You are on page 1of 17

WASTE MANAGEMENT, Vol. 13, pp.

361-377, 1993
Printed in the U.S.A. All rights reserved.

0956-053X/93 $6.00 + .00


Copyright 1993 Pergamon Press Ltd.

ORIGINAL CONTRIBUTION

A D V A N C E D C H E M I C A L O X I D A T I O N : ITS P R E S E N T
R O L E A N D P O T E N T I A L F U T U R E IN H A Z A R D O U S
WASTE TREATMENT
C. P. Huang* Chengdi Dong, and Zhonghung Tang
Department of Civil Engineering, Universityof Delaware, Newark, DE 19716

ABSTRACT.Chemical

oxidation reactions involving hydroxyl radicals have been extremely effective in the destruction of organic
pollutants. These advanced chemical oxidation processes (AOP) generally use a combination of oxidation agents (such as H202 or
O3), irradiation (such as uv or ultrasound), and catalysts (such as metal ions or photocatalysts) as a means to generate hydroxyl radical. The hydroxyl radical is one of the strongest inorganic oxidants next to elemental fluorine. The hydroxyl radical is stable over a
wide pH range, up to pH 10. The hydroxyl radical reacts with organic by three major mechanisms: hydroxy addition, hydrogen abstraction, and electron transfer. Several AOP systems are reviewed first. The merits as well as limitations of these systems are discussed. The potential of AOP for future hazardous wastes treatment is then demonstrated by four AOP systems, H202/Fe 2+,
TiO2/uv/O2, H2Oz/uv/O2, and TiO2/uv/H202, exemplified by chlorophenols. A reaction scheme can be generalized for the oxidation of halogenated phenols by advanced chemical oxidation, specifically, ones involving hydroxyl radicals. Upon the attack of a
halogenated phenol, ArXnOH, by a hydroxyl radical, OH., a free radical, Ar(OH)2Xn., is formed. This free radical can undergo two
reaction paths: (l)hydroxylation without dechlorination (Type A) and (2) hydroxylation with deehlorination (Type B). It has been
observed that mono-halogenated phenols (n=l) only follow Type A path; dichlorophenols (n = 2 ) and trichlorophenols (n = 3 ) can
have both Type A and Type B reaction pathways; tetrahalogenated (n = 4) and pentahalogenated (n = 5) phenols only follow Type B
reaction pathway.

for the treatment o f complex wastes with a good degree


o f success. However, the presence o f toxic chemicals in
the waste streams can cause unwanted operational difficulties. Much has been reported on the difficulties in the
operation of biological oxidation systems in treating municipal wastewater let alone hazardous wastes. These
difficulties are intrinsic to the process in that it is sensitive to ambient conditions as well as waste characteristics. Moreover, the generally long retention time and
start-up time of the biological oxidation process make it
unattractive for the treatment of toxic and refractory organic pollutants.
Thermal destruction of specific chemicals at extreme
temperature, i.e. incineration, has been successfully applied to the treatment o f solid waste, sludge, and gases.
While simple in principle, thermal systems have suffered
from many difficulties in operations. Additionally, thermal processes are rather limited in treating aqueous
wastes as considerable energy is required to heat and vaporize the water mass before specific organic pollutants
are destroyed at relatively elevated temperatures (usually
greater than 850C).
To overcome this difficulty, a thermal catalytic oxidation process, wet-air oxidation, has been developed. In

INTRODUCTION
Oxidative destruction o f chemicals provides ultimate solutions for the treatment o f hazardous wastes. Oxidative
destruction o f organic pollutants can be accomplished by
various means: biological, chemical and physical (thermal). While there are differences in the mode of operation, the basic principle o f energy dissipation process,
intrinsic to oxidative destruction o f organic matters, is
largely the same among these systems.
Traditionally, attention received by biological oxidation processes far exceeds that o f the chemical and physical oxidation systems, due in part to its cost effectiveness
and versatility in handling a wide variety o f organic pollutants. Since its introduction in the 1900s, several biological treatment systems have evolved and been adopted
*To whom correspondence may be addressed.
is supported by a grant (No.
R815081) from the Exploratory Research Program, US
Environmental Protection Agency. Contents of this publication
do not necessarily reflect the views or policy of the funding
agency. Any conclusions or mention of chemicals and
processes are made by the authors only and should not be
implied as their endorsement by the funding agency.

Acknowledgements-This work

361

362

wet-air oxidation, organic compounds are decomposed


with atmosphere air or pure oxygen at 150 to 370C and
10 to 220 bar pressure. Organic substrates are first carbonized(l). The dissolved oxygen reacts catalytically on
the surface of the carbon center so as to yield hydrogen
peroxide which then decomposes to form oxygen and hydroxyl radicals. These radicals then react with the carbon
to yield carbon dioxide (autocatalytic oxidation). Substrates such as acetic acid, which does not carbonize below 300C, will be degraded by wet-air oxidation.
Above 300C, acetic acid is reported to be oxidized very
slowly. Substrates such as sugars, which are easy to carbonize, will be completely degraded.
Several bench scale studies have been conducted to
determine the susceptibility of specific compounds to
wet-air oxidation. The following compounds can be destroyed by wet-air oxidation (2,3): aliphatic compounds
including those with multiple halogen atoms can be oxidized; residual oxygenated compounds of low molecule
weight such as alcohols, aldehydes, ketones, and carboxylic acids might be formed; aromatic hydrocarbons such
as toluene and pyrene are easily oxidized; halogenated
aromatics with at least one nonhalogen functional groups
of electron donating nature present on the ring can be destroyed. These substitutes include hydroxyl, amino, or
methyl group. Halogenated aromatics such as 1,2-dibenzene and PCBs are resistant to degradation unless catalysts are employed.
A specific catalyst system based on acidic solutions of
bromide, nitrate, and manganese ions has been developed
(4). Katox process is a catalytic oxidation technology
employing special contact media. The organics are
bathed in oxygen-containing water and oxidized at the
surface (in particular in the internal surface) of the catalyst. Copper ions have a significant catalytic effect.
Mn(V) catalysts and activated chromium compounds
have also been investigated (5). The major problems associated with these processes are the separation of the
catalysts.
The wet-air oxidation process has greatly improved
the efficiency of the incineration process as well as the
ease of operation. However, due to the severity of the reaction, construction materials for the system become the
major cost factor. Due mainly to its high cost, wet-air
oxidation process remains an operation of the preliminary treatment nature without total destruction of the organic pollutants.
Moreover, the wet-air oxidation process is limited by
the solubility of oxygen in water. This gas (02) to liquid
(H20) mass limitation can be eliminated at supercritical
water temperatures. In the supercritical region, water exhibits a density of 0.1 to 0.5 g/cm 3 and a dielectric constant of 3 to 10. Most importantly, it has almost 100%
solubility of organics, because thermal energy contained
in the water molecule reduces hydrogen bonding. Supercritical water becomes an excellent solvent for nonpolar

C.P. HUANG, C. DONG, AND Z. TANG


organics. The surface tension of the water is zero, which
allows oxygen to penetrate even the smallest pores and
oxidize any organic materials present. When temperatures reach 500C, near total insolubility of inorganics
take place. The major advantages of supercritical water
oxidation (SCWO) are: (1) enhanced solubility of oxygen and air in water and elimination of the mass transfer
between two phases; (2) rapid oxidation of organics in
short residence time; (3) complete oxidation of organics;
(4) potential removal of inorganic constituents as solids
or brine; (5) recovery of the heat of combustion as high
temperature process heat or power. Thus, this process
has been shown as a viable technology for the thermal
oxidation of hazardous wastes (6). It has been demonstrated that SCWO process is promising in oxidizing
concentrated hazardous wastes. However, like the wet-air
process, the severity of the system makes it unattractive
for the treatment of dilute wastes. In his recent study,
Thornton reported that high molecular weight products
such as dibenzofuran and dibenzo-p-dioxin may be produced in the SCWO treatment of phenol and that chlorinated dibenzofuran and dibenzo-p-dioxin may be
produced (7).
In light of the increasing concern over the contamination of the environment by hazardous chemicals, there is
great need to develop innovative technologies for the
safe destruction of toxic pollutants. The processes must
be cost effective, easy to operate, and capable of achieving a total or near-total mineralization.
This has
prompted researchers to investigate innovative chemical
oxidation technologies. This report gives an overview of
recent developments in advanced oxidation processes,
with an emphasis on the chemical systems. Four AOP
systems are illustrated using chlorophenols.

