Next Article in Journal
Novel Cobalt Dichloride Complexes with Hindered Diphenylphosphine Ligands: Synthesis, Characterization, and Behavior in the Polymerization of Butadiene
Next Article in Special Issue
The High Efficiency of Anionic Dye Removal Using Ce-Al13/Pillared Clay from Darbandikhan Natural Clay
Previous Article in Journal
Synthesis, Crystal Structure, and Biological Evaluation of Fused Thiazolo[3,2-a]Pyrimidines as New Acetylcholinesterase Inhibitors
Previous Article in Special Issue
Interaction of Arsenic Species with Organic Ligands: Competitive Removal from Water by Coagulation-Flocculation-Sedimentation (C/F/S)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Different Role of Bisulfite/Sulfite in UVC-S(IV)-O2 System for Arsenite Oxidation in Water

1
Department of Environmental Science and Engineering, School of Resource and Environmental Sciences, Wuhan University, Wuhan 430079, China
2
School of Civil Engineering, Engineering Research Center of Urban Disasters Prevention and Fire Rescue Technology of Hubei Province, Wuhan University, Wuhan 430072, China
3
Voevodsky Institute of Chemical Kinetics and Combustion, Institutskaya 3, 630090 Novosibirsk, Russia
4
Novosibirsk State University, Pirogova str. 2, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(12), 2307; https://doi.org/10.3390/molecules24122307
Submission received: 16 May 2019 / Revised: 12 June 2019 / Accepted: 14 June 2019 / Published: 21 June 2019
(This article belongs to the Special Issue Advanced Materials and Technologies for Wastewater Treatment)

Abstract

:
It is of interest to use UV-sulfite based processes to degrade pollutants in wastewater treatment process. In this work, arsenic (As(III)) has been selected as a target pollutant to verify the efficacy of such a hypothesized process. The results showed that As(III) was quickly oxidized by a UV-sulfite system at neutral or alkaline pH and especially at pH 9.5, which can be mainly attributed to the generated oxysulfur radicals. In laser flash photolysis (LFP) experiments (λex = 266 nm), the signals of SO3•− and eaq generated by photolysis of sulfite at 266 nm were discerned. Quantum yields for photoionization of HSO3 (0.01) and SO32− (0.06) were also measured. It has been established that eaq does not react with SO32−, but reacts with HSO3 with a rate constant 8 × 107 M−1s−1.

