Вы находитесь на странице: 1из 6

Environ. Sci. Technol.

2010, 44, 72327237

Chromate-Induced Activation of Hydrogen Peroxide for Oxidative Degradation of Aqueous Organic Pollutants
ALOK D. BOKARE AND WONYONG CHOI* School of Environmental Science and Engineering, Pohang University of Science and Technology (POSTECH), Pohang 790-784, Korea

Received December 27, 2009. Revised manuscript received April 2, 2010. Accepted April 9, 2010.

The oxidation of organic compounds in water was investigated with using chromate as an activator of H2O2. 4-chlorophenol (4-CP) was used as a main model substrate, and its degradation was successfully achieved at circumneutral pH. Unlike the traditional Fenton-based activation of H2O2 that is mainly limited to acidic condition, the oxidative capacity of the proposed Cr(VI)/ H2O2 system is active over a wide range of pH 3-11. H2O2 substitutes the oxo ligands of chromate by the peroxo ligands and, subsequently, converts chromate(VI) into a tetraperoxochromate(V) complex. The instantaneous disproportionation between chromium-coordinated peroxo ligands initiates the generation of HO that are responsible for the degradation of organic compounds in the Cr(VI)/H2O2 system. The oxidation rate of 4-CP and the in situ generated concentration of peroxochromate(V) decreased with increasing pH. The generation of HO in the Cr(VI)/H2O2 solution was conrmed by monitoring the production of p-hydroxybenzoic acid from the oxidation of benzoic acid as a probe reaction and by quenching the degradation of 4-CP in the presence of methanol as a radical scavenger. The oxidation of 4-CP investigated at different H2O2 concentrations and pH indicated the pH-dependent competition between peroxo ligand exchange and dissociation reactions. The proposed Cr(VI)/H2O2 process can be ideally suited for the treatment of chromate-contaminated wastewaters with recalcitrant organic compounds. The degradation of 4-CP in actual Cr(VI)-contaminated wastewater was successfully demonstrated in the presence of added H2O2. The Cr(VI)/H2O2 system is proposed as a viable advanced oxidation process.

use of hydrogen peroxide (H2O2) as a precursor of HO has received great attention because it is an environmentally benign oxidant. H2O2 decomposes to HO upon activation by transition metal ions (e.g., Fe2+), UV irradiation, and TiO2 photocatalyst (9, 10). Further, the easy availability (million metric ton-scale) and low price (ca. 1.0 $/kg of 100% H2O2) (11) have facilitated large-scale applications of the H2O2based oxidation processes. The success of the Fentons reagent (Fe2+ + H2O2) in the oxidation of various organic contaminants has motivated continuous research efforts on the transition-metal-induced generation of HO from H2O2. Modications of the classical Fenton reaction such as photo-Fenton have been investigated but the precipitation of ferric ions at pH > 3 limits the practical applicability to acidic conditions only (12, 13). To extend the Fentons chemistry to neutral and/or alkaline solutions, synthetic chelating ligands like aminopolycarboxylates and N-heterocyclic carboxylates, or macrocyclic iron-chelating agents (e.g., porphyrins, tetraamido macrocycles) are used to stabilize Fe3+ and prevent precipitation (14, 15). Ferric precipitation problems can be also reduced by immobilizing Fe(II)/Fe(III) on inert supports like Naon (16, 17) or by synthesizing Fe-clay heterogeneous nanocomposites using laponite or bentonite (18). Alternatively, copper-based Fenton-like systems containing organic or amino acids efciently oxidize organic dyes up to pH 9 via the catalytic decomposition of H2O2 (19, 20). However, these iron-chelating organic ligands used in Fenton-like systems lower the efciency of pollutant oxidation by scavenging HO, and also cause additional pollution by increasing the total organic content in water (21). From this point of view, it is highly desirable that HO should be efciently generated from the metal-induced decomposition of H2O2 over a wide pH range without the addition of secondary organic compounds. The redox states of the involved metal species should be stable over a wide pH range. Compared to iron or copper, chromium exists in a wider range of possible oxidation states (from -2 to +6), with the trivalent (Cr3+) and hexavelant (Cr6+ or chromate) species commonly found in water. Being an oxyanion, chromate is completely soluble over the entire pH range (22). The reduction of Cr(VI) proceeds through generating the reactive chromium intermediates like Cr(V) and Cr(IV), which are capable of generating HO from H2O2 via HaberWeiss-type reaction (23). The reactivity of Cr(V) is known from studies of chromates biological toxicity, which is ascribed to HO generation from the reaction of Cr(V) with enzyme-generated H2O2 present in living cells (24). Cr(VI) + reductant f Cr(V) f Cr(IV) f Cr(III) (1)

