You are on page 1of 18

Atmospheric Environment 140 (2016) 117e134

Contents lists available at ScienceDirect

Atmospheric Environment
journal homepage: www.elsevier.com/locate/atmosenv

Review article

Catalytic oxidation of volatile organic compounds (VOCs) e A review


Muhammad Shahzad Kamal b, Shaikh A. Razzak a, Mohammad M. Hossain a, *
a
b

Department of Chemical Engineering, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia
Center for Integrative Petroleum Research, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia

h i g h l i g h t s
 Common VOCs, their sources and impacts.
 Alternative techniques for VOCs destruction/separation.
 Catalytic oxidation of VOCs.
 Mechanism of catalytic oxidation of VOCs.
 Recent development of VOCs oxidation catalysts.

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 22 January 2016
Received in revised form
27 April 2016
Accepted 17 May 2016
Available online 18 May 2016

Emission of volatile organic compounds (VOCs) is one of the major contributors to air pollution. The main
sources of VOCs are petroleum reneries, fuel combustions, chemical industries, decomposition in the
biosphere and biomass, pharmaceutical plants, automobile industries, textile manufacturers, solvents
processes, cleaning products, printing presses, insulating materials, ofce supplies, printers etc. The most
common VOCs are halogenated compounds, aldehydes, alcohols, ketones, aromatic compounds, and
ethers. High concentrations of these VOCs can cause irritations, nausea, dizziness, and headaches. Some
VOCs are also carcinogenic for both humans and animals. Therefore, it is crucial to minimize the emission
of VOCs. Among the available technologies, the catalytic oxidation of VOCs is the most popular because of
its versatility of handling a range of organic emissions under mild operating conditions. Due to that fact,
there are numerous research initiatives focused on developing advanced technologies for the catalytic
destruction of VOCs. This review discusses recent developments in catalytic systems for the destruction
of VOCs. Review also describes various VOCs and their sources of emission, mechanisms of catalytic
destruction, the causes of catalyst deactivation, and catalyst regeneration methods.
2016 Elsevier Ltd. All rights reserved.

Keywords:
VOCs
Oxidation
Oxidation process
Catalysts
Multicomponent catalysts
Mechanism
Kinetics

1. Introduction
Volatile organic compounds (VOCs) are those organic compounds with a Reid vapor pressure of over 10.3 Pa at normal temperature (293.15 K) and pressure (101.325 kPa). The VOCs are a
large group of carbon-based chemicals that easily evaporate at
room temperature (Li et al., 2009; Ojala et al., 2011; Olsen and
Nielsen, 2001). Table 1 lists the physicochemical and thermodynamic properties of common VOCs (Ihsan, 1995). VOCs are classied as major contributors to air pollution. They contribute both
indirectly as ozone/smog precursors and directly as substances
toxic to the environment (Amann and Lutz, 2000; Finlayson-Pitts

* Corresponding author.
E-mail address: mhossain@kfupm.edu.sa (M.M. Hossain).
http://dx.doi.org/10.1016/j.atmosenv.2016.05.031
1352-2310/ 2016 Elsevier Ltd. All rights reserved.

and Pitts, 1997; Lakshmanan et al., 2010; Molina and Rowland,


1974; Peng and Wang, 2007; Rodhe, 1990). Rapid urbanization and
industrialization contribute to the growing emissions of VOCs into
the environment. Emission of VOCs can be from a wide range of
outdoor and indoor sources. Outdoor sources include but are not
limited to chemical industries, paper production, food processing,
paint drying, transportation, petroleum reneries, automobile
manufacturers, metal degreasing, textile manufacturers, electronic
component plants, solvents, and cleaning products. Indoor sources
include household products, ofce supplies, printers, heatexchanger systems, insulating materials, pressed woods, wood
stoves, and leaks from piping (Drobek et al., 2015; Liotta, 2010;
 and Liotta, 2012).
Ozturk and Yilmaz, 2006; Scire
Type and nature of VOCs depend on the source of emission.
Examples of VOCs are halogenated hydrocarbons, alcohols,

118

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

Table 1
Physical, chemical, and thermodynamic properties of common VOCs (Ihsan, 1995).
Compound

Formula

Cp (J/mol  C)

B.P. ( C)

DGf (kJ/mol)

DHf (kJ/mol)

DHc (kJ/mol) @ 25  C

Toluene
Propane
Benzene
Ethylbenzene
o-Xylene
Acetylene
Acetone
Formaldehyde
n-Hexane
Dicholoromethane
Tricholoromethane
Tetracholoromethane
Tricholoroethylene
Tetracholoroethylene
Acetaldehyde
Ethylene

C7H8
C3H8
C6H6
C8H10
C8H10
C2H2
C3H6O
CH2O
C6H14
CH2Cl2
CHCl3
CCl4
C2HCl3
C2Cl4
C2 H4 O
C2 H4

166.0
73.8
136.1
185.9
188.8
44.1
125.0
35.4
195.0
50.8
65.8
133.9
80.02
95.6
53.7
42.9

110.6
42
80.1
136
144
84
56
19
68
39.6
61.2
76.72
87.2
121.1
20.2
103.7

114.09
23.4
124.5
120.0
110.8
209.2
155.3
109.9
4.0
68.9
68.5
62.5
6.7
20.6
133.2
68.4

12.0
103.8
49.0
12.5
24.4
226.7
248.1
115.9
198.8
95.5
101.3
132.8
19.1
14.2
166.4
52.5

3909.8
2220.0
3267.6
4564.7
4552.8
1299.6
1789.9
563.4
4163.1
583.8
435.2
260.7
910.8
772.8
1192.3
1411.1

aldehydes, aromatics, alkanes, ketones, olens, ethers, esters, parafns, and sulfur containing compounds (Carpentier et al., 2002;
Doggali et al., 2012; Khan and Kr Ghoshal, 2000; Miranda-Trevino
and Coles, 2003; Ozturk and Yilmaz, 2006; Soylu et al., 2010;
Yose et al., 2015).
The effect of VOCs on the atmosphere depends on the nature of
VOCs, their concentration, and emission sources. However, they
have been identied as been responsible for stratospheric ozone
depletion, tropospheric ozone formation, ground level smog formation, climate change, sick building syndrome, decay of plants,
toxicity of the atmosphere, and carcinogenic effects in humans
(Carabineiro and Thompson, 2007; Masui et al., 2010; Ozturk and
Yilmaz, 2006). As most of the countries have imposed strict conditions to limit VOCs in the environment, there are numerous
research initiatives around the world developing efcient technologies to meet the stringent environment regulations.
The emission of VOCs can be controlled using methods based on
recovery and destruction. The techniques based on recovery
include absorption, adsorption, membrane separation, and
condensation. High concentrations of VOCs, especially watersoluble compounds can be removed from ue gas streams by absorption using a suitable solvent. The disposal of VOCs and the
spent solvent from an absorber are the common problems faced by
the absorption processes. Physical and/or chemical adsorption is
another technique in which a suitable adsorbent (activated carbon,
zeolite, polymeric adsorbents etc.) is employed to selectively
adsorb the VOCs. Adsorption approach is economical only in situations where VOCs are highly diluted in a ue gas stream. The high
cost of the adsorbents and the necessity of frequent regeneration of
the adsorbent are the major limitations of adsorption processes
(Kujawa et al., 2015). The VOCs can be removed using condensation
induced by increasing the system pressure at a given temperature
or lowering the temperature at a given pressure. One of the limitations of the condensation processes is the disposal of the spent
coolants. Due to the energy intensive nature of the process,
condensation is limited to only evaporative solvents (Shah et al.,
2000). Membrane separation is another possible alternative for
the removal of VOCs. Silicon rubber membranes are the most
commonly used membrane for the separation of VOCs. In the bioltration process, VOCs in the air is removed biologically in a solid
phase reactor (Leson and Winer, 1991). In this approach, the
contaminated wet-gas is fed at the bottom of the lter and the
contaminants in air diffuse into a wet, biologically active layer (i.e.,
biolm) on the surface of the lter particles. The biolm with aerobic bacteria degrades the target pollutant(s) and produces CO2,
water, and microbial biomass. Both the membranes and

bioltration process are costly and their operation/maintenance is


expensive.
In the methods based on destruction, the VOCs are converted
into carbon dioxide and water. The destruction processes can be
thermal, catalytic, or biological oxidation. Thermal oxidation or
thermal incineration is suitable for removing VOCs from ue gas
streams with a high ow rate and a high concentration of VOCs.
More than 99% of the VOCs can be burned by thermal oxidation,
typically at high temperatures (>1000  C), which requires additional fuel and temperature resistant materials. Incomplete thermal
combustion produces undesirable byproducts such as dioxins and
carbon monoxide in the incinerator ue gas. Moreover, noxious
byproducts are formed as a result of thermal incineration (Moretti,
2002). The maximum concentration of the VOCs must be less than
the lower explosive limit (LEL) of a particular compound in order to
avoid any explosions (Moretti, 2002). In practice, a maximum
concentration of 25% of the LEL is used, which can be achieved by
diluting with ambient air.
Catalytic oxidation is one of the most effective and economically
feasible techniques for the oxidation of VOCs into CO2, water, and
other relatively less harmful compounds. The catalytic oxidation
targets the complete destruction of VOCs rather than transferring it
to another phase as in other techniques, such as in condensation
and adsorption. In this approach, the VOCs are oxidized in presence
of a suitable catalyst at much lower temperatures (250e500  C)
than thermal oxidation processes (Carabineiro et al., 2015b; Chen
et al., 2008b, 2013, 2014b, 2014c; Konsolakis et al., 2013; Larsson
and Andersson, 1998; Papaefthimiou et al., 1998a; Santos et al.,
2010). Catalytic combustion is a more thermally efcient process
than the other non-catalytic thermal oxidation processes and can
be used for dilute efuent streams of VOCs (<1% VOCs). Furthermore, it can be more energy efcient if used in the recuperative
mode, coupling with a heat exchanger after the catalytic combustion chamber. As catalytic combustion takes place at lower operating temperatures and its startup fuel requirement is lower.
Therefore, moderate volumes of contaminated air can be treated at
much lower fuel costs (Musialik-Piotrowska, 2007; Peng and Wang,
2007; Pope et al., 1978; Van der Vaart et al., 1991). Although catalytic oxidation can be applied effectively to treat waste gas streams
with varying concentrations of VOCs and ow rates, it is most
suitable for moderate ow rates and a low concentration of VOCs
(Ojala et al., 2011; Spivey, 1987). The technique is relatively more
environmentally friendly due to its low temperature operation and
the formation of less dioxins and noxious products (Abbasi et al.,
2011; Jamalzadeh et al., 2013). By the proper design of the catalyst system, formation of other toxic reaction intermediates can

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

also be minimized. As catalytic oxidation is the most suitable for


efuent streams with low concentrations of VOCs, it is highly
suitable for end-of-pipe VOCs pollution control (Solsona et al.,
2011). One of the main challenges of catalytic destruction of VOCs
is the selection of the proper catalysts from the large number of
available catalysts. Due to their wide variety and nature of the range
of mixtures of VOCs, it is most often difcult to identify the best
possible catalysts. This article presents a comprehensive review of
the recent developments of catalytic oxidation of VOCs. This
manuscript is divided into the following sections; (i) Major VOCs
and their sources, (ii) Mechanisms and kinetics of catalytic oxidation of VOCs, (iii) Catalysts for the oxidation of VOCs, (iv) Deactivation and regeneration of catalysts, and (v) Conclusion.

119

Exposure to a low level of aldehyde may cause throat irritation,


shortness of breath, eye irritation, and chest tightness (Zhu and Wu,
2015). Exposure to a high level of aldehyde increases the risk of
acute poisoning, and long-term exposure can have adverse effects
on human health that may lead to chronic toxicity (Andersen et al.,
2008; Liotta, 2010; Main and Hogan, 1983; Zhu and Wu, 2015).
Long-term exposure can also cause nasal tumors and irritation of
the respiratory tract, mucous membranes of the eyes, and skin
(Collins, 2001; Yu and Crump, 1998). Due to the increased use of
ethanol-based biofuels in transportation vehicles, aldehydes from
biofuels in the atmosphere have signicantly increased in recent
 and Liotta, 2012). Owing to these reasons, the reduction
years (Scire
of aldehyde in the atmosphere is of signicant interest, particularly
through low-temperature processes.

2. Major VOCs and their sources


2.3. Aromatic compounds
2.1. Halogenated VOCs
Halogenated VOCs are hazardous compounds due to their
strong bioaccumulation potential, acute toxicity, and resistance to
degradation (Alonso et al., 2002). Among the halogenated VOCs,
polychloromethanes (PCMs) are the most common and widely
used in various applications. Main sources of PCMs are the waste
materials originating from water purication and evaporation
systems (Huang et al., 2014). The commonly encountered PCMs are
chloroform, dichloromethane, trichloromethane, and carbon tetrachloride (Cappelletti et al., 2012; Huang et al., 2014; Doherty,
2000; Malaguarnera et al., 2012; Alvarez-Montero et al., 2011).
Other halogenated VOCs include chlorobenzene, dichloroethane,
trichloroethane, tetrachloroethane, trichloroethylene, and tetrachloroethylene (Giraudon et al., 2008). Generally, these compounds
are used in formulations and processing of chemical extractants,
paints, and adhesives, drugs manufacturing, polymer syntheses, as
solvents in chemical reactions, and as cleaning agents (Aranzabal
et al., 2014). These compounds are highly volatile and often have
long atmospheric lifetimes (Aranzabal et al., 2014). Humans can be
exposed to halogenated VOCs by drinking water, inhalation, and
adsorption during swimming (Huang et al., 2014). Halogenated
VOCs have a signicant impact on the destruction of the ozone
layer and as a source of radicals in the atmosphere, which
contribute to the greenhouse gas effects. The 100-year global
warming potential (GWP) of halogenated VOCs range from 10 to
1800, which is much higher than that of CO2 with a GWP of only
one (Abedi et al., 2015).
2.2. Aldehydes
Aldehydes are among one of the largest components of the total
reactive VOCs in the atmosphere. Formaldehyde and acetaldehyde
are two of the most commonly encountered aldehyde VOCs. These
belong to one of the groups acting as a source of new radicals,
which control the formation of ozone. The most important source
of formaldehyde is the degradation of VOCs (alcohols and other
hydrocarbons) in multiple steps oxidations (Ivanova et al., 2013).
Aldehydes are one of the major indoor pollutants, which are
emitted from decorative materials and trapped in airtight buildings
(Klett et al., 2014; Mitsui et al., 2008). They are also released from
many industrial products, such as treated wood resins, cosmetics,
plastic adhesives, construction materials, cleaning agents, disinfectants, particleboard, medium-density breboard, plywood,
carpeting, cigarette smoke, and fabrics (Li et al., 2014b). Aldehydes
can also be formed by the breakage of the C-C bond of alkoxy
radicals (Atkinson, 1997). Biogenic aldehyde sources include live
and decaying plants, bio-waste decomposition, biomass burning,
and seawater (Luecken et al., 2012).