Graphical Abstract

1. Introduction

Recently, advanced oxidation processes (AOPs) using sulfite (S(IV)), as effective strategies for the removal of contaminants, have attracted a lot of attention [1,2,3]. Though there have been some reports on the UVC(254 nm lamp)-S(IV) system, the hydrated electrons (eaq) generated have mainly been used for reductive dehalogenation [4,5]. The eaq could be only formed in sufficiently high concentration under oxygen-free conditions according to reactions (1) and (2). However, in most natural aqueous environments, dissolved oxygen (ca. 0.2 mM) is invariably present, and this may impose some limits on reductive dehalogenation using UVC-S(IV) system. SO3•−, another product of S(IV) photolysis, a relatively mild oxidant [6], reacts rapidly with oxygen to produce SO5•− radicals and then form SO4•− and even HO• (reactions (3)–(7)) [7,8,9,10,11,12]. These radicals possess higher redox potential and can effectively degrade pollutants [2,13]. Hence, it may also be possible to oxidize or degrade contaminants in an oxygen-containing sulfite system under UVC irradiation (namely, a UVC-S(IV)-O2 system).
SO32− + hν → SO3•− + eaq
eaq + O2 → O2•−    k1 = 1.9 × 1010 M−1 s−1
SO3•− + O2 → SO5•−    k2 = (1.5–2.5) × 109 M−1 s−1
SO5•− + ΗSO3 → SO42− + SO4•− + H+   k3 = 2.5 × 104 M−1 s−1
SO5•− + SO32− → SO42− + SO4•−   k4 = 3.8 × 106 M−1 s−1
SO4•− + H2O → SO42− + HO• + H+  k5 = 6.6 × 102 s−1
SO4•− + OH → SO42− + HO•   k6 = 1.4 × 107 M−1 s−1
The environmental chemistry and pollution control of arsenic have been extensively studied due to its high toxicity and carcinogenicity. In an investigation of 445,000 well water samples in China, it has been estimated that 5.6 million people are exposed to high concentrations of arsenic in drinking water (>50 μg L−1) and that some 14.7 million are exposed to arsenic concentrations of >10 μg L−1 [14]. A similar problem has happened in the United States, The US geological survey reported that 1% of the 54,000 US public water supplies exceed 50 µg L−1 of arsenic concentrations, 3% exceed 20 µg L−1, 8% exceed 10 µg L−1, and 14% exceed 5 µg L−1 (Water Resources Investigations Report 99-4279) [15]. Many As(III) oxidation processes have been devised, such as direct photolysis [16,17] and homogeneous or heterogeneous metal-sulfite systems [2,18,19,20]. However, the former is time-consuming while the latter introduces secondary pollution in the form of the metal in the metal-S(IV) systems. Inorganic As(III) is much more toxic than inorganic As(V) and accounts for about 20% of the arsenic present in the natural environment [21]. Therefore, it is very attractive to develop a method for oxidizing As(III) to As(V) quickly and safely. Zhang et al. [22] reported As(III) oxidation with SO2/O2 and UV light, which was mainly accomplished through the action of oxysulfur radicals (especially SO4•−). However, the reaction mechanism in UVC-S(IV)-O2 system is still unclear with respect to the performance of two different S(IV) species (HSO3/SO32− ions) under UVC light.
In this work, As(III) oxidation by UVC-S(IV)-O2 system has been investigated with the aim of revealing the different role/performances of HSO3/SO32− at pH 7 and 9.5, respectively. Laser flash photolysis has been utilized to obtain the quantum yields of photolysis of S(IV) species under 266 nm irradiation and the rate constant for the reaction between eaq and S(IV).

2. Materials and Methods

2.1. Materials

All chemicals were of analytical grade and were used without further purification. NaAsO2(99.5% is Xiya Reagent Center, Sichuan, China), Na2HAsO4·7H2O (99%, Alfa Aesar, A Johnson Matthey Co., China) served as As(III) and As(V) sources, respectively. Ethanol (EtOH), tert.-butyl alcohol (TBA), NaOH, H2SO4, KBH4, HCl, and Na2SO3 were purchased from Sinopharm Chemical Reagent Co. Ltd. (Shanghai, China). Ultrapure water of 18.2 MΩ cm resistivity was obtained through a water purification system (Youpu Super Pure Technology Co., Ltd. Sichuan, China) and was used in all experiments. All the stock solutions were kept refrigerated at 4 °C.

2.2. Experimental Procedures

All photodegradation experiments were conducted in a 400 mL cylindrical reactor cooled by external jacket water circulation at a constant temperature of 25 °C (Figure S1). A 254 nm lamp was used as the excitation source and was placed at the center of the reactor. A Na2SO3 stock solution was freshly prepared prior to the reaction, using cooled boiling water to prevent the oxidation of sulfite. Predetermined amounts of As(III) were added to the reactor, and the solution was constantly stirred with a poly tetra fluoroethylene(PTFE)-coated magnetic stirrer. A pH meter (Mettler Toleod LE409) was used to measure the pH value before the reaction. After the pH and temperature of solution had stabilized, dilute H2SO4 or NaOH was used to adjust the pH to the desired value. Each experiment was started by spiking a certain volume of fresh Na2SO3 solution and immediately switching on the lamp. The pH was not controlled during the reaction. Aliquots (2 ml) of the solution were withdrawn at fixed time intervals, and then a specific amount of HCl (1:1) was added to terminate the reaction. The As(III) concentration were measured by liquid chromatography-hydride generation–atomic fluorescence spectrometry (LC−HG−AFS, Bohui Innovation Technology Co., Ltd., Beijing, China).