Introduction
Advanced oxidation processes (AOPs) based on the in situ formation and the subsequent reaction of hydroxyl radicals (HO) have emerged as innovative technologies for the rapid destruction of recalcitrant organic pollutants (1). The nonselective reactivity of HO toward most organic pollutants with near diffusion-limited bimolecular rate constants (108-109 M-1 s-1) is responsible for its ability to mineralize toxic organic compounds to water and CO2 (2, 3). Various physicochemical methods of generating HO radicals in AOPs include Fenton reaction (4), sonolysis (5), photocatalysis (6), ozonation (7) and their combination (8). In particular, the
* Corresponding author phone: +82-54-279-2283; fax: +82-54279-8299; e-mail: wchoi@postech.edu.
7232
9

Cr(n) + H2O2 f Cr(n + 1) + HO + HO-(n ) IV or V) (2) In biological systems, reductants like ascorbic acid, cysteine, or fructose convert Cr(VI) to active Cr(V) species. However, H2O2 can directly activate Cr(VI) as well. Although H2O2 is a strong oxidant [E0(H2O2/H2O) ) 1.77 V], it can also serve as a reductant of the stronger oxidant, Cr(VI) [E0(O2/ H2O2) ) 0.68 V and E0(Cr6+/Cr3+) ) 1.35 V] (25). Several investigations have been conducted on the reduction of Cr(VI) by H2O2 to conrm the production of peroxychromium complexes under different experimental conditions (pH, reactant ratios, temperature, ionic strength, etc.) (26-28). The reduction of Cr(VI) by H2O2 is also used for removing Cr(VI) from wastewaters (29). However, the redox chemistry
10.1021/es903930h 2010 American Chemical Society

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 44, NO. 19, 2010

Published on Web 04/21/2010

of H2O2-mediated Cr(VI) reduction and the accompanied generation of HO have not been exploited for the degradation of pollutants. In this work, we have investigated the oxidation capacity of the Cr(VI)/H2O2 redox system for the degradation of aquatic organic pollutants. The oxidation mechanism involves the formation of a tetraperoxochromate(V) complex and its subsequent reaction leads to the generation of hydroxyl radicals. The oxidation reaction is active over a wide range of pH 3-11 unlike the Fenton system. Although Cr(VI) is a highly toxic species and not suitable as a reactant that should be added to water, the proposed method should be proper for the treatment of industrial wastewaters that are contaminated with Cr(VI) and recalcitrant organic pollutants.

Experimental Section
Chemicals and Materials. Chemicals that were used as received in this study include: sodium chromate (Sigma), sodium dichromate (Sigma), hydrogen peroxide (30%, Kanto), 4-chlorophenol (4-CP, Sigma), phenol (Aldrich), orange II (Aldrich), rhodamine B (Aldrich), nitrobenzene (Aldrich), benzoic acid (Aldrich), p-hydroxy benzoic acid (p-HBA, Aldrich), methanol (Daejung), and acetonitrile (Merck). All solutions were prepared in ultrapure water (18 M cm) prepared by a Barnstead purication system. Chromatecontaminated wastewater ([Cr(VI)] 45 mM and pH 1.6) was obtained from a local electroplating industry (SermaGard Corporation, Korea). Procedure and Analytical Methods. The reactions were carried out in 50-mL glass beakers stirred on a magnetic stirrer. An aliquot of stock solution of 4-CP (or other substrate, 1 mM) was added to make a desired concentration (typically 0.1 mM), and then the reaction was initiated by the sequential addition of chromate and H2O2. Unless otherwise mentioned, the chromate concentration was xed at 2 mM. The initial pH of the solution was adjusted to a desired value with 1 N HClO4 or 1 N NaOH standard solution. Sample aliquots (1 mL) were withdrawn at regular time intervals from the reactor and injected into 4-mL glass vials containing 50 L sodium sulte (Na2SO3, 2 M) to quench residual H2O2. All experiments were carried out in triplicate for a given condition. Quantitative analysis of substrates was done using a high performance liquid chromatograph (HPLC Agilent 1100) equipped with a C-18 column (Agilent Zorbax 300SB) and a diode-array detector. The eluent compositions were as follows: (a) 0.1% phosphoric acid aqueous solution and acetonitrile (80:20 v/v) for 4-CP, (b) water, acetonitrile and acetic acid (78:20:2 v/v) for phenol, (c) water and methanol (50:50 v/v) for nitrobenzene, and (d) 0.1% phosphoric acid aqueous solution and acetonitrile (85:15 v/v) for benzoic acid. Quantication of ionic intermediates/products was performed using an ion chromatograph (IC, Dionex DX-120) equipped with Dionex IonPac AS-14 column and a conductivity detector. The eluent composition was 3.5 mM Na2CO3 + 1 mM NaHCO3. The concentrations of dyes (orange II and rhodamine B) and in situ generated peroxochromate(V) were monitored using a UV-visible spectrophotometer (Agilent 8453). Chromate concentration was determined using the diphenylcarbazide (DPC) method (30). Total organic carbon (TOC) was measured using a TOC analyzer (TOC-VCSH, Shimadzu).

