You are on page 1of 7

Journal of Electroanalytical Chemistry 507 (2001) 96 102

www.elsevier.com/locate/jelechem

Production of hydroxyl radicals by electrochemically assisted


Fentons reagent
Application to the mineralization of an organic micropollutant,
pentachlorophenol
Mehmet A. Oturan a,*, Nihal Oturan b, Claude Lahitte c, Stephane Trevin c
a
Laboratoire dElectrochimie Moleculaire, Uni6ersite Paris 7, 2 place Jussieu, 75251 Paris Cedex 05, France
Laboratoire de Geomateriaux, Uni6ersite de Marne la Vallee, 5 Boule6ard Descartes, bat. IFI, Champs/Marne,
77454 Marne la Vallee Cedex 2, France
c
EDF, Di6ision Recherche et De6eloppement, Site des Renardie`res, Route de Sens-Ecuelles, 77818 Moret sur Loing Cedex, France
b

Received 14 September 2000; received in revised form 20 December 2000; accepted 26 December 2000

Abstract
Hydroxyl radicals are very powerful oxidizing agents. They are involved in hydroxylation reactions, in biological and
atmospheric phenomena. A recent application of these radicals is their use in decontamination of water polluted by toxic organic
substances like pesticides. Chemically, these radicals are produced by the use of a mixture of (H2O2 + Fe2 + ), the so called
Fentons reagent. In this work Fentons reagent is generated by electrochemistry in a catalytic way. The reaction of the hydroxyl
radicals with pentachlorophenol (PCP) was studied. These radicals generated in situ in aqueous solution react with PCP and thus
lead to its degradation. The evolution of the composition of the solution was followed by chromatographic analysis, COD analysis
and the measurement of the total organic carbon (TOC) of the studied aqueous solution. Tetrachloro-o-benzoquinone and
tetrachloro-p-benzoquinone (TCBQ) are the only aromatic intermediates identified. They result from the oxidation of the
corresponding tetrachlorohydroquinones (TCHQ). Just like PCP, the TCBQs are degraded and disappear in their turn. The
mineralization of the initial toxic substrate is confirmed on the one hand, by the regular decrease in quantity of the total organic
carbon of the solution (TOC analysis) and on the other hand, by the quantitative release of chloride ions according to the
electrical charge passed during electrolysis. The degradation of PCP appears relatively slow compared to that of the other organic
pollutants studied by the electro-Fenton process and also to the degradation of other less substituted chlorophenols. The kinetic
rate of the appearance of chloride ions is slower than that in the degradation of PCP. This phenomenon highlights the formation
of chlorated aliphatic intermediates. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Fentons reagent; Hydroxyl radicals; Pentachlorophenol; Ion chromatography; Total organic carbon; Chemical oxygen demand

1. Introduction
Pentachlorophenol (PCP) is a molecule which has a
broad spectrum of pesticidal efficacy.
Since its introduction around 1930, PCP has been
produced and used in large quantities. Its main uses
before being restricted were as a wood preservative but

* Corresponding author. Tel.: + 33-1-60957230; fax: + 33-160957392.


E-mail address: oturan@univ-mlv.fr (M.A. Oturan).

also as a pesticide (herbicide, fungicide, molluscicide,


and insecticide), disinfectant and bactericide [1]. As its
biological degradation is very slow, this intense use
caused pollution of soils and groundwater [2]. It is thus
not surprising that today we find PCP in drinking water
resources. In May 1997, the US Environmental Protection Agency (EPA) noted 59 sites polluted by PCP in
the USA [3]. Several site contaminated were detected in
Northern Europe (England, Denmark Sweden, etc).
Thus, such a situation requires the development of
economic and ecological techniques to eliminate PCP
and PCP-like compounds in contaminated water.