2.3. Analysis

Arsenic speciation was simultaneously analyzed by LC−HG−AFS. Phosphate buffer (45 mM, pH 5.6) was used as mobile phase to separate inorganic As(III) and As(V) on a Hamilton PRP-X100 anion-exchange column (Switzerland) in LC. Solutions of 5% HCl–2% KBH4 were used for the determination of arsenic species concentration in HG–AFS. Argon (99.99%) was used as the carrier gas and shielding gas during the determination.
All laser flash photolysis (LFP) experiments were conducted in a 1 cm quartz cell in air-equilibrated or argon-saturated solutions at an initial pH of 7 or 9.5, at 298 K, under atmospheric pressure. Argon-saturated solutions were obtained by constantly bubbling argon through the sample. The LFP setup in the time-resolved experiments was based on an LS-2137U Nd:YAG laser (Lotis TII, Belarus) with an excitation wavelength of 266 nm, a pulse duration of 5–6 ns, an illumination spot area of 0.03 cm2 and an energy per pulse of up to 10 mJ. The time resolution of the setup was ca. 50 ns. Solutions in LFP experiments were refreshed after every 100–150 pulses to maintain their degradation less than 15% during the measurements. Spectra of the sulfite solution were recorded on an Agilent 8453 spectrophotometer (Agilent Technologies) using a 1 cm cell.

3. Results and Discussion

3.1. As(III) Oxidation in the UVC-S(IV)-O2 System

Figure 1 shows the efficiencies of As(III) oxidation in the UVC-S(IV)-O2 system and the related control systems at pH 7 and 9.5, respectively. Whether in neutral or alkaline solution, As(III) alone under UVC irradiation showed no obvious oxidation in 10 min. In a dark experiment with sulfite, only about 13% of As(III) was oxidized at pH 7, but this amount was doubled at pH 9.5 (about 26%), suggesting that alkaline pH may activate sulfite in some way to oxidize As(III). In the UVC-S(IV) system, along with rapid S(IV) oxidation caused by UVC irradiation (Figure S2), As(III) could also be oxidized to As(V) to some extent at pH 9.5 or 7. Indeed, we found that alkaline solution strongly facilitated As(III) oxidation from only 23% at pH 7 to 72% at pH 9.5 within 10 min. Three aspects could explain this marked difference. First, the critical oxidation-reduction potential (ORP) of As(V)/As(III) couples drop with decreasing pH value, such that the oxidation of As(III) to As(V) was more feasible in alkaline than in acidic solutions [23,24]. In addition, Hayon etc. [25] reported the pKa of HSO3 as 7.2, and so at pH 7 about 40% of sulfite should be present as SO32−, increasing to almost 100% at pH 9.5. SO32− has a better quantum yield under 254 nm irradiation compared to HSO3 (see Section 3.2 for details). Lastly, we noticed that in the experiment started at neutral pH, the solution became more acidic during the reaction time, whereas the alkaline pH was well maintained (Figure S3), consistent with the results of As(III) oxidation at pH 7 or 9.5.
High sulfite concentration (2 mM) could induce an anaerobic environment in solution within an extremely short time, since SO5•−, SO4•−, HO• formation and self-oxidation of sulfites all consume oxygen [26,27]. Once the dissolved oxygen concentration dropped to a low level, reaction (3) could be a rate-controlling step in the chain reactions and hence influence the As(III) oxidation. We conducted pumping experiments with synthetic gas to prove the influence of oxygen on As(III) oxidation. As shown in Figure 1, when synthetic gas (21% O2/79% N2) was constantly pumped into the reaction solution, As(III) oxidation efficiency was greatly enhanced at both pH 7 and 9.5, demonstrating that oxygen was indeed necessary for the chain reaction process and promoted the formation of radicals for the As(III) oxidation.
Radical-scavenging experiments were employed to prove the existence of relevant radicals in the UVC-S(IV)-O2 system (Figure 2). As demonstrated above, oxygen was necessary for the chain reaction process and hence synthetic gas was constantly pumped into the reaction solution. Commonly, alcohols (EtOH and TBA) have been selected as probes of SO4•− and HO•, because they have no obvious absorption at 254 nm and the rate constants for the reactions of EtOH and SO4•−/HO• have no significant difference (kEtOH, SO4•− = (1.6–6.2) × 107 M−1 s−1 [26], kEtOH, HO• = (1.8–2.8) × 109 M−1 s−1 [26]), whereas TBA is inert toward SO4•− in comparison with HO• (kTBA, SO4•− = 9.1 × 105 M−1 s−1 [2], kTBA, HO• = (3.8–7.6) × 108 M−1 s−1 [26]). In Figure 2, it can be seen that As(III) oxidation was not inhibited in the presence of TBA at pH 7, but the initial oxidation rate (r) decreased from 0.227 min−1 to 0.214 min−1 following addition of the same TBA concentration at pH 9.5. An alkaline solution could promote the HO• formation according to reactions (6) and (7). Besides, EtOH (5 mM or 177 mM) only partly inhibited As(III) oxidation, especially at pH 9.5 (23%–28% inhibition). Hence, other reactive species must be responsible for the As(III) oxidation. SO5•−, the precursor of SO4•− and HO•, also has a relatively high ORP (1.1 V [6]) and could possibly oxidize As(III). Unfortunately, aniline, the common radical probe for SO5•−, shows a strong absorption at 254 nm. Additionally, oxygen is necessary for the formation of SO5•− according to reaction (3). Hence, anaerobic experiments were carried out to prove this indirectly. As shown in Figure 2, anaerobic environment greatly hindered As(III) oxidation (with the rate decreasing from 0.121 min−1 to 0.018 min−1 at pH 7 and from 0.227 min−1 to 0.031 min−1 at pH 9.5). These results demonstrated that oxysulfur radicals generated in the UVC-sulfite system were the main reason for As(III) oxidation. Notably, about 18% and 26% of As(III) were still oxidized under anaerobic conditions at pH 7 and pH 9.5. This partial oxidation was clearly not due to SO3•− or eaq because of the weak oxidation ability of the former and the reducing capacity of the latter. However, SO3•− could generate dithionate (S2O62−) according to reaction (8) [28] and then oxidize As(III). This is relevant because SO3•− accumulates in anaerobic environments, as proved in our previous work [26].
SO3•− + SO3•− → S2O62−  k7 = 1.8 × 108 M−1 s−1
According to the results of radical-scavenging experiments and subsequent LFP experiments, a transformation process of S(IV) under UVC irradiation is proposed (Scheme 1), which includes three sections. Under anaerobic conditions, SO3•− and eaq are the main products of S(IV) photolysis, which would be proved by LFP experiments. Under aerobic conditions, SO3•− forms reactive sulfur species (RSS) and eaq forms reactive oxygen species (ROS), respectively. These reactive species showed enough ability to oxidize As(III). Hence, all three sections contribute to As(III) oxidation.