Aromatic compounds such as benzene, toluene, and ethylbenzene occur naturally in petroleum products, including gasoline
and other fuels produced from crude oil. Consequently, the
incomplete combustion of liquid fuels in automobiles releases
signicant amounts of aromatic compounds into the atmosphere.
These compounds are also used in various products and formulations such as petrochemicals, paint, medicine, and detergents

(Ozelik
et al., 2009). Aromatic compounds are not only toxic and
carcinogenic, but also cause severe damage to the ozone layer,
produce photochemical smog, and pose mutagenic hazards (Kim
and Shim, 2010). The aromatic solvents commonly used in paints,
thinners, gums, adhesives, lacquers, and printing inks, are classied
as priority pollutants (Shahna et al., 2010). At low-level exposure,
aromatic VOCs can cause weakness, confusion, nausea, loss of
appetite and memory, tiredness, and loss of sight. Inhalation of a
high level of aromatic compounds can cause unconsciousness,
dizziness, and even death (Kim, 2002; Suib, 2013). The maximum
safe limit of aromatic compounds is 1 mg/L in drinking water and
200 ppm in air (Alifanti et al., 2007; Blanco et al., 2007; Chen et al.,
2008a; de Rivas et al., 2006; Delimaris and Ioannides, 2008b;
Huang et al., 2008a; Luo et al., 2007; Musialik-Piotrowska and
 et al., 2001b; Scire
Landmesser, 2008; Palacio et al., 2008; Scire
et al., 2003; Tidahy et al., 2007, 2008; Wang et al., 2008; Wu et al.,
2000; Zhang et al., 2008; Li et al., 2015).
2.4. Polycyclic aromatic hydrocarbons
Polycyclic aromatic hydrocarbons (PAH) are a family of VOCs,
and they are mainly released by combustion processes (VarelaGanda et al., 2013). These compounds contain several benzene
rings in their structure. Some typical examples of PAH are naphthalene (two rings), phenanthrene (three rings), and pyrene (four
rings). The main sources of PAH are incomplete fuel combustion,
asphalt transformation plants, and coal power plants (Diehl et al.,
2010; Puertolas et al., 2010). PAH have been identied as carcinogenic and therefore, it is extremely important to reduce their
emission (Kim et al., 2007b).
2.5. Alcohols and ketones
Alcohols and ketones are commonly used in cosmetics and
personal care products such as nail polish, nail polish removers,
colognes, perfumes, rubbing alcohol, and hair spray. Ketones are
used in aerosols, varnishes, window cleaners, paint thinners, and
adhesives (Makshina et al., 2008; Santos et al., 2011). Common
alcohol based VOCs include ethyl alcohol, isopropyl alcohol, benzyl
alcohol, while acetone, methacrylates (methyl or ethyl), ethyl acetate, and methyl ethyl ketone are ketone based VOCs.

120

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

Alcohols (mainly ethanol) increase the formation of aldehydes


by secondary reactions which cause throat irritation, shortness of
breath, eye irritation, and chest tightness (Zhu and Wu, 2015).
Long-term exposure to short-chain alcohols (ethanol, isopropanol,
and n-butanol) may cause central nervous system depression (U.S.
Department of Health and Human Services. Hazardous Substances
Data Bank (HSDB, online database). National Toxicology Information
Program, National Library of Medicine, Bethesda, MD. 1993). High
concentrations of ketones can cause the irritation of eyes, nose, and
throat. Other effects reported from acute inhalation exposure in
humans include central nervous system depression, headache, and
nausea. Some studies have reported nerve damage due to exposure
to ketones (U.S. Environmental Protection Agency. Updated Health
Effects Assessment for Methyl Ethyl Ketone. EPA/600/8-89/093. Environmental Criteria and Assessment Ofce, Ofce of Health and Environmental Assessment, Ofce of Research and Development,
Cincinnati, OH. 1990.)
2.6. Miscellaneous VOCs
Some alkenes also belong in the category of VOCs. Propylene is
considered to be extremely polluting due to its large photochemical
ozone creativity potential (POCP). Ethylene is of considerable
importance as it is being used as a raw material in many petrochemical syntheses (Patdhanagul et al., 2012). Ethylene released
during storage of agricultural products, can induce physical and
chemical changes of vegetables and fruits (Patdhanagul et al., 2010;
Sue-aok et al., 2010; Trinh et al., 2015). Ethylene is extensively used
as a solvent in the production of varnishes, synthetic resins, adhesives, printing ink, organic intermediates of pharmaceutical
products, and perfumes due to its low toxicity, good solubility, and
good volatility (Yuan et al., 2011; Zhang et al., 2009). Methyl tertbutyl ether (MTBE) belongs in the category of semi-volatile VOCs,
and is used as a fuel enhancer to improve octane ratings (ZadakaAmir et al., 2012; Zhou et al., 2013) and the oxidation ability of
the fuel (Kujawa et al., 2015). The presence of MTBE in drinking
water sources has become a major issue in Europe and USA (Ji et al.,
2009; Rossner and Knappe, 2008). The main sources of MTBE are
leakage from underground storage tanks, air deposition, and exhausts from watercrafts. MTBE is potentially carcinogenic and
adversely affects the odor and taste of drinking water. The safe limit
for MTBE in drinking water issued by the US Environmental Protection Agency is 20e30 ppb (Zadaka-Amir et al., 2012).
3. Mechanisms and kinetics of catalytic oxidation of VOCs
As discussed above, there are a range of volatile compounds,
which play a direct role in atmospheric chemistry. Although the
catalytic oxidation of VOCs has been extensively studied, it is still
difcult to provide a generalized correlation and a single reaction
mechanism due to the different properties of pollutants and reaction conditions. Hermia and Vigneron have established a correlation between the oxidizability of VOCs and their molecular weight
(Hermia and Vigneron, 1993). According to them, the higher the
molecular weight, the more difcult it is to oxidize VOCs. This
correlation agrees well with the order established by Palazzolo and
Tichenor
regarding
the
ease
of
oxidation:
alcohols < aldehydes < aromatics < ketones < acetates < alkanes
(Tichenor and Palazzolo, 1987).
At sufciently high reaction temperatures, the catalytic oxidation mainly gives CO2 and H2O according to the following reaction:



1
y
Cx Hy x y O2 /xCO2 H2 O
4
2

The mechanisms proposed for the complete catalytic oxidation


of VOCs generally fall into three main categories: (i) Marsevan
Krevelen (MVK) model, (ii) Langmuir-Hinshelwood (L-H) model,
and (iii) Eley-Rideal (E-R) model.
Mars-van Krevelen model considers that the reaction occurs
between the adsorbed VOCs and the lattice oxygen of the catalyst
rather than the oxygen in the gas phase. This model assumes that
the oxidation of the VOCs takes place in two steps as shown in Fig. 1.
In the rst step, the adsorbed VOCs react with oxygen in the
catalyst, resulting in the reduction of the metal oxide. In the second
step, the reduced metal oxide is reoxidized by the gas phase oxygen
present in the feed (Song et al., 2001). As the catalyst is rst reduced
and then reoxidized, this mechanism is also known as the redox
mechanism. In the steady state, rates of the reduction and oxidation
steps must be equal. This model has been widely used for kinetics
modeling of oxidation reactions of hydrocarbons over metal oxide
n
~ ez et al., 2002). According to the Mars-van Krevecatalysts (Ordo
len mechanism, the rate of oxidation of the VOCs can be expressed
by Equation (2).

rVOC

kO2 kvoc Pvoc PO2

(2)

gkvoc Pvoc kO2 PO2

where, rVOC : reaction rate (mol/m3/s); PVOC: partial pressure of the


VOCs; PO2: partial pressure of oxygen; kVOC: rate constant of VOCs
oxidation; kO2: rate constant of catalyst re-oxidation; g is the stoichiometry coefcient of O2 in the oxidation.
The Langmuir-Hinshelwood (L-H) mechanism assumes that the
reaction takes place between the adsorbed VOCs and the adsorbed
oxygen. Therefore, it is essential for both the VOCs and oxygen
molecule (species) to be adsorbed on the surface of the catalyst
(Everaert and Baeyens, 2004). The VOCs and oxygen may adsorb on
similar type of active sites (single site L-H model) or two different
types of active sites (duel site L-H model). Accordingly, the rate
expressions for the single site L-H model and duel site L-H model
are different as expressed by Equations (3) and (4), respectively.

rVOC

rVOC

k Kvoc KO2 Pvoc PO2


1 Kvoc Pvoc KO2 PO2

2

k Kvoc KO2 Pvoc PO2



1 Kvoc Pvoc 1 KO2 PO2

(3)

(4)

where, KO2 is the equilibrium constant for the adsorption of oxygen


and Kvoc is the equilibrium constant for the adsorption of VOCs. The
advantage of this model is that it not only includes the reaction rate
but also takes into consideration of the adsorption of VOCs and
oxygen.
According to the EleyeRideal (EeR) mechanism, the reaction
occurs between the adsorbed oxygen species and reactant

(1)
Fig. 1. Schematic of oxidation of VOCs (benzene) on MnO (Li et al., 2011).

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

molecules in the gas phase. The controlling step is the reaction


between an adsorbed molecule and a molecule in the gas phase
(Behar et al., 2015). The following equation represents the kinetic
expression based on the E-R mechanism.

rVOC

k Kvoc Pvoc PO2


1 Kvoc Pvoc

(5)

The validity of each mechanism strongly depends on the properties of the catalyst (active metal and the support) as well as on the
nature of the VOCs. Generally, one of the above models provides a
good t for the experimental data for the oxidation of the VOCs
over metal-oxide and noble-metal catalysts.
Table 2 summarizes the kinetics models and the estimated
activation energies for the oxidation of a range of VOCs reported in
the literature. The activation energy varies between 10 and 50 kcal/
mol. Choudhary and Deshmukh conducted a kinetic study of the
combustion of propane and methyl ethyl ketone (MEK) over Crdoped ZrO2 (Choudhary and Deshmukh, 2005). Although the
experimental data t both the power law and MVK model
reasonably well, the MVK model gave the best t. Activation energy
for the combustion of propane and MEK was determined to be
16.5 kcal mol1 and 13.2 kcal mol1, respectively. Ordonez et al.
showed that the kinetics of the oxidation of aromatic compounds
(benzene and toluene) over Pt/g-Al2O3 cab be expressed using the
MVK model. The estimated activation energy of benzene and
toluene are 24 and 25 kcal/mol, respectively (Ordonez et al., 2010).
Radic et al. studied the kinetics of the oxidation of n-hexane and
toluene over 0.12% Pt/Al2O3 (Radic et al., 2004). MVK type rate
expression was used to calculate the reaction rate constant. The
activation energy for the chemisorption decreases with increasing
Pt crystallite size. On the other hand, the activation energy for the
surface reaction is independent of the Pt crystallite size. Gangwal
et al. also applied the MVK mechanism to study the kinetics of the
oxidation of n-hexane (Gangwal et al., 1988). Pei et al. tted the
experimental data of the catalytic oxidation of formaldehyde over
copper manganese oxide to three models including rst order kinetics, LH, and the MVK models (Pei et al., 2015). They proposed the
following models and tted the experimental data to them: (i)
single-variable (surface concentration) kinetic model; (ii) twovariable kinetic model (surface concentration and water vapor
concentration); and three variable model (surface concentration,
water vapor concentration, and temperature). The best t of the
experimental data was obtained for the LH model that considers
two adsorbed reactants with competitive adsorption. The goodness

121

of the t was found to be better at higher reaction temperatures.


Bedia et al. performed a kinetic study on a catalyst based on Pd
supported on carbon for toluene oxidation (Bedia et al., 2010).
Oxidation temperature was in the range of 250e400  C. Model
based on the LH mechanism provided the best t, with a minimum
standard deviation and the lowest error. The rate-limiting step was
reported to be the surface reaction between adsorbed toluene and
dissociatively adsorbed oxygen. Behar et al. studied the kinetics of
toluene oxidation over nano-sized Cu-Mn spinels (Behar et al.,
2015). Results showed that the experimental data did not t the
power law model or the Langmuir model well. MVK model provided a good t for the experimental data.
Miranda et al. tested several kinetic models for the oxidation of
trichloroethene over different metal oxide catalysts (Miranda et al.,
2007). They showed that the experimental data based on the Mn
catalyst follows rst-order kinetics, while the results obtained with
the Cr catalyst follows zeroth order kinetics. The kinetic behavior of
Cr can be best described by the LH mechanism, taking into
consideration the inhibition effect of Cl2. Tseng et al. determined
that the power law model provided the best t for the catalytic
oxidation of dichloroethane on the Mn2O3/g-Al2O3 catalyst (Tseng
et al., 2010). Xin et al. studied the catalytic oxidation of ethylene
over a palladium catalyst and the experimental data were t to
several kinetic models (Xin et al., 2015). The reaction is rst order in
the ethylene concentration with a total activation energy of 48.2 kJ/
mol. Hu studied the combustion kinetics of acetone over the
Cu0.13Ce0.87Oy catalyst using several models (Hu, 2011). Results
indicate that the power model does not provide a good t for the
acetone combustion. Experimental data t the kinetic equation
based on Langmuir-Hinshelwood model well.
The above discussion indicates that the mechanism of VOCs
oxidation mainly depends on the type of VOCs compounds and the
characteristics of the catalysts. Therefore, the understanding of the
characteristics of the catalyst is essential to determine the reaction
mechanism and develop the kinetics models representing the
destruction reactions for the treatment of a given type of VOCs.
4. Catalysts for the oxidation of VOCs
Catalysts used for the oxidation of VOCs can be classied into
three major groups: (i) noble metals catalysts (Papaefthimiou et al.,
1997, 1998c; Tsou et al., 2005); (ii) non-metal oxide catalysts
(Barbero et al., 2006; Bastos et al., 2012; Burgos et al., 2002b;
Centeno et al., 2005; Cordi et al., 1997b; Garcia et al., 2010;
Krishnamoorthy et al., 2000; Lahousse et al., 1998b; Parida and