0022-0728/01/$ - see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 0 2 2 - 0 7 2 8 ( 0 1 ) 0 0 3 6 9 - 2

M.A. Oturan et al. / Journal of Electroanalytical Chemistry 507 (2001) 96102

PCP is the most toxic representative of the


chlorophenols. It is classified in the priority list of the
organic micropollutants [4,5]. In 1998 the EPA announced additional restrictions in the use of PCP. It is
toxic for plants, animals and humans. Even in very
weak concentration (0.1 mg l 1) it can lead to the
deterioration of the ecosystem [6].
Advanced oxidation processes (AOPs) based on the
chemistry of hydroxyl radicals (OH) are currently used
for the destruction of organic pollutants [7]. These
radicals react in a non-selective way on organic compounds leading finally to the mineral end-products. The
degradation of PCP has been the subject of several
studies in this context: Fentons reagent [8,9], direct
photochemistry or the photolysis of hydrogen peroxide
[10], photochemistry assisted by Fentons Reagent [11],
photocatalysis [12 14] using TiO2 as the heterogeneous
catalyst, sonochemistry [15]. Direct photochemistry is a
very slow process. Assisted photochemistry (photolysis
of hydrogen peroxide and the photo-Fenton process)
requires the addition of chemical reagents in wastewater to be treated. Photocatalysis is an efficient process,
but the use of powder TiO2, imposes the need of
post-processing for elimination of these particles.
Recently, a new advanced oxidation process induced
by electrochemistry was proposed as an alternative
process [1620]. It is based on hydroxyl radical production taking place from Fentons reagent [21], itself
generated in situ by electrochemistry. By analogy with
the photo-Fenton process, this process is called electroFenton. We have already applied this technique successfully to the total destruction of organic pollutants
such as chlorophenoxyacetic acid herbicides (2,4-D,
2,4-DP, CPMP, 2,4,5-T) [16,18] and nitrophenols [17].
The electro-Fenton process can be regarded as equivalent to the photo-Fenton process with the difference
that Fentons reagent is produced in situ in a catalytic
way and controlled by electrochemistry. In this work
we present a study of the degradation of PCP by this
new process. HPLC and ion chromatographic analysis
and TOC measurements confirm the mineralization of
PCP during the electro-Fenton treatment of aqueous
solutions containing PCP at saturated concentration.

2. Experimental

2.1. Materials
Pentachlorophenol (PCP), tetrachlorohydroquinone
(TCHQ), tetrachlorobenzoquinone (TCBQ), chemicals
used as standards for chromatographic analyses and all
other products were commercially available (Across,
Aldrich, Prolabo, Sigma, Fluka). They were of the
highest purity and were used as received without further purification. Mohrs salt, (NH4)Fe(SO4)26H2O,

97

was used to obtain ferrous ion. Carbon felt (working


electrode) was obtained from Carbone-Lorraine, molecular oxygen from Carboxique. Solutions and HPLC
eluents were prepared with deionized water (18.2 mV)
obtained from a Millipore Milli-Q system. The pH of
initial solutions was set at 3 by addition of aqueous
H2SO4 (Prolabo-Titrinorm grade).