3.2. LFP Studies of SO3•− and Hydrated Electrons

As shown in Figure S4, both SO32− and HSO3 exhibited similar spectra, in which the intensities decrease sharply at wavelengths > 250 nm and there was only weak absorption at 266 nm. Therefore, for LFP experiments, a sulfite concentration of at least 50 mM was needed to detect the signals of eaq or SO3•−.
Flash excitation of SO32− ions at pH 9.5 in argon-saturated solutions generated a transient absorption in the region 250–780 nm with a maximum at about 720 nm (Figure 3), which was mainly attributed to hydrated electrons [29]. The lifetimes of hydrated electrons (eaq) under these conditions were about 7–9 μs and did not depend on sulfite concentration, in agreement with literature estimates (k(eaq + SO32−) < 1.5 × 106 M−1s−1 [2]). The eaq absorbance at 720 nm (Figure S5) showed a good linear dependence on the excitation energy, which allowed estimation of the quantum yields of monophotonic ionizations of SO32−ion266nm = 0.06) and HSO3 ions (φion266nm = 0.01). From this, it could be concluded that HSO3 ions produced far fewer eaq due to photoionization, in full agreement with the lower degradation efficiency of As(III) at pH 7 (Figure 1).
Flash excitation of SO32− ions at pH 9.5 in air-saturated solutions (Figure 4) also allowed detection of the absorption spectrum of the SO3•− radical, with a maximum at 255 nm. Waygood etc. [30] used S2O62− as SO3•− radical source at pH 4.3 and observed an absorption maximum at 260 nm. Thus, the main photochemical process for sulfite system at 266 nm excitation is monophotonic photoionization (Reaction (1)).

