV would reduce the contribution in ZVI/PDS as result of the decrease of available FeII [72]. However, US/Fe0f-OSC at near-neutral pH still kept well activity, therefore, the introduction of ClO2 not only cooperated with PDS to generate co-activation process (SO4 and ClO2) but also promoted the production and circulation of FeIV in presence of Fe0f (with the cycle of FeII, FeIII, O = FeIV+, and O = FeIV). Meanwhile, US could play another role on the circulation by polishing the surface of Fe0f to provide the fresh ZVI source. Meanwhile, the generated ClOx (0−3) would accelerate corrosion of Fe0f phase interface to release ionic Fe. Importantly, US would also promote production and transformation of O = FeIV/V, and free radical chain transfer (from SO4 to •OH).

The main reaction processes were shown in Fig. 8 and Section S12.

Fig. 8.

Fig. 8

The possible reaction mechanism of US/Fe0f-OSC.

3.6. Evaluation of energy consumption based on US

The electrical efficiency per log order (EE/O) was applied to evaluate the application prospect of US/Fe0f-OSC (seeing Section S13 for details). Table S2 listed the calculated EE/O and total cost values of different AOPs for treating wastewater. By economic comparison, the US/Fe0f-OSC with EE/O value 13.44 kWh⋅m−3 and total cost value 3.64 $⋅m−3.

4. Conclusions

Due to the limitation of ability, the determination (EPR and ROC/MS) of active species in the complex US/Fe0f-OSC, and -like systems still face the challenges and should be further explored and analyzed (e.g. isotope detection (18O, 17O, and 16O) or (H, D), etc.). The establishment and calculation of US cavitation model involving bubble–liquid dynamics should be further enhanced.

In this study, the EPR was applied to detect the radicals species of different systems for comparison. The ROC-MS (substitution of EPR) was a preliminary attempt to track the radicals transformation. The US apparent roles to removal reaction were expounded (e.g. de-order reaction, widening the applied range of pH and fluid eddy). The inherent roles of US were stated as follows:

i) US could induce ClO2 to co-activate PDS, low the bubble-water interface temperature, and promote the production and circulation of FeII/FeIV; ii) US could enrich the kinds of contaminant intermediates, polish the surface the Fe0f to avoid surface passivation to provide fresh Fe source, promote the production and transformation of O = FeIV/V, and push free radical chain transfer from SO4 to •OH.

In addition, the US/Fe0f-OSC may reveal other roles: relatively economical treating cost and potential (ClO2) disinfection.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

This work was supported by State Key Laboratory of Urban Water Resource and Environment, Harbin Institute of Technology (2019DX08) and National Natural Science Foundation of China (21904029).

Footnotes

Appendix A

Supplementary data to this article can be found online at https://doi.org/10.1016/j.ultsonch.2021.105750.

Appendix A. Supplementary data

The following are the Supplementary data to this article:

Supplementary data 1
mmc1.docx (2.4MB, docx)