2.2. Procedures and equipment


Electrolyses were carried out in a three-electrode
electrochemical cell controlled by a potentiostatgalvanostat EG&G (Model 273A). The working electrode
was a 10 cm2 carbon felt piece (CarboneLorraine), the
counter electrode was a 1 cm2 platinum sheet (Radiometer) and the reference electrode was a saturated
calomel electrode (SCE) from Radiometer. Aqueous
solutions of PCP were prepared at 0.03 mM concentration. This value constitutes its limit of solubility in
water. Prior to the electrolyses, dioxygen was bubbled
for 5 min to saturate the aqueous solutions which were
agitated continuously by a magnetic stirrer. A catalytic
quantity of ferrous ion was introduced into the solution
before the beginning of electrolysis. The potential of the
working electrode was fixed at 0.5 V (SCE) in the
case of controlled potential electrolyses. The current
controlled electrolyses were carried out in an undivided
cell by application of a current of 50 mA. The current
and the amount of charge passed through the solution
were measured and displayed continuously during
electrolyses.
Samples were withdrawn from the electrolysis solution at regular interval of coulometric charge (0, 25, 50,
100, 200 C, etc.). They were analyzed by reverse-phase
chromatography using a Gilson HPLC system (Gilson
pumps 305 and 306+ Dynamic mixer 811C+Manometric module 805) equipped with a UVvis detector
model 119 and a reverse-phase Hypersil C-18 column
(250 mm 4.6 mm I.D., 5 mm). The mobile phase was
a mixture of water+methanol (25:75 v/v) containing
1% of acetic acid. The flow rate of the mobile phase
was 1.0 ml min 1. Detection was carried out at 280
nm. Calibration curves were prepared for the quantitative analysis of PCP and the intermediate products
formed during electrolysis. The identification of the
intermediates was conducted by retention time comparison and internal standard addition methods using
standard solutions.
The concentration of chloride ions released during
electrolysis was measured by ion chromatography
(Dionex 100 equipped with a conductivity detector), on
an anionic exchanger column (IonPack AS14-Dionex).
The volume of injection was 10 ml and the mobile phase
was a mixture of 3.5 mM sodium carbonate and 10 mM
sodium hydrogencarbonate solution with a flow rate of
1.2 ml min 1.

M.A. Oturan et al. / Journal of Electroanalytical Chemistry 507 (2001) 96102

98

The chemical oxygen demand (COD) was determined


by the AFNOR NFT-90-101 method. Oxidation of PCP
was carried out by potassium dichromate in concentrated acidic medium, in the presence of silver sulfate
(catalyst) and mercury sulfate (for chelating chloride
ions). The dichromate excess was measured by a normalized solution of ammonium and ferrous sulfate
(Mohrs salt) in the presence of ferroin.
The total organic carbon (TOC) of the initial and
electrolysed samples was determined with a Shimadzu
5000 TOC analyzer equipped with an autosampler Shimadzu (ASI-5000A Autosampler). The samples were
acidified by sulfuric acid addition in order to obtain
pH B4 (just below) and to give no free carbonate
anions in solution. The catalyst used to carry out the
combustion reaction was platinum. The carrier gas was
oxygen with a flow rate of 150 ml min 1. The detection
was carried out by using an infra-red detector NDIR.
Calibration of the analyzer was achieved with potassium
hydrogen phthalate standards.

3. Results and discussion

3.1. Electrochemically monitored Fentons reaction:


hydroxyl radicals generation
The well-known Fentons reaction:
H2O2 +Fe2 + Fe3 + +OH + OH

(1)

constitutes a source of hydroxyl radicals production by


chemical means. This reaction can be entirely monitored
by electrochemistry. To do this it is sufficient to introduce a catalytic quantity of ferric ions into the initial
solution. The hydrogen peroxide and the ferrous ions
are simultaneously generated on the cathodic working
electrode at a potential of 0.5 V (SCE), according to
the following electrochemical reactions:

O2 + 2H+ + 2e H2O2
Fe

3+

+e

X Fe

2+

E=0.69 V (SHE)
E=0.77 V (SHE)

(2)
(3)

Fentons reaction (Eq. (1)) takes place then in homogeneous medium leading to the formation of hydroxyl
radicals.
The anodic reaction is the oxidation of water to
molecular oxygen (Eq. (4)) which is used for optimal
production of hydrogen peroxide (Eq. (2)) necessary for
Fentons reaction.
2H2O X O2 + 4H+ + 4e

(4)

Fig. 1 shows two catalytic cycles taking place during this


process. The chemical equation corresponding to the
electro-Fenton process can be obtained by the sum of
the reactions (1)(4):
electrical
(5)
1/2O2 + H2O 
2OH
energy