3.3. Decay of eaq in Aqueous Solution

Lowering the pH from 9.5 to 7.0 led to a decrease not only of photoionization quantum yield, but also lifetime of eaq (Figure 5). Moreover, the observed rate constant (kobs720 nm) of eaq decay at pH 7 exhibited linear dependence on sulfite concentration (Figure 5), indicating that this species was consumed by the reaction with HSO3 ions. Therefore, using the data of Figure 4 and the fact that at pH 7 about 60% of S(IV) was in the form of HSO3 ions, one could calculate the rate constant for eaq quenching by HSO3 (k = 8 × 107 M−1s−1) which is consistent with that in a previous literature report [28].

4. Conclusions

The UV-sulfite system has been successfully used to oxidize As(III). Photolysis of sulfite by 254 nm lamp irradiation induced the production of SO3•− and secondary SO5•−, SO4•− and HO•. Oxysulfur radicals were responsible for As(III) oxidation at neutral or alkaline pH. Oxygen played a vital role in promoting As(III) oxidation. Through LFP experiments, we observed the signals of eaq and SO3•− at 720 and 255 nm, respectively, providing evidence for sulfite photoionization. SO32− ions (φion266nm = 0.06) exhibited a much higher quantum yield of photoionization than HSO3 ions (φion266nm = 0.01), indicating that alkaline pH was more favorable for application of the UV-S(IV)-O2 system. The rate constant (k = 8 × 107 m−1 s−1) for reaction between eaq and HSO3 has been measured.

Supplementary Materials

The following are available online.

Author Contributions

T.L. and I.P.P. performed the LFP experiments; T.L., Z.W. and Y.W. performed the As(III) oxidation experiments; Z.L. and I.P.P. conceived and designed the experiments; Z.L. and I.P.P. analyzed the data; T.L. and I.P.P. participated in drafting the article; Z.L. revised the article.