Table 2
Summary of the activation energy of the oxidation of some VOCs, and the catalysts and kinetics models used.
VOCs

Catalyst

Model type

Activation energy (kcal/mol)

Reference

Methyl ethyl ketone


Methyl ethyl ketone
Ethylene
n-hexane
toluene
Benzene
n-Hexane
Formaldehyde
Toluene
Toluene
Trichloroethene
Trichloroethene
Dichloroethane
Acetone
Benzene
Toluene
n-Hexane

Cr-doped ZrO2
Cr-doped ZrO2
Pd
Pt/Al2O3
Pt/Al2O3
Pt,Ni/g-Al203
Pt,Ni/g-Al203
CuO/MnO2
Pd/activated- C
Cu1.5Mn1.5O4
MnO2
Cr2O3
Mn2O3/g-Al2O3
Cu0.13Ce0.87Oy
Pt/g-Al2O3
Pt/g-Al2O3
Pt/g-Al2O3

Mars-van Krevelen
Power law
Power law
Mars-van Krevelen
Mars-van Krevelen
Mars-van Krevelen
Mars-van Krevelen
Langmuir-Hinshelwood
Langmuir-Hinshelwood
Mars-van Krevelen
Power Law
Power Law
Power Law
Langmuir-Hinshelwood
Mars-van Krevelen
Mars-van Krevelen
Mars-van Krevelen

16.5
13.2
48.2
22.4
17.4
22.9
19.6
13.2
19.8
26.6
9.3
10.5
10.4
26.1
24.1
25.3
31.8

Choudhary and Deshmukh, 2005


Choudhary and Deshmukh, 2005
Xin et al., 2015
Radic et al., 2004
Radic et al., 2004
Gangwal et al., 1988
Gangwal et al., 1988
Pei et al., 2015
Bedia et al., 2010
Behar et al., 2015
Miranda et al., 2007
Miranda et al., 2007
Tseng et al., 2010
Hu, 2011
n
~ ez et al., 2002)
(Ordo
n
~ ez et al., 2002)
(Ordo
n
~ ez et al., 2002)
(Ordo

122

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

Samal, 1999); and (iii) mixed-metal catalysts (Gangwal et al., 1988;


Garcia et al., 2006). Metal oxide based catalysts are cheaper and
more resistant to poisoning, but they are less durable and less
efcient compared to supported noble-metal catalysts in the
oxidation of VOCs (Bastos et al., 2009). Typically, the active metal
oxides are deposited on a suitable support. Several types of supports with good thermal stability and high surface area have been
used, which include the following: alumina (Al2O3), zirconia (ZrO2),
CeO2, SiO2, titania (TiO2),SnO2, CuO, Fe2O3, La2O3, MgO, montmorillonite, zeolites, and carbon based materials (Delimaris and
rrez-Ortiz et al.,
Ioannides, 2008a, 2009; Garcia et al., 2004; Gutie
2006; Heynderickx et al., 2010; Li et al., 2009; Masui et al.; Mitsui
et al., 2008; Wyrwalski et al., 2006). The importance and
different functions of support materials were reviewed recently by
several groups (Corma and Garcia, 2008; Jntgen, 1986; Roth et al.,
2014; Trimm and Stanislaus, 1986). The use of nanotubular materials as catalyst supports, in which the active centers are nanoparticles inside the tubes, has also been investigated. Carbon
nanotubes, boron nitride nanotubes, and halloysite nanotubes are
some typical examples of nanotubular supports. Halloysite is a
hydrated phase of kaolinite which is cheaper and abundantly
available and can acquire different morphologies (Carrillo and
Carriazo, 2015).

2011). Also they are less stable in the presence of chloride compounds (Everaert and Baeyens, 2004). Several investigations on
noble-metal catalysts on different supports have been reported
(Cant and Hall, 1970). Performance of these catalysts depends on
the method of preparation, precursor type, particle size, metal
loading, concentration of the VOCs, reactor type, and the overall gas
ow rate (Cant et al., 1998; Carballo and Wolf, 1978; Radic et al.,
2004). Larger the heat of formation of an oxide the lesser the activity of the catalyst. Use of noble-metal catalysts in propylene
oxidation indicates that the total reaction rate increases with the
oxygen pressure and decreases with the olen pressure. For Rh and
Ir the oxidation rate increases with the olen pressure and decreases with the oxygen concentration. On the contrary, Ru catalysts show higher oxidation rate with increased oxygen pressure
while the rate is independent of the olen pressure (Cant and Hall,
1970; Carballo and Wolf, 1978). Paulis et al. studied the effect of
chlorine, which is normally added as a catalyst precursor, on the
oxidation of toluene on the Pt/Al2O3 catalyst (Paulis et al., 2001).
Chlorinated precursors increase the surface acidity and decrease
the adsorption of carbonates and bicarbonates on the surface of the
support during CO oxidation. Moreover, using chlorinated precursors will result in platinum oxychloride species, which have a
higher reduction temperature.

4.1. Noble-metal-based catalysts

4.1.1. Gold-based catalysts


For a long time, gold was considered a poor catalyst owing to its
inertness to reactive compounds. In 1987 Haruta et al. found that
the reactivity of gold catalysts can be improved by forming defective structures (Haruta, 1997b; Haruta et al., 1989). Gold catalysts
can be prepared by deposition-precipitation, chemical vapor
deposition, and cation adsorption methods (Liotta, 2014). When
deposited on a metal oxide support, the gold complexes and gold
nanoparticles show good reactivity towards oxidation reactions (de
Almeida et al., 2013; Hashmi et al., 2010). Performance of gold
catalysts depends on many factors, including the type and properties of the support, loading of gold particles, methods of

Supported noble metals (Pt, Pd, Rh etc.) are attractive as catalysts due to their high efciency for the removal of VOCs at low
temperatures (Avgouropoulos et al., 2006; Liotta, 2010;
dar et al., 2004; Papaefthimiou et al., 1999; Tidahy
Maldonado-Ho
et al., 2007). The most common support materials for the noble
metal catalysts are ceramics or metallic material in monolithic or
honeycomb form. Table 3 summarizes some noble metal catalyst
systems reported recently. Noble-metal-based catalysts are
expensive and can be deactivated by sintering or poisoning (Liotta,
2010) and alone they are not normally selective enough (Ojala et al.,

Table 3
Reported Noble-metal catalyst systems for the oxidation of VOCs.
Catalyst

Support

VOCs

Temp (oC)

Conversion %

References

Pd
Au
Au
Pd
Pt
Pt
Pt
Pt
Au
Au
Au
Au
Au
Au
Au
Au
Au
Au
Au
Au
Au
Pt
Pt
Pt
Pt
Au
Pd
Pd

Nb2O5
CeO2-ZrO2-TiO2
CeO2/Fe2O3
Co3O4
Al2O3
CeO2
CeO2
Al2O3
CuO
Fe2O3
La2O3
MgO
NiO
CuO
Fe2O3
La2O3
MgO
NiO
Co3O4
Co3O4
Co3O4
Activated carbon
Activated carbon
Activated carbon
Activated carbon
Fe2O3
Co3O4
Al2O3

Toluene
Propane
Benzene
Xylene
Toluene
n-butanol
n-butanol
n-butanol
Ethyl acetate
Ethyl acetate
Ethyl acetate
Ethyl acetate
Ethyl acetate
Toluene
Toluene
Toluene
Toluene
Toluene
Benzene
Toluene
Xylene
Benzene
Toluene
Ethylbenzene
Xylene
Toluene
o-Xylene
o-Xylene

440
360
200
249
200
167
135
165
311
354
325
290
345
315
345
>400
387
320
189
138
162
112
109
106
104
260
254
145

90
50
100
90
95
90
90
90
100
100
100
100
100
100
100
100
100
100
90
90
90
100
100
100
100
90
90
90

(Rooke et al., 2015)


(Ali et al., 2015)
(Tabakova et al., 2015)
(Wang et al., 2015)
(Rui et al., 2014)
(Sedjame et al., 2014)
(Sedjame et al., 2014)
(Sedjame et al., 2014)
(Carabineiro et al., 2015b)
(Carabineiro et al., 2015b)
(Carabineiro et al., 2015b)
(Carabineiro et al.,2015b)
(Carabineiro et al., 2015b)
(Carabineiro et al.,2015b)
(Carabineiro et al.,2015b)
(Carabineiro et al., 2015b)
(Carabineiro et al., 2015b)
(Carabineiro et al., 2015b)
(Liu et al., 2015)
(Liu et al., 2015)
(Liu et al., 2015)
(Joung et al., 2014)
(Joung et al., 2014)
(Joung et al., 2014)
(Joung et al., 2014)
(Han et al., 2014)
(Wang et al., 2013b)
(Huang et al., 2008b)

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

preparation, size and shape of the gold nanoparticles, and con and Liotta, 2012).
centration of VOCs (Scire
Carabineiro et al. used Au catalysts supported on La2O3, MgO,
NiO, and Fe2O3 for the oxidation of toluene (Carabineiro et al.,
2015b). They showed that the reducibility and the crystal size are
major factors affecting the catalytic activity, while the oxidation
state of Au has no effect on the catalytic activity (Carabineiro et al.,
2015b).
Au/Co3O4 system has been reported as the most active system
for the combustion of light alkanes (Haruta, 1997a; Waters et al.,
1994). Au/CeO2 and Au/MnOx have been used to oxidize different
VOCs such as toluene and ethyl acetate (Bastos et al., 2012). At
250  C, both Au/CeO2 and Au/MnOx oxidize ethyl acetate into CO2.
Au/Mn5O10 catalyst system completely oxidizes ethanol at 230  C.
Toluene oxidation is the toughest and it has been achieved at
300  C with Au/Mn3O4. Gold catalysts supported on ceria have been
used for the oxidation of benzene, toluene, propanol, propene, and
formaldehyde (Lakshmanan et al., 2010; Pinna, 1998; Scire et al.,
2003; Shen et al., 2008).
Solsona et al. investigated a catalyst based on gold supported on
nickel-cerium oxide for the oxidation of propane (Solsona et al.,
2011). These catalysts showed good activity, which is due to the
high surface area of the catalyst, low bond strength (Ni-O), and the
high reducibility of the nickel sites (Solsona et al., 2011). Shen et al.
studied several Au/CeO2 catalysts with a gold content of less than
0.85%, prepared by different routes for formaldehyde oxidation
(Shen et al., 2008). The results indicate that gold catalysts with a
high dispersion are more active in the oxidation of formaldehyde,
while the formation of larger crystals decreases the activity of the
catalyst.
4.1.2. Palladium catalysts
Palladium catalysts have a higher thermal and hydrothermal
resistance compared to other noble metal catalysts (Huang et al.,
2008b). There are different views on the contribution of the
metallic state and oxidized state of Pd to the oxidation reactions of
VOCs. Some authors believe that the oxidized state is more
lin and Primet, 2002), while others
important (Burch et al., 1996; Ge
consider that the metallic state is more important in oxidation
ge
 et al., 2000; Huang et al., 2008b; Ihm et al., 2004).
reactions (De
Some results suggest that both states are equally important
(Yazawa et al., 1998). The support plays an important role in the
activity of Pd-based catalysts. Normally the porous supports increase the activity of Pd catalysts due to the stabilization of active
Pd species on porous structures (Centi, 2001; Wang et al., 2013b).
Acid-base properties of the supports are known to inuence the
activity of Pd catalysts. Weakly acidic supports provide high catalytic activity (Okumura et al., 2003). Pd-based catalysts are more
efcient than noble-metal catalysts and metal-oxide catalysts in
removing benzene, toluene, xylene, and methane (Centi, 2001;
Huang et al., 2008b; Kim and Shim, 2009).
Wang et al. showed that Pd/Co3O4 catalysts possess high efciency for the oxidation of xylene. Efciency depends on the particle size, oxidized Pd species, and the oxygen vacancies on the
support (Wang et al., 2013b). The comparison of two Pd/Co3O4
catalysts synthesized by in situ nanocasting and post-impregnation
methods indicates that the catalyst synthesized by in situ nanocasting has a high catalytic activity. The high activity is due to the
more ordered structure of the catalyst and the well dispersed PdO
species. Huang et al. found that the catalytic activity of the Pd/Al2O3
catalysts increased with the increasing Pd loading (Huang et al.,
2008b). Ihm et al. showed that the pre-treatment methods
strongly affect the activity of the Pd/Al2O3 catalysts [140]. Catalysts
treated in hydrogen showed higher activity compared to the catalyst treated in air.