This equation summarizes the catalytic properties of the


electro-Fenton process: the ferrous ion does not appear
in the equation and 75% of oxygen used by the system
is generated by the anodic reaction. The production of
2 mol of hydroxyl radicals requires only the consumption of a half mole of dioxygen.
The hydroxyl radicals thus formed react on the PCP:
PCP + OH products
(6)
Generally hydroxyl radicals can react on aromatic compounds by three types of reactions: addition on a double
bond, abstraction of a hydrogen atom (if the molecule
has a side group Cn Hm X) or electron transfer. In the
case of PCP, all the positions of the aromatic cycle being
occupied, one can expect an ipso-attack on the chlorines position. The abstraction of a hydrogen atom being
excluded, the ipso-addition can be competed by the
electron transfer reaction. In a recent study [22] Fang et
al. highlighted this competition in the case of the reactions of hydroxyl radicals produced by pulse radiolysis
with the pentachlorophenolate ion (in basic medium):
53% of PCP reacted by electron transfer and 47% by
electrophilic ipso-addition.

3.2. En6ironmental approach: TOC and chloride ion


e6olution

Fig. 1. Schematic presentation of the electrocatalytic production of


hydroxyl radicals by the electro-Fenton process.

The TOC measurement and chloride ion analysis


during electrolysis of PCP aqueous solutions indicate its
mineralization progress. The TOC evolution for an
initially 0.03 mM PCP solution (V=125 cm3) as a
function of total charge passed during controlled potential electrolysis is illustrated in Fig. 2. The TOC of the
solution decreases regularly with a kinetic rate slightly
higher at the beginning of electrolysis. The TOC abatement is 30% at 200 C, 54% at 600 C and 82% at 1500
C.
Ion chromatography analyses highlight the release of
chloride ions during electrolysis. Fig. 3 shows that this

M.A. Oturan et al. / Journal of Electroanalytical Chemistry 507 (2001) 96102

99

3.3. PCP degradation analysis

Fig. 2. TOC decay during potential controlled electrolysis (Eappl =


0.50 V vs. SCE) for a 0.03 mM PCP aqueous solution (V=125 cm3,
pH 3).

Fig. 3. Chloride ions formation during potential controlled electrolysis for a 0.03 mM PCP solution (125 cm3) as a function of charge
passed.

Fig. 4. Evolution of PCP concentration during the electro-Fenton


process. Potential controlled electrolysis at 0.5 V vs. SCE. Volume
of solution =125 cm3. The 0.10 mM PCP solution is prepared in a
hydro-organic (water acetonitrile 95:5 v/v) mixture.

release is quantitative. The rate of formation which is


high at the beginning of electrolysis, decreases with the
disappearance of the PCP and its aromatic derivatives.
A comparison with Fig. 4 permits one to note that the
release of chlorides continue even after the total disappearance of aromatics (Q \200 C), which indicates that
the dechlorination continues with aliphatic compounds
formed by aromatic ring opening reactions.

The HPLC analysis of the PCP concentration during


electrolysis confirms that its degradation is accompanied by the appearance of quinone derivatives. Indeed,
the medium being strongly oxidizing, primary hydroxylated products, 1,2-tetrachlorocatechol and tetracholoro-p-hydroquinone are oxidized to tetrachloroo-benzoquinone
and
tetrachloro-p-benzoquinone
(Scheme 1). The fact that no accumulation of degradation products of PCP in the presence of the hydroxyl
radicals seems to indicate that the kinetic rate of reaction of these intermediates with hydroxyl radicals is
higher than that of PCP. The benzoquinones formed
disappear then in their turn by reaction with hydroxyl
radicals. An oxidative ring opening reaction at the level
of the CC bonds between adjacent hydroxyl and/or
ketones groups, which are destabilized by the strong
electrophilic effect of this group, leads to the formation
of (partially chlorinated) aliphatic compounds.
HPLC retention time comparison and internal standard addition methods using standard solutions of the
authentic products (commercially available) permitted
us to identify tetrachloro-o-benzoquinone (tR =4.5
min) and tetrachloro-p-benzoquinone (tR = 5.4 min) as
intermediates. These two compounds very quickly reach
the stationary state (formation rate from PCP equal to
consumption rate by hydroxyl radicals) with a concentration close to the detection limit for tetrachloro-obenzoquinone. Their corresponding chromatographic
peaks disappear in the same time order as that for PCP.
According to the ortho and para directing effect of
the OH group [2224], the TOC decrease of the
solution and the formation of chloride ions during the
electro-Fenton process make it possible to propose the
following mineralization reaction mechanism for PCP
as shown in Scheme 1.
The progressive TOC removal and the quantitative
release of chloride ions throughout electrolysis constitute obvious evidence of the mineralization of PCP.
The global mineralization reaction, in other words the
cold combustion, of the PCP by hydroxyl radicals, can
be written:
Cl5C6OH + 18OH 6CO2 + 5HCl +7H2O