Funding

The work was financially supported by a joint project from the NSFC and the Russian Foundation for Basic Research (NSFC-RFBR No. 51811530099 and No. 18-53-53006_GFEN) and Russian Foundation for Basic Research project No. 17-03-00252.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Han, D.; Wan, J.; Ma, Y.; Wang, Y.; Li, Y.; Li, D.; Guan, Z. New insights into the role of organic chelating agents in Fe(II) activated persulfate processes. Chem. Eng. J. 2015, 269, 425–433. [Google Scholar] [CrossRef]
  2. Zhou, D.; Yuan, Y.; Yang, S.; Gao, H.; Chen, L. Roles of oxysulfur radicals in the oxidation of acid orange 7 in the Fe(III)–sulfite system. J. Sulfur Chem. 2015, 36, 373–384. [Google Scholar] [CrossRef]
  3. Deng, W.; Zhao, H.; Pan, F.; Feng, X.; Jung, B.; Abdel-wahab, A.; Batchelor, B.; Li, Y. Visible-light-driven photocatalytic degradation of organic water pollutants promoted by sulfite addition. Environ. Sci. Technol. 2017, 51, 13372–13379. [Google Scholar] [CrossRef]
  4. Li, X.; Ma, J.; Liu, G.; Fang, J.; Yue, S.; Guan, Y.; Chen, L.; Liu, X. Efficient reductive dechlorination of monochloroacetic acid by sulfite/UV process. Environ. Sci. Technol. 2012, 46, 7342–7349. [Google Scholar] [CrossRef] [PubMed]
  5. Song, Z.; Tang, H.; Wang, N.; Zhu, L. Reductive defluorination of perfluorooctanoic acid by hydrated electrons in a sulfite-mediated UV photochemical system. J. Hazard. Mater. 2013, 262, 332–338. [Google Scholar] [CrossRef] [PubMed]
  6. Neta, P.; Huie, R.E. Free-radical chemistry of sulfite. Environ. Health Perspect. 1985, 64, 209–217. [Google Scholar] [CrossRef] [PubMed]
  7. Park, H.; Vecitis, C.D.; Cheng, J.; Choi, W.; Mader, B.T.; Hoffmann, M.R. Reductive defluorination of aqueous perfluorinated alkyl surfactants: Effects of ionic headgroup and chain length. J. Phys. Chem. A 2009, 113, 690–696. [Google Scholar] [CrossRef]
  8. Huie, R.E.; Neta, P. Chemical behavior of SO3•− and SO5•− radicals in aqueous solutions. J. Phys. Chem. 1984, 88, 5665–5669. [Google Scholar] [CrossRef]
  9. Buxton, G.V.; Mcgowan, S.; Salmon, G.A.; Willians, J.E.; Wood, N.D. A study of the spectra and reactivity of oxysulphur-radical anions involved in the chain oxidation of S(IV):a pulse and -radiolysis study. Atmos. Environ. 1996, 30, 2483–2493. [Google Scholar] [CrossRef]
  10. Huie, R.E.; Neta, P. Rate constants for some oxidations of S(IV) by radicals in aqueous solutions. Atmos. Environ. 1987, 21, 17431747. [Google Scholar] [CrossRef]
  11. Deister, U.; Warneck, P. Photooxidation of SO32− in aqueous solution. J. Phys. Chem. 1990, 94, 2191–2198. [Google Scholar] [CrossRef]
  12. Herrmann, H.; Reese, A.; Zellner, R. Time-resolved UV/VIS diode array absorption spectroscopy of SOx•−(x = 3, 4, 5) radical anions in aqueous solution. J. Mol. Struct. 1995, 348, 183–186. [Google Scholar] [CrossRef]
  13. Wine, P.H.; Tang, Y.; Thorn, R.P.; Wells, J.R.; Davis, D.D. Kinetics of aqueous phase reactions of the SO4•− radical with potential importance in cloud chemistry. J. Geophys. Res. Atmos. 1989, 94, 1085–1094. [Google Scholar] [CrossRef]
  14. Rodríguez-lado, L.; Sun, G.; Berg, M.; Zhang, Q.; Xue, H.; Zheng, Q.; Johnson, C.A. Groundwater arsenic contamination throughout China. Science 2013, 341, 866–869. [Google Scholar] [CrossRef] [PubMed]
  15. Bissen, M.; Frimmel, F.H. Arsenic-A review. Part II: Oxidation of arsenic and its removal in water treatment. Acta Hydrochim. Hydrobiol. 2003, 31, 97–107. [Google Scholar] [CrossRef]
  16. Ryu, J.; Monllor-Satoca, D.; Kim, D.H.; Yeo, J.; Choi, W. Photooxidation of arsenite under 254 nm irradiation with a quantum yield higher than unity. Environ. Sci. Technol. 2013, 47, 9381–9387. [Google Scholar] [CrossRef] [PubMed]