References

  • 1.Wu X.H., Si S.Y., Tan W., Lu X., Ye F.G., Zhao S.L. Preparation of magnetic mesoporous metal-phenolic coordination spheres for extraction of crystal violet and leuco-metabolites in fish. J. Chromatogr. A. 2021;1636 doi: 10.1016/j.chroma.2020.461776. [DOI] [PubMed] [Google Scholar]
  • 2.Yakout S.M., Hassan M.R., Abdeltawab A.A., Aly M.I. Sono-sorption efficiencies and equilibrium removal of triphenylmethane (crystal violet) dye from aqueous solution by activated charcoal. J. Clean. Prod. 2019;234:124–131. [Google Scholar]
  • 3.Hurtaud-Pessel D., Couëdor P., Verdon E. Liquid chromatography-tandem mass spectrometry method for the determination of dye residues in aquaculture products: Development and validation. J. Chromatogr. A. 2011;1218(12):1632–1645. doi: 10.1016/j.chroma.2011.01.061. [DOI] [PubMed] [Google Scholar]
  • 4.Yu Y., Theerthagiri J., Lee S.J., Muthusamy G., Ashokkumar M., Choi M.Y. Integrated technique of pulsed laser irradiation and sonochemical processes for the production of highly surface-active NiPd spheres. Chem. Eng. J. 2021;411 [Google Scholar]
  • 5.Choi J., Lee H., Son Y. Effects of gas sparging and mechanical mixing on sonochemical oxidation activity. Ultrason. Sonochem. 2021;70 doi: 10.1016/j.ultsonch.2020.105334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Choi J., Cui M., Lee Y., Ma J., Kim J., Son Y., Khim J. Hybrid reactor based on hydrodynamic cavitation, ozonation, and persulfate oxidation for oxalic acid decomposition during rare-earth extraction processes. Ultrason. Sonochem. 2019;52:326–335. doi: 10.1016/j.ultsonch.2018.12.004. [DOI] [PubMed] [Google Scholar]
  • 7.Theerthagiri J., Lee S.J., Karuppasamy K., Arulmani S., Veeralakshmi S., Ashokkumar M., Choi M.Y. Application of advanced materials in sonophotocatalytic processes for the remediation of environmental pollutants. J. Hazard. Mater. 2021;412 doi: 10.1016/j.jhazmat.2021.125245. [DOI] [PubMed] [Google Scholar]
  • 8.Lee J., von Gunten U., Kim J.H. Persulfate-based advanced oxidation: critical assessment of opportunities and roadblocks. Environ. Sci. Technol. 2020;54:3064–3081. doi: 10.1021/acs.est.9b07082. [DOI] [PubMed] [Google Scholar]
  • 9.Xu Q.H., Shi F., You H., Wang S.T. Integrated remediation for organic-contaminated site by forcing running-water to modify alkali-heat/persulfate via oxidation process transfer. Chemosphere. 2021;262 doi: 10.1016/j.chemosphere.2020.128352. [DOI] [PubMed] [Google Scholar]
  • 10.Duan X., Sun H., Wang S. Metal-free carbocatalysis in advanced oxidation reactions. Accounts Chem. Res. 2018;51(3):678–687. doi: 10.1021/acs.accounts.7b00535. [DOI] [PubMed] [Google Scholar]
  • 11.Matzek L.W., Carter K.E. Activated persulfate for organic chemical degradation: A review. Chemosphere. 2016;151:178–188. doi: 10.1016/j.chemosphere.2016.02.055. [DOI] [PubMed] [Google Scholar]
  • 12.Pan Y.W., Zhang Y., Zhou M.H., Cai J.J., Tian Y.S. Enhanced removal of antibiotics from secondary wastewater effluents by novel UV/pre-magnetized Fe0/H2O2 process. Water Res. 2019;153:144–159. doi: 10.1016/j.watres.2018.12.063. [DOI] [PubMed] [Google Scholar]