FIGURE 1. (a) Degradation of 4-CP in the presence of chromate and H2O2. (b) The concurrent variation in the chromate concentration during the course of 4-CP degradation ([4-CP]0 ) 100 M, [Cr(VI)]0 ) 2 mM, [H2O2]0 ) 20 mM, pHi (unadjusted) ) 8.0, pHf ) 8.1). In panel b, 1 N HCl (3 mL) and H2O2 (20 mM) were added after completing the removal of 4-CP. The nal [Cr(VI)] was below the detection limit of the DPC method (<0.05 M).The error bars in all gures represent the standard deviation among triplicate measurements. ruled out the possibility of 4-CP removal by volatilization. The fact that Cr(VI) or H2O2 alone did not degrade 4-CP implies that a reaction between Cr(VI) and H2O2 should generate a reactive species that is responsible for the degradation of 4-CP. The concurrent production of chloride ions (Figure 1) accounted for 85% of the removed 4-CP. The minor decit may be ascribed to the generation of chlorinated intermediates that were not determined in this work. TOC was also reduced by 28((3)% in 6 h reaction, which indicates that some fraction of 4-CP was actually mineralized. This clearly demonstrates that 4-CP can be oxidatively degraded in the Cr(VI)/H2O2 system even at circumneutral pH. The oxidative capacity of Cr(VI)/H2O2 was further tested for other organic pollutants. Figure 2 shows that the degradation of phenol, nitrobenzene, orange II, and rhodamine B was successfully achieved at pH 7 in the Cr(VI)/H2O2 solution. The TOC was also signicantly removed, conrming the oxidative degradation of all substrates. The decrease of TOC in 6 h reaction was 53((2)% for phenol, 56((1)% for nitrobenzene, 67((5)% for orange II, and 37((2)% for rhodamine B.
VOL. 44, NO. 19, 2010 / ENVIRONMENTAL SCIENCE & TECHNOLOGY
9

Results and Discussion


Oxidation in Cr(VI)/H2O2 System. To evaluate the oxidative capacity of the Cr(VI)/H2O2 system, 4-CP was selected as a model substrate, and its degradation in aqueous solution was investigated. As shown in Figure 1, the complete degradation of 0.1 mM 4-CP was achieved in 6 h in the airequilibrated solution of pHi 8. The decomposition was not observed at all in the absence of either Cr(VI) or H2O2, which