ADVANCED C H E M I C A L OXIDATION SYSTEMS


The oxidation potential of an oxidant is related to its oxidation-reduction potential, E. An oxidant with a high E
value is a strong oxidation agent. Table 1 lists the E
value for a host of oxidation agents against that of molecular 02 (8). All chemical species shown in Table 1 are
stronger oxidation agents than 02, although it must be
mentioned that kinetic factors often outweigh thermodynamic properties in controlling the oxidation reactions.
Nevertheless, a stronger oxidant will generally exhibit a
faster oxidation reaction. Based on this simple premise,
F2 is the strongest among the list. However, F2, a halogen, may produce halogenated compounds during the
oxidation process. Among these oxidants, 03 has received the greatest attention. However, the mechanism
of ozonation was not fully understood until the late seventies, when Hoigne and coworkers (9,10) first reported
the mechanism of ozonation, although over one hundred
public water treatment facilities were built in Europe dur-

ADVANCED CHEMICAL OXIDATION

363

TABLE 1

Oxidation-Reduction Potentials of Oxidation Agents


Redox Reaction

(8)

Eo (NHE)
volt, 25C

F2 + 2e = 2F-

2.87

OH + H+ + e = H20

2.33

03 + 2H + 2e = 0 2 + H 2 0

2.07

H202 + 2H + 2e = H302+

1.76

MnO4- + 4 H + + 3e = MnO2 + 2 H 2 0

1.68

HCIO2 + 3 H + 4e = C l + 2 H 2 0

1.57

MnO4- + 8H+ + 5e = Mn 2+ + 4 H 2 0

1.49

HOCI + H + 2e = C I + H 2 0

1.49

C12 + 2e = 2CI-

1.36

H B r O + H + 2e = B r + H 2 0

1.33

03 + H 2 0 + 2e = 0 2 + 2 O H

1.24

molecular 03. The rate constants are usually on the order


of 10s-10 l (Mlsec z) (13). Chemically, hydroxyl radicals are much less selective than molecular 03.
The concept of"advanced oxidation processes" (AOP)
was established by Glaze et al. (14,15). AOP was defined as the oxidation processes which generate hydroxyl
radicals in sufficient quantity to affect water treatment
(14,15). Many systems are qualified under this broad
definition of AOP. Most of these systems use a combination of strong oxidants, e.g. 03 and H202, catalysts, e.g.
transition metal ions or photocatalyst, and irradiation,
e.g. ultraviolet (uv), ultrasound (us), or electron beam
(eb). Table 2 lists typica! AOP systems currently been
reported in the literature.

CIO2(g) + e = CIO2-

1.15

Br2 + 2e = 2 B r

1.07

HIO + H + 2e = I + H 2 0

0.99

Hydroxyl Radical

CIO2(aq) + e = CIO2-

0.95

C I O + 2 H 2 0 + 2e = C I + 2OH-

0.90

The hydroxyl radical is one of the most reactive free radicals and one of the strongest oxidants (16):

H202 + 2H + + 2e = 2 H 2 0

0.87

CIO 2 + 2 H 2 0 + 4e = C I + 4OH-

0.78

BrO" + H 2 0 + 2e = Br" + 4 O H

0.70

12 + 2e = 2 1

0.54

13 + 3e = 3 1

0.53

I O + H 2 0 + 2e = I- + 2 O H

0.49

ing the early seventies (11). It was thought that molecular ozone was the major oxidation species. Hoigne et al.
(9,10) proposed that 03 can have two reaction modes: direct ozonation reactions and free radical (such as hydroxyl radical) decomposition reactions (9,10). The
direct 03 reaction involves molecular 03 which is highly
selective and produces a relatively slow reaction. The
typical rate constants are on the order of 1-103 (M-1 sec-1).
As a result, there is no complete mineralization and a
variety of intermediates could be formed (12). The hydroxyl radical formed during the ozone decomposition
reacts much more rapidly with organic compounds than

OH + H + + e" = H20; E = 2.33 v

[1]

Thus, in acidic conditions, the hydroxyl radical will have


higher oxidation potential. In basic solution, i.e. pH 11,
the hydroxyl radical and H202 react to give an oxide and
a peroxide ion (17). Irradiation of water generates OH
relatively independent of pH until pH 10. The reaction
of OH with organic compounds can be classified into
three mechanisms: hydroxy addition, hydrogen abstraction, and electron transfer. Organic compounds containing aromatic systems or carbon-carbon multiple bonds
undergo additional reaction with OH due to the rich nelectron cloud on the aromatic ring. For example:

OH. + C6H6 ~

.C6H6OH

[2]

TABLE 2

List of Typical AOP Systems


H o m o g e n e o u s System
with irradiation
O3/ultraviolet (uv)

HzOz/uv
electronbeam
ultrasound(us)

Hydrogen abstraction is the usual reaction with unsaturated organic compounds:

OH. + CH3COH3 -+ CH2COCH3 + H20

[3]

H202/us
uv/us
without irradiation

O3/[t202
O3/OH

Electron transfer is usually found in reactions between


hydroxyl radical and inorganic ions. For example:

HEO2/Fe2+ (Fenton's)
Heterogeneous Systems
with irradiation

Fe 2+ + OH. --~ OH- + Fe 3+

[4]

TiO2/O2/ultraviolet
TiO2/HEOE/ultraviolet
without irradiation
electro-Fenton

This is important to the Fenton's reagent reaction. Since

364

C.P. HUANG, C. DONG, AND Z. TANG

OH has high and indiscriminate reactivity, it can react


with almost all types of organics, such as ethylenic, lipid,
aromatic, and aliphatic, and inorganics, such as anions
and cations. The addition of OH to benzene is found to
be very fast. The process produces the cyclohexadienyl
radical with an unpaired electron. The dienyl radical disproportionates to a complex mixture of products including phenol, cyclohexadienes, hydrocyclohexadienes,
biphenyl, and hydroxylated biphenols r the ring is
opened, oxidation by OH can result in the final products
such as CO2, H20 through a serial reaction:

HOOC-C4H~COOH+OH.
+H2OCO2+H20

drogen peroxide. The deprotonated form of hydrogen


peroxide (HO2-) can react with ozone to produce ozonide
(O3-), then hydroxyl radical (OH) (27). These species
initiate the chain reaction. As a result, compounds normally refractory to ozonation alone are rapidly converted
to CO2 and water.
Additionally, poly-chlorinated
biphenyl (PCBs), which are very stable to ozone, are destroyed rapidly to less than 0.1 ppm by the combination
03 and uv method. This process was initially developed
by Houston Research Inc. and has been specified as Best
Practicable Control Technology Currently Available for
the treatment of PCBs by the U. S. EPA.

HOOC-COOH
H202/uv system. UV photolysis of H202 generates OH':
[Sl
H202 + hv = 2'OH

The above reaction involves both hydroxy addition and


hydrogen abstraction for the mineralization of benzene.
The hydroxyl radical can be efficiently generated
from H202 and catalysts such as transition metal ions
without uv irradiation. One such system is Fenton's reagent, in which H202 is decomposed into a hydroxyl
radical and hydroxide ion in the presence of Fe 2+ ions.
Fenton's reagent was discovered by Fenton in 1894
(18,19). It was not accepted widely until 50 years later in
organic chemistry, where selectivity is emphasized, because of the unselective oxidation property of hydroxyl
radicals. The use of Fenton's reagent to oxidize toxic organics only began at the late sixties. The major advantages of Fenton's reagent as a hazardous waste treatment
technology are: (1) there are no chlorinated organic
compounds formed during the oxidation processes as in
chlorination or ozonation; (2) both iron and H202 are
cheap and nontoxic; (3) there is no mass transfer limitation due to its homogeneous catalytic nature; (4) there is
no light involved as catalyst so that the reactor design is
much easier than those uv light systems.
Due to these advantages, Fenton's reagent has been
widely applied in the treatment of hazardous organics.
The direct oxidation of phenolic wastes has been studied
by Eisenhauer (20). Bishop et al. (21) have investigated
the feasibility of Fenton's oxidation of refractory organics in municipal waste water. Recently, Fenton's oxidation has also been used as the pretreatment of
nonbiodegradable industrial wastes to render the organics
more biodegradable for both aerobic processes (22) and
anaerobic processes (23). Finally, Fenton's reagent has
been attempted to oxidize pentachlorophenol in a simplified soil system (24) and polychlorinated biphenyls
(PCBs) and chlorobenzene (25, 26).

Homogeneous Systems with Irradiation


OJuv System. Photolysis of aqueous ozone produces hy-

[6]

Sundstrom et al. (28,29,30) have studied the destruction of individual aliphatic and aromatic compounds by
uv catalyzed oxidation with hydrogen peroxide. The results demonstrated that the system can destroy a wide variety of hazardous compounds present in water at low
concentration levels. For chlorinated aliphatic compounds, the organic chlorine was converted to chloride
ion, indicating that the chlorinated structures were effectively destroyed. In the case of aromatic compounds,
many intermediates were formed which could be eliminated by extending the treatment time. It was found that
benzene and trichloroethylene have similar reaction rates,
with trichloroethylene (TCE) having a slightly faster rate
than benzene. The oxidation rate increases with increasing hydrogen peroxide and decreases with increasing organic concentration. The rate constant for hydroxyl
radical addition to benzene ring is 7.8109 M-lsec~. The
addition of hydroxyl radical to TCE is 4.O109 Mlsec ~
(31). Thus, hydroxyl radical addition to benzene is about
twice as fast as that to TCE. The TCE oxidation rate
should be expected to be reduced by half at the constant
organic carbon concentration. The decomposition of
H202 by uv has much lower efficient than that of ozone
due to the much lower extinction coefficient of H202 by
uv photolysis.
Processes such as HzO2/uv, O3/uv, and H202/O3/uv
have been shown to be effective for groundwater decontamination and soil remediation. Because the activation
energy between active radicals, such as hydroxyl radicals
and organics is close to zero, oxidation occurs rapidly.
However, these systems are not without disadvantages.
First, when the processes involve ozone, mass transfer
between gaseous ozone and aqueous solution becomes
the major limiting factor (31). Second, although H202
can eliminate the mass transfer limitation, the decomposition of H202 by uv has been proved inefficient. This is

ADVANCED CHEMICAL OXIDATION

365

because the extinction coefficient of H202 at 254 nm is


only 19.6 (M~cm ~) (33) compared with about 3,000 ( M
~sec~) for ozone (34). Third, the uv penetration depth is
still difficult to match the reactor size. If a precise match
could not be satisfied, the reaction rate could be reduced
due to the area where uv is not intensive enough to accelerate the decomposition of oxidants. At the same time, a
significant amount of uv could be absorbed by water
molecules where the light intensity is higher than necessary and the energy could be wasted in the form of heat.
Since the reaction rate is directly proportional to light intensity, turbidity and color will also significantly reduce
the efficiency of the systems.