123

4.1.3. Platinum catalysts


Platinum-based catalysts are widely used in the oxidation of
VOCs due to their high activity and stability (Joung et al., 2014).
Although conventional platinum catalysts are supported on metal
oxides, activated carbon has also been used as the support (Joung
et al., 2014; Wu and Chang, 1998). Industrial applications of platinum catalysts are limited due to their high cost and the likelihood
of poisoning, especially when dealing with chlorinated products
(Jones and Ross, 1997; Krishnamoorthy et al., 2000; Petrosius et al.,
1993).
In humid environments, Pt-based catalysts suffer deactivation
owing to the adsorption of water on the support. Sedjame et al.
used a Pt catalyst supported on CeO2-Al2O3 for the complete
oxidation of n-butanol and acetic acid (Sedjame et al., 2014). The
addition of ceria to alumina decreases the surface area and changes
the physicochemical properties of the catalytic material and the
activity. Addition of ceria enhanced the performance of the catalyst
for the oxidation of acetic acid, while the oxidation of n-butanol
was not signicantly affected, which shows that the nature of the
VOCs inuences the performance of the catalyst.
4.1.4. Mixed noble-metal catalysts
Hosseini et al. studied toluene oxidation over a series of bimetallic Pd/Au/TiO2-ZrO2 catalysts (Hosseini et al., 2009). They found
that the Au/Pd catalysts showed higher activity than the monometallic Au or Pd catalysts. The highest toluene oxidation performance was found to be at a Pd/Au ratio of 4. The enhanced
performance of the bimetallic catalyst is related to the synergistic
effects between palladium and gold. Barakat et al. demonstrated
the synergistic effects by using the Pd-Nb-V/TiO2 catalyst for
toluene oxidation (Barakat et al., 2012). They showed that doping
by a hierarchically structured titania support with Nb and V enhances the catalytic activity of the support in the oxidation of VOCs.
Modication with vanadium increases the efciency of titaniasupported palladium catalysts for the oxidation of short chain alkenes (Garcia et al., 2004).
4.2. Non-noble metal oxides
Non-noble metal based catalysts can be either supported or
unsupported metal oxides (Carabineiro et al., 2015a). The supported catalysts show better activity and performance in the
oxidation of VOCs due to greater dispersion of the active component. The transition and rare earth metal oxides are used as nonnoble-metal catalysts. The most commonly used metal-oxide catalysts include copper oxide, manganese dioxide, iron oxide, nickel
oxide, chromium oxide, and cobalt oxide (Galvita et al., 2014; Garcia
et al., 2010; Huang et al., 2010; Morales et al., 2013; Solsona et al.,
2011; Xia et al., 2009). Table 4 shows some commonly used nonnoble-metal catalysts. The non-noble-metal catalysts have several
advantages, including being readily available and their low price
that make them a good alternative to expensive noble-metal catalysts. Although non-noble metal oxides have relatively lower activity than the noble-metal catalysts, they are commonly used for
the oxidation of VOCs due to their low cost (Ciuparu et al., 2002;
lin and Primet, 2002; Kim,
Delimaris and Ioannides, 2008a; Ge
 et al., 2001a). Other advantages
2002; Rotter et al., 2004; Scire
include the following: long lifetime, masking tolerance, capability
of regeneration, and the availability of a range of metal oxides in
different sizes and shapes (Zimowska et al., 2007).
Support materials and the preparation methods are crucial in
determining the performance of metal-oxide catalysts. Support
plays an important role in determining the physicochemical
properties of the active phase (Lin et al., 2015). Porous materials
with a high surface area and large pores favor high dispersion and

124

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

Table 4
Some reported metal-oxide catalysts for the oxidation of VOCs.
Catalyst

Support

VOCs

Temp

Conversion %

References

Co3O4
Co3O4
Co3O4
Co3O4
CuO
Co3O4
CeO2
Nb2O5

Clay
Clay
e
e
e
e
Aluminosilicate
e

Acetylene
Propylene
1,2-Dichloroethane
Ethyl acetate
Ethyl acetate
Propane
Acetone
Toluene

360
460
350
245
280
250
200
400

100
100
100
100
e
99
85
90

(Assebban et al., 2015)


(Assebban et al., 2015)
(de Rivas et al., 2012)
(Chen et al., 2013)
(Chen et al., 2013)
(Solsona et al., 2008)
(Lin et al., 2015)
(Rooke et al., 2015)

good catalytic activity in the oxidation of VOCs (Chen et al., 2014a;


He et al., 2012; He et al., 2010; Jeong et al., 2014; Soylu et al., 2010;
Yan et al., 2014). In the open literature there are numerous research
articles reporting attempts to enhance the activity of metal-oxide
catalysts. The methods used include making a porous structure
(Tang et al., 2014a; Xia et al., 2010), forming a solid solution
(Hanfeng et al., 2011; Liotta et al., 2008; Wang et al., 2013a) or
perovskites crystal structure (Huang et al., 2008a; MusialikPiotrowska and Syczewska, 2000; Pena and Fierro, 2001), and
doping (Chang et al., 2012). Cheaper metal oxides are also more
tolerant to deactivation by poisoning compared to catalysts based
on noble metals (Baldi et al., 1998; Carabineiro et al., 2015a; Li et al.,
2009; Morales et al., 2008; Rezlescu et al., 2015; Wenxiang et al.,
2015).
4.2.1. Cobalt-based catalysts
Co3O4 is one of the most active low-cost metal oxides, which has
been used for a wide range of reactions (Busca et al., 1997;
Choudhary and Deshmukh, 2005; Chen et al., 2013; de Rivas
et al., 2012; Garcia et al., 2010; Solsona et al., 2008). High activity
of Co3O4 is associated with the presence of mobile oxygen inside
their spinel type structure (Koodziej et al., 2012; Milt et al., 2002;
Solsona et al., 2008; Wyrwalski et al., 2007; Yuranov et al., 2002).
The catalyst has excellent reduction ability and oxygen vacancies as
well as a high concentration of electrophilic oxide species (Liu et al.,
2009). However, activity of Co3O4 depends on the method of
preparation, treatment conditions, oxidation state, and surface area
(Kovanda et al., 2006). Co3O4 is the most efcient catalyst available
for the total oxidation of toluene and propane (Wyrwalski et al.,
2006).
Co3O4 supported on TiO2 and Al2O3 was investigated for the
total oxidation of 1,2-dichlorobenzene (Krishnamoorthy et al.,
2000). For the supported Co3O4, the activity depends mainly on
the nature of the support and the metal oxide-support interactions.
de Rivas et al. synthesized a series of Co3O4 catalysts for the
oxidation of 1,2-dichloroethane using different routes (de Rivas
et al., 2011). The catalyst prepared by the precipitation method
with a particle size of 10 nm gave the highest activity. The highly
dispersed nano-sized crystals helped improve the activity of these
catalysts. According to de Rivas et al., the activity of the supported
Co3O4 catalyst is higher than that of the supported noble-metal
catalysts for the oxidation of 1,2-dichloroethane (de Rivas et al.,
2011).
4.2.2. Nickel-based catalysts
NiO is another active metal oxide used for various catalytic applications including the oxidation of VOCs (Solsona et al., 2011). The
catalyst is highly active because of its p-type semi-conductor
properties with an electron deciency in the lattice (Heracleous
et al., 2005). This allows electrons to be easily removed from the
metal cations, resulting in the formation of active species such as
O. The formation of Ni3 in NiO has been identied by many

researchers (Heracleous and Lemonidou, 2006; Salagre et al., 1996).


The addition of one heteroatom such as cerium, which forms NiO
with a high surface area and a smaller crystal size (Solsona et al.,
2012), enhances the activity of NiO.
4.2.3. Titanium-based catalysts
Titania is a low cost, readily available, and chemically stable
catalyst suitable for the removal of a range of VOCs, especially in
photocatalytic oxidation of VOCs. Titania can degrade VOCs into
CO2, acids, and H2O under near UV light (Fresno et al., 2008;
Hussain et al., 2010; Sleiman et al., 2009; Zou et al., 2006). Photocatalytic oxidation with titania occurs at low temperature, and
therefore, is suitable for removing indoor pollutants (Fujishima and
Zhang, 2006; Periyat et al., 2008).
4.2.4. Manganese-based catalysts
Manganese oxide is among low cost active catalysts for the
~ o et al., 2015a). They have been evaluated
oxidation of VOCs (Castan
for the destruction of many VOCs, including n-hexane (Lahousse
et al., 1998a), acetone (Parida and Samal, 1999), benzene (Luo
et al., 2000), ethanol (Aguero et al., 2009; Bastos et al., 2009;
Lamaita et al., 2005; Peluso et al., 2008), toluene (Aguero et al.,
2009), propane (Finocchio and Busca, 2001) and NOx (Kang et al.,
2007). The manganese-based catalysts are attractive due to their
high efciency in the oxidation and low toxicity. Performance of
manganese oxide depends mainly on the catalyst structure,
methods of preparation, surface area, nature of the support material, and the oxidation state (Einaga et al., 2015). The activity of the
MnO2 catalyst for the oxidation of ethyl acetate and n-hexane is
even higher than the activity of the Pt/TiO2 catalyst (Sun et al.,
2015). The high efciency of the catalyst is attributed to the coexistence of mixed valence states Mn2/Mn3 or Mn3/Mn4 and
lattice oxygen (Gandhe et al., 2007; Kim and Shim, 2010). The
reactivity of Mn2O3 is highest followed by MnO2 and MnO for CO
oxidation (Ramesh et al., 2008). Manganese-based catalysts are
deactivated due to the deposition of Cl species formed by the
combustion reactions. Therefore, Mn-based catalysts are best
suited in Cl-free environments (Finocchio et al., 2000).
4.2.5. Copper-based catalysts
Cupric oxide is also a highly active catalyst for the deep oxidation of CO, methane, methanol, ethanol, and acetaldehyde (Cordi
et al., 1997a; Hutchings and Taylor, 1999). CuO is also used to
catalyze the oxidation of methanol to methyl formate and propylene to acrolein (Cordi et al., 1997b). Several factors affect the performance of the CuO catalyst in the total oxidation of VOCs.
Catalysts prepared by different methods can demonstrate varying
activity (Heynderickx et al., 2010). The oxidation state of Cu, which
determines the mechanism of oxidation, plays an important role.
The lattice oxygen in CuO plays an active role in the oxidation and
with the depletion of lattice oxygen; the rate of oxidation becomes
limited by the diffusion of lattice oxygen to the surface. VOCs can

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

extract nearly all the lattice oxygen from CuO.


The presence of CuO in the CuO/Al2O3 catalyst signicantly
enhances the decomposition rate (Cordi et al., 1997b). Supported
CuO can dehydrogenate and oxidize various VOCs such as methanol, acetaldehyde, and formic acid. The dehydrogenation rate is
faster on CuO than on the supported Cu catalyst system (Cordi et al.,
1997b). The addition of other metals such as ceria can enhance its
catalytic properties (Heynderickx et al., 2010). Among the various
alumina supported catalysts, the activity of Cu is highest followed
by Co, Fe, and Ni for the total oxidation of toluene (Kim et al., 2014).
4.2.6. Chromium-based catalysts
Chromium oxide catalysts are a group of very active catalysts,
particularly for the removal of halogenated VOCs (Krishnamoorthy
et al., 2000; Petrosius et al., 1993). Crystalline chromium oxide is
more suitable than amorphous chromium oxide as it favors the
formation of CO2 (Sinha and Suzuki, 2007). Sinha et al. synthesized
mesoporous chromium oxide with high activity for the oxidation of
VOCs. The catalyst was prepared by the neutral templating route
and the presence of mixed oxidation states of 2 and 5 was
conrmed by XPS analysis (Sinha and Suzuki, 2007).
Several studies have shown that chromium oxide catalyst is the
most effective catalyst for the removal of VOCs. For example, the
activity of TiO2-supported catalysts for the catalytic oxidation of
trichloroethylene was highest for chromium oxide (98%) followed
by manganese oxide (79%), cobalt oxide (58%), and iron oxide (54%)
(Rotter et al., 2005). Chromium oxide catalyst supported on various
supports such as silica, alumina, porous carbon, clay, titania, and
clay was effective in the removal of carbon tetrachloride, chloromethane, trichloroethylene, ethyl chloride, chlorobenzene, and
perchloroethylene (Rotter et al., 2005).
Commercial application of the catalyst is limited due to the loss
of active Cr during long-term reaction. The loss of activity of Cr may
be due to the formation of other volatiles such as CrO2Cl2, by
cocking, by water, or attrition in uidized beds (Padilla et al., 1999;
Yim et al., 2001). High Cl content in ue gas can result in the loss of
chromium by attack of nascent Cl2 (Padilla et al., 1999). Chromium
trioxide supported on carbon is a very active catalyst for a range of
VOCs. However, high toxicity of chromium causes serious catalyst
disposal problems (Li et al., 2009).
4.2.7. Vanadium-based catalysts
Vanadium-based catalysts originally designed for the removal of
nitrogen oxides have proven to be active in the destruction of
various polychlorinated pollutants (Cho and Ihm, 2002).
Vanadium-based catalysts have a high resistance to sulfur dioxide
poisoning, in addition to high activity and selectivity (Jones and
Ross, 1997). Furthermore, these catalysts have a reasonable stability in Cl2-HCl environments. Consequently, the vanadium-based
catalysts are suitable for the oxidation of gas streams containing
chlorinated VOCs and NOx.
Delaigle et al. investigated the VOx/TiO2 catalytic system for the
abatement of different aromatic pollutants and arrived at important conclusions (Delaigle et al., 2009). Best performance was
achieved using sulphated titania as it exhibits a large number of
strong Lewis acid sites. The higher number of Lewis acid sites
improve the spreading of active VOx. The higher number of
Brnsted acid sites also helps in promoting the limited adsorption
of aromatics.
Krishnamoorthy et al. studied the catalytic oxidation of 1,2dichlorobenzene over a series of vanadium/TiO2-based catalysts
(Krishnamoorthy et al., 1998). Although the catalyst on a TiO2
support was active for the oxidation of 1,2-dichlorobenzene, the
addition of V2O5 considerably increases its activity. The oxidation of
1,2-dichlorobenzene produces CO and CO2, and other incomplete

125

combustion products were not detected (Krishnamoorthy et al.,


1998).
Co-pollutants of chlorobenzene such as CO, CO2, and NOx can
affect the performance of vanadium-based catalysts. Overall effect
of water on the VOx catalyst is a sum of three dynamic effects: (i)
retrieval of chlorine species, (ii) reduction of VOx active species,
and (iii) reduction in the number of adsorption sites. CO can affect
the acidity of the catalyst but it has no overall inuence. Chlorobenzene conversion is enhanced when NO is added in the presence
of oxygen (Delaigle et al., 2009).
In summary, due to the reasonable stability of vanadium-based
catalysts in Cl2-HCl environment, they have been proposed for the
catalytic oxidation of chlorinated VOCs.
4.2.8. Cerium-based catalysts
Cerium is the most abundant among the rare earth elements.
Cerium-based catalysts have unique properties due to their abundant oxygen vacancies associated with strong interactions with
other metals, oxygen storage capacity, and ready shuttling between
the Ce3 and Ce4 states (Gorte, 2010; Zimmer et al., 2002). Owing
to their outstanding oxygen storage capacity, cerium-based catalysts are often used as structural and electronic promoters (Huang
et al., 2015). CeO2 and CeO2-based catalysts are cheaper, environmental friendly, and are effective for non-chlorinated VOCs such as
methane, CO, methanol, and propane (Dai et al., 2007). CeO2 catalysts have shown good activity in the catalytic oxidation of phenolic
wastewater and the catalytic incineration of aromatic hydrocarbons
(Lin et al., 2002, 2003; Wang and Lin, 2004).
Cerium based catalysts have been tested for the catalytic
oxidation of a range of chlorinated VOCs. Dai et al. investigated the
catalytic activity of CeO2 for the oxidation of various chlorinated
VOCs (Dai et al., 2007). They concluded that the catalytic oxidation
of chlorinated alkanes is easier than chlorinated alkylenes. Dai et al.
also investigated the catalytic combustion of trichloroethylene
using the CeO2 catalyst (Dai et al., 2008). The CeO2 catalyst showed
a high activity for the catalytic combustion of trichloroethylene,
which is attributed to the high mobility of oxygen, basicity, and
oxygen-supplying ability of CeO2 (Dai et al., 2008).
In summary, cerium-based catalysts are more suitable for nonchlorinated VOCs, as they can be deactivated by the adsorption of
HCl or Cl2 (Li et al., 2010). However, research has now moved to
increase the efciency of cerium-based catalysts for destruction of
chlorinated VOCs.
4.2.9. Mixed-metal catalysts
Generally, the destruction efciency of single-metal-oxide catalysts is lower than that of precious metal catalysts for the removal
of VOCs. The performance of metal-oxide catalysts can be improved
by combining two or more oxides to impart a synergistic effect.
Table 5 presents some mixed metal catalysts reported in the open
literature. Generally, a synergistic effect is observed in a range of
composite oxides, including Mn-Ce oxides (Delimaris and
Ioannides, 2008a; Yu et al., 2010), Mn-Cu oxides (Morales et al.,
2006), Co-Ce oxides (Liotta et al., 2008), Sn-Ce oxides (Vasile
et al., 2013), Mn-Co oxides (Tang et al., 2014b), and Ce-Cu oxides.
A study has shown that the rate determining step in the total
oxidation of VOCs is the rate of oxygen removal from the metal
oxide, indicating that the reducibility of the metal oxide is the most
critical aspect (Bastos et al., 2009). Reducibility of a metal oxide can
be improved by adding a second cation, i.e., using mixed metal
oxides (Saqer et al., 2011). The activity of ceria containing catalysts
is greatly enhanced by other metal oxides such as nickel, copper,
and vanadium (Asgari et al., 2013; Liu et al., 2015; Trovarelli, 1996).
The high catalytic activity is associated with the dual oxidation
states of ceria (Ce3 and Ce4), which facilitate oxygen storage and