7233

FIGURE 2. Degradation of phenol, nitrobenzene, orange II and rhodamine B (separate single-component experiments) in the presence of chromate and H2O2. [Cr(VI)]0 ) 2 mM, [H2O2]0 ) 20 mM, [substrate]0 ) 100 M and pHi ) 7. Although the Cr(VI)/H2O2 system exhibits an efcient oxidative capacity at neutral pH and can be a good candidate for ex situ oxidation treatment method, the toxic nature of Cr(VI) raises serious concerns regarding its utilization as a water treatment reagent. The homogeneous Fentons process requires 50-80 mg/L of Fe ions and about 2 mg/L of Fe ions in treated waters is permitted to be discharged directly into the environment (31). In case of Cr(VI), however, the maximum level permitted in the treated wastewaters is only 0.05 mg/L (or 1 M) (32). Figure 1b shows the variation in [Cr(VI)] during the oxidation of 4-CP at pHi ) 8. The concentration rapidly decreases by 40% after H2O2 addition, but does not change further over a 6 h reaction period. This decrease in [Cr(VI)] is attributed to its reduction to Cr(III) (reaction 3). 2CrO42- + 3H2O2 + 10H+ f 2Cr3+ + 3O2 + 8H2O (3) The conversion from Cr(VI) to Cr(III) in the presence of H2O2 depends on pH as the reaction 3 demands protons as a reagent (26). At pH 8, the complete reduction to Cr(III) cannot be attained as shown in Figure 1b. After the complete removal of 4-CP was achieved in 6 h, the residual Cr(VI) could be fully reduced to Cr(III) by the addition of 1 N hydrochloric acid and 20 mM H2O2. Therefore, the toxic Cr(VI) species can be initially utilized for activating H2O2 and then removed after completing the oxidation of organics. Figure 3 shows the effect of initial pH on the kinetics of 4-CP degradation. The oxidation of 4-CP decreased with increasing pHi and was completely inhibited at pHi ) 13. This clearly indicates that the Cr(VI)-induced activation of H2O2 is favored at acidic conditions. The effects of [H2O2] on 4-CP degradation are compared between pH 8 and 4 in Figure 4. At pHi ) 8, the efciency of 4-CP degradation decreased signicantly when [H2O2] was reduced from 20 to 2 mM (Figure 4a). In contrast, the degradation of 4-CP at pHi ) 4 was less affected by the decrease in [H2O2] (Figure 4b). This also conrms that the Cr(VI)-induced activation of H2O2 is more efcient at acidic pH. Less amount of H2O2 is required at acidic pH. As reaction 3 indicates, the reaction between Cr(VI) and H2O2 is highly favored at lower pH. This implies that the reactive intermediate oxidants that are generated in situ in the course of reaction 3 could be also favored at lower pH. The chemical interaction between Cr(VI) and H2O2 depends on the speciation and redox potential of both reactants. The major Cr(VI) species at pH < 6 is HCrO4-,
7234
9

FIGURE 3. Effect of initial pH on the degradation of 4-CP in the Cr(VI)/H2O2 system. [4-CP]0 ) 100 M, [Cr(VI)]0 ) 2 mM and [H2O2]0 ) 20 mM. The pHf values increased by <0.4 unit (for all pHi < 7) and by <0.1 unit (for all pHi > 7). while CrO42- is the major species at pH > 6. On the other hand, H2O2 remains stable in the pH range 1-10 but deprotonates at pH > 11 (pKa ) 11.65) to form HO2- (see Supporting Information Figure S1). Since CrO42- is a much weaker oxidant than HCrO4- [E0(CrO42-/Cr(OH)3 ) -0.13 VNHE vs E0(HCrO4-/Cr3+ ) 1.35 VNHE) (33)], CrO42- (dominant at pH > 6) is less efcient than HCrO4- for the Cr(VI)-induced oxidative decomposition of H2O2. At the condition of [H2O2] ) 2 mM and pHi ) 4, the degradation rate of 4-CP rapidly increased with increasing [Cr(VI)] from 0.02 to 2 mM (Figure 4c). In addition, dichromate(Cr2O72-) was also tested as an activator of H2O2 instead of chromate and demonstrated similar reactivity for the oxidation of 4-CP (see Supporting Information Figure S2). Generation of HO and Oxidative Reaction Mechanism. To ascertain whether chromate-mediated H2O2 decomposition generates hydroxyl radicals, the oxidative conversion of benzoic acid (BA) to p-hydroxybenzoic acid (p-HBA) was used as a probe reaction. The formation of p-HBA from the reaction of (BA + HO) has been used as an indirect method to detect the formation of HO (34, 35). Figure 5 shows the production of p-HBA from BA in the Cr(VI)/H2O2 system at different pH values, wherein the rate of p-HBA formation is initially rapid and then slowly approaches saturation over 6 h period. The initial rate of p-HBA formation decreases with increasing pH, which is similar to the pH-dependent behavior of 4-CP degradation (see Figure 3). The observation indicates that the concentration of HO generated from the reaction of Cr(VI) and H2O2 is higher at lower pH. Furthermore, the addition of methanol as a hydroxyl radical scavenger completely inhibited the degradation of 4-CP (Figure 6). This conrms that H2O2 acts as the precursor of hydroxyl radicals, which are primarily responsible for 4-CP oxidation in the presence of Cr(VI). The reaction mechanism between chromate and H2O2 has been extensively studied (over almost a century!) to explain its complicated kinetic behavior and pH-dependent pathways of chromate reduction (28, 36). A critical understanding of chromate-H2O2 redox process was prompted by the implication of H2O2 in chromate toxicity and carcinogenecity (23, 24) and potential applications of H2O2-mediated chromate removal in wastewaters (29). Mechanistic investigations employing electron spin resonance spectroscopy and spectrophotometric studies have conrmed that H2O2 can function as either a reducing or an oxidizing agent and the nal oxidation state of chromium (ranging from +3 to +6) depends on the acidity/basicity of the solution (28). All