  • 13.Wu J., Zhao J., Hou J., Zeng R.J., Xing B. Degradation of tetrabromobisphenol a by sulfidated nanoscale zerovalent iron in a dynamic two-step anoxic/oxic process. Environ. Sci. Technol. 2019;53(14):8105–8114. doi: 10.1021/acs.est.8b06834. [DOI] [PubMed] [Google Scholar]
  • 14.Wu Y.W., Chen X.T., Han Y., Yue D.T., Cao X.D., Zhao Y.X., Qian X.F. Highly efficient utilization of nano-Fe(0) embedded in mesoporous carbon for activation of peroxydisulfate. Environ. Sci. Technol. 2019;53:9081–9090. doi: 10.1021/acs.est.9b02170. [DOI] [PubMed] [Google Scholar]
  • 15.Lv H.Z., Niu H.Y., Zhao X.L., Cai Y.Q., Wu F.C. Carbon zero-valent iron materials possessing high-content fine Fe0 nanoparticles with enhanced microelectrolysis-Fenton-like catalytic performance for water purification. Appl. Catal. B-Environ. 2021;286 [Google Scholar]
  • 16.Huang Y.X., Jiang J.W., Ma L.M., Wang Y.W., Liang M.L., Zhang Z.G., Li L. Iron foam combined ozonation for enhanced treatment of pharmaceutical wastewater. Environ. Res. 2020;183 doi: 10.1016/j.envres.2020.109205. [DOI] [PubMed] [Google Scholar]
  • 17.Wu Z., Zhao Y., Wu H., Gao Y., Chen Z., Jin W., Wang J., Ma T., Wang L. Corrosion engineering on iron foam toward efficiently electrocatalytic overall water splitting powered by sustainable energy. Adv. Funct. Mater. 2021;31(17):2010437. doi: 10.1002/adfm.v31.1710.1002/adfm.202010437. [DOI] [Google Scholar]
  • 18.Cosson H., Ernst W.R. Photodecomposition of chlorine dioxide and sodium chlorite in aqueous solution by irradiation with ultraviolet light. Ind. Eng. Chem. Res. 1994;33(6):1468–1475. [Google Scholar]
  • 19.Leitner N.K.V., Delaat J., Dore M. Photodecomposition of chlorine dioxide and chlorite by UV-irradiation.1. photo-products. Water Res. 1992;26:1665–1672. [Google Scholar]
  • 20.Hao R., Mao X., Qian Z., Zhao Y.i., Wang L., Yuan B.o., Wang K., Liu Z., Qi M., Crittenden J. Simultaneous removal of SO2 and NO using a novel method of ultraviolet irradiating chlorite-ammonia complex. Environ. Sci. Technol. 2019;53(15):9014–9023. doi: 10.1021/acs.est.8b06950. [DOI] [PubMed] [Google Scholar]
  • 21.Hao R.L., Li C., Wang Z., Gong Y.P., Yuan B., Zhao Y., Wang L.D., Crittenden J. Removal of gaseous elemental mercury using thermally catalytic chlorite-persulfate complex. Chem. Eng. J. 2020;391 [Google Scholar]
  • 22.Wenk J., Aeschbacher M., Salhi E., Canonica S., von Gunten U., Sander M. Chemical oxidation of dissolved organic matter by chlorine dioxide, chlorine, and ozone: effects on its optical and antioxidant properties. Environ. Sci. Technol. 2013;47:11147–11156. doi: 10.1021/es402516b. [DOI] [PubMed] [Google Scholar]
  • 23.Xu Q.H., Leng H.R., You H., Wang S.T., Li H.Y., Yu Y.B. A novel co-catalyzed system between persulfate and chlorite by sonolysis for removing triphenylmethane pharmaceutical. J. Environ. Sci. 2022;112:291–306. doi: 10.1016/j.jes.2021.05.011. [DOI] [PubMed] [Google Scholar]
  • 24.Tan X., Zhang D., Parajuli K., Upadhyay S., Jiang Y., Duan Z. Comparison of four quantitative techniques for monitoring microalgae disruption by low-frequency ultrasound and acoustic energy efficiency. Environ. Sci. Technol. 2018;52(5):3295–3303. doi: 10.1021/acs.est.7b05896. [DOI] [PubMed] [Google Scholar]