126

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

Table 5
Some reported mixed-oxide catalysts for the oxidation of VOCs.
Catalyst

VOC

Temp ( C)

Conversion (%)

References

Ce-Co
La-Co
Mn-Ce
Mn-Ce
Mn-Ce
Mn0.5Ce0.5-O2
Cu-Ce
Cu0.3Ce0.7Ox
Cu0.3Ce0.7Ox
Mn-Co
Mn-Co
Mn-Co
Mn-Co
Mn-Co
MnOx-TiO2
MnOx-TiO2-SnOx
Ce-Zr
Cu-Co

Toluene
Toluene
Benzene
Toluene
Ethyl acetate
Formaldehyde
Chlorobenzene
Toluene
Propanol
Toluene
Ethylbenzene
Ethyl acetate
n-hexane
Toluene
Chlorobenzene
Chlorobenzene
1-2 Dichloroethane
Benzene

250
300
260
245
180
270
328
212
192
250
250
194
210
250
177
177
120
290

100
100
90
90
90
100
99
90
90
98.7
90
90
90
100
90
90
90
90

(Carabineiro et al., 2015a)


(Carabineiro et al., 2015a)
(Wenxiang et al., 2015)
(Wenxiang et al., 2015)
(Wenxiang et al., 2015)
(Li et al., 2014b)
(He et al., 2015)
(He et al., 2014)
(He et al., 2014)
(Zhou et al., 2015)
(Zhou et al., 2015)
(Tang et al., 2014b)
(Tang et al., 2014b)
(Qu et al., 2014)
(Li et al., 2014a)
(Li et al., 2014a)
(de Rivas et al., 2013)
(Li et al., 2015)

release from the catalyst matrix (Yose et al., 2015). For example,
CeO2 catalysts achieved 90% conversion of trichloroethylene, but is
not thermally stable and deactivates within a few hours due to
adsorption of HCl and Cl2 (Dai et al., 2008). However, modifying
CeO2 with other metals can improve the oxygen storage capacity,
thermal resistance, and enhance its activity (Gluhoi et al., 2005;
Trovarelli et al., 1999).
CuO-CeO2 mixed-metal catalysts have been investigated in the
combustion reactions of CO and CH4, the reduction of NO, and the
wet oxidation of phenol (Avgouropoulos and Ioannides, 2003; Liu
and Flytzanistephanopoulos, 1995; Liu et al., 1994). The high efciency is attributed to the promoting effect of ceria due to the high
oxygen storage capacity. Cu-Ce oxides have been studied for the
oxidation of the following VOCs: ethyl acetate, ethanol (Larsson and
Andersson, 1998), benzene (Hu et al., 2008), and toluene (Delimaris
and Ioannides, 2009). This catalyst system has been synthesized
using combustion (Rao et al., 2003), thermal decomposition
(Xiaoyuan et al., 2001), impregnation, co-precipitation, and sol-gel
method (Zheng et al., 2005). He et al. synthesized Cu-Ce mixedoxide catalysts using the self-precipitation approach and investigated their efciency for toluene and propanol oxidation (He et al.,
2014). Cu and Ce have a synergistic effect on the reducibility of the
mixed-oxide catalyst. Low temperature reducibility is the highest
for Cu0.3Ce0.7Ox followed by Cu0.15Ce0.85Ox, Cu0.4Ce0.6Ox, and
CuCeOx.
In the oxidation of ethyl acetate, ethanol, and toluene the mixed
catalyst system has a lower activity compared to the individual
metal oxides due to the suppression of the intrinsic activity by
combining the two metal oxides (Delimaris and Ioannides, 2009).
MnOx-CeO2 catalyst system has been investigated for the abatement of different types of VOCs, such as ethanol (Rao et al., 2007),
formaldehyde (Tang et al., 2006), hexane (Picasso et al., 2007),
phenol (Chen et al., 2001), ethyl acetate (Delimaris and Ioannides,
2008a), and toluene (Delimaris and Ioannides, 2008a). These catalysts are mainly prepared by the sol-gel method (Rao et al., 2007;
Tang et al., 2006), co-precipitation (Chen et al., 2001; Picasso et al.,
2007; Qi et al., 2004; Silva et al., 2004; Tang et al., 2006), and
combustion (Delimaris and Ioannides, 2008a).
CeO2-CrOx mixed catalysts show excellent performance in the
removal of various halogenated VOCs (Yang et al., 2015b). Yang et al.
(2015a) showed that the formation of Cr6 species with its strong
oxidizing ability and the oxygen defects are enhanced due to the
good synergism between CeO2 and CrOx. All the mixed catalysts
prepared under different conditions show better performance

compared to the individual metal oxide catalysts.


~o
Mn-Co are efcient systems for the removal of VOCs (Castan
et al., 2015b; Liotta et al., 2013). Mixed-oxide catalysts of these
two metals are more efcient compared to the individual metal
catalysts due to the synergistic effects. Tang et al. showed that the
Mn and Co mixed-oxide catalyst performed better in the removal of
ethyl acetate and n-hexane as compared to the individual MnOx
and Co3O4 catalysts (Tang et al., 2014b). This high efciency is due
to the higher surface area, better low temperature reducibility, and
the porous structure of the mixed oxide.
Li et al. investigated the removal efciency of MnOx-TiO2 and
MnOx-TiO2-SnOx catalysts for the oxidation of chlorobenzene (Li
et al., 2014a). Both catalysts showed excellent performance with
90% conversion below 180  C. The stability of MnOx-TiO2-SnOx is
higher than that of MnOx-TiO2, indicating that the addition of Sn
increases the stability of Mn-Ti oxides. The high stability of MnOxTiO2-SnOx is due to the low energy required to desorb Cl species
and the absence of MNOxCly species on active sites. Mn is the active
part of the mixed catalyst, while Sn has a minimal effect on the
activity. Sn acts as an additive and increases the stability of the
catalyst. TiO2 acts as a support and has very low activity and it helps
to disperse active Mn (Li et al., 2014a). Genty et al. developed a
different mixed-metal-oxide catalyst (Fe, Cu, Zn, Ni, Co, Mn or Mg)
for the catalytic oxidation of toluene and CO (Genty et al., 2012).
Among the synthesized catalysts, Mn-Co is a very promising
candidate for CO and toluene oxidation. Li et al. investigated the
catalytic combustion of toluene over Mn-containing mixed oxides
(Li et al., 2004). They found that the catalytic conversion of toluene
on the Mn-Zr mixed oxides prepared by the reverse microemulsion
method is much higher than that on other mixed oxides of Mn.
They have also reported that the catalytic activity gradually increases with increasing Mn loading. Li et al. also reported the catalytic oxidation of toluene using Fe-Mn, Co-Mn, and Cu-Mn mixed
oxides. Among these metal oxides, the Cu-Mn catalyst provides the
highest activity (Li et al., 2004).
Gennequin et al. demonstrated the doping effects of Co on MgCo-Al hydrotalcite catalysts in the oxidation of toluene (Gennequin
et al., 2010). The addition of cobalt has a benecial effect on the
activity. The reconstruction of the layered structure enhances the
interaction between the cobalt species and the support, which
improves the performance of the catalyst. De Rivas et al. studied the
oxidation of 1,2-dichloroethane over Ce-Zr mixed-oxide catalysts
after treating them with H2SO4 and HNO3 (de Rivas et al., 2013).
Although mixed catalysts treated with HNO3 did not show any

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

change in activity, a signicant increase in the activity was


observed with sulphated catalysts. This increase in the activity is
attributed to the increase in the total acidity and the concentration
of active sites with a moderate/strong acid strength. In addition, the
treatment with H2SO4 does not change the surface area of the
catalyst or the redox properties associated with Ce4/Ce3.
4.3. Catalytic oxidation of mixtures of VOCs
The composition of VOCs signicantly varies depending on their
source. Generally, the industrial ue gas streams contain a mixture
of VOCs rather than a pure compound. The concentration of the
different VOCs mainly depends on the source of the emission. The
physical and chemical properties of these VOCs may also be
different. Therefore, it is very important to develop technology/
catalysts, which can decompose mixtures of VOCs effectively.
Studies on the effects of mixtures on the catalytic oxidations of
VOCs are reported in the open literature. Normally the presence of
other molecules inhibits the destruction of VOCs. Reports of mixtures promoting the destruction of VOCs are rare (Ivanova et al.,
2013). The inhibition effects are mainly due to the competition
among the different species of VOCs and the reaction intermediates
cot et al., 1994; Papaefthimiou et al., 1997;
for adsorption sites (Mare
Tsou et al., 2005). Burgos et al. demonstrated that the catalytic
oxidation rate of 2-propanol over a platinum-based catalyst
signicantly decreases in the presence toluene and methyl ethyl
ketone (Burgos et al., 2002a). They proposed that both toluene and
methyl ethyl ketone compete for active Pt sites, leaving much lower
number of active sites for 2-propanol. Santos et al. reported the
inhibition effects of toluene in the oxidation of ethyl acetatetoluene and ethanol-toluene mixtures (Santos et al., 2011). Inhibition by toluene is more pronounced on ethyl acetate-toluene oxidations. Piotrowska et al. also reported the inhibition effects of
toluene and n-hexane during the oxidation of n-butyl-acetate on
the perovskite catalyst (Musialik-Piotrowska and Syczewska, 2000).
Tsou et al. showed that the oxidation of methylisobutyl ketone over
a Pt/zeolite catalyst can be severely affected due to the presence of
o-xylene (Tsou et al., 2005). The afnity of o-xylene for the sites on
the catalyst causes strong adsorption of o-xylene on the catalyst,
which affects the adsorption/reaction of methylisobutyl ketone
n
~ ez et al., 2002). Ordonez et al. showed that the presence of
(Ordo
n-hexane does not affect the conversion of benzene and toluene.
However, the presence of benzene or toluene inhibits the combustion of hexane, and the aromatic compounds inhibit each other
n
~ ez et al., 2002).
when present together (Ordo
Papaefthimiou et al. studied the oxidation of ethyl acetate and
benzene over Pt/Al2O3 and Pt/TiO2 (W6) catalysts (Papaefthimiou
et al., 1998b). The oxidation of benzene is strongly inhibited by ethyl
acetate in binary benzene-ethyl acetate mixtures, while ethyl acetate oxidation is not inuenced by the presence of benzene, suggesting different interaction mechanisms of benzene and ethyl
acetate with the catalyst.
Barakat et al. reported the oxidation of a mixture of methylethyl
ketone-toluene over a Pd5%V-TiO2 catalyst (Barakat et al., 2014).
They showed that the presence of toluene enhanced the oxidation
performance of the catalyst unlike in other mixtures. As the
oxidation of MEK is easier than toluene, they proposed that the
toluene adsorption is decreased by MEK. They also suggested that
the variations in the surface temperature of the catalyst caused by
adsorption enhances the oxidation of toluene molecules present in
the gaseous stream.
There are some studies reported in the literature on the effects
of other compounds on the oxidation of VOCs. Such conditions are
commonly present in ue gas streams. As water molecules generally compete for the adsorption sites, water can also inhibit the

127

catalytic processes involving VOCs. In some cases water can have


poisoning effects by decreasing the activity of the catalysts (Ivanova
et al.,). For example, the addition of water vapor signicantly decreases the oxidation of propane over Pt- and Pd-based catalysts
cot et al., 1994). Papaefthimiou et al. demonstrated the effects
(Mare
of water vapor on the catalytic destruction of ethyl acetate and
benzene over Pt/Al2O3 and Pt/TiO2 catalysts (Papaefthimiou et al.,
1998b). They found that the oxidation of benzene is drastically
decreased when water was added to the feed streams. Strong
adsorption of water on the metal sites is believed to be responsible
for the decreased oxidation. In fact, the adsorbed water molecules
react with ethyl acetate, converting it into ethanol and acetic acid.
5. Deactivation and regeneration of catalysts
Several factors are responsible for the change in the activity and
selectivity of catalysts with the passage of time, leading to their
deactivation. Catalyst deactivation is a signicant additional
expense as the cost of the catalyst may be as high as 28% of the
operating cost of an oxidizer unit (Gallastegi-Villa et al., 2014).
Bartholomew reviewed the deactivation of catalysts and classied
the deactivation process into the following six groups: fouling or
coking, poisoning, vapor-solid and/or solid-solid reaction, thermal
degradation, crushing, and vapor compound formation
(Bartholomew, 2001). Poisoning is the loss of active sites of the
catalyst due to chemisorption of impurities on the catalyst (Forzatti
and Lietti, 1999; Neyestanaki et al., 2004). A poison can block the
active sites or can alter the activity of the catalyst. In addition, as the
adsorption coefcient of a poison is higher than that of the reactants, the poison can limit the access of the reactants to the active
sites. A catalyst can be deactivated by the loss of active sites due to
the structural modication of the catalyst caused by sintering,
which is a thermally induced process (Neyestanaki et al., 2004).
Coke formation is the result of side reactions taking place on the
catalyst surface when carbonaceous byproducts deactivate the
catalyst either by covering the catalyst surface or by pore blocking.
Mechanisms of coke formation have been described in several reviews (Rostrup-Nielsen and Trimm, 1977; Trimm, 1983).
Catalysts used for the oxidation of VOCs are also deactivated by
additional factors that are normally not commonly encountered in
other catalytic reactions. Phosphorus can penetrate alumina catalysts and form a glassy phase (Libanati et al., 1998). Various
byproducts, which adsorb on titania catalysts, such as alcohol,
aldehyde, and carboxylic acid can also deactivate them. Interaction
of chlorine with catalysts leading to their deactivation is a major
problem in the oxidation of chlorinated compounds. Coke formation and chlorine deposition are the main reasons of catalyst
deactivation in the oxidation of trichloroethylene using zeolites.
Chlorine atoms attack the Bronsted acid sites which causes structural changes of the catalyst (Aranzabal et al., 2012). Almost all
types of catalysts are affected by chlorine deposition (Abbas et al.,
2011). Following are some of the catalysts which can be subjected
to chlorine poisoning: cerium oxide based catalysts (Dai et al.,
2008), supported noble-metal catalysts (Abdullah et al., 2006;
Guillemot et al., 2007), transition metal oxides, vanadia based
_ n
 ski et al., 2002), perovskites (Kieling et al., 1998),
catalysts (Kuazy
and mixed-metal catalysts. Catalyst deactivation can also be due to
water present in almost all exhaust gases and formed in oxidation
reactions. Catalyst deactivation in the ozone-catalytic oxidation of
benzene using MnO2 nanoparticles supported on zeolites is due to
the occupation of active sites and adsorption centers by water.
Deactivation can be suppressed by reducing the amount of water
vapor and increasing the reaction temperature (Teh et al., 2015).
However, water also helps remove the chlorine which can be
deposited on the surface of the catalyst during the oxidation