Electron-beam (eb) system. Irradiation of water with


high-energy electrons results in the formation of the reactive free radicals e- (aq.), H', and "OH. These radicals are
formed through the excited state molecules such as
H20*, H20 +, and e'. When the excited molecules and
electrons interact and transfer their energy to other molecules, several secondary reactive species are formed.
The relative efficiencies (G) are defined as the number of
radicals, excited states, or other products formed or lost
in a system absorbing 100 eV of energy. According to
their production abundance, the following order exists:

'OH (2.7) = H30 + (2.7) > e" (2.6) > H202 (0.71)
> H' (0.55)

I7]

The e and H are the reducing species. O f these radicals,


e" and O H make up 90% of the reactive species and are
the major radicals to render organic destruction. H202 is
also an important species in oxidizing toxic organics, although it is not as abundant as the other two species. The
interaction between aqueous electrons and specific organics and inorganics has been extensively studied (35).
The e (aq.) is a powerful reducing agent with a potential
E of 2.77 v. as:

e" + H + = 1/2H2

H + 02 = HO2; k = 2.11010 ( M l s e c 1)

[9]

Ultrasound (us) system. Ultrasound (us) can affect organic oxidation through three mechanisms: nucleation,
growth, and cavitation. Cavitation which is created by
the collapse of a gas- or vapor-filled bubble in a body of
liquid is the most important phenomenon. The instantaneous pressure and temperature at the center of the collapsing bubble has been estimated to be about 75,000 psi
and 13,000F respectively (36). Due to this high local
pressure and temperature, it has been recognized that
cavitation can enhance the rate of chemical reactions significantly. Next, solvent will be compressed and refracted by the rapid movement of fluids caused by a
variation of sonic pressure. Then, a large amount of energy will be put into a small volume with little heating.
The phenomenon is called microstreaming.
US radiation is known to decompose water vapor
molecules in the bubbles into free radicals, such as hydroxyl radicals, hydrogen radicals, and hydroperoxyl
radicals (37,38). Prasad et al. studied the liberation of a
chloride ion in a saturated aqueous solution of chlorobenzene and m-dichlorobenzene by ultrasound (US) (40).
More than 90% dechlorination of chlorobenzene and mdichlorobenzene occurs in 60 and 100 minutes, respectively (40). It is believed that the hydroxyl radical
formed by the reaction between H and O2 is responsible
for the oxidation of chlorobenzene and m-dichlorobenzene in water. Although hydrogen peroxide can be produced by application of US alone to a diluted aqueous
solution, the amount may be too small to be significant.
Hydrogen peroxide can be added to this process as an initiator to increase free radical concentration in the solution.

H202/us system. Chen et al. (41) investigated the oxidation of chloroform by us/H202 process. The optimal molar ratios of H202 to CHCI3 are between 30 and 50.
When the ratio of H202 to CHCI3 is 50:1, the best percentage removal (94%) was achieved. The overall reaction can be presented as follows:

[8]
CHCI3 + 2'OH ~

The aqueous electrons can result in the dehaiogenation of


organohalogen compounds. Further reaction of the organic radicals leads to the complete destruction of the
compounds. Hydrogen radicals can undergo hydrogen
addition or hydrogen abstraction. These reactions are
considered as the minor ones due to the following extremely fast reaction between the hydrogen radical and
oxygen:

CO2 + 3Cl" + 3H +

[10]

The reaction is first order with a rate constant of 0.0177


(l/min). Since the concentration of hydroxyl radicals
generated is much higher than that of chloroform, the order of reaction with respect to the free radical is zero.
However, the optimal ratio of 50:1 for H202 to chloroform is quite high for practical application.
It appears that an excess amount of H202 would

C. P. HUANG, C. DONG, AND Z. TANG

366

slightly retard the decomposition of chloroform. That


excess H202 can retard the oxidation reaction has been
reported by Ho et al. in the study of photooxidation of
2,4-dinitrotoluene in the presence of H202 (42). This is
because H202 itself is also a hydroxyl radical scavenger.

US/uv system. Ultrasound (us) can be combined with uv


to destroy toxic organics. The chemical consequence of
photolysis and sonolysis of a simple haloalkane is the
cleavage of the carbon-halogen bond in liberating halogen ions (36). The principal products of water under
sonolysis are H202, H2, OH, and "H. When sonochemical decomposition of aqueous CC14 solution occurs, the
radicals such as "OH and .H react with CC14 and form
C12, CO2, HC1, C2C16, and HOC! (39, 43, 44).
GC/MS analysis indicates that olefins such as
CH2=CHCI and CHCI=CHCI were formed during the decomposition of CC13CH3. In an oxidative environment,
e.g. in the presence of H202, olefins are readily oxidized
to epoxides. Both olefins and epoxides are reactive
monomers which polymerize to polymers in the presence
of O H and CIO radicals. The pH of photosonolysis is
very acidic in the range of 1. However, less acidic solutions of pH 3 were found under sonolysis and photolysis
separately.
The order of CCI3CH3 disappearance and CI appearance is the following for oxidation of immiscible
CCI3CH3 in water: Photosonolysis ---> Photolysis --->
Sonolysis.
For a miscible sample of CCI3CH3 in water, the efficiency for different processes follows the different order:
Photosonolysis --->Sonolysis --->Photolysis.
The aqueous solutions exhibit lower vapor pressure
than the higher concentration of the immiscible pair. The
lower vapor pressure of the liquid causes higher cavitation bubble collapse intensity. Since the temperature and
the pressure can reach to thousands of Kelvins and Atmospheres, respectively, respectively, the increasing
cavitation intensity is considered as the most important
factor to enhance the dissociation of the covalent bonds
of the organics and their extent of decomposition(36,45).

Homogeneous Systems without Irradiation

OYH202 system. Buhler et al. (27), Form et al. (46), and


Sehested et al. (47) have reported that H202 can initiate
the decomposition of 03 by single electron transfer
wherein the initiating species is the hydroperoxide ion
HO2-:

H202 = HO2- + H+; Ka = 1.610 -12

[11]

The hydroperoxide ion reacts with ozone to produce the


ozonide ion 03- and hydroperoxide radical HO2.

HO2- + 03 ---> 03- + HO2;


k = 2.8'106 M'lsec -1

[121

These products can form OH radicals through the following initiation steps:

HO2 = H + + 02-; Ka = 1.6"10-5 M

02- + 03 ~

03- + 02; k = 1.6'109 M'lsec 1

[13]

[14]

O3- + H + ---> HO3; k = 5.2"1010 M'lsec -1

[15]

HO3 ~ .OH + 02; k = 1.1"105sec"1

[16]

Once the hydroxyl radical is formed, the following


propagation steps generate hydroxyl radicals by autocatalytic mechanism:

03 + 'HO ---> 02 + "HO2;

[17]

03- + "HO2 ~

[181

202 + H O

The chain mechanism generates hydroxyl radicals by


consuming H202 and 03. It is terminated by recombination of different radicals. Hydroxyl radicals abstract a
hydrogen atom from saturated hydrocarbon compounds
or add to unsaturated organics to form an organic carboncentered radical in less than 10 -6 second. The carboncentered radical reacts quickly with oxygen to yield a
peroxyl radical when oxygen is present. This radical can
decompose unimolecularly to produce a superoxide ion
(O2-, the deprotonated form of the HOz radical). Predominant hydrogen peroxide and minor superoxide can
also be formed if the carbon-centered radical decomposes
bimolecularly. The latter process is considered to take
place through a tetroxide structure. Superoxide reacts
quickly with ozone, yielding ozonide and continuing the
chain reaction. When ozone is present, this reaction is
much faster than superoxide disproportionate to hydrogen peroxide.

ADVANCED CHEMICAL OXIDATION

367

03/01-1- system. At neutral to high pH, ozone decomposes into hydroxyl radicals through the following initiation steps (9, 48):

03 + OH" --~ -O2- + 02

[19]

03 + H20 --~ 2"HO + 0 2

[2Ol

2Fe 2+ + H202 + 2H + = Fe 3+ + 2H20

This equation indicates that Fenton's reaction is strongly


dependent on solution pH. In fact, only in acidic conditions, is O H the predominant reactive oxidant.
In the presence of organic compounds, organic oxidation proceeds via addition of OH, or via hydrogen atom
abstraction. Eq. [24] is for O H addition:

-OH + R ~
The above initial steps result in the hydroxyl radicals
as the major oxidation species. This is the reason that alkaline oxidation rates by ozone are several orders of
magnitude higher than those in acidic media (49). Several investigators classified the ozonation at alkaline condition as advanced oxidation process (50). The catalytic
effect of a hydroxide ion has been recognized for a long
time in the ozonation of organic pollutants. Niegowski
reported pH of 11.8 as the optimal value (51). Jones indicated that phenol oxidation proceeds most efficiently at
a pH of 11.4(52).

