Rate equation for the degradation of nitrobenzene by ‘Fenton-like’ reagent

https://doi.org/10.1016/S1093-0191(02)00024-2Get rights and content

Abstract

This paper describes the effect of temperature and initial concentration of H2O2, Fe(II), PhNO2 and dissolved oxygen on the degradation rate of PhNO2 in homogeneous aqueous solution by ‘Fenton-like’ reagent ([H2O2]o≫[Fe(II)]o). The oxidation products o-, m- and p-nitrophenol were found as intermediates in the ratio 1:1.3–2.8:1.4–2.7 as compared with PhNO2 when conversion of the latter was less than 25%. This fact suggests that hydroxylation of PhNO2 was promoted by HOradical dot radicals. The reaction was investigated in a completely mixed-batch reactor under a wide range of experimental conditions: pH ∼3.0; 278–318 K; 1.5<[H2O2]o<26.5 mM; 0.04<[Fe(II)]o<1.1 mM; 0.3<[PhNO2]o<2.5 mM; and 0<[O2]o<1.4 mM. The activation energy for the degradation of PhNO2 was determined to be 59.7 kJ mol−1. The degradation rate of PhNO2 follows pseudo-first-order kinetics. The results of this study demonstrate that the degradation rate of PhNO2 in the ‘Fenton-like’ system could be predicted with sufficient precision by the equation RD=1.05×1011exp(−59.7/RT)[H2O2]o0.68[Fe(II)]o1.67[PhNO2]o−0.32. The second-order rate constant for the overall rate of H2O2 decomposition by Fe(III) was found to be 0.83 M−1 s−1 at 298 K. The value of the steady-state HOradical dot radical concentration in the ‘Fenton-like’ reaction was found to be ∼10−13 M, as estimated by two independent methods.

Introduction

Nitroaromatic compounds are widely used as raw materials in many industrial processes, such as the preparation of pesticides, explosives, textiles and paper. Consequently, these compounds are found as water pollutants as a result of their release in industrial wastewater (Beltran et al., 1998, Sarasa et al., 1998). Nitroaromatic compounds are not only important as constituents in industrial waste streams, but also as hazardous wastes that have contaminated soils, groundwater, sludges and other hazardous solid wastes regulated under the Resource Conservation and Recovery Act (US EPA, 1980, Sittig, 1985). Remediation of wastewaters containing these pollutants is very difficult, since they are usually resistant to biological degradation (O'Connor and Young, 1989).

Oxidation processes that involve the generation of the highly reactive hydroxyl radical (HOradical dot) are of current interest for the destruction of organic pollutants in surface and groundwaters, and industrial wastewaters. Generation of HOradical dot radicals by the dark reaction of H2O2 with ferrous salt (known as Fenton's reagent) has been the subject of numerous studies during the last decade (Arnold et al., 1995, Pignatello and Day, 1996, Chen and Pignatello, 1997, Hislop and Bolton, 1999). The oxidizing species generated in the Fenton reaction have been discussed by many investigators, but are still controversial (Walling, 1975, Stubbe and Kozarich, 1987, Bossmann et al., 1998, MacFaul et al., 1998, Goldstein and Meyerstein, 1999, Kremer, 1999). The recognition of the HOradical dot radical as the active intermediate is not yet universal, and doubts as to its very existence in the system have even been raised (Bossmann et al., 1998, Kremer, 1999). In other studies of Fenton's reagent, it is generally considered that the reaction between H2O2 and Fe(II) in acidic aqueous medium (pH ≤3) produces HOradical dot radicals [Eq. (1)] and can involve the steps presented below [Eq. (1)–Eq. (6)] (Haber and Weiss, 1934, Barb et al., 1951a, Barb et al., 1951b, Walling, 1975). The rate constants are reported at 298 K in M−1 s−1 for a second-order reaction rate (Lin and Gurol, 1998, De Laat and Gallard, 1999).FeII+H2O2FeIII+HO+HOk1=63FeIII+H2O2FeII+H++HOOk2=0.002−0.01FeII+HOFeIII+HOk3=3×108HO+H2O2HOO+H2Ok4=2.7×107FeII+HOOFeIII+HOOk5=1.2×106FeIII+HOOFeII+H++O2k6<2×103