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 44, NO. 19, 2010

FIGURE 5. Comparison of the time proles of p-HBA formation during the oxidation of benzoic acid (BA) in the Cr(VI)/H2O2 system. [BA]0 ) 10 mM, [Cr(VI)]0 ) 2 mM, and [H2O2]0 ) 20 mM.

FIGURE 6. Effect of methanol (OH radical scavenger) on the degradation of 4-CP in the Cr(VI)/H2O2 system. [4-CP]0 ) 100 M, [Cr(VI)]0 ) 2 mM, [H2O2]0 ) 20 mM, [CH3OH]0 ) 100 mM, pHi (not adjusted) ) 8.0 (pHf ) 8.1). reducing agent. The tetraperoxochromate(V) is metastable and ultimately undergoes the disproportionation of the peroxo ligands within the chromium coordination sphere, which produces superoxide (O2-) and singlet oxygen (1O2) with regenerating chromate (reaction 5) (29, 38-40). This redox process also occurs for chromium in the solid state wherein the thermal decomposition of potassium tetraperoxochromate(V) generates chromate and superoxide KO2 (41). [CrV(O2)4]3- f [CrVIO4]2- + O2- + 1O2 (5)

FIGURE 4. Effect of (a, b) [H2O2] and (c) [Cr(VI)] on the degradation of 4-CP (100 M). (a) pHi ) 8.0 (pHf ) 8.1) and [Cr(VI)]0 ) 2 mM. (b) pHi ) 4.0 (pHf ) 4.3) and [Cr(VI)]0 ) 2 mM. (c) pHi ) 4.0 (pHf ) 4.3) and [H2O2] ) 2 mM. redox pathways entail one-electron transfer processes but the rate of ligand substitution and stability of intermediate species is controlled by pH. This complex chemistry between chromate and H2O2 is initiated by the sequential substitution of oxo ligands by peroxo groups. This sequential ligand exchange and one-electron reduction leads to tetrakis(2peroxo)chromate(V) anion complex formation (eq 4) (28, 37). [Cr O4]
VI 2-

Hydroxyl radicals are then subsequently produced through the reaction of superoxide with the peroxochromate(V) complex (reaction 6). Alternatively, direct decomposition of the chromate(V) complex into chromate(VI) can also generate hydroxyl radicals (reaction 7) (41, 42). [CrV(O2)4]3- + nO2- + nH+ f [CrV(O2)4-n(O)n]3- + nHO + nO2 (n ) 1 - 4) [CrV(O2)4]3- + H+ f [CrVI(O2)3(O)]2- + HO (6) (7)

+ nH2O2 f [Cr (O)3(O2)]

VI

2-

f f

[CrVI(O)(O2)3]2- f [CrV(O2)4]3- (4) H2O2 plays a dual role in this process by providing four peroxide units for ligand exchange and also acting as a

The above equations indicate that the reaction of tetraperoxochromate(V) complex should be also favored at low pH like reaction 3.
VOL. 44, NO. 19, 2010 / ENVIRONMENTAL SCIENCE & TECHNOLOGY
9