H202/Fe2+ (Fenton's Reagen 0 system. The powerful


oxidizing properties of a mixture of H202 and Fe 2+ salts
was first observed by Fenton at the end of the last century (18,19). Forty years later, Haber and Weiss established the oxidizing species as hydroxyl radical (53). In
recent years, the hydroxyl radical has been observed directly by electron spin resonance spectroscopy (30).
From much literature, organic compounds of nearly all
types could be oxidized by this Fenton's reagent.
Fenton's reaction is given by Eq. [21] (54,55):

Fe 2+ + H202 = Fe 3+ + O H + O H

[21]

The products of the reaction are ferric ion, the hydroxyl


radical (OH), and the hydroxide ion. The rate is first order with rate constant ofkj = 76 (Mlsec "~) (54,55).
In the absenceof added substrate, the hydroxylradical will oxidize a
second moleculeof ferrous ion, as shown in Eq. [22]:

.ROH --+ hydroxylated products [24]

Eq. [25] is for hydrogen atom abstraction:

OH + R ~

'R + H20 ~

oxidized products

[25]

In both equations, the oxidation process is extremely


fast. The rate constants are as high as 107 - 10m (M-~sec1) (5).
Organic free radicals are formed as transient intermediates. The intermediate radicals are further oxidized by
Fe 3+, oxygen, H202, OH, or other intermediates or added
reagents to form a final, stable, oxidized product. The
following equations illustrate these processes:

Fe 3+ + R --~ Fe 2+ + products

[26]

2'R -+ R-R (dimerization)

[27]

Fe 2+ + R + H + --~ Fe 3+ + RH (reduction)

[28]

Fenton's oxidations can be classified into two groups:


(1) chain reactions, in which only a small amount of reducing agent is needed, and (2) non-chain reactions, in
which all the oxidation is effected by the hydroxyl radical and there is considerable loss of the hydroxyl radical
due to the following reaction:

Fe2++OH ~
Fe 2+ + O H = Fe 3+ + OH"

[231

(Fe-OH) 2+

[29]

[22]

The rate constant is k2 = 3108 ( M l s e c q) (54,55). The expected stoichiometry of the reaction is given in Eq. [23].
Protons have been added to show formation of water:

The decomposition rate of hydrogen peroxide reaches


the maximum at a pH of 3.5. This phenomenon is attributed to the progressive hydrolysis of the ferric ion, which
provides a relative large catalytically active surface for

368

C.P. HUANG, C. DONG, AND Z. TANG

contact with the H202. The accelerator in H202 decomposition will yield more free hydroxyl radicals.
If a small amount of F e E+ ion is needed, a chain reaction occurs through regeneration of Fe E+of Eq. [6]. If all
the oxidation is effected by hydroxyl radical formed by
reaction 1, non-chain reactions take place. Since there
are extra Fe 2+ present, considerable loss of the hydroxyl
radical proceeds via equation 2.
As the pH is raised, there is some evolution of oxygen
due to the reactions:

HO-OH = HO-O" + H +

[301

The reaction arises from the nature of H202, because


H202 is an extremely weak acid (ka = 1.55 x 10"12 at

20C). Its hydroxyl groups generally behave like those


of alcohols (7). In the presence of Fe 3*, 02 is evolved
through the following steps:

Fe 3+ + H O - O ~

HO-O ~

Fe 2+ + H O - O

H+ + ( O - O f

Fe 3+ + ( O - O f ~

H202 + (.O-O)- ~

[31]

[321

Fe 2+ + 02

[331

.HO + OH" + 02

[34]

Superoxide can be produced through the following reaction:

O H + H202 ---) H20 + H O - O

tion of electricity from solar irradiation. Upon irradiation, electrons are produced at the conduction band (cb)
and positive holes are formed in the valence band (vb)
(56). If the semiconductor is in an aqueous suspension,
the electrons, which are reducing agents, can migrate to
dark surfaces and react with oxidizing chemicals. The
positive holes, which are oxidizing agents, will remain at
the surface of the semiconductor. In the absence of oxidizable substances such organic compounds, the positive
holes will react with the semiconductor and corrode the
solar cells (57). A coating of organic paint on the semiconductor is commonly applied to protect the solar cells
(56,57). The organic compound will consume the positive holes while being oxidized. Another way to harvest
the solar energy is with the production of hydrogen and
oxygen gases from water by semiconductor material irradiated with sunlight or an artificial light source (58). In
contrast to electricity generation, the solar energy is converted into chemical energy, in the form of hydrogen.
Since the separation of hydrogen from the gas mixture is
difficult, researchers have tried to suppress the oxidation
of water by adding organic chemicals to water (58). The
following illustration shows that in the presence of organic chemicals, the positive holes will be used by the
organic compounds rather than water. In fact this concept
has been used as corrosion control of solar cells.

[351

Haber and Weiss (53) suggest that this reaction contributed to the maintenance of a chain reaction sequence
under the acidic conditions of their study. Under neutral
pH, the sum of Eqs. [21] and [30] gives Eq. [33]. In this
superoxide-driven Fenton's reaction, iron functions as
the true catalyst as the result of the recycling mechanism
where iron has been canceled after summation of Eq.
[10] and [1].

Heterogeneous Systems with Irradiation


Photocatalytic oxidation. Semiconductors such as CdS,
CdSe, and Si have been used as solar cells for the genera-

n2o

C02

Org~aies
Three kinds of semiconductors can be recognized: (1)
R-type, (2) O-type and (3) RO-type. The classification
is made on the bandgap position of the semiconductors
with respect to the water reduction (H+/H:) and the water
oxidation (OdH20) potentials. The O-type semiconductors, such as WO3, Fe203, and MoS2, have valance bands
located below the EMF of the water oxidation and can
oxidize water to oxygen. The R-type semiconductors,
such as CdTe, CdSe, and Si, have conduction band located above the EMF of the water reduction and can
readily reduce water to hydrogen. The RO-type semiconductors, such as TiO2, CdS and SrTiO3, a combination of the above two types, can split water into O: and
H2. It must be noted that the presence of organic compounds can greatly enhance the decomposition of water
by one to three orders of magnitude.

ADVANCED CHEMICAL OXIDATION


Studies have also shown that trace amounts of Pt and
RuO2 can drastically improve the production of hydrogen
in an organic compound-water mixture (58). Apparently
the Pt enhances the rate of water reduction while the
RuO2 catalyzes further the oxidation of the organic matters.
Kawai and Sakata (59,60,61) have demonstrated a
photocatalytic reaction with different media such as chlorine and nitrogen compounds. The decomposition of
some organic molecules and the production of hydrogen
gas occurred at the same time. Five percent platinum was
deposited photoelectrochemically on the surface of TiO2.
They concluded that the reaction can proceed at room
temperature in a solution without showing any thermal
effect. A commercial-grade titanium dioxide was pretreated by rinsing with 1M HC104 and distilled water
several times until the conductivity dropped to less than
10 !umho/cm. The titanium dioxide samples were centrifuged at 10,000 rpm for 30 minutes and then dried overnight at 105C and ground to a fine powder. In two
studies, Kawai and Sakada have some very interesting
findings (59,61). They have studied the photocatalytic
oxidation of natural products, e.g. glucose, ethanol, cellulose and lignin, food stuffs, e.g. potato, fatty oil, and
herbs, wood, e.g. cherry wood, white dutch clover and
water hyacinth, green algae, dead animals and excrement
using TiO2 under a xenon lamp. As shown in Table 3, in
neutral and in 5M NaOH solution, total oxidation was almost possible. They have also found that nitrogen and
chlorine are converted to NH3 and HC1, respectively, instead of other organic chemicals.
It is noticed that in contrast to what was reported by
Fujihira et al. (62,63), Kawai and Sakada (59,61) have
demonstrated a rather complete oxidation of some halogenated hydrocarbons with the production of HC1, HF,
CO2 and H2. This is interesting indeed. However, the
authors provide no details of the reaction kinetics. All
experiments were conducted in neutral solution or in 5M
NaOH media using TiO2 catalyst and a 500 watt xenon
lamp. Table 4 summarizes the end-products produced
during photocatalytic oxidation in uv-irradiated TiO2 suspensions by various researchers.
Barbeni et al. (64,65) have studied the degradation of
chlorinated hydrocarbons, 2,4,5-trichlorophenoxyacetate

Organic Matter
acetic acid
propionic acid
n-butylic acid
v-valeric acid
pivalic acid
tluene
benzene
acetopbenone
lactam

369
TABLE 3
Photocatalytic Dissociation of
Organic-Water Mixture by TiO2-Pt.

Organic Matter
glucose
ethanol
cellulose
pyruvic acid
gylcine
polyethylene
polyvinyl alcohol
polyvinyl chloride (PVC)
proline

End-Product
CH3CHO; C2HsOH; (CH3)2CO; H2
C2H6;CH3CHO; CH3COOH;H2
C2H6; C2H5OH; (CH3)2CO; H2
20H; H2
NH3; H2
CO2; H2
CO2; H2
CO2; HCI; H2
CO2; NH3; H2

stearic acid

CO2; H2

potato
fatty oil
cherry wood
white dutch clover
golden rod
water hyacinth
dead cockroach
human excrement
teflon
trichlorbenzene
trichloroethylene

CH3OH; (CH3)2CO; H2
C2H6; H2
C2H6; CH3OH; (CH3)zCO;H2
CH4; CH3OH; C2HsOH; NH3; H2
CH3OH; NH3; H2
NH3; H2
NH3; H2
NH3; CH3OH
HF; CO2; H2
HCI; CO2; H2
HCI; CO2; H2

and 2,4,5-trichlorophenol, using TiO2 as photocatalyst.