The hypothesis of Haber and Weiss (1934) that the Fenton reaction involves the formation of HOradical dot radicals as the actual oxidants has been proved by many techniques, including EPR spectroscopy. Although a considerable number of investigators, using the electron paramagnetic resonance (EPR) spin-trapping technique, have found evidence for the formation of HOradical dot radicals from Fenton's reagent (Dixon and Norman, 1964, Buettner, 1987, Rosen et al., 2000), it has also been reported by others (Rush and Koppenol, 1987, Rahhal and Richter, 1988) that this species is not the only oxidizing intermediate, but that some type of high-valent iron-oxo intermediates also exist (Groves and Watanabe, 1986, Kean et al., 1987, Sychev and Isak, 1995, Bossmann et al., 1998, Kremer, 1999). Using EPR spin-trapping, three types of oxidizing species (free HOradical dot, bound HOradical dot and high-valence iron species, which is probably a ferryl ion, FeIVO) were detected by Yamazaki and Piette (1991). In the present work, the principal oxidant is assumed to be HOradical dot radical, but others, such as iron-oxo species, cannot be ruled out.

Nitrobenzene is one of the most representative nitroaromatic compounds present in several wastewaters and it is considerably soluble at room temperature. A number of studies on the degradation of PhNO2 in aerated aqueous solutions by Fenton's reagent have been reported (Lipczynska-Kochany, 1991, Lipczynska-Kochany, 1992), although they are actually ‘Fenton-like’ processes because [H2O2]o≫[Fe(II)]o. However, in a system containing PhNO2, H2O2 and Fe(II), the effects of temperature and initial concentrations of these components and dissolved oxygen on the degradation rate of PhNO2 have not been investigated in detail. Nitrobenzene half-lives of 360 min at [PhNO2]o=0.1 mM, [H2O2]o=3.9 mM, [FeCI2]o=0.035 mM (Lipczynska-Kochany, 1991) and of 250 min at [PhNO2]o=0.1 mM, [H2O2]o=8.0 mM, [FeCI2]o=0.035 mM (Lipczynska-Kochany, 1992) have been determined, but these were studied only at room temperature and were reported without any experimental details. In spite of the numerous studies of Fenton's reagent, the chemistry and kinetics of PhNO2 oxidation has not been well elucidated. Due to its importance as an environmental pollutant, this compound was selected for a study on its chemical degradation by ‘Fenton-like’ reagent. The study was conducted with the aim of determining some intermediate products and kinetic parameters for PhNO2 removal in pure water. Knowledge of the kinetic information is necessary for better the planning of better methods of degradation of the pollutant, as well as for mechanistic studies.

Section snippets

Materials and reagents

Nitrobenzene 99% (Probus), o-, m- and p-nitrophenol >99% (Merck), hydrogen peroxide 30% (Merck), oxygen 99.99% (AlphaGas), nitrogen 99.99% (AlphaGas), acetonitrile 99.8%, isocratic grade for HPLC (Merck), FeSO4·7H2O 98% (Panreac) and sodium hydrogen sulfite solution 40% w/v (Panreac) were used as received. All solutions of PhNO2, H2O2 and ferrous salt were prepared in Millipore water.

Degradation experiments

All experiments were conducted in a thermostatted batch glass reactor (1 l) equipped with a magnetic stirrer in

Intermediates in the degradation of PhNO2 by HPLC

To clarify the reaction pathways for PhNO2 after the dark ‘Fenton-like’ reaction, we worked on the identification of reaction products by HPLC. Identification of nitrophenols as reaction intermediates suggests that the degradation of PhNO2 by the ‘Fenton-like’ process takes place by reaction with HOradical dot radical, which is assumed to be the main oxidant. Reaction of PhNO2 with HOradical dot radicals to form HOradical dot–PhNO2 adducts (nitrohydroxycyclohexadienyl radicals) has a significantly high rate constant:HO+PhNO2

Conclusions

Nitrobenzene removal has been investigated by applying ‘Fenton-like’ reagent ([H2O2]o≫[Fe(II)]o), which was found to be an appropriate method to efficiently remove this compound from aqueous solutions. The study and mathematical treatment of typical operating parameters in this system resulted in the following statements related to the kinetics of the reaction:

  • The degradation rate of PhNO2 can be expressed as a pseudo-first-order reaction with respect to PhNO2 concentration. This reaction rate

Acknowledgements

M. Rodriguez expresses his gratitude to the ULA-CONICIT collaboration (Venezuela) for financial support and V. Timokhin thanks the Ministerio de Education y Cultura (Spain) for a NATO Science Fellowship. The authors wish to express their gratitude for the financial support given by the Ministerio de Ciencia y Tecnologı́a (Spain, Project AMB 99-0442). We gratefully acknowledge Profs M.L. Kremer (Hebrew University, Israel), M. Bekbolet (Bogazici University, Turkey) and N. Restrepo (Universidad

References (51)

  • J. Sarasa et al.

    Treatment of a wastewater resulting from dye manufacturing with ozone and chemical coagulation

    Water Res.

    (1998)
  • S.M. Arnold et al.

    Degradation of atrazine by Fenton's reagent: condition optimization and product quantification

    Environ. Sci. Technol.

    (1995)
  • W.G. Barb et al.

    Reactions of ferrous and ferric ions with hydrogen peroxide. Part I—the ferrous ion reaction

    Trans. Faraday Soc.

    (1951)
  • W.G. Barb et al.

    Reactions of ferrous and ferric ions with hydrogen peroxide. Part II—the ferric ion reaction

    Trans. Faraday Soc.

    (1951)
  • F.J. Beltran et al.

    Nitroaromatic hydrocarbon ozonation in water. 1. Single ozonation

    Ind. Eng. Chem. Res.

    (1998)
  • B. Bohn et al.

    Gas-phase reaction of the OH–benzene adduct with O2: reversibility and secondary formation of HO2

    Phys. Chem. Chem. Phys.

    (1999)
  • S.H. Bossmann et al.

    New evidence against hydroxyl radicals as reactive intermediates in the thermal and photochemically enhanced Fenton reactions

    J. Phys. Chem. A

    (1998)
  • R. Chen et al.

    Role of quinone intermediates as electron shuttles in Fenton and photo-assisted Fenton oxidations of aromatic compounds

    Environ. Sci. Technol.

    (1997)
  • J. De Laat et al.

    Catalytic decomposition of hydrogen peroxide by Fe(III) in homogeneous aqueous solution: mechanism and kinetic modeling

    Environ. Sci. Technol.

    (1999)
  • Dixon, W.T., Norman, R.O.C., 1964. Electron spin resonance studies of oxidation. Part III. Some alicyclic compounds. J....
  • S. Goldstein et al.

    Comments on the mechanism of the ‘Fenton-like’ reaction

    Acc. Chem. Res.

    (1999)
  • J.T. Groves et al.

    Oxygen activation by metalloporphyrins related to peroxidase and cytochrome P-450. Direct observation of the OO bond cleavage step

    J. Am. Chem. Soc.

    (1986)
  • F. Haber et al.

    The catalytic decomposition of hydrogen peroxide by iron salts

    Proc. R. Soc. Lond. A

    (1934)
  • T.J. Hardwick

    The rate constant of the reaction between ferrous ions and hydrogen peroxide in acid solution

    Can. J. Chem.

    (1957)
  • K.A. Hislop et al.

    The photochemical generation of hydroxyl radicals in the UV-vis/ferrioxalate/H2O2 system

    Environ. Sci. Technol.

    (1999)
  • Cited by (61)

    • The profound review of Fenton process: What's the next step?

      2025, Journal of Environmental Sciences (China)
    • Impact and implications of nanocatalyst in the Fenton-like processes for remediation of aquatic environment contaminated with micro-pollutants: A critical review

      2022, Journal of Water Process Engineering
      Citation Excerpt :

      The speciation diagram of iron shows that pH 2.8 to 3.2 is the optimum pH range for the Fenton process (Fig. 2) [46]. Some other researchers also showed that the pH range of 2 to 4 could provide an optimum treatment efficiency in the Fenton process [53–56]. Since wastewater samples are not found within this pH range, to attain the optimal pH range of the Fenton process, a high amount of acid is added to adjust the pH of wastewater samples prior to operating the Fenton process.

    View all citing articles on Scopus
    View full text