7235

FIGURE 8. Degradation of 4-CP in Cr(VI)-contaminated wastewater obtained from an electroplating industry in the presence or absence of H2O2. The raw wastewater was diluted and adjusted to pH 7 to rule out the possible contribution from Fenton reaction of iron impurity. ([Cr(VI)]0 ) 2 mM, [H2O2]0 ) 20 mM and pHi ) 7). production of chromium(V) is reduced with increasing pHi. The production of chromium(V) was immediate at pHi ) 3, but was markedly reduced at pHi ) 6 and negligible at pH g 8. Since the oxidation of 4-CP was observed even above pH 8 (see Figure 3), peroxochromate(V) species seems to be generated at concentrations below the absorbance detection limit. This pH-dependent formation of peroxochromate(V) is consistent with the pH-dependent oxidation of 4-CP (see Figure 3), which supports that the peroxochromate(V) species is the precursor of HO (reactions 6-7). In conclusion, the overall oxidation mechanism in the Cr(VI)/H2O2 system can be summarized as follows: chromate(VI) is reduced to peroxochromate(V) by H2O2 and the subsequent dissociation of peroxochromate(V) accompanies the production of HO, which is highly favored at acidic condition. Cr(VI)/H2O2 Process for Wastewater Treatment. Practical applications of the H2O2-based oxidation processes, such as Fenton and Fenton-like reactions, for wastewater remediation are severely limited by the precipitation of Fe3+ ions at pH > 3 (12, 13), which limits the process to acidic conditions and requires large amount of ferrous salts. This creates an additional problem of iron sludge and its disposal. In contrast, chromate as an activator of H2O2 is soluble over a wide pH range. However, the extreme toxicity of chromate(VI) is a major hindrance toward practical applications despite the wide working pH range. The deliberate addition of chromate(VI) into wastewaters would not be sensible even if the post-treatment can remove chromate(VI) from water. The proposed process should be ideally suited for chromatecontaminated wastewaters with recalcitrant organic compounds. The chromate wastewaters are ususally very acidic, which is optimal for the proposed Cr(VI)/H2O2 process. Every year, the widespread use of chromate(VI) in electroplating, leather tanning, metal nishing, and petroleum rening industries generates huge volume of chromate wastewaters (more than 50 000 tons in the United States alone), which require extensive treatments to prevent groundwater and soil contamination. The addition of H2O2 into these industrial efuents, before pretreatment, would facilitate in situ oxidative degradation of organic pollutants present therein. We have successfully established the process viability by using Cr(VI)-contaminated wastewater from an electroplating industry for the complete removal of 4-CP at neutral pH (Figure 8). The contribution of Fe present in the wastewater ([Fe] < 2.5% of total [Cr(VI)]) toward the oxidation of 4-CP

FIGURE 7. (a) UV-visible absorption spectra of chromate solution before and after the addition of H2O2 at pHi ) 3. (b) Effect of pHi on the time prole of peroxochromate(V) formation and decomposition. [Cr(VI)]0 ) 2 mM, [H2O2]0 ) 20 mM and the ordinate values at time zero represent the absorbance immediately after H2O2 addition. To investigate the effect of pH on the formation and dissociation of the peroxochromate(V) complex, UV-visible absorption spectra of the chromate solution were recorded. Figure 7a shows the variation of the absorption spectra of chromate solution before and after the addition of H2O2. The peak at 352 nm corresponds to the ligand(oxygen)-to-metal charge-transfer band of chromate. When H2O2 is added in the solution, this peak decreases and the absorbance above 500 nm increases. The absorbance change at this wavelength during the reaction of Cr(VI) with H2O2 may indicate the appearance of the chromium(V) complex. Several other studies have conrmed that this absorbance spectrum belongs to the peroxochromate(V) anion complex (28, 43, 44). The presence of an isosbestic point at 404 nm conrms that chromium(VI) and chromium(V) directly convert into each other as reactions 4 and 5 indicate. Figure 7b shows the time proles of the absorbance at 500 nm in the Cr(VI)/H2O2 system at various pH. At low pH (pHi ) 3), the chromium(V) complex was initially formed upon reacting with H2O2 and then decayed. This decrease in absorbance is simultaneously accompanied by the increase in the absorbance at 370 nm (see Figure 7a), which indicates the regeneration of Cr(VI) through the conversion of the peroxochromate(V) complex (reaction 5). This explains why [Cr(VI)] remained constant throughout the reaction of Cr(VI) and H2O2 despite the continued oxidation of 4-CP (see Figure 1b). The initial
7236
9

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 44, NO. 19, 2010

can be ruled out at neutral pH. The Cr(VI)/H2O2 system is proposed as a viable AOP.

Acknowledgments
This work was supported by KOSEF NRL program (No. R0A2008-000-20068-0), the KOSEF EPB center (Grant No. R112008-052-02002), and KCAP (Sogang Univ.) funded by MEST through NRF (2009-C1AAA001-2009-0093879).

Supporting Information Available


Additional information as described in the text. This material is available free of charge via the Internet at http:// pubs.acs.org.