128

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

lez-Velasco et al., 2000). As wet


reaction of chlorinated VOCs (Gonza
air can remove both coke and chlorine, wet air is more effective in
regenerating catalysts compared to dry air (Gallastegi-Villa et al.,
2014). Intermediates produced during the oxidation are also
another major source of deactivation. Hsieh at al. investigated the
deactivation of Pt catalysts during the oxidation of streams containing wastewater (Hsieh et al., 2002) due to the agglomeration of
Pt clusters. Thevenet et al. investigated the deactivation of the
catalyst in acetylene conversion using TiO2 nanoparticles (Thevenet
et al., 2014). Deposited organic acids are a major source of deactivation of catalysts and they can be regenerated using synthetic air
(Aranzabal et al., 2012, 2014; Gallastegi-Villa et al., 2014). Synthetic
air removed more than 80% of the adsorbed acid and converted it
into CO2. Ihm et al. found that the deactivation during oxidation of
n-hexane using Pd/Al2O3 catalyst is mainly due to the formation of
carbonaceous intermediates (Ihm et al., 2004). Dissociative
adsorption of the reactant causes the deactivation of the catalyst at
low temperature in the oxidation of trichloroethylene over the Co
and Cr oxide catalyst. Oliveira et al. investigated the oxidation of
chlorobenzene and xylene over chromium and Pd-impregnated Alpillored bentonite (Oliveira et al., 2008). The catalyst was deactivated mainly due to the loss of Cr owing to the formation of the
volatile compound CrO2Cl2, decrease in the surface area, and formation of coke. Apart from the catalyst itself, the temperature and
space velocity can also cause catalyst deactivation.
Regeneration of a catalyst depends on the reversibility of the
deactivation process. For example, coke formation can easily be
reversed while sintering is typically irreversible. Some catalyst
poisons can be selectively removed by chemical washing, mechanically or by oxidation (Argyle and Bartholomew, 2015). The
decision to regenerate or discard a catalyst depends mainly on the
rate of deactivation and cost of the catalytic systems. Most of the
available catalysts for the removal of VOCs are expensive and they
are frequently deactivated. Therefore, regeneration of catalysts is of
utmost importance in terms of operation and economics. Several
methods available for the regeneration of catalysts, such as heat
treatment methods, pressure swing, chemical regeneration, oxygen
plasma, ozone injection, pin-to-plate dielectric barrier discharge,
and radio frequency (RF) plasma (HafezKhiabani et al., 2015; Kim
et al., 2007a; Sultana et al., 2015; Zhu et al., 2015). The surface
carbonate species on the deactivated catalyst can be removed using
conventional heat treatment methods. However, the method has
the negative effect of aggregating nanoparticles. (Konova et al.,
2004a, 2004b). Oxygen plasma and ozone injection was found to
be effective in regenerating the Au/TiO2 catalyst (Kim et al., 2007a).
Air is much cheaper and have easier access compared to oxygen as a
discharge gas for plasma regeneration. However, the presence of
nitrogen can cause extra poisoning due to the formation of nitrogen
oxide. Cold plasma of humid air can solve the problem of nitrogen
oxide formation due to the presence of water vapor (Zhu et al.,
2015). In general, modern catalyst regeneration techniques are
replacing conventional heat treatment or mechanical techniques.
6. Conclusion
Oxidation using catalysts is one of the most promising techniques for the removal of VOCs. Toxic VOCs can be oxidized into
CO2, water, and other relatively less harmful compounds over a
catalyst at relatively low temperatures (250e500  C) compared to
thermal oxidation. Operations at low temperatures result in
lowering of the fuel cost and the formation of less dioxins and
noxious products. Catalytic oxidation is most suitable for efuent
streams with a low concentration of VOCs and therefore most
suitable for pollution control due to end-of-pipe VOCs. This article
reviewed various catalytic systems that have been used in the

removal of a range of VOCs. These catalytic systems are based on


noble metal catalysts, metal oxides, and mixed-metal catalysts.
Supported noble-metal catalysts such as Pt, Pd, Au, and Ag are the
most promising due to their high efciency for the removal of VOCs
at low temperature. The activity of the supported noble metals
depends on several factors such as the type and nature of the
support, metal loading, and metal type. Although supported noblemetal catalysts have the highest efciency, they are expensive and
can be deactivated by sintering or poisoning. In addition, noblemetal catalysts alone are not selective enough. Metal-oxide catalysts have several advantages that make them a good alternative to
expensive noble-metal catalysts. Metals oxides are more tolerant to
deactivation by poisoning compared to noble metals. Other advantages include long lifetime, masking tolerance, capability of
regeneration, and availability of a range of metal oxides in different
sizes and shapes. However, the efciency of single-metal-oxide
catalysts in removing different types of VOCs is lower than that of
the noble-metal catalysts. The synergistic effect of two or more
metal oxides can have a higher efciency compared to a singlemetal-oxide catalyst. Even though, catalytic oxidation has achieved considerable success in removing VOCs, it still faces several
challenges. Catalyst deactivation is a major problem associated
with catalytic oxidation. Catalysts can be easily deactivated by
adsorption of water vapor and other intermediates. Future work
should focus on the development of efcient catalytic systems that
can oxidize a range of VOCs at much lower temperatures.
Acknowledgements
This project was funded by the National Plan for Science,
Technology and Innovation (MAARIFAH) of the King Abdulaziz City
for Science and Technology (KACST) through the Science & Technology Unit at the King Fahd University of Petroleum & Minerals
(KFUPM), Kingdom of Saudi Arabia (award number 13-ENV196804).
References
Abbas, N., Hussain, M., Russo, N., Saracco, G., 2011. Studies on the activity and
deactivation of novel optimized TiO2 nanoparticles for the abatement of VOCs.
Chem. Eng. J. 175, 330e340.
Abbasi, Z., Haghighi, M., Fatehifar, E., Saedy, S., 2011. Synthesis and physicochemical
characterizations of nanostructured Pt/Al 2 O 3eCeO 2 catalysts for total
oxidation of VOCs. J. Hazard. Mater. 186, 1445e1454.
Abdullah, A.Z., Abu Bakar, M.Z., Bhatia, S., 2006. Combustion of chlorinated volatile
organic compounds (VOCs) using bimetallic chromium-copper supported on
modied H-ZSM-5 catalyst. J. Hazard. Mater. 129, 39e49.
Abedi, K., Ghorbani-Shahna, F., Jaleh, B., Bahrami, A., Yarahmadi, R., Haddadi, R.,
Gandomi, M., 2015. Decomposition of chlorinated volatile organic compounds
(CVOCs) using NTP coupled with TiO2/GAC, ZnO/GAC, and TiO2eZnO/GAC in a
plasma-assisted catalysis system. J. Electrost. 73, 80e88.
Aguero, F.N., Scian, A., Barbero, B.P., Cads, L.E., 2009. Inuence of the support
treatment on the behavior of MnOx/Al2O3 catalysts used in VOC combustion.
Catal. Lett. 128, 268e280.
Ali, A.M., Daous, M.A., Khamis, A.A., Driss, H., Burch, R., Petrov, L.A., 2015. Strong
synergism between gold and manganese in an AueMn/triple-oxide-support
(TOS) oxidation catalyst. Appl. Catal. A Gen. 489, 24e31.
^rvulescu, V.I., 2007. Ceria-based oxides as supports for
Alifanti, M., Florea, M., Pa
LaCoO3 perovskite; catalysts for total oxidation of VOC. Appl. Catal. B Environ.
70, 400e405.
Alonso, F., Beletskaya, I.P., Yus, M., 2002. Metal-mediated reductive hydrodehalogenation of organic halides. Chem. Rev. 102, 4009e4092.
Alvarez-Montero, M., Gomez-Sainero, L., Mayoral, A., Diaz, I., Baker, R., Rodriguez, J.,
2011. Hydrodechlorination of chloromethanes with a highly stable Pt on activated carbon catalyst. J. Catal. 279, 389e396.
Amann, M., Lutz, M., 2000. The revision of the air quality legislation in the European
Union related to ground-level ozone. J. Hazard. Mater. 78, 41e62.
Andersen, M.E., Clewell, H.J., Bermudez, E., Willson, G.A., Thomas, R.S., 2008.
Genomic signatures and dose-dependent transitions in nasal epithelial responses to inhaled formaldehyde in the rat. Toxicol. Sci. 105, 368e383.
lez-Marcos, M.P., Gonza
lez-Marcos, J.A., Lo
pezAranzabal, A., Pereda-Ayo, B., Gonza
Fonseca, R., Gonz
alez-Velasco, J.R., 2014. State of the art in catalytic oxidation of
chlorinated volatile organic compounds. Chem. Pap. 68, 1169e1186.

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134


lezAranzabal, A., Romero-S
aez, M., Elizundia, U., Gonz
alez-Velasco, J.R., Gonza
Marcos, J.A., 2012. Deactivation of H-zeolites during catalytic oxidation of
trichloroethylene. J. Catal. 296, 165e174.
Argyle, M.D., Bartholomew, C.H., 2015. Heterogeneous catalyst deactivation and
regeneration: a review. Catalysts 5, 145e269.
Asgari, N., Haghighi, M., Shaei, S., 2013. Synthesis and physicochemical characterization of nanostructured Pd/ceria-clinoptilolite catalyst used for p-xylene
abatement from waste gas streams at low temperature. J. Chem. Technol. Biotechnol. 88, 690e703.
Assebban, M., Tian, Z.-Y., El Kasmi, A., Bahlawane, N., Harti, S., Chak, T., 2015.
Catalytic complete oxidation of acetylene and propene over clay versus cordierite honeycomb monoliths without and with chemical vapor deposited cobalt
oxide. Chem. Eng. J. 262, 1252e1259.
Atkinson, R., 1997. Gas-phase tropospheric chemistry of volatile organic compounds: 1. Alkanes and alkenes. J. Phys. Chem. Ref. Data 26, 215e290.
Avgouropoulos, G., Ioannides, T., 2003. Selective CO oxidation over CuO-CeO 2
catalysts prepared via the ureaenitrate combustion method. Appl. Catal. A Gen.
244, 155e167.
Avgouropoulos, G., Oikonomopoulos, E., Kanistras, D., Ioannides, T., 2006. Complete
oxidation of ethanol over alkali-promoted Pt/Al2 O 3 catalysts. Appl. Catal. B
Environ. 65, 62e69.
Baldi, M., Finocchio, E., Milella, F., Busca, G., 1998. Catalytic combustion of C3 hydrocarbons and oxygenates over Mn3 O4. Appl. Catal. B Environ. 16, 43e51.
Barakat, T., Rooke, J., Cousin, R., Lamonier, J.-F., Giraudon, J.-M., Su, B.-L., Siffert, S.,
2014. Investigation of the elimination of VOC mixtures over a Pd-loaded Vdoped TiO2 support. New J. Chem. 38, 2066e2074.
Barakat, T., Rooke, J.C., Franco, M., Cousin, R., Lamonier, J.F., Giraudon, J.M., Su, B.L.,
Siffert, S., 2012. Pd-and/or Au-Loaded Nb-and V-Doped macro-mesoporous TiO2
supports as catalysts for the total oxidation of VOCs. Eur. J. Inorg. Chem. 2012,
2812e2818.
Barbero, B.P., Gamboa, J.A., Cads, L.E., 2006. Synthesis and characterisation of La1
xCaxFeO3 perovskite-type oxide catalysts for total oxidation of volatile organic
compounds. Appl. Catal. B Environ. 65, 21e30.
Bartholomew, C.H., 2001. Mechanisms of catalyst deactivation. Appl. Catal. A Gen.
212, 17e60.
 a
~o, J., Pereira, M., Delgado, J., Figueiredo, J., 2012. Total
Bastos, S., Carabineiro, S., Orf
oxidation of ethyl acetate, ethanol and toluene catalyzed by exotemplated
manganese and cerium oxides loaded with gold. Catal. Today 180, 148e154.
 ~
Bastos, S., Orf
ao, J., Freitas, M., Pereira, M., Figueiredo, J., 2009. Manganese oxide
catalysts synthesized by exotemplating for the total oxidation of ethanol. Appl.
Catal. B Environ. 93, 30e37.
Bedia, J., Rosas, J., Rodrguez-Mirasol, J., Cordero, T., 2010. Pd supported on mesoporous activated carbons with high oxidation resistance as catalysts for toluene
oxidation. Appl. Catal. B Environ. 94, 8e18.