They have proposed a formation of OH free radicals as
the major mechanism for the oxidation of chlorinated organics. Ollis has examined the photo-oxidation of chlorinated hydrocarbons with TiO2 with promising results
(66,67). Ollis and associates have reported that the order
of ease of conversion is chloro-olefins ~ chloroparaffins
--~ chloroacetic acids (66,67). Matthews has used a thin
film of TiO2 to study the oxidation of some organic impurities in water. He has reported that the kinetics of
photo-oxidation reactions can be described by the Langmuir type adsorption equation (68,69).
In summary, there is strong evidence to suggest that
organic compounds such as those toxic chemicals found
in contaminated groundwater and wastewaters can be decomposed (oxidized) by photocatalytic oxidation. The
major products will be hydrogen and CO2. According to
the results of Kawai and Sakada, the overall reaction can
be expected as follows:

TABLE 4
Other Work on the Photocatalytic Dissociation of Organic Matters on TiO2
End-Product
Reference
CH4; CO2; C2H6; H2
Kruatler and bard (72)
ethane; ethylene; CO2; H2
propane; CO2; H2
n-butane; CO2; H2
isobutane; isobutylene; CO2; H2
benzaldehyde; biphenyl; cresols
Fujihira et al. (62, 63)
phenol; bephenyl
Kawai and Sakata (59-61)
hydroxyacetophenone; phenol
imides
Pavlik and Tantayanon (73)

C. P. HUANG, C. DONG, AND Z. TANG

370

CaHbOcNdCle + (2a - c)H20 ~

aCO2 + dNH3 + eriC1

+ 0.5(4a+b-2c-3d-e)H2

[39]

C7F15COO + H + = C7F15COOH

[40]

[361

For a complete photocatalytic oxidation reaction, CI will


become HCI and N will be converted to NH3.

Heterogeneous System without Irradiation


Electrolysis has a broad definition in that it includes all
reactions occurring at an electrode surface. The reactions
are forced to take place by an externally imposed voltage.
An important development in electrochemical oxidation
process is the generation of hydrogen peroxide by reduction of oxygen at the cathode. In the presence of ferrous
ions, the hydrogen peroxide so generated can form a hydroxyl radical--a well known Fenton' s reagent (18,19).
Hydrogen peroxide was generated at the graphite electrode in acidic conditions. In the presence of Fe 2+ ions,
the oxidation reaction can take place readily through the
hydroxyl radicals (54). An electron Fenton's reagent can
be produced by the addition of an appropriate amount of
Fe+2 ion to the catholyte. The oxygen reduction potential
is a function of pH and it is found that pH 3 is the most
favorable (at a cathodic potential of-0.6 V vs. saturated
calomel electrode). In the presence of organic compounds, the hydroxyl radicals formed will be rapidly consumed and the organic compounds of interest will be
oxidized (21,70,71).
In a study of electrogenerated Fenton's reagent oxidation of phenol, Sudoh et al. also concluded that pH is optimal at 3 (74-78). At pH 4, the degradation of phenol
hardly proceeded even with increasing electric current.
At pH 2, the rates of carbon dioxide production and COD
reduction of phenol were lower than those at pH 3. At
pH 1, the concentrations of the products (such as carbon
dioxide) were extremely low. COD decreased slowly in
comparison with the decrease of phenol concentration.
Since tarry precipitates were observed in the reaction solution, the oxidation of phenol might proceed to produce
the dimer or trimer of phenol.
In a recent study, Huang and coworkers (79) have
demonstrated that a surfactant, perfluoro-octanoate (C8),
can be decomposed effectively by indirect electrochemical oxidation(79). Results show that in 15 minutes, the
total C8 concentration was decreased from 100 ppm to
10 and from 50 ppm to 1 ppm. They have also proposed
a reaction pathway for the oxidation of C8 by the Fenton's reagent reaction:

02 + 2H + +2e --~ H202 (cathodic reduction) [37]

H202 + Fe 2+ --~ OH" + Fe 3+ + OH"

C7F15COONH4 = C7F15COO" + NH4+

[38]

C7F15COOH + OH' --~ C7F15. + CO2 + H20 [41]

C7F15.+ OH" --+ C7F15OH (CnF2n+IOH)

CnF2n+IOH + O H ~

[42]

CnF2n(OH)2 or

CnF2n+I(OH)

[43]

Reaction [37] is a reduction of oxygen at the cathode


electrode; reaction [38] is a generation of hydroxyl radical (Fenton's reagent); reaction [39] is a dissociation reaction of C8; reaction [40] is an acid-base equilibrium
reaction; reaction [41] is a Kolbe reaction; reaction [42]
is a free radical reaction with the formation of perfluoroheptanol; reaction [43] shows hydroxyl radical attacks on
the perfluoroheptanol to form simple perfluoro alcohols
such as C3F6(OH)2, C3F7(OH), C4Fs(OH)2,or C4F9(OH).
These alcohols are extremely volatile with a boiling point
around 0 C. Laboratory results clearly show the presence of simple perfluoroalcohol groups in the gas phase
of the reactor (79).

CASES OF ADVANCED C H E M I C A L
OXIDATION OF C H L O R O P H E N O L S
The oxidation of chlorophenols has been studied by
Huang and co-workers using Fenton's reagent and photocatalytic oxidation processes (80). Fig. 1 shows the results of the decomposition of the parent compounds,
mono-, di-, tri-, tetra- and penta-chlorophenois by the
Fenton's reagent. The results clearly indicate that the
degradation of chlorophenols is very fast and follows a
first order kinetic expression in terms of the parent compound concentration. Under the experimental conditions,
a total removal of chlorophenols is possible in less than 5
minutes, except perhaps for tetra- and penta-chlorophenols. The order in the destruction of these chlorophenols
is as follows:

ADVANCED CHEMICAL OXIDATION

371

1o0%

100%

-.<>-2-chlorophenol
--c3-3-dalorophenol
--<>-4-chlorophenol

80%
o

60%

k . . . , . . . . , . . . . ,

80%
~~1~

2o~

....

i....i

....

,....,

....

-o-2,34ichlomphenol
-o-2,44ich10rophenol
--o-23<lichlorophen01
"~-2'6 4ichlrphenl
--o-3,44ichloro~nol
3.54ichlorophenol

20%

0%

0%
1

Time (min)

Time (min)

--o-2j,4-trichl0mpl~01
8o~ ~
~\

-o-2.~,~-~d~o~h~ot

60% ~'~

"*-2,4,5-ttichJ~ophmol

---23,6-trich10t0pheml

2o%~icl~bmol

It~
80%~k

--o-phenol
i
--o-2,3,4,6-tetrachl0rophenol ~
.-o--pentachlorophenol
:

60%

"
@ 20%

o%
I

Time (rain)

0%

Time (min)

F I G U R E 1. Destruction o f Chlorophenols by Fenton's Reagent Oxidation. Experimental conditions: Concentration of organic was 10 "3 M except
2,3,4,6-tetrachlorophenol and pentachlorophen, for which concentration was 5x10 "5 M; Fe +2 = 10 -3 M; H202 = 6.7x10 4 M per minute; pH = 3.0;
reaction volume = 2 liter; ionic strength = 5x10 2 M NaNO3; room temperature.

4CP > 3CP > 2CP (for monochlorophenol)

35DCP > 23DCP > 25DCP > 34DCP > 24DCP -26 DCP (for dichlorophenols)

4CP ~_ 3CP = 2CP (for monochlorophenols)

23DCP > 35DCP - 25DCP > 34DCP > 24DCP >


26 DCP (for dichlorophenols)

235TCP ~ 345TCP > 236TCP > 234TCP


245TCP > 246DCP (for trichiorophenols)

345TCP -~ 235TCP > 234TCP ~ 245TCP


236TCP > 246DCP (for trichlorophenols)

Except tetrachlorophenol and pentachlorophenol, the


number o f chlorine substitutions appears to bear no effect
on the destruction o f the chlorophenol. This is contradictory to the conclusion made by Eisenhauer (20) that the
rate decreases with increasing the number of chlorine
substitution.
The decomposition o f these chlorophenols was also
studied under otherwise similar experimental conditions
except using a photocatalyst, TiO2 and in the presence of
atmospheric oxygen (Fig. 2). One major observation can
be noted: the photocatalytic oxidation o f chlorophenols
takes place at a rate that is one order of magnitude
smaller than the Fenton's reagent process. The order o f
oxidation follows:

This order is slightly different from that o f the Fenton's reagent process, except that in both processes,
26DCP and 246TCP are the most difficult ones for oxidation.
For the purpose of comparison, the oxidation of these
chlorophenols was also conducted with the H202/uv system (Fig. 3). It must be mentioned that the uv light
source used in this experiment was the same as that used
in the photocatalytic oxidation process reported above;
the wavelength (350 nm) o f this uv light source is longer
than other systems such as H202/uv and O J u v o f other
researchers. The rate o f oxidation is the same as that o f

C . P . H U A N G , C. D O N G , A N D Z. T A N G

372

100%

100%

-o-2~orophenol
--o-3-chl0r0phen01

80~

=
o
=

~.