Literature Cited
(1) Gogate, P. R.; Pandit, A. B. A review of imperative technologies for wastewater treatment I: oxidation technologies at ambient conditions. Adv. Environ. Res. 2004, 8, 501551. (2) Huang, C. P.; Dong, C.; Tang, Z. Advanced chemical oxidation: Is present role and potential future in hazardous waste treatment. Waste Manage. 1993, 13, 361377. (3) Andreozzi, R.; Caprio, V.; Insola, A.; Marotta, R. Advanced oxidation processes (AOP) for water purication and recovery. Catal. Today 1999, 53, 5159. (4) Neyens, E.; Baeyens, J. A review of classic Fentonss peroxidation as an advanced oxidation technique. J. Hazard. Mater. 2003, B98, 3350. (5) Adewuyi, Y. G. Sonochemistry: environmental science and engineering applications. Ind. Eng. Chem. Res. 2001, 40, 4681 4715. (6) Hoffmann, M. R.; Martin, S. T.; Choi, W.; Bahnemann, D. W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 6996. (7) Lima, A. A.; Montalvao, A. F.; Dezotti, M.; SantAnna, G. L. Ozonation of a complex industrial efuent: oxidation of organic pollutants and removal of toxicity. Ozone: i. Eng. 2006, 28, 38. (8) Gogate, P. R.; Pandit, A. B. A review of imperative technologies for wastewater treatment II: Hybrid methods. Adv. Environ. Res. 2004, 8, 553597. (9) Bhandari, A.,Surampalli, R. Y.,Champagne, P.,Ong, S. K.,Tyagi, R. D.,Lo, I. M. C.,Eds. Remediation Technologies for Soils and Groundwater; American Society of Civil Engineers: Reston, VA, 2007. (10) Strukul, G., Ed. Catalytic Oxidations with Hydrogen Peroxide as Oxidant; Kluwer Academic Publishers: Amsterdam, 1992. (11) Magerlein, W.; Dreisbach, C.; Hugl, H.; Tse, M. K.; Klawonn, M.; Bhor, S.; Beller, M. Homogeneous and heterogeneous ruthenium catalysts in the synthesis of ne chemicals. Catal. Today 2007, 121, 140150. (12) Pignatello, J. J.; Oliveros, E.; MacKay, A. Advanced oxidation processes for organic contaminant destruction based on the Fenton reaction and related chemistry. Crit. Rev. Environ. Sci. Technol. 2006, 36, 184. (13) Laine, D. F.; Cheng, I. F. The destruction of organic pollutants under mild reaction conditions: a review. Microchem. J. 2007, 85, 183193. (14) Nam, S.; Renganathan, V.; Tratnyek, P. G. Substituent effects on azo dye oxidation by the FeIII-EDTA-H2O2 system. Chemosphere 2001, 45, 5965. (15) Collins, T. J. TAML oxidant activators: a new approach to the activation of hydrogen peroxide for environmentally signicant problems. Acc. Chem. Res. 2002, 35, 78290. (16) Fernandez, J.; Bandara, J.; Lopez, A.; Buffat, Ph.; Kiwi, J. Photoassisted Fenton degradation of nonbiodegradable azo dye (Orange II) in Fe-free solutions mediated by cation transfer membranes. Langmuir 1999, 15, 185192. (17) Balanosky, E.; Fernandez, J.; Kiwi, J.; Lopez, A. Degradation of membrane concentrates of the textile industry by Fenton like reactions in iron-free solutions at biocompatible pH values (pH 7-8). Water Sci. Technol. 1999, 40, 417424. (18) Feng, J.; Hu, X.; Yu, P. L.; Qiao, S. Photo Fenton degradation of high concentration Orange II (2 mM) using catalysts containing Fe: A comparative study. Sep. Purif. Technol. 2009, 67, 213217.