  17. Yeo, J.; Choi, W. Iodide-mediated photooxidation of arsenite under 254 nm irradiation. Environ. Sci. Technol. 2009, 43, 3784–3788. [Google Scholar] [CrossRef] [PubMed]
  18. Yuan, Y.; Yang, S.; Zhou, D.; Wu, F. A simple Cr(VI)–S(IV)–O2 system for rapid and simultaneous reduction of Cr(VI) and oxidative degradation of organic pollutants. J. Hazard. Mater. 2016, 307, 294–301. [Google Scholar] [CrossRef] [PubMed]
  19. Liu, Z.; Yang, S.; Yuan, Y.; Xu, J.; Zhu, Y.; Li, J. A novel heterogeneous system for sulfate radical generation through sulfite activation on a CoFe2O4 nanocatalyst surface. J. Hazard. Mater. 2017, 324, 583–592. [Google Scholar] [CrossRef]
  20. Chen, L.; Luo, T.; Yang, S.; Xu, J.; Liu, Z.; Wu, F. Efficient metoprolol degradation by heterogeneous copper ferrite/sulfite reaction. Environ. Chem. Lett. 2018, 16, 599–603. [Google Scholar] [CrossRef]
  21. Cullen, W.R.; McBride, B.C.; Reglinski, J. The reduction of trimethylarsine oxide to trimethylarsine by thiols: a mechanistic model for the biological reduction of arsenicals. J. Inorg. Biochem. 1984, 21, 45–60. [Google Scholar] [CrossRef]
  22. Zhang, W.N.; Singh, P.; Muir, D. Kinetics of oxidation of As(III) with SO2/O2 and UV light. In Minor Elements 2000: Processing and Environmental Aspects of As, Sb, Se, Te, and Bi; Young, C., Ed.; Society for Mining, Metallurgy, and Exploration: Littleton, Co, USA, 2000; pp. 333–343. [Google Scholar]
  23. Yang, J.Q.; Chai, L.Y.; Li, Q.Z.; Shu, Y. De Redox behavior and chemical species of arsenic in acidic aqueous system. Trans. Nonferrous Met. Soc. China 2017, 27, 2063–2072. [Google Scholar] [CrossRef]
  24. Müller, K.; Ciminelli, V.S.T.; Dantas, M.S.S.; Willscher, S. A comparative study of As(III) and As(V) in aqueous solutions and adsorbed on iron oxy-hydroxides by Raman spectroscopy. Water Res. 2010, 44, 5660–5672. [Google Scholar] [CrossRef] [PubMed]
  25. Hayon, E.; Treinin, A.; Wilf, J. Electronic Spectra, Photochemistry, and Autoxidation Mechanism of the Sulfite-Bisulfite-Pyrosulfite Systems. the SO2•−, SO3•−, SO4•−, and SO5•− Radicals. J. Am. Chem. Soc. 1972, 94, 47–57. [Google Scholar] [CrossRef]
  26. Xu, J.; Ding, W.; Wu, F.; Mailhot, G.; Zhou, D.; Hanna, K. Rapid catalytic oxidation of arsenite to arsenate in an iron(III)/sulfite system under visible light. Appl. Catal. B Environ. 2016, 186, 56–61. [Google Scholar] [CrossRef]
  27. Yuan, Y.; Luo, T.; Xu, J.; Li, J.; Wu, F.; Brigante, M.; Mailhot, G. Enhanced oxidation of aniline using Fe(III)-S(IV) system: Role of different oxysulfur radicals. Chem. Eng. J. 2019, 362, 183–189. [Google Scholar] [CrossRef]
  28. Mcelroy, W.J.; Waygood, S.J. Oxidation of Formaldehyde by the Hydroxyl Radical in Aqueous Solution. J. Chem. Soc. Faraday Trans. 1991, 87, 1513–1521. [Google Scholar] [CrossRef]
  29. Hart, J.; Radical, H.; Michael, B.D.; Hart, E.J. The Rate Constants of Hydrated Electron, Hydrogen Atom, and Hydroxyl Radical Reactions with Benzene, 1,5-Cyclohexadiene, 1,4-Cyclohexadiene, and Cyclohexene. J. Phys. Chem. 1969, 74, 2878–2884. [Google Scholar]
  30. Waygood, S.J.; Mcelroy, W.J. Spectroscopy and Decay Kinetics of the Sulfite Radical Anion. J. Chem. Soc. Faraday Trans. 1992, 88, 1525–1530. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds mentioned in this paper are available from the authors.
Figure 1. As(III) oxidation efficiency by the UVC-S(IV)-O2 system and related control systems after a reaction time of 10 min. Initial conditions: [S(IV)]0 = 2 mM, [As(III)]0 = 5 μM. For most experiments, reaction solutions were stirred and exposed to the air, except in experiments with bubbling of synthetic air at a flow rate of about 500 mL min−1.