60%

40%

4O%

2O%

2O%

~.
I~.~

0~
40

80

120

160

200

40

80

100%
"

'

"

-o-23,5-trich10rophenol
...-2,3,6-tfichlorophenol
....-a-2,4,5-tddaloropheno[
2,4,6adchlorophenol

60%

60%

40%

2O%

20%

80

120

200

-o-2,3,4,6.tetrachl0r0phenol.'
--o-PCP

e~ 80~

4O%

40

160

' -~enol ..... ' :

" -~-2,3,4~@[alo~phe~]

80%

120

Time (rain)

100%

0%
0

\~,-~ +~,4~!ch~o~o~i

Time (min)

_o

--o- 2,Ydichlorophenol:
"-o-2,4-dichlo~ophenol~
"o"2,5-dichlor~phenol
'.
~2,6-dichlorophenol.

80%

60%

0%
0

160

0%
0

200

40

Time (rain)

80

120

160

200

Time (rain)

FIGURE 2. Destruction of Chlorophenols by UV/TiO2 Oxidation. Experimental conditions: Concentration of organic was 103 M except
2,3,4,6-tetrachlorophenol and pentachlorophen, for which concentration was 5x10 -5 M; TiO2 = 1 g/L; pH = 4.0; ionic strength = 5x10 2 M NaNO3;
room temperature.
100% I . . .

, . ..

, . ..

,...

'00% ]k .....

,...

-o- 2-chlomphenol
-t~-34a10r0~nol
--e-4-ch]rl~nl

s0%
=g 60% ~

,, 80%[.['~

-o-2,4-dichlorophenol:

[~l k

60%tl ~ ,

0%

. . . . .

40

80

120

:=,...
160

0%
0

200

40

Time (min)

100%

~-~i~:

-o-23~.

80%

-,- 3,4 dichlo~ot~nol

80

120

160

200

160

200

Time (min)

100%

' "

--*- 2,6-dicMompheaol.

L/ ~ ' ~

--*-2~oro~

--~--2,4~,

80%

-o-Z3,4~

6O%
40%
20%

0%
40

80

120

Time (min)

160

200

0%
0

40

g0

1211

Time (min)

FIGURE 3. Destruction of Chlorophenols by UV/H202 Oxidation. Experimental conditions: Concentration of organic was 10-3 M except
2,3,4,6-tetrachlorophenol and pentachlorophenol, for which concentration was 5xl 0.5 M; H202 = 0.1 M per minute; pH = 4.0; reaction volume = 1

373

ADVANCED CHEMICAL OXIDATION


the TiOz/uv/O2 system. The order of the oxidation follows:

4CP > 3CP > 2CP (for monochlorophenols)

35DCP > 23DCP ~_ 34DCP > 25DCP > 24DCP >


26 DCP (for dichlorophenols)

345TCP > 235TCP > 234TCP = 236TCP >


245TCP > 246DCP (for trichlorophenols)

It is interesting to note that the most difficult chlorophenols for oxidation by this method are the same as the
above two methods, 26DCP and 246TCP.
The oxidation of these chlorophenols by a modified
photocatalytic oxidation process, TiO2/uv/H202, was also
conducted. Fig. 4 shows that the rate of chlorophenol
oxidation is extremely fast. Tseng has reported that the
rate of chlorophenol oxidation follows a first-order expression of the Langmuir-Hinshelwood type as adsorption onto the TiO2 surface controls the oxidation process
(82). Depending on the degree of chlorine substitution,
the rate of oxidation generally follows the order:

4CP > 3CP > 2CP (for monochlorophenols)

26DCP _=_ 25DCP > 34DCP _=_24DCP > 23DCP


(for dichlorophenols)

Generally, monochlorophenols are more readily oxidized


than dichlorophenols or trichlorophenols.

The mechanism of phenol oxidation by these AOP


systems is not totally clear, although there are several
models suggested. Recently, Dong and Huang (81) reported a reaction pathway for the oxidation of 4CP by the
TiO2/uv/O2 system. It is believed that OH attacks the ortho- and para-positions more readily than the meta-position. A series of reactions involving OH produces
chloro-dihydroxycyciohexadienyl radical which can then
react with OH to form 4-chlorocatechol, 4-chlororesorcinol, and hydroquinone as the major intermediates which
can react with OH to yield hydroxyhydroquinone. A series of reactions with OH eventually leads to the formation of CO2 and H20. It is expected that the reaction
pathways for the oxidation of chlorophenols of higher order than the monochlorophenols will be extremely complex. However, laboratory results all indicate that a total
mineralization of chlorophenols is possible by AOP
processes such as the Fenton's reagent.
Based on a recent study, Dong and Huang (80) have
proposed the following generalized reaction pathway,
shown below, for the oxidation of halogenated phenol.
Upon the attack of a halogenated phenol, (ArXnOH)
by OH, a free radical, Ar(OH)2X,. (chlorodihydroxycyclohexadienyl, CIDHCD) is formed. This free radical
can undergo two reaction paths: (1) hydroxylation without dechlorination (Type A) and (2) hydroxylation with
dechlorination (Type B). For pathway Type A, C1DHCD
dehydrated to form a free radical, ArOX~. (chlorophenoxy, C1PO) which is converted to Ar(OH)2X (chlorodihydroxybenzene, CIDHB). It is noted that one of the OH
group has a preference for the ortho- and the para-position of the aromatic ring. For pathway Type B, the free
radical CIDHCD is dechlorinated to yield Ar(OH)OX._,.
free radical (chlohydroxyphenoxy, CIHPO) which can
undergo further reaction with a hydroperoxide radical,
HO2, to form Ar(OH)2Xn., (chlorodihydroxybenzene
CIDHB). As indicated by the following reaction pathway, for mono-halogenated phenol ( n = 1), only Type A
path has been confirmed (80). Both Type A and Type B
reaction pathways are possible for dihaiogenated (n = 2)

O"

OH

=
OH

X~
(n --1

x.

OH

to 5)

OH

XR.HX

OH

TYPE A

(hydroxylation without
dechlorination)

OH(o,p) Xn

(Y

OH
H O 2.

-0 2
OH

Xn.1

TYPE B

(hydroxylation with
dechlorlnation)
OH(o,p) Xn-1

374

C.P. HUANG, C. DONG, AND Z. TANG


n = 3 (trlchlorophenols)

n = 0 (phenol)
OH
~

OH
@ O

OH

OH
H

OH
c1_5%~,

~
OH

OH
~CI

or
n -- I ( m o n o c h i o r o p h e n o l s )

OH

OH

CI

OH

cI

..
El

n = 4 (tetrachlorophenol$)

CI

OH

OH

C~[CI
n ~ 2 (dichiorophenols)
OH
OH
~CI
A O~CI
el

C~OH

T C l
Cl

or

OH
OH
o~.,.,,Cll~.)j~ or C ~ C '

~ CI
CI

T C I
CI

~I~ -El
OH

n = 5 (pentaehlorophenol)

CI
OH

OH
C~[CI

OH

~ H cI

or

Ci ~f
ci

Cl

and trihalogenated (n = 3) phenols. As for tetrahalogenated (n = 4) and pentahalogenated (n = 5) phenols, only


Type B reaction pathway is possible.
To further demonstrate the versatility of the Fenton's
reagent process, a sanitary landfill leachate was treated

Ci

i 8o%I

c]

c i ~f- c!
oH

with this system. Fig. 5 indicates that significant COD


removal can be accomplished by the H202/Fe ~+ system.
The COD of the leachate decreases from an original
8,500 ppm to 2,000 ppm after treatment with 510 .2 M
H202 and 2 10 3 M FeSO4 at pH 3. The results also show

--o-2-dal0r0phen01

--o3-chlorophenol
-o-4<hl0~phenol

80%

" 60%

OH
C~,~C|

or

CI ~ f
ci

100%

OH
C~.OH

60%

I .

40%

. i

..~-dichloro~nol.
I

-o-2,3-dichlor0phenol
--o-2,44ichlor01~erml
--o-2,54ich10r0~nol
-a--2,6-dich/orophenol
-4-3,44ich10~lSen01

20%
. . . .

40

~"..., . . .
80 120 160 200

0%
40

80 120 160
Time (min)

Time (rain)

100%lf.m . . , . . . . . . . . . . . . . . .