(19) Lin, T. Y.; Wu, C. H. Activation of hydrogen peroxide in coppe(II)/ amino acid/H2O2 systems: effects of pH and copper speciation. J. Catal. 2005, 232, 117126. (20) Shah, V.; Verma, P.; Stopka, P.; Gabriel, J.; Baldrian, P.; Nerud, F. Decolorization of dyes with copper(II)/organic acid/hydrogen peroxide systems. Appl. Catal., B 2003, 46, 287292. (21) Lee, C.; Sedlak, D. L. A novel homogeneous Fenton-like system with Fe(III)-phosphotungstate for oxidation of organic compounds at neutral pH values. J. Mol. Catal. A 2009, 311, 16. (22) Beverskog, B.; Puigdomenech, I. Revised Pourbaix diagrams for chromium at 25-300 C. Corros. Sci. 1997, 39, 4357. (23) Shi, X.; Chiu, A.; Chen, C. T.; Halliwell, B.; Castranova, V.; Vallyathan, V. Reduction of chromium(VI) and its relationship to carcinogenesis. J. Toxicol. Environ. Health B 1999, 2, 87104. (24) Bagchi, D.; Stohs, S. J.; Downs, B. W.; Bagchi, M.; Preuss, H. G. Cytotoxicity and oxidative mechanisms of different forms of chromium. Toxicology 2002, 180, 522. (25) Cotton, F. A.; Wilkinson, G.; Murillo, C. A.; Bochmann, M. Advanced Inorganic Chemistry; Wiley-Interscience: New York, 1999. (26) House, D. A. Recent developments in chromium chemistry. Adv. Inorg. Chem. 1996, 44, 341373. (27) Perez-Benito, J.; Arias, C. A kinetic study of the chromium(VI)hydrogen peroxide reaction. Role of diperoxochromate(VI) intermediates. J. Phys. Chem. A 1997, 101, 47264733. (28) Vander Griend, D. A.; Golden, J. S.; Arrington, C. A., Jr. Kinetics and mechanism of chromate reduction with hydrogen peroxide in base. Inorg. Chem. 2002, 41, 70427048. (29) Pettine, M.; Campanella, L.; Millero, F. Reduction of hexavalent chromium by H2O2 in acidic solutions. Environ. Sci. Technol. 2002, 36, 901907. (30) APHA, AWWA, WEF. Standard Methods for the Examination of Water and Wastewater, 20th ed.; APHA: Washington, DC, 1988; pp 3-65-3-68. (31) Sabhi, S.; Kiwi, J. Degradation of 2,4-dichlorophenol by immobilized iron catalysts. Water Res. 2001, 8, 19942002. (32) Acar, F. N.; Malkoc, E. The removal of chromium(VI) from aqueous solutions by Fagus orientalis L. Bioresour. Technol. 2004, 94, 1315. (33) CRC Handbook of Chemistry and Physics, 85th ed.; CRC Press: Boca Raton, FL, 2004. (34) Mopper, K.; Zhou, X. Hydroxyl radical photoproduction in the sea and its potential impact on marine processes. Science 1990, 250, 661664. (35) Joo, S. H.; Feitz, A. Z.; Sedlak, D. L.; Waite, T. D. Quantication of the oxidizing capacity of nanoparticulate zero-valent iron. Environ. Sci. Technol. 2005, 39, 12631268. (36) Baxendale, J. H. Decomposition of hydrogen peroxide by catalysts in homogeneous aqueous solution. Adv. Catal. 1952, 4, 3186. (37) Dickman, M. H.; Pope, M. T. Peroxo and superoxo complexes of chromium, molybdenum, and tungsten. Chem. Rev. 1994, 94, 569584. (38) Wiberg, K. B.; Schafer, H. J. Direct observation of intermediates in the chromic acid oxidation of alcohols. J. Am. Chem. Soc. 1967, 89, 455457. (39) Zhang, L.; Lay, P. L. EPR spectroscopic studies on the formation of chromium(V) peroxo complexes in the reaction of chromium(VI) with hydrogen peroxide. Inorg. Chem. 1998, 37, 1729 1733. (40) Hodgson, E. K.; Fridovich, I. Production of superoxide radical during the decomposition of potassium peroxochromate(V). Biochemistry 1974, 13, 38113815. (41) Peters, J. W.; Bekowies, P. J.; Winer, A. M.; Pitts, J. N., Jr. Potassium perchromate as a source of singlet oxygen. J. Am. Chem. Soc. 1975, 97, 32993306. (42) Kawanishi, S.; Inoue, S.; Sano, S. Mechanism of DNA cleavage induced by sodium chromate(VI) in the presence of hydrogen peroxide. J. Biol. Chem. 1986, 261, 59525958. (43) McGarvey, B. R. ESR and optical spectrum of potassium perchromate. J. Chem. Phys. 1962, 37, 20012004. (44) Quane, D.; Bartlett, B. Optical spectra of potassium tetraperoxychromate(V) and the violet diperoxychromate(VI) ion. J. Chem. Phys. 1970, 53, 44044405.

ES903930H

VOL. 44, NO. 19, 2010 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

7237

Вам также может понравиться