  • 25.Theerthagiri J., Madhavan J., Lee S.J., Choi M.Y., Ashokkumar M., Pollet B.G. Sonoelectrochemistry for energy and environmental applications. Ultrason. Sonochem. 2020;63 doi: 10.1016/j.ultsonch.2020.104960. [DOI] [PubMed] [Google Scholar]
  • 26.Madhavan J., Theerthagiri J., Balaji D., Sunitha S., Choi M.Y., Ashokkumar M. Hybrid advanced oxidation processes involving ultrasound: an overview. Molecules. 2019;24:3341. doi: 10.3390/molecules24183341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Wei Z.S., Villamena F.A., Weavers L.K. Kinetics and mechanism of ultrasonic activation of persulfate: An in situ EPR spin trapping study. Environ. Sci. Technol. 2017;51:3410–3417. doi: 10.1021/acs.est.6b05392. [DOI] [PubMed] [Google Scholar]
  • 28.Johnson R.L., Tratnyek P.G., Johnson R.O. Persulfate persistence under thermal activation conditions. Environ. Sci. Technol. 2008;42(24):9350–9356. doi: 10.1021/es8019462. [DOI] [PubMed] [Google Scholar]
  • 29.Suslick K.S., Hammerton D.A., Cline R.E. The sonochemical hot-spot. J. Am. Chem. Soc. 1986;108(18):5641–5642. [Google Scholar]
  • 30.Suslick K.S. Sonochemistry. Science. 1990;247(4949):1439–1445. doi: 10.1126/science.247.4949.1439. [DOI] [PubMed] [Google Scholar]
  • 31.Misik V., Miyoshi N., Riesz P. EPR spin-trapping study of the sonolysis of H2O/D2O mixtures-probing the temperatures of cavitation regions. J. Phys. Chem. 1995;99(11):3605–3611. [Google Scholar]
  • 32.Cui D., Mebel A.M., Arroyo-Mora L.E., Holness H., Furton K.G., O'Shea K. Kinetic, product, and computational studies of the ultrasonic induced degradation of 4-methylcyclohexanemethanol (MCHM) Water Res. 2017;126:164–171. doi: 10.1016/j.watres.2017.09.005. [DOI] [PubMed] [Google Scholar]
  • 33.Neta P., Huie R.E., Ross A.B. Rate constants for reactions of inorganic radicals in aqueous solution. J. Phys. Chem. Ref. Data. 1988;17(3):1027–1284. [Google Scholar]
  • 34.Buxton G.V., Greenstock C.L., Helman W.P., Ross A.B. 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. 1988;17(2):513–886. [Google Scholar]
  • 35.Zamora P.L., Villamena F.A. Theoretical and experimental studies of the spin trapping of inorganic radicals by 5,5-Dimethyl-1-pyrroline N-oxide (dmpo). 3. sulfur dioxide, sulfite, and sulfate radical anions. J. Phys. Chem. A. 2012;116(26):7210–7218. doi: 10.1021/jp3039169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Murray S., Rendell N.B., Taylor G.W. Microbore high-performance liquid chromatography electrospray ionisation mass spectrometry of steroid sulphates. J. Chromatogr. A. 1996;738(2):191–199. doi: 10.1016/s0378-4347(97)00083-2. [DOI] [PubMed] [Google Scholar]
  • 37.Lamb D.J., Wang H.M., Mallis L.M., Linhardt R.J. Negative-ion fast-atom-bombardment tandem mass-spectrometry to determine sulfate and linkage position in glycosaminoglycan-derived disaccharides. J. Am. Soc. Mass. Spectr. 1992;3(8):797–803. doi: 10.1016/1044-0305(92)80002-3. [DOI] [PubMed] [Google Scholar]
  • 38.Nha Tran T.T., Wang T., Hack S., Bowie J.H. Fragmentations of [M-H](-) anions of peptides containing Ser sulfate. A joint experimental and theoretical study. Rapid Commun. Mass. Sp. 2013;27(21):2287–2296. doi: 10.1002/rcm.6686. [DOI] [PubMed] [Google Scholar]