100%

--o-2,3,4-trichl0tophenol
--o-2,334fichlor0phenol
- ' ~ 2,3,6-tfrdll~phenol

80~ [~\
I.
Lil

I~

- -

. i

--o-phenol

8O%

-,s'-2,4,5-trichlotophend

40% [ ~

'

200

-~- 2,4,646ch10mphen01
2,4,646ch10mtSen01

6o%
4o%
20%

40

80 120 160
Time (rain)

200

0%
40

80 120 160
Time (rain)

200

FIGURE 4. Destruction of Chlorophenols by UV/TiO2/H202 Oxidation. Experimental conditions: Concentration of organic was 10.3 M except
2,3,4,6-tetrachlorophenol and pentachlorophen, for which concentration was 5x 10-5 M; TiO2 = 1 g/Liter; H202 = 0.1 M per minute; pH = 4.0; ionic
strength = 5x102 M NaNO3; room temperature.

ADVANCED CHEMICAL OXIDATION

375

the O3/uv system will be limited by the ozone transfer


process between the gas and the liquid phase. Photocatalytic oxidation, especially along the use of H202, becomes an extremely competitive process against the
H2Oz/Fe2+ system. However, the process is operationally
limited in that the development of a fixed-bed photocatalytic reactor is yet to be perfected commercially. The application of photocatalytic oxidation is most readily
applied in suspensions.

tO000

0000

6000

4000

2000

Raw Leachate

H202 treated Fe*2+H202 treated

REFERENCES
FIGURE 5. COD removal by H202 and Fe +2 + H202 with and without
precipitation. Experimental conditions: Initial COD = 8,500 ppm,
H202 = 5x10 "2 M; FeSO4 = 2x10 "3 M, pH = 3.0 (with H2SO4).

that the extent of mineralization is high when the


leachate is treated by the Fenton's reagent process (Fig. 6).

CONCLUSION

Advanced chemical oxidation processes, especially the


H202/Fe 2+ systems, are effective in decomposing organic
pollutants. Much has been learned about the mechanism
of the oxidation of organic compounds by AOP systems,
especially those involving OH reactions. While OH can
be generated by many systems, it appears that the homogeneous systems using H202 and catalytic metal ions,
specifically Fe 2+, will be the most attractive ones, both
from process control and process performance points of
view. The OJuv process also depends on OH to drive
the oxidation reaction. Upon irradiation with uv, 03 is
converted to H202, which is then converted to OH upon
uv irradiation. Therefore, the O3/uv system will not be
competitive against the HEO2/Fe 2+ process. Moreover,

10

o~
0

20

40

60

80

I00

120

Time (rain)

FIGURE 6. A comparison on CO 2 production by various chemical


treatment methods. Experimental conditions: Initial COD = 8,500
ppm, H/O2 = 5x10 "2 M; FeSO4 = 2x10 "3 M, pH = 3.0 (with H2SO4).

1. Lohamann, D. U., and Tilly, C. A. Naverbrennung. Chemie- lng.


Tech, 37:913 (1965).
2. Dietrich, M. J. Randall, T. L., and Canney, P. J. Wet air oxidation
of hazardous organics in wastewater. Environ. Progress. 4(3): 171
(1985).
3. Randall, T. L. and Knopp, P. V. Detoxification of specific organic
substances by wet air oxidation. J. Water Poll. Contr. Fed. 52:389
(1980).
4. Miller R. A. et al. Evaluation of catalyzed wet Air oxidation for
treating hazardous wastes, in Proceedings of the 7th Annual
Research Symposium, US EPA, Philadelphia, PA (1981).
5. Martinetz, D. Detoxification and decomposition, in: Chemical
Wastes:
Handling
and
Treatment.
Springer-Verlag
Berlin-Heidelberg, p267-176 (1986).
6. Modell, M. Supercritical water oxidation, in H. M. Freeman (ed),
Standard Handbook of Hazardous Waste Treatment and Disposal
McGraw Hill, Inc., p153-168 (1989).
7. Thornton, T. D. Phenol oxidation in supercritical water: Reaction
kinetics, products, and pathways. Ph.D. Dissertation, University of
Michigan, Ann Arbor, MI, p198 (1991).
8. CRC, Handbook of Chemistry and Physics. 56th ed. CRC Press
Inc., p141-143, Cleveland, OH (1975).
9. Hoigne, J. and Bader, H. The role of hydroxyl radical reactions in
ozonation processes in aqueous solutions. Water Resources 10:377
(1976).
10. Hoigne' J. and Bader, H. Rate constants of reactions of ozone with
organic and inorganic compounds in waste. Water Research
17:173 (1983).
11. Posselt, H. S., and Weber, W. J., Jr. Chemical oxidation, in:
Physicochemical Processes for Water Quality Control. John Wiley
& Sons,, p.363-412, New York (1972).
12. Rice, R. G. and Gomez-Taylor, M. Oxidation by-products from
drinking water treatment, in: P. M. Huck and P. Toft (ed),
Treatment of Drinking Water for Organic Contaminants. Pergamon
Press, New York, pl07-134 (1987).
13. Farhataziz, P. C. and Ross, A. B., Selected Specific Rates of
Radicals of Transients from Water in Aqueous Solutions. National
Bureau of Standards, Washington, D. C. NSRDS-NBS59, (1977).
14. Glaze, W. H. Drinking water treatment with ozone. Environ. Sci.
and TechnoL 21:224 (1987).
15. Glaze, W. H., Kang, J. W., and Chapin, D. H. The chemistry of
water treatment involving ozone, hydrogen peroxide and ultraviolet
radiation. Ozone Sci. & Technol. 9(4):335 (1987).
16. Buettner, G. R. Spin trapping of hydroxyl radicals, in: R. A.
Greenwald (ed), CRC Handbook of Methods for Oxygen Radical
Research. CRC Press, Boca Raton, p 151-155, FL (I 985).
17. Spinks, J. W. T. and Woods, R. J. An Introduction to Radiation
Chemistry. John Wiley & Sons, New York, NY (1990).
18. Fenton, H. J. H. Oxidation of tartaric acid in the presence of iron..Z
Chem. Soc. 65:899 (1894).
19. Fenton, H. J. H. Oxidation of certain organic acids in the presence
of ferrous salts. Proc. Chem. Soc. 15:224 (1899).

376
20. Eisenhauer, H. R. Oxidation of phenolic wastes. J. Water Poll.
ControlFed 36:1116 (1964).
21. Bishop, D. F., Stem, G., Fleischman, M., and Marshall, L. S.
Hydrogen peroxide catalytic oxidation of refractory organics in
municipal wastewaters. Ind. & Eng. Chem. Design and Develop.
7:110 (1968).
22. Bowers, A. R., Gaddipati, P., Eckenfelder, W. W., Jr., and Monsen,
R. M. Treatment of toxic or refractory wastewaters with hydrogen
peroxide. Water Sci. and Tech. 21:477(1989).
23. Wang, Y. T., and Latchaw, J. L. Anaerobic biodegradability and
toxicity of hydrogen peroxide oxidation products of phenols.
Research ,1. Water Poll.Cont. Fed. 6:234(1990).
24. Rauch, P. A., Watts, R. J., and Miller, G. C. Optimum conditions
for the Fenton's reagent treatment of pentachlorophenol in a
simplified soil system. J. of Environ. Eng., (1991).
25. Sedlak, D. L. and Andre, A. W. Aqueous phase oxidation of
polychlorinated biphenyl by hydroxyl radical. Environ. Sci. &
Technol. 25:1419(1991).
26. Sedlak, D. L. and Andre, A. W. Oxidation of chlorobenzene with
Fenton's reagent. Environ. Sci. & Technol. 25:777(2, 1991.
27. Buhler, R. F. Staehelin, J. Hoigne, J. Ozonation decomposition in
water studied by pulse radiolysis. J. Phys. Chem. 80:2560(1984).
28. Sundstorm, D. W. Klei, H. E., Nalette, T. A. Reidy, D. J., and Weir,
B. A.. Destruction of halogenated aliphatics by ultraviolet catalyzed
oxidation with hydrogen peroxide. Hazardous Waste Hazard. Mat.
3: 101(1986).
29. Sundstrom, D. W., Weir, B. A. and Redig, K. A., Destruction of
mixtures of Aromatic pollutants by UV light catalyzed oxidation
with hydrogen peroxide, in (Eds.) Tedder, D. W. and Pohland, F.
G., Emerging Technologies in Hazardous Waste Management:
Chapter 5. ACS Symposium Series, 422:67 (1990).
30. Weir, B. A, Sundstrom, D. W., and Klei, H. E. Destruction of
benzene by ultraviolet catalyzed oxidation with hydrogen peroxide.
Hazardous Waste & Hazard, Mater. 4:165(1987).
31. Chrostowski, P. C., Dietrich, A. M., and Suffer, I. H., Ozone and
oxygen induced oxidative coupling of aqueous phenols. Water Res.,
17:1627(1983).
32. Glaze, W. H. and Kang, J. W. Advanced oxidation processes for
treating groundwater contaminated with TEC and PCE: Laboratory
studies. ,1. Am. Water Works. Assoc. 5:57(1988).
33. Baxendale, T. A. and Wilson, J. A. The photolysis of hydrogen
peroxide at high light intensities. J. Chem. Soc. Faraday Trans.
53:344(1957).
34. Hart, E. J. Shested, K. Holcman, P. Molecular absorption of
ultraviolet and visible bands of ozone in aqueous solution. Anal.
Chem. 55:46(1983).
35. Cooper, W. J., Nickelson, M. G., Meacham, D. M. Cadavid, E. M.,
Waite, T. D., and Kurucz, C. N. High energy electron beam
irradiation: qualitative evaluation of factors affecting removal of
toxic chemicals from aqueous solution, in Proceedings of the 11th
Annual National Conference & Exhibition, Superfund-90,
Washington, D.C., p753-759(1990).
36. Suslick, K. S. Ultrasound, Its Chemical, Physical and Biological
Effects. VCR Publ. New York, NY,, p 138-150(1988).
37. Mason, T. J. Practical Sonochemistry. Ellis Horwood Publ. New
York, NY., p186(1991).
38. Mason, T. J. Chemistry with Ultrasound. Elsevier Appl'd. Sci.,
New York, NY. p 195(1990).
39. Mason, T. J. and Lomier, J. P. Sonochemistry. Ellis Horwood Ltd.
New York, NY., p252(1988).
40. Prasad, P. B. Sharma, P. K. Chemic Analytique 2:160(1970).
41. Chen, J. R., Xu, X. W., Lee, A. S., and Yen, T. F. A Feasibility
study of dechlorination of chloroform in water by ultrasound in the
presence of hydrogen peroxide. Environ. Technol. 11:829(1990).
42. Ho, P. C. Photooxidation of 2,4-dinitrotoulene in aqueous solution
in the presence of hydroxy radical. Environ. Sci. & Technol
20:260(1986).