(7)

which is equivalent to the electrochemical oxidation


reaction:
Cl5C6OH + 11H2O 6CO2 + 5HCl +18H+ + 18e
(8)

3.4. Kinetic of PCP degradation


Fig. 4 shows the evolution of the PCP concentration
for two different initial concentrations, as function of

100

M.A. Oturan et al. / Journal of Electroanalytical Chemistry 507 (2001) 96102

Scheme 1. Mineralization of PCP by hydroxyl radicals produced by the electro-Fenton process

the charge passed during electrolysis. The solubility of


the PCP being very weak in water, we prepared a
relatively concentrated solutions (0.10 mM) using a
hydro organic solvent (5% acetonitrile in water). PCP
has a retention time (tR) of 17.0 min (see Section 2 for
conditions). In both cases, the concentration of the
PCP decreases exponentially until total disappearance.
In order to determine the rate constant for hydroxylation of PCP, electrolyses were carried out in the
presence of equal concentrations of PCP and salicylic
acid which gave kPCP =0.16 kSA. The kinetic constant
for hydroxylation of the salicylic acid being well known
(kSA =2.21010 M 1 s 1) [25] we have calculated the
absolute rate constant kPCP =3.6 109 M 1 s 1 for
hydroxylation of PCP (reaction (6)).
Considering a stationary state of hydroxyl radical
concentration, we have also estimated the apparent rate
constant (kapp =kPCP [OH]) for reaction (6), which
becomes pseudo first-order (Fig. 5). We found 8.5
10 3 s 1 in aqueous medium and 2.8 10 3 s 1 in a
hydro-organic medium containing 5% of acetonitrile.
The kinetic rate of degradation is three times slower in
the presence of acetonitrile. In spite of a relatively weak
kinetic constant (2.2107 M 1 s 1) [25], the reaction
of hydroxyl radicals with acetonitrile becomes domi-

nant because its concentration is relatively very high


compared to that of PCP.
A quick examination of kinetic data reveals that the
degradation rate of the PCP as that of its mineralization are weaker than that of other organic pollutants
such as 2,4-D [16] or p-nitrophenol [17]. This result can
be explained by the absence of a free site for hydroxyl
radical attack on the aromatic cycle in the case of PCP,
because the addition of a hydroxyl radical on a position
already occupied by a substituent will be more difficult
compared to an unsubstituted site. Moreover, Atkinson

Fig. 5. Determination of the apparent rate constant for hydroxylation


reaction: PCP + OHproducts. Electrolysis at controlled current
(I =50 mA), [PCP]0 =0.03 mM, pH 3.

M.A. Oturan et al. / Journal of Electroanalytical Chemistry 507 (2001) 96102

[23] has already noted that in the case of PCB the


reactivity of the hydroxyl radicals fell proportionately
with the number of chloride substituents.