  • 39.Giorio C., Campbell S.J., Bruschi M., Tampiere F., Barbon A., Toffoletti A., Tapparo A., Paijens C., Wedlake A.J., Grice P., Howe D.J., Kalberer M. Online quantification of criegee intermediates of alpha-pinene ozonolysis by stabilization with spin traps and proton-transfer reaction mass spectrometry detection. J. Am. Chem. Soc. 2017;139:3999–4008. doi: 10.1021/jacs.6b10981. [DOI] [PubMed] [Google Scholar]
  • 40.Guo Q., Qian S.Y., Mason R.P. Separation and identification of DMPO adducts of oxygen-centered radicals formed from organic hydroperoxides by HPLC-ESR, ESI-MS and MS/MS. J. Am. Soc. Mass. Spectr. 2003;14(8):862–871. doi: 10.1016/S1044-0305(03)00336-2. [DOI] [PubMed] [Google Scholar]
  • 41.Bauer N.A., Hoque E., Wolf M., Kleigrewe K., Hofmann T. Detection of the formyl radical by EPR spin-trapping and mass spectrometry. Free Radical Bio. Med. 2018;116:129–133. doi: 10.1016/j.freeradbiomed.2018.01.002. [DOI] [PubMed] [Google Scholar]
  • 42.Domingues P., Domingues M.R.M., Amado F.M.L., Ferrer-Correia A.J. Detection and characterization of hydroxyl radical adducts by mass spectrometry. J. Am. Soc. Mass. Spectr. 2001;12(11):1214–1219. doi: 10.1016/S1044-0305(01)00310-5. [DOI] [PubMed] [Google Scholar]
  • 43.Gao H.-Y., Huang C.-H., Mao L.i., Shao B.o., Shao J., Yan Z.-Y., Tang M., Zhu B.-Z. First direct and unequivocal electron spin resonance spin-trapping evidence for pH-dependent production of hydroxyl radicals from sulfate radicals. Environ. Sci. Technol. 2020;54(21):14046–14056. doi: 10.1021/acs.est.0c04410. [DOI] [PubMed] [Google Scholar]
  • 44.Li F.F., Xiao L.Q., Liu L.J. Metal-diazo radicals of alpha-carbonyl diazomethanes. Sci. Rep. 2016;6:22876. doi: 10.1038/srep22876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Kondo T., Mišík V., Riesz P. Sonochemistry of cytochrome c. Evidence for superoxide formation by ultrasound in argon-saturated aqueous solution. Ultrason. Sonochem. 1996;3(3):S193–S199. [Google Scholar]
  • 46.Li F., Sun L.W., Liu Y.B., Fang X.F., Shen C.S., Huang M.H., Wang Z.W., Dionysiou D.D. A ClO center dot-mediated photoelectrochemical filtration system for highly-efficient and complete ammonia conversion. J. Hazard. Mater. 2020;400 doi: 10.1016/j.jhazmat.2020.123246. [DOI] [PubMed] [Google Scholar]
  • 47.Zheng Z.X., Ng Y.H., Tang Y.M., Li Y.P., Chen W.R., Wang J., Li X.K., Li L.S. Visible-light-driven photoelectrocatalytic activation of chloride by nanoporous MoS2@BiVO4 photoanode for enhanced degradation of bisphenol A. Chemosphere. 2021;263 doi: 10.1016/j.chemosphere.2020.128279. [DOI] [PubMed] [Google Scholar]
  • 48.Ozawa T., Miura Y., Ueda J.-I. Oxidation of spin-traps by chlorine dioxide (ClO2) radical in aqueous solutions: First ESR evidence of formation of new nitroxide radicals. Free Radical Bio. Med. 1996;20(6):837–841. doi: 10.1016/0891-5849(95)02092-6. [DOI] [PubMed] [Google Scholar]
  • 49.Nishikawa H., Nakamura R., Ohki Y., Nagasawa K., Hama Y. Characterization of ClOx radicals in vacuum-ultraviolet-irradiated high-purity silica glass. Phys. Rev. B. 1992;46:8073–8079. doi: 10.1103/physrevb.46.8073. [DOI] [PubMed] [Google Scholar]