C . P . H U A N G , C. D O N G , AND Z. T A N G
43. Jennings, B. H. and Townsend, S. N. The sonochemical reaction of
carbon tetrachloride and chloroform in aqueous suspension in an
inert atmosphere. 3~ Phys. Chem. 65:1574(1961).
44. Chen, B. and White, J. M. Behavior of Ti3+ centers in the
low-temperature reduction of Pt/TiO2 system. , Phys. Chem.
87:1326(1983).
45. Edmonds, P. D. Ultrasonics. Academic Press, New York, NY.
p355-358 (1981).
46. Formi, L. Bahnemann, D., and Hart, E. J. Mechanism of the
hydroxide Ion initiated decomposition of ozone in aqueous solution.
Phys. Chem. 86:255(1982).
47. Sehested, K. Holeman, J. and Hart, E. J. Ultraviolet spectrum and
decay of the ozonide Ion radical, 03 in strong alkaline solutions. J.
Phys. Chem. 86:2066(1982).
48. Peyton, G. R. Modeling advanced oxidation processes water
treatment, in: Tedder D. W. and Pohland, G. F. (eds), Emerging
Technologies in Hazardous Waste Management, ACS Symposium
Series, 422:100(1991).
49. Games, L. M. and Staubach, J. a. Reaction of nitrilotriacetate with
ozone in model and natural waters. Environ. Sci. & Technol.
14:571(1980).
50. Haag, W. R.and Hoigne, J. Dynamics of the ozonation of phenol. Ii.
Mathematic simulation. Environ. Sci. & TechnoL 17(5): 261
(1983).
51. Niegoski, S. J. Destruction of phenols by oxidation with ozone.
lndustr. & Eng. Chem. 45:632(1953).
52. Jones, H. R. Environmental Control in the Organic and
Petrochemical Industries, Noyes Data Corp., Park Ridge, NJ
(1971).
53. Haber, F. and Weiss, J. The catalytic decomposition of hydrogen
peroxide by iron salts. Proc. R. Soc. A 147:332(1934).
54. Walling, C. Fenton's reagent revisited. Aec. Chem Res.
8:125(1975).
55. Walling, C. Fenton chemistry revisited.
Acc. Chem. Res.
9:175(1976).
56. Frank, A. J. and Honda, K. Oxygen and hydrogen generation of
polymer protected CdS photoanodes, a~ Electroanal. Chem.150:
673(1983).
57. Gerischer, H. On the stability of semiconductor electrode against
photodecomposition", 3~ Eleetroanal. Chem. 58: 263(1975).
58. Gratzel, M. Energy Resource Thorough Photochemistry and
Catalysis, Academic Press, New York, NY (1983).
59. Kawai, T. and Sakata, T. Photocatalytic decomposition of gaseous
water over TiO2 and TiO2-RuO2 surface, Chem. Phys. Letter, 72(1):
87(1980).
60. Kawai, T., and Sakada, T. Conversation of carbohydrate into
hydrogen fuel by a photocatalytic process", Nature, 286:
474(1980).
61. Kawai, T. and Sakada, T. Photocatalytic hydrogen production from
water by the decomposition of polyvinychloride, protein, algae,
dead insects and excrement. Chem. Letters, 81: 34(1981).
62. Fujihira, M. et al. Heterogeneous photocatalytic oxidation of
aromatic compounds on semiconductor materials.
The
photo-Fenton type reaction. Chem. Letter (Chem. Soc. Japan) 1053
(1981).
63. Fujihira, M. et al. Heterogeneous photocatalytic reaction on
semiconductor materials", J. Electroanal. Chem. 126:277(1981).
64. Barbeni, M Pramauro, E., Pelizelli, E., Vincenti,M., Borgarello, E.,
and Serpone, N. Photodegradation of pentachlorophenol catalyzed
by semiconductor particles. 14:195 0985).
65. Barbeni, M., Morello, M. Pramauro, E., Pelizelli, E. Vincenti, M,
Borgarello, E., Graetzel, M., and Serpone, N. Photodegradation of
4-chlorophenol catalyzed by titanium dioxide particle. Nouv. ,~
Chim. 8:547 (1987).
66. Ollis, D. F., Hsiao, C. Y., Budiman, L., & Lee, C. L. Heterogeneous
photoassisted catalysis:
conversions of perchloroethylene,

ADVANCED CHEMICAL OXIDATION


dichloroethane, chloroacetic acids, and chlorobenzenes. J
Catalysis. 88:89 (1984).
67. Pruden, A. L., & Ollis, D. F. Photoassisted heterogeneous catalysis:
The degradation oftrichloroethylene in water. J. Catalysis 82:404
(1983).
68. Matthews, R. W. Photo-oxidation of organic material in aqueous
suspensions of titanium dioxide. Water Res. 20: 569(1986).
69. Matthews, R. W. Photo-oxidation of organic material in aqueous
suspensions of titanium dioxide. Water Res. 20(5): 569(1986).
70. Kakvoda, R, Oxidation of organic compounds with electrolytically
generated oxidant I. Oxidation of methanol ethanol and formic acid.
Electroanal. Chem., 24:53(1970).
71. Matsue, T., Fujihira, M., and Osa, T. Oxidation of alkylbenzenes by
electrogenerated
hydroxyl
radicaI.J.
Electrochem.
Soc.
128:2565(1981).
72. Kruatler, B. and Bard, A. J., Heterogeneous photocatalytic
decomposition of organic acids. J. Am. Chem. Soc. 100:
5985(1978).
73. Pavlik, J. W. and Tantayanon, S. Photocatalytic oxidations of
lactams and n-acylamines. ,Z Am. Chem. Soc., 103:6755(1981).
74. Sudoh, M., Kodera, T., Sakai, K. Zhang, J. Q. and Kqide, K.
Oxidation degradation of aqueous phenol effluent with
electrogenerated Fenton's reagent. J. Chem. Eng. of Japan.
19:513(1986).
75. Sudoh M., Kodera, T., Hino H., and Shimamura, H. Effect of
anodic and cathodic reaction on oxidative degradation of phenol in

377
an undivided bipolar electrolyzer. J Chem. Eng. Japan.
21:198(1988).
76. Sudoh M., Hino, K. Shimamura, H. Oxidation degradation rate of
phenol in an undivided bipolar electrolyzer. ~ Chem. Eng. Japan.
21:536(1988).
77. Sudoh, M, Kitaguchi, H. and Koide, K. Polarization characteristics
of packed bed electrode reactor for electroreduction of oxygen to
hydrogen peroxide. J of Chem Eng. Japan. 18:364(1985)
78. Sudoh, M., Kitaguchi, H. and Koide, K. Electrochemical production
of hydrogen peroxide by reduction of oxygen. J of Chem. Eng. of
Japan. 18:409(1985).
79. Huang, C. P., C. S. Chu, L. Takiyama, and D. Dong. The
Development of Electrochemical Processes For ln-situ Treatment
Of Ammonium Perfluoro-octanoate Contaminated Groundwater,
Technical Report, E. I. Du Pont de Nemours & Company, 106pp
(1992).
80. Dong, C. and Huang, C. P. Photocatalytic degradation of
4-chlorophenol by TiO2. in: C. P. Huang, C. R. O'Melia and
Morgan, J. J. (eds.) Aquatic Chemistry. ACS Advances in
Chemistry Series, (1993). 1993.
81. Tseng, J. M. and Huang, C. P. Mechanistic aspects of the
photocatalytic oxidation of phenol in aqueous solutions, in: Tedder
D. W. and Poland, G. F. (eds), Emerging Technologies in
Hazardous Waste Management, ACS Symposium Series,
422:12(1991).

You might also like