3.5. Comparison with other methods


The degradation of the PCP in order to detoxify
contaminated water has already been the subject of
several studies by various techniques. The biodegradation of PCP in aqueous medium appeared very slow
and incomplete [26]. Beattie et al. [27] showed that the
complete oxidation of the PCP, in alkaline medium,
was possible by the use of strong oxidants like permanganate, but this technique is neither ecologically, nor
economically sound. Electrochemical technology was
also used to eliminate the toxicity of the PCP. The
direct electrochemical reduction of chlorinated phenols
involved sequential dechlorination [2830]. This technique allows a decrease of the toxicity (toxicity being
proportional to the number of chloride carried by the
chlorophenol) of the solution but not its elimination.
For example, in the case of the electrochemical reduction of PCP [31], the products of the reaction are the
chlorophenol derivatives less substituted than PCP:
2,3,4,6-tetrachlorophenol, 2,4,6-trichlorophenol, 2,4,5trichlorophenol, etc. Even in the case of a total dechlorination, the solution would always contain phenols.
Direct electrochemical oxidation of chlorophenols [32
35] has been postulated to involve formation of a
chlorophenol radical cation that easily deprotonates.
Subsequent reactions include further oxidation to a
benzoquinone derivative and ring-opening to carboxylic
acids. The faradic yields are low and the process requires the use of specific electrodes because of the high
overpotential for water oxidation.

4. Conclusion
We showed by this work that the electro-Fenton
process, which leads to the total degradation of PCP
with a quantitative release of chloride ions in solution,
can constitute an interesting alternative for the decontamination of wastewater polluted by PCP. The effective mineralization of a molecule such as PCP
consolidates the non selective character of this process.
This result constitutes a new proof of the generalization
of this process with respect to depollution of wastewater. The electrocatalytic in situ generation of Fentons
reagent constitutes, from the ecological and economic
point of view, a significant advantage compared to the
photo-Fenton or chemical Fenton processes.
The PCP concentration in solution after treatment by
the electro-Fenton process is lower than 0.5 ppm which
is the detection limit of our HPLC analysis apparatus.
This value is higher than the limit set by the Environ-

101

mental Protection Agency for drinking water (0.22 mg


l 1) but it is close to the regulatory limits for surface
and ground water and far below the regulatory limit for
wastewater [36]. However this technique which is not
economically interesting for very dilute solutions is
eminently suitable for wastewater treatment.
Consequently, the approach used in this work seems
to be able to offer an appropriate solution to wastewater treatment and especially for industrial and agricultural effluents contaminated by organic micropolluants.

Acknowledgements
The authors thank Mr Olivier Bulteau (EDF, Division Recherche et Developpement) for running the
TOC analyses.