  • 50.Marcon J., Mortha G., Marlin N., Molton F., Duboc C., Burnet A. New insights into the decomposition mechanism of chlorine dioxide at alkaline pH. Holzforschung. 2017;71:599–610. [Google Scholar]
  • 51.Zhu B.-Z., Zhao H.-T., Kalyanaraman B., Frei B. Metal-independent production of hydroxyl radicals by halogenated quinones and hydrogen peroxide: An ESR spin trapping study. Free Radical Biol. Med. 2002;32(5):465–473. doi: 10.1016/s0891-5849(01)00824-3. [DOI] [PubMed] [Google Scholar]
  • 52.Zalibera M., Rapta P., Stasko A., Brindzova L., Brezova V. Thermal generation of stable spin trap adducts with super-hyperfine structure in their EPR spectra: An alternative EPR spin trapping assay for radical scavenging capacity determination in dimethylsulphoxide. Free Radical Res. 2009;43:457–469. doi: 10.1080/10715760902846140. [DOI] [PubMed] [Google Scholar]
  • 53.Jiang B., Wang X.L., Liu Y.K., Wang Z.H., Zheng J.T., Wu M.B. The roles of polycarboxylates in Cr(VI)/sulfite reaction system: Involvement of reactive oxygen species and intramolecular electron transfer. J. Hazard. Mater. 2016;304:457–466. doi: 10.1016/j.jhazmat.2015.11.011. [DOI] [PubMed] [Google Scholar]
  • 54.St Denis T.G., Vecchio D., Zadlo A., Rineh A., Sadasivam M., Avci P., Huang L., Kozinska A., Chandran R., Sarna T., Hamblin M.R. Thiocyanate potentiates antimicrobial photodynamic therapy: In situ generation of the sulfur trioxide radical anion by singlet oxygen. Free Radical Bio. Med. 2013;65:800–810. doi: 10.1016/j.freeradbiomed.2013.08.162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Ranguelova K., Rice A.B., Khajo A., Triquigneaux M., Garantziotis S., Magliozzo R.S., Mason R.P. Formation of reactive sulfite-derived free radicals by the activation of human neutrophils: An ESR study. Free Radical Bio. Med. 2012;52(8):1264–1271. doi: 10.1016/j.freeradbiomed.2012.01.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Kishore K., Asmus K.D. Radical cations from one-electron oxidation of aliphatic sulfoxides in aqueous solution-a radiation chemical study. J. Chem. Soc., Perkin Trans. 1989;2:2079–2084. [Google Scholar]
  • 57.Zhu K., Xu H., Chen C., Ren X., Alsaedi A., Hayat T. Encapsulation of Fe0 dominated Fe3O4/Fe0/Fe3C nanoparticles into carbonized polydopamine nanospheres for catalytic degradation of tetracycline via persulfate activation. Chem. Eng. J. 2019;372:304–311. [Google Scholar]
  • 58.Fukui S., Hanasaki Y., Ogawa S. High-performance liquid-chromatographic determination of methanesulphinic acid as a method for the determination of hydroxyl radicals. J. Chromatogr. 1993;630(1-2):187–193. doi: 10.1016/0021-9673(93)80455-h. [DOI] [PubMed] [Google Scholar]
  • 59.Kondo T., Kirschenbaum L.J., Kim H., Riesz P. Sonolysis of dimethyl-sulfoxide water mixtures - a spin-trapping study. J. Phys. Chem. 1993;97(2):522–527. [Google Scholar]
  • 60.Misik V., Riesz P. Peroxyl radical formation in aqueous solutions of N, N-dimethylformamide, N-methylformamide, and dimethylsulfoxide by ultrasound: Implications for sonosensitized cell killing. Free Radical Bio. Med. 1996;20:129–138. doi: 10.1016/0891-5849(95)02009-8. [DOI] [PubMed] [Google Scholar]