mez-Mendoza, N.-A., Go
 mez-Garca, M.-A.,
 ski, D.,
Behar, S., Go
Swierczy
n
Quignard, F., Tanchoux, N., 2015. Study and modelling of kinetics of the
oxidation of VOC catalyzed by nanosized CueMn spinels prepared via an
alginate route. Appl. Catal. A Gen. 504, 203e210.
Blanco, J., Petre, A., Yates, M., Martin, M., Martin, J., Martin-Luengo, M., 2007. Tailormade high porosity VOC oxidation catalysts prepared by a single-step procedure. Appl. Catal. B Environ. 73, 128e134.
Burch, R., Loader, P., Urbano, F., 1996. Some aspects of hydrocarbon activation on
platinum group metal combustion catalysts. Catal. Today 27, 243e248.
Burgos, N., Paulis, M.A., Antxustegi, M.M., Montes, M., 2002a. Deep oxidation of VOC
mixtures with platinum supported on Al2O3/Al monoliths. Applied Catal. B
Environ. 38, 251e258.
Burgos, N., Paulis, M.a., Mirari Antxustegi, M., Montes, M., 2002b. Deep oxidation of
VOC mixtures with platinum supported on Al2O3/Al monoliths. Appl. Catal. B
Environ. 38, 251e258.
Busca, G., Daturi, M., Finocchio, E., Lorenzelli, V., Ramis, G., Willey, R., 1997. Transition metal mixed oxides as combustion catalysts: preparation, characterization and activity mechanisms. Catal. Today 33, 239e249.
Cant, N.W., Angove, D.E., Patterson, M.J., 1998. The effects of residual chlorine on the
behaviour of platinum group metals for oxidation of different hydrocarbons.
Catal. Today 44, 93e99.
Cant, N.W., Hall, W.K., 1970. Catalytic oxidation: II. Silica supported noble metals for
the oxidation of ethylene and propylene. J. Catal. 16, 220e231.
Cappelletti, M., Frascari, D., Zannoni, D., Fedi, S., 2012. Microbial degradation of
chloroform. Appl. Microbiol. Biotechnol. 96, 1395e1409.
 ~
Carabineiro, S., Chen, X., Konsolakis, M., Psarras, A., Tavares, P., Orf
ao, J., Pereira, M.,
Figueiredo, J., 2015a. Catalytic oxidation of toluene on CeeCo and LaeCo mixed
oxides synthesized by exotemplating and evaporation methods. Catal. Today
244, 161e171.
Carabineiro, S., Chen, X., Martynyuk, O., Bogdanchikova, N., Avalos-Borja, M.,
 a
~o, J., Pereira, M., Figueiredo, J., 2015b. Gold
Pestryakov, A., Tavares, P., Orf
supported on metal oxides for volatile organic compounds total oxidation.
Catal. Today 244, 103e114.
Carabineiro, S.A., Thompson, D.T., 2007. Catalytic applications for gold nanotechnology. Nanocatalysis 377e489. Springer.
Carballo, L.M., Wolf, E.E., 1978. Crystallite size effects during the catalytic oxidation
of propylene on Ptg-Al2O3. J. Catal. 53, 366e373.
Carpentier, J., Lamonier, J., Siffert, S., Zhilinskaya, E., Aboukas, A., 2002. Characterisation of Mg/Al hydrotalcite with interlayer palladium complex for catalytic
oxidation of toluene. Appl. Catal. A Gen. 234, 91e101.
Carrillo, A., Carriazo, J., 2015. Cu and Co oxides supported on halloysite for the total

129

oxidation of toluene. Appl. Catal. B Environ. 164, 443e452.


~ o, M.H., Molina, R., Moreno, S., 2015a. Catalytic oxidation of VOCs on
Castan
MnMgAlOx mixed oxides obtained by auto-combustion. J. Mol. Catal. A Chem.
398, 358e367.
~ o, M.H., Molina, R., Moreno, S., 2015b. Cooperative effect of the CoeMn mixed
Castan
oxides for the catalytic oxidation of VOCs: inuence of the synthesis method.
Appl. Catal. A Gen. 492, 48e59.
Centeno, M., Paulis, M., Montes, M., Odriozola, J., 2005. Catalytic combustion of
volatile organic compounds on gold/titanium oxynitride catalysts. Appl. Catal. B
Environ. 61, 177e183.
Centi, G., 2001. Supported palladium catalysts in environmental catalytic technologies for gaseous emissions. J. Mol. Catal. A Chem. 173, 287e312.
Chang, H., Li, J., Chen, X., Ma, L., Yang, S., Schwank, J.W., Hao, J., 2012. Effect of Sn on
MnO xeCeO 2 catalyst for SCR of NO x by ammonia: enhancement of activity
and remarkable resistance to SO 2. Catal. Commun. 27, 54e57.
Chen, H., Sayari, A., Adnot, A., Larachi, F., 2001. Compositioneactivity effects of
MneCeeO composites on phenol catalytic wet oxidation. Appl. Catal. B Environ.
32, 195e204.
Chen, H., Zhang, H., Yan, Y., 2014a. Fabrication of porous copper/manganese binary
oxides modied ZSM-5 membrane catalyst and potential application in the
removal of VOCs. Chem. Eng. J. 254, 133e142.
Chen, M., Ma, Y., Li, G., Zheng, X., 2008a. Support effect, thermal stability, and
structure feature of toluene combustion catalyst. Catal. Commun. 9, 990e994.
Chen, M., Qi, L., Fan, L., Zhou, R., Zheng, X., 2008b. Zirconium-pillared montmorillonite and their application in supported palladium catalysts for volatile organic
compounds purication. Mater. Lett. 62, 3646e3648.
 a
~o, J., Pereira, M., Figueiredo, J.,
Chen, X., Carabineiro, S., Bastos, S., Tavares, P., Orf
2013. Exotemplated copper, cobalt, iron, lanthanum and nickel oxides for catalytic oxidation of ethyl acetate. J. Environ. Chem. Eng. 1, 795e804.
 a
~o, J., Pereira, M., Figueiredo, J.,
Chen, X., Carabineiro, S., Bastos, S., Tavares, P., Orf
2014b. Catalytic oxidation of ethyl acetate on cerium-containing mixed oxides.
Appl. Catal. A Gen. 472, 101e112.
 a
~o, J., Pereira, M., Figueiredo, J., 2014c.
Chen, X., Carabineiro, S.A., Tavares, P.B., Orf
Catalytic oxidation of ethyl acetate over La-Co and La-Cu oxides. J. Environ.
Chem. Eng. 2, 344e355.
Cho, C.-H., Ihm, S.-K., 2002. Development of new vanadium-based oxide catalysts
for decomposition of chlorinated aromatic pollutants. Environ. Sci. Technol. 36,
1600e1606.
Choudhary, V., Deshmukh, G., 2005. Kinetics of the complete combustion of dilute
propane and methyl ethyl ketone over Cr-doped ZrO2 catalyst. Chem. Eng. Sci.
60, 1575e1581.
Ciuparu, D., Lyubovsky, M.R., Altman, E., Pfefferle, L.D., Datye, A., 2002. Catalytic
combustion of methane over palladium-based catalysts. Catal. Rev. 44,
593e649.
Collins, J., 2001. Tuberculosis in cattle: new perspectives. Tuberculosis 81, 17e21.
Cordi, E.M., ONeill, P.J., Falconer, J.L., 1997a. Transient oxidation of volatile organic
compounds on aCuO/Al2O3 catalyst. Appl. Catal. B Environ. 14, 23e36.
Cordi, E.M., ONeill, P.J., Falconer, J.L., 1997b. Transient oxidation of volatile organic
compounds on aCuO/Al2O3 catalyst. Appl. Catal. B Environ. 14, 23e36.
Corma, A., Garcia, H., 2008. Supported gold nanoparticles as catalysts for organic
reactions. Chem. Soc. Rev. 37, 2096e2126.
Dai, Q., Wang, X., Lu, G., 2007. Low-temperature catalytic destruction of chlorinated
VOCs over cerium oxide. Catal. Commun. 8, 1645e1649.
Dai, Q., Wang, X., Lu, G., 2008. Low-temperature catalytic combustion of trichloroethylene over cerium oxide and catalyst deactivation. Appl. Catal. B Environ.
81, 192e202.
de Almeida, M.P., Martins, L., Carabineiro, S., Lauterbach, T., Rominger, F., Hashmi, A.,
Pombeiro, A., Figueiredo, J., 2013. Homogeneous and heterogenised new gold Cscorpionate complexes as catalysts for cyclohexane oxidation. Catal. Sci. Technol. 3, 3056e3069.
rrez-Ortiz, J.I., Lo
pez-Fonseca, R., Gonz
de Rivas, B., Gutie
alez-Velasco, J.R., 2006.
Analysis of the simultaneous catalytic combustion of chlorinated aliphatic
pollutants and toluene over ceria-zirconia mixed oxides. Appl. Catal. A Gen. 314,
54e63.
pez-Fonseca, R., Jime
nez-Gonza
lez, C., Gutie
rrez-Ortiz, J.I., 2011.
de Rivas, B., Lo
Synthesis, characterisation and catalytic performance of nanocrystalline Co 3 O
4 for gas-phase chlorinated VOC abatement. J. Catal. 281, 88e97.
pez-Fonseca, R., Jime
nez-Gonz
rrez-Ortiz, J.I., 2012.
de Rivas, B., Lo
alez, C., Gutie
Highly active behaviour of nanocrystalline Co3O4 from oxalate nanorods in the
oxidation of chlorinated short chain alkanes. Chem. Eng. J. 184, 184e192.
 pez-Fonseca, R., Gutie
rrez-Ortiz, J.,
de Rivas, B., Sampedro, C., Garca-Real, M., Lo
2013. Promoted activity of sulphated Ce/Zr mixed oxides for chlorinated VOC
oxidative abatement. Appl. Catal. B Environ. 129, 225e235.
g e
, P., Pinard, L., Magnoux, P., Guisnet, M., 2000. Catalytic oxidation of volatile
De
organic compounds: II. Inuence of the physicochemical characteristics of Pd/
HFAU catalysts on the oxidation of o-xylene. Appl. Catal. B Environ. 27, 17e26.
Delaigle, R., Debecker, D.P., Bertinchamps, F., Gaigneaux, E.M., 2009. Revisiting the
behaviour of vanadia-based catalysts in the abatement of (chloro)-aromatic
pollutants: towards an integrated understanding. Top. Catal. 52, 501e516.
Delimaris, D., Ioannides, T., 2008a. VOC oxidation over MnOxeCeO2 catalysts prepared by a combustion method. Appl. Catal. B Environ. 84, 303e312.
Delimaris, D., Ioannides, T., 2008b. VOC oxidation over MnOxeCeO2 catalysts prepared by a combustion method. Appl. Catal. B Environ. 84, 303e312.
Delimaris, D., Ioannides, T., 2009. VOC oxidation over CuOeCeO2 catalysts prepared
by a combustion method. Appl. Catal. B Environ. 89, 295e302.

130

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

Diehl, F., Barbier, J., Duprez, D., Guibard, I., Mabilon, G., 2010. Catalytic oxidation of
heavy hydrocarbons over Pt/Al2O3. Inuence of the structure of the molecule on
its reactivity. Appl. Catal. B Environ. 95, 217e227.
Doggali, P., Teraoka, Y., Mungse, P., Shah, I.K., Rayalu, S., Labhsetwar, N., 2012.
Combustion of volatile organic compounds over CueMn based mixed oxide
type catalysts supported on mesoporous Al2O3, TiO2 and ZrO2. J. Mol. Catal. A
Chem. 358, 23e30.
Doherty, R.E., 2000. A history of the production and use of carbon tetrachloride,
tetrachloroethylene, trichloroethylene and 1, 1, 1etrichloroethane in the United
States: part 1-historical background; carbon tetrachloride and tetrachloroethylene. Environ. Forensics 1, 69e81.
s, N., Motuzas, J., Simone, S., Algieri, C.,
Drobek, M., Figoli, A., Santoro, S., Navascue
Gaeta, N., Querze, L., Trotta, A., 2015. PVDF-MFI mixed matrix membranes as
VOCs adsorbers. Microporous Mesoporous Mater 207, 126e133.
Einaga, H., Yamamoto, S., Maeda, N., Teraoka, Y., 2015. Structural analysis of manganese oxides supported on SiO2 for benzene oxidation with ozone. Catal.
Today 242, 287e293.
Everaert, K., Baeyens, J., 2004. Catalytic combustion of volatile organic compounds.
J. Hazard. Mater. 109, 113e139.
Finlayson-Pitts, B.J., Pitts, J.N., 1997. Tropospheric air pollution: ozone, airborne
toxics, polycyclic aromatic hydrocarbons, and particles. Science 276, 1045e1051.
Finocchio, E., Baldi, M., Pistarino, C., Romezzano, G., Bregani, F., Toledo, G., 2000.
A study of the abatement of VOC over V2O5-WO3-TiO2 and alternative SCR
catalysts. Catal. Today 59, 261e268.
Finocchio, E., Busca, G., 2001. Characterization and hydrocarbon oxidation activity
of coprecipitated mixed oxides Mn3O4/Al2O3. Catal. Today 70, 213e225.
Forzatti, P., Lietti, L., 1999. Catalyst deactivation. Catal. Today 52, 165e181.
ndez-Alonso, M.D., Tudela, D., Coronado, J.M., Soria, J., 2008. PhoFresno, F., Herna
tocatalytic degradation of toluene over doped and coupled (Ti, M) O2 (M Sn or
Zr) nanocrystalline oxides: inuence of the heteroatom distribution on deactivation. Appl. Catal. B Environ. 84, 598e606.
Fujishima, A., Zhang, X., 2006. Titanium dioxide photocatalysis: present situation
and future approaches. C. R. Chim. 9, 750e760.
ez, M., Gonz
lezGallastegi-Villa, M., Aranzabal, A., Romero-Sa
alez-Marcos, J., Gonza
Velasco, J., 2014. Catalytic activity of regenerated catalyst after the oxidation of
1, 2-dichloroethane and trichloroethylene. Chem. Eng. J. 241, 200e206.
Galvita, V.V., Filez, M., Poelman, H., Bliznuk, V., Marin, G.B., 2014. The role of
different types of CuO in CuOeCeO2/Al2O3 for total oxidation. Catal. Lett. 144,
32e43.
Gandhe, A.R., Rebello, J.S., Figueiredo, J., Fernandes, J., 2007. Manganese oxide OMS2 as an effective catalyst for total oxidation of ethyl acetate. Appl. Catal. B Environ. 72, 129e135.
Gangwal, S., Mullins, M., Spivey, J., Caffrey, P., Tichenor, B., 1988. Kinetics and
selectivity of deep catalytic oxidation of n-hexane and benzene. Appl. Catal. 36,
231e247.
nchez-Royo, J.F., Murillo, R., Mastral, A.M., Aranda, A.,
Garcia, T., Agouram, S., Sa
zquez, I., Dejoz, A., Solsona, B., 2010. Deep oxidation of volatile organic
Va
compounds using ordered cobalt oxides prepared by a nanocasting route. Appl.
Catal. A Gen. 386, 16e27.
 Taylor, S.H., 2006. Total
 s, D., Linares-Solano, A.,
Garcia, T., Solsona, B., Cazorla-Amoro
oxidation of volatile organic compounds by vanadium promoted palladiumtitania catalysts: comparison of aromatic and polyaromatic compounds. Appl.
Catal. B Environ. 62, 66e76.
Garcia, T., Solsona, B., Taylor, S.H., 2004. The oxidative destruction of hydrocarbon
volatile organic compounds using palladiumevanadiaetitania catalysts. Catal.
Lett. 97, 99e103.
lin, P., Primet, M., 2002. Complete oxidation of methane at low temperature over
Ge
noble metal based catalysts: a review. Appl. Catal. B Environ. 39, 1e37.
Gennequin, C., Barakat, T., Tidahy, H., Cousin, R., Lamonier, J.-F., Aboukas, A.,
Siffert, S., 2010. Use and observation of the hydrotalcite memory effect for
VOC oxidation. Catal. Today 157, 191e197.
Genty, E., Cousin, R., Capelle, S., Gennequin, C., Siffert, S., 2012. Catalytic oxidation of
toluene and CO over nanocatalysts derived from hydrotalcite-like compounds
(X62 Al23): effect of the bivalent cation. Eur. J. Inorg. Chem. 2012,
2802e2811.
Giraudon, J.-M., Nguyen, T., Leclercq, G., Siffert, S., Lamonier, J.-F., Aboukas, A.,
Vantomme, A., Su, B.-L., 2008. Chlorobenzene total oxidation over palladium
supported on ZrO2, TiO2 nanostructured supports. Catal. Today 137, 379e384.
Gluhoi, A.C., Bogdanchikova, N., Nieuwenhuys, B.E., 2005. The effect of different
types of additives on the catalytic activity of Au/Al2O3 in propene total oxidation: transition metal oxides and ceria. J. Catal. 229, 154e162.
lez-Velasco, J., Aranzabal, A., Lo
pez-Fonseca, R., Ferret, R., Gonz
Gonza
alez-Marcos, J.,
2000. Enhancement of the catalytic oxidation of hydrogen-lean chlorinated
VOCs in the presence of hydrogen-supplying compounds. Appl. Catal. B Environ.
24, 33e43.
Gorte, R.J., 2010. Ceria in catalysis: from automotive applications to the wateregas
shift reaction. AlChE J. 56, 1126e1135.
Guillemot, M., Mijoin, J., Mignard, S., Magnoux, P., 2007. Mode of zeolite catalysts
deactivation during chlorinated VOCs oxidation. Appl. Catal. A Gen. 327,
211e217.
rrez-Ortiz, J.I., De Rivas, B., Lo
pez-Fonseca, R., Gonz
Gutie
alez-Velasco, J.R., 2006.
Catalytic purication of waste gases containing VOC mixtures with Ce/Zr solid
solutions. Appl. Catal. B Environ. 65, 191e200.
HafezKhiabani, N., Fathi, S., Shokri, B., Hosseini, S.I., 2015. A novel method for
decoking of PteSn/Al 2 O 3 in the naphtha reforming process using RF and pin-