References
[1] M.C. Kiefer, S. Hengraprom, S. Knuteson, Environmental Organic Chemistry, Springer, Berlin, 1998.
[2] G. Engelhardt, P.R. Wallno fer, W. Mu cke, G. Renner, Toxicol.
Environ. Chem. 11 (1986) 233.
[3] M.A. Engwall, J.J. Pignatello, D. Grasso, Water Res. 33 (1999)
1151.
[4] US Environmental Protection Agency, Health and Environmental Effects Profile No. 135, Washington DC, 1980.
[5] CE Directive 98/83/CE of 3/12/1998, JOCE L 330, December 5,
1998.
[6] E.D. Packham, C.L. Duxbury, C.I. Mayfeld, V.E. Innis, J.
Kruuv, J.E. Thopson, Bull. Environ. Contam. Toxicol. 29 (1982)
739.
[7] J.R. Bolton, S.R. Cater, in: Aquatic and Surface Photochemistry, G.R. Heltz, R.G. Zepp, D.G. Crosby (Eds.), Levis Publishers, Boca Raton, FL, 1994, pp. 467 490.
[8] M. Barbeni, C. Minero, E. Pelizetti, E. Borgarello, N. Serpone,
Chemosphere 16 (1987) 2225.
[9] O. Kyama, Y. Kamagata, K. Nakamura, Water Res. 28 (1994)
865.
[10] T.-E.L. Ho, J.R. Bolton, Water Res. 32 (1998) 489.
[11] M.A. Engwall, J.J. Pignatello, D. Grasso, Water Res. 33 (1999)
1151.
[12] N. Serpone, R. Terziyan, D. Lawless, A.M. Pelletier, C. Minero,
E. Pelizetti, in: G.R. Heltz, R.G. Zepp, D.G. Crosby (Eds.),
Aquatic and Surface Photochemistry, Levis Publishers, Boca
Raton, FL, 1994, pp. 387 398.
[13] M. Barbeni, E. Pramauro, E. Pelizetti, E. Borgarello, N. Serpone, Chemosphere 14 (1985) 195.
[14] G. Mills, M.R. Hoffmann, Environ. Sci. Technol. 27 (1993)
1681.
[15] C. Petrier, Y. Jiang, M.-F. Lamy, Environ. Sci. Technol. 32
(1998) 1316.
[16] M.A. Oturan, J. Appl. Electrochem. 30 (2000) 477.
[17] M.A. Oturan, J. Peiroten, P. Chartrin, A.J. Aurel, Environ. Sci.
Technol. 34 (2001) 3474.
[18] M.A. Oturan, J.J. Aaron, N. Oturan, J. Pinson, Pestic. Sci. 55
(1999) 558.
[19] E. Brillas, E. Mur, R. Sauleda, L. Sa`nchez, J. Peral, X.
Dome`nech, J. Casado, Appl. Catal. B: Environ. 16 (1998) 31.
[20] A.A. Gallegos, D. Pletcher, Electrochim. Acta 44 (1999) 2483.
[21] H.J.F. Fenton, J. Chem. Soc. 65 (1894) 8234.

102

M.A. Oturan et al. / Journal of Electroanalytical Chemistry 507 (2001) 96102

[22] X. Fang, H.-P. Schuchmann, C.V. Sonntag, J. Chem. Soc.


Perkin Trans. 2 (2000) 1391.
[23] R. Atkinson, Environ. Sci. Technol. 27 (1987) 305.
[24] X. Chen, R.H. Schuler, J. Phys. Chem. 97 (1993) 421.
[25] N.V. Raghavan, S. Steenken, J. Am. Chem. Soc. 102 (1980)
3495.
[26] J.G. Mueller, D.P. Middaugh, S.E. Lantz, P.J. Chapman, Appl.
Environ. Microbiol. 57 (1991) 1227.
[27] J.K. Beattie, J.A. De Martin, B.J. Kennedy, Aust. J. Chem. 47
(1994) 1859.
[28] I.F. Cheng, Q. Fernando, N. Korte, Environ. Sci. Technol. 31
(1997) 1074.
[29] J.D. Rodgers, W. Jedral, N.J. Bunce, Environ. Sci. Technol. 33
(1999) 1453.

[30] A.I. Tsyganok, K. Otsuka, Appl. Catal. B: Environ. 22 (1999)


15.
[31] C.-H. Lin, S.-K. Tseng, Chemosphere 39 (1999) 2375.
[32] E. Brillas, R. Sauleda, J. Casado, J. Electrochem. Soc. 145 (1998)
759.
[33] A.M. Polcaro, S. Palmas, F. Renoldi, M. Mascia, J. Appl.
Electrochem. 29 (1999) 147.
[34] Ch. Comninellis, C. Pulgarin, J. Appl. Electrochem. 21 (1991)
108.
[35] M. Gattrell, D.W. Kirk, J. Electrochem. Soc. 140 (1993) 1534.
[36] German Federal Ministry for Economic Cooperation and Development, Deutsche Gesellschaft fu r Technische Zusammenarbeit
(GZT) GmbH (Ed.), Environmental Handbook, vol. 3, Compendum of Environmental Standards, Eschborn, 1995.

You might also like