  • 61.Wang Z., Jiang J., Pang S.Y., Zhou Y., Guan C.T., Gao Y., Li J., Yang Y., Qu W., Jiang C.C. Is sulfate radical really generated from peroxydisulfate activated by iron(II) for environmental decontamination? Environ. Sci. Technol. 2018;52:11276–11284. doi: 10.1021/acs.est.8b02266. [DOI] [PubMed] [Google Scholar]
  • 62.Pestovsky O., Stoian S., Bominaar E.L., Shan X.P., Munck E., Que L., Bakac A. Aqueous Fe-IV=O: Spectroscopic identification and oxo-group exchange. Angew. Chem. Int. Ed. 2005;44:6871–6874. doi: 10.1002/anie.200502686. [DOI] [PubMed] [Google Scholar]
  • 63.Ghosh A., Mitchell D.A., Chanda A., Ryabov A.D., Popescu D.L., Upham E.C., Collins G.J., Collins T.J. Catalase-peroxidase activity of Iron(III)-Taml activators of hydrogen peroxide. J. Am. Chem. Soc. 2008;130(45):15116–15126. doi: 10.1021/ja8043689. [DOI] [PubMed] [Google Scholar]
  • 64.Jiang N., Xu H.D., Wang L.H., Jiang J., Zhang T. Nonradical oxidation of pollutants with single-atom-Fe(III)-activated persulfate: Fe(V) being the possible intermediate oxidant. Environ. Sci. Technol. 2020;54:14057–14065. doi: 10.1021/acs.est.0c04867. [DOI] [PubMed] [Google Scholar]
  • 65.Prat I., Mathieson J.S., Güell M., Ribas X., Luis J.M., Cronin L., Costas M. Observation of Fe(v)=O using variable-temperature mass spectrometry and its enzyme-like C-H and C=C oxidation reactions. Nat. Chem. 2011;3(10):788–793. doi: 10.1038/nchem.1132. [DOI] [PubMed] [Google Scholar]
  • 66.Chen K., Que L. Stereospecific alkane hydroxylation by non-heme iron catalysts: Mechanistic evidence for an FeV=O active species. J. Am. Chem. Soc. 2001;123:6327–6337. doi: 10.1021/ja010310x. [DOI] [PubMed] [Google Scholar]
  • 67.Li H., Shan C., Pan B. Fe(III)-Doped g-C3N4 mediated peroxymonosulfate activation for selective degradation of phenolic compounds via high-valent iron-oxo species. Environ. Sci. Technol. 2018;52(4):2197–2205. doi: 10.1021/acs.est.7b05563. [DOI] [PubMed] [Google Scholar]
  • 68.Zdilla M.J., Lee A.Q., Abu-Omar M.M. Concerted dismutation of chlorite ion: water-soluble iron-porphyrins as first generation model complexes for chlorite dismutase. Inorg. Chem. 2009;48:2260–2268. doi: 10.1021/ic801681n. [DOI] [PubMed] [Google Scholar]
  • 69.Kurt L., Czako B. Elsevier; 2005. Strategic Applications Of Named Reactions In Organic Synthesis. [Google Scholar]
  • 70.Xie L., Zhang X.-P., Zhao B., Li P., Qi J., Guo X., Wang B., Lei H., Zhang W., Apfel U.-P., Cao R. Enzyme-Inspired Iron Porphyrins for Improved Electrocatalytic Oxygen Reduction and Evolution Reactions. Angew. Chem. Int. Ed. 2021;60(14):7576–7581. doi: 10.1002/anie.202015478. [DOI] [PubMed] [Google Scholar]
  • 71.Bataineh H., Pestovsky O., Bakac A. pH-induced mechanistic changeover from hydroxyl radicals to iron(IV) in the Fenton reaction. Chem. Sci. 2012;3(5):1594. doi: 10.1039/c2sc20099f. [DOI] [Google Scholar]
  • 72.Wang Z., Qiu W., Pang S.Y., Gao Y., Zhou Y., Cao Y., Jiang J. Relative contribution of ferryl ion species (Fe(IV)) and sulfate radical formed in nanoscale zero valent iron activated peroxydisulfate and peroxymonosulfate processes. Water Res. 2020;172 doi: 10.1016/j.watres.2020.115504. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary data 1
mmc1.docx (2.4MB, docx)

Articles from Ultrasonics Sonochemistry are provided here courtesy of Elsevier

RESOURCES