Figure 1. As(III) oxidation efficiency by the UVC-S(IV)-O2 system and related control systems after a reaction time of 10 min. Initial conditions: [S(IV)]0 = 2 mM, [As(III)]0 = 5 μM. For most experiments, reaction solutions were stirred and exposed to the air, except in experiments with bubbling of synthetic air at a flow rate of about 500 mL min−1.
Molecules 24 02307 g001
Figure 2. As(III) oxidations in solutions containing scavengers for the relevant radicals generated by UVC-S(IV)-O2 system at (a) pH 7 and (b) 9.5. Initial conditions: [S(IV)]0 = 2 mM, [As(III)]0 = 5 μM, [tert.-butyl alcohol (TBA)]0 = 2 mM, [EtOH]0 = 5 or 177 mM. The flow rate of N2 was about 500 mL min−1. Except for the N2 experiments, other results were obtained under aerobic conditions by constantly pumping synthetic air into the solution.
Figure 2. As(III) oxidations in solutions containing scavengers for the relevant radicals generated by UVC-S(IV)-O2 system at (a) pH 7 and (b) 9.5. Initial conditions: [S(IV)]0 = 2 mM, [As(III)]0 = 5 μM, [tert.-butyl alcohol (TBA)]0 = 2 mM, [EtOH]0 = 5 or 177 mM. The flow rate of N2 was about 500 mL min−1. Except for the N2 experiments, other results were obtained under aerobic conditions by constantly pumping synthetic air into the solution.
Molecules 24 02307 g002
Scheme 1. Proposed mechanism of action of the UVC/S(IV) system with/without oxygen.
Scheme 1. Proposed mechanism of action of the UVC/S(IV) system with/without oxygen.
Molecules 24 02307 sch001
Figure 3. Laser flash photolysis (LFP) of 50 mM sulfite at pH 9.5 in argon-saturated solution. Transient absorption spectra at different time delays after excitation. The solid line is the literature spectrum of eaq taken from [29]. Insert:kinetic curve at 720 nm with the first-order best fit with lifetime 8 μs.
Figure 3. Laser flash photolysis (LFP) of 50 mM sulfite at pH 9.5 in argon-saturated solution. Transient absorption spectra at different time delays after excitation. The solid line is the literature spectrum of eaq taken from [29]. Insert:kinetic curve at 720 nm with the first-order best fit with lifetime 8 μs.
Molecules 24 02307 g003
Figure 4. LFP of 50 mM sulfite at pH 9.5 in air-saturated solution. Transient absorption spectra at 1.2 μs delay (points) after excitation corresponding to the SO3•− radical.
Figure 4. LFP of 50 mM sulfite at pH 9.5 in air-saturated solution. Transient absorption spectra at 1.2 μs delay (points) after excitation corresponding to the SO3•− radical.
Molecules 24 02307 g004
Figure 5. Kinetics of decay of eaq absorption at 720 nm at pH 7 and 9.5. Smooth black curves are the monoexponential best fits with lifetimes of 0.43 and 8.6 μs, respectively. Insert: dependence of kobs720 nm at pH 7 upon sulfite concentration.
Figure 5. Kinetics of decay of eaq absorption at 720 nm at pH 7 and 9.5. Smooth black curves are the monoexponential best fits with lifetimes of 0.43 and 8.6 μs, respectively. Insert: dependence of kobs720 nm at pH 7 upon sulfite concentration.
Molecules 24 02307 g005

Share and Cite

MDPI and ACS Style

Luo, T.; Wang, Z.; Wang, Y.; Liu, Z.; P. Pozdnyakov, I. Different Role of Bisulfite/Sulfite in UVC-S(IV)-O2 System for Arsenite Oxidation in Water. Molecules 2019, 24, 2307. https://doi.org/10.3390/molecules24122307

AMA Style

Luo T, Wang Z, Wang Y, Liu Z, P. Pozdnyakov I. Different Role of Bisulfite/Sulfite in UVC-S(IV)-O2 System for Arsenite Oxidation in Water. Molecules. 2019; 24(12):2307. https://doi.org/10.3390/molecules24122307

Chicago/Turabian Style

Luo, Tao, Zhenhua Wang, Yi Wang, Zizheng Liu, and Ivan P. Pozdnyakov. 2019. "Different Role of Bisulfite/Sulfite in UVC-S(IV)-O2 System for Arsenite Oxidation in Water" Molecules 24, no. 12: 2307. https://doi.org/10.3390/molecules24122307

Article Metrics

Back to TopTop