to-plate DBD plasma systems. Appl. Catal. A Gen. 493, 8e16.


Han, W., Deng, J., Xie, S., Yang, H., Dai, H., Au, C.T., 2014. Gold supported on iron
oxide nanodisk as efcient catalyst for the removal of toluene. Ind. Eng. Chem.
Res. 53, 3486e3494.
Hanfeng, L., Ying, Z., Huang, H., Zhang, B., Yinfei, C., 2011. In-situ synthesis of
monolithic Cu-Mn-Ce/cordierite catalysts towards VOCs combustion. J. Rare
Earths 29, 855e860.
Haruta, M., 1997a. Novel catalysis of gold deposited on metal oxides. Catal. Surv.
Asia 1, 61e73.
Haruta, M., 1997b. Size-and support-dependency in the catalysis of gold. Catal.
Today 36, 153e166.
Haruta, M., Yamada, N., Kobayashi, T., Iijima, S., 1989. Gold catalysts prepared by
coprecipitation for low-temperature oxidation of hydrogen and of carbon
monoxide. J. Catal. 115, 301e309.
Hashmi, A.S.K., Lothschtz, C., Ackermann, M., Doepp, R., Anantharaman, S.,
Marchetti, B., Bertagnolli, H., Rominger, F., 2010. Gold catalysis: in situ EXAFS
study of homogeneous oxidative esterication. Chem. A Eur. J. 16, 8012e8019.
He, C., Li, J., Zhang, X., Yin, L., Chen, J., Gao, S., 2012. Highly active Pd-based catalysts
with hierarchical pore structure for toluene oxidation: catalyst property and
reaction determining factor. Chem. Eng. J. 180, 46e56.
He, C., Li, Q., Li, P., Wang, Y., Zhang, X., Cheng, J., Hao, Z., 2010. Templated silica with
increased surface area and expanded microporosity: synthesis, characterization, and catalytic application. Chem. Eng. J. 162, 901e909.
He, C., Xu, B.-T., Shi, J.-W., Qiao, N.-L., Hao, Z.-P., Zhao, J.-L., 2015. Catalytic
destruction of chlorobenzene over mesoporous ACeOx (A Co, Cu, Fe, Mn, or
Zr) composites prepared by inorganic metal precursor spontaneous precipitation. Fuel Process. Technol. 130, 179e187.
He, C., Yu, Y., Yue, L., Qiao, N., Li, J., Shen, Q., Yu, W., Chen, J., Hao, Z., 2014. Lowtemperature removal of toluene and propanal over highly active mesoporous
CuCeOx catalysts synthesized via a simple self-precipitation protocol. Appl.
Catal. B Environ. 147, 156e166.
Heracleous, E., Lee, A., Wilson, K., Lemonidou, A., 2005. Investigation of Ni-based
alumina-supported catalysts for the oxidative dehydrogenation of ethane to
ethylene: structural characterization and reactivity studies. J. Catal. 231,
159e171.
Heracleous, E., Lemonidou, A., 2006. NieNbeO mixed oxides as highly active and
selective catalysts for ethene production via ethane oxidative dehydrogenation.
Part I: characterization and catalytic performance. J. Catal. 237, 162e174.
Hermia, J., Vigneron, S., 1993. Catalytic incineration for odour abatement and VOC
destruction. Catal. Today 17, 349e358.
Heynderickx, P.M., Thybaut, J.W., Poelman, H., Poelman, D., Marin, G.B., 2010. The
total oxidation of propane over supported Cu and Ce oxides: a comparison of
single and binary metal oxides. J. Catal. 272, 109e120.
Hosseini, M., Siffert, S., Cousin, R., Aboukas, A., Hadj-Sadok, Z., Su, B.-L., 2009. Total
oxidation of VOCs on Pd and/or Au supported on TiO2/ZrO2 followed by
operando DRIFT. C. R. Chim. 12, 654e659.
Hsieh, C.-C., Lee, J.-F., Liu, Y.-R., Chang, J.-R., 2002. Structural investigation of catalyst
deactivation of Pt/SDB for catalytic oxidation of VOC-containing wastewater.
Waste Manag. 22, 739e745.
Hu, C., 2011. Catalytic combustion kinetics of acetone and toluene over
Cu0.13Ce0.87Oy catalyst. Chem. Eng. J. 168, 1185e1192.
Hu, C., Zhu, Q., Jiang, Z., Zhang, Y., Wang, Y., 2008. Preparation and formation
mechanism of mesoporous CuOeCeO2 mixed oxides with excellent catalytic
performance for removal of VOCs. Microporous Mesoporous Mater. 113,
427e434.
Huang, B., Lei, C., Wei, C., Zeng, G., 2014. Chlorinated volatile organic compounds
(Cl-VOCs) in environmentdsources, potential human health impacts, and current remediation technologies. Environ. Int. 71, 118e138.
Huang, H., Liu, Y., Tang, W., Chen, Y., 2008a. Catalytic activity of nanometer La1
xSrxCoO3(x 0, 0.2) perovskites towards VOCs combustion. Catal. Commun. 9,
55e59.
Huang, H., Xu, Y., Feng, Q., Leung, D.Y.C., 2015. Low temperature catalytic oxidation
of volatile organic compounds: a review. Catal. Sci. Technol. 5, 2649e2669.
Huang, S., Zhang, C., He, H., 2008b. Complete oxidation of o-xylene over Pd/Al2O3
catalyst at low temperature. Catal. Today 139, 15e23.
Huang, Y.C., Luo, C.H., Yang, S., Lin, Y.C., Chuang, C.Y., 2010. Improved removal of
indoor volatile organic compounds by activated carbon ber lters calcined
with copper oxide catalyst. Clean Soil Air Water 38, 993e997.
Hussain, M., Ceccarelli, R., Marchisio, D., Fino, D., Russo, N., Geobaldo, F., 2010.
Synthesis, characterization, and photocatalytic application of novel TiO2
nanoparticles. Chem. Eng. J. 157, 45e51.
Hutchings, G.J., Taylor, S.H., 1999. Designing oxidation catalysts. Catal. Today 49,
105e113.
Ihm, S.-K., Jun, Y.-D., Kim, D.-C., Jeong, K.-E., 2004. Low-temperature deactivation
and oxidation state of Pd/g-Al2O3 catalysts for total oxidation of n-hexane.
Catal. Today 93, 149e154.
Ihsan, B., 1995. Thermochemical Data of Pure Substances and 934, 587.
rez, A., Centeno, M., Odriozola, J.A., Structured catalysts for volatile
Ivanova, S., Pe
organic compound removal. New and future developments in catalysis. Catal.
Remediat. Environ. Concerns, 233e256.
 Odriozola, J.A., 2013. Structured catalysts for
rez, A., Centeno, M.A.,
Ivanova, S., Pe
volatile organic compound removal. New Future Dev. Catal. Catal. Remediat.
Environ. Concerns 233e256.
Jamalzadeh, Z., Haghighi, M., Asgari, N., 2013. Synthesis, physicochemical characterizations and catalytic performance of Pd/carbon-zeolite and Pd/carbon-CeO2

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134


nanocatalysts used for total oxidation of xylene at low temperatures. Front.
Environ. Sci. Eng. 7, 365e381.
Jeong, M.-G., Park, E.J., Jeong, B., Kim, D.H., Kim, Y.D., 2014. Toluene combustion over
NiO nanoparticles on mesoporous SiO 2 prepared by atomic layer deposition.
Chem. Eng. J. 237, 62e69.
Ji, B., Shao, F., Hu, G., Zheng, S., Zhang, Q., Xu, Z., 2009. Adsorption of methyl tertbutyl ether (MTBE) from aqueous solution by porous polymeric adsorbents.
J. Hazard. Mater. 161, 81e87.
Jones, J., Ross, J.R., 1997. The development of supported vanadia catalysts for the
combined catalytic removal of the oxides of nitrogen and of chlorinated hydrocarbons from ue gases. Catal. Today 35, 97e105.
Joung, H.-J., Kim, J.-H., Oh, J.-S., You, D.-W., Park, H.-O., Jung, K.-W., 2014. Catalytic
oxidation of VOCs over CNT-supported platinum nanoparticles. Appl. Surf. Sci.
290, 267e273.
Jntgen, H., 1986. Activated carbon as catalyst support: a review of new research
results. Fuel 65, 1436e1446.
Kang, M., Park, E.D., Kim, J.M., Yie, J.E., 2007. Manganese oxide catalysts for NOx
reduction with NH 3 at low temperatures. Appl. Catal. A Gen. 327, 261e269.
Khan, F.I., Kr Ghoshal, A., 2000. Removal of volatile organic compounds from
polluted air. J. Loss Prev. process Ind. 13, 527e545.
Kieling, D., Schneider, R., Kraak, P., Haftendorn, M., Wendt, G., 1998. Perovskitetype oxidesecatalysts for the total oxidation of chlorinated hydrocarbons. Appl.
Catal. B Environ. 19, 143e151.
, M., Ogata, A., Futamura, S., 2007a. Catalyst regeneration
Kim, H., Tsubota, S., Date
and activity enhancement of Au/TiO 2 by atmospheric pressure nonthermal
plasma. Appl. Catal. A Gen. 329, 93e98.
Kim, S.C., 2002. The catalytic oxidation of aromatic hydrocarbons over supported
metal oxide. J. Hazard. Mater. 91, 285e299.
Kim, S.C., Nahm, S.W., Shim, W.G., Lee, J.W., Moon, H., 2007b. Inuence of physicochemical treatments on spent palladium based catalyst for catalytic oxidation
of VOCs. J. Hazard. Mater. 141, 305e314.
Kim, S.C., Park, Y.-K., Nah, J.W., 2014. Property of a highly active bimetallic catalyst
based on a supported manganese oxide for the complete oxidation of toluene.
Powder Technol. 266, 292e298.
Kim, S.C., Shim, W.G., 2009. Properties and performance of Pd based catalysts for
catalytic oxidation of volatile organic compounds. Appl. Catal. B Environ. 92,
429e436.
Kim, S.C., Shim, W.G., 2010. Catalytic combustion of VOCs over a series of manganese oxide catalysts. Appl. Catal. B Environ. 98, 180e185.
lez, A.,
Klett, C., Duten, X., Tieng, S., Touchard, S., Jestin, P., Hassouni, K., Vega-Gonza
2014. Acetaldehyde removal using an atmospheric non-thermal plasma combined with a packed bed: role of the adsorption process. J. Hazard. Mater. 279,
356e364.
Koodziej, A., ojewska, J., Tyczkowski, J., Jodowski, P., Redzynia, W., Iwaniszyn, M.,
Zapotoczny, S., Kustrowski, P., 2012. Coupled engineering and chemical
approach to the design of a catalytic structured reactor for combustion of VOCs:
cobalt oxide catalyst on knitted wire gauzes. Chem. Eng. J. 200, 329e337.
Konova, P., Naydenov, A., Tabakova, T., Mehandjiev, D., 2004a. Deactivation of
nanosize gold supported on zirconia in CO oxidation. Catal. Commun. 5,
537e542.
Konova, P., Naydenov, A., Venkov, C., Mehandjiev, D., Andreeva, D., Tabakova, T.,
2004b. Activity and deactivation of Au/TiO2 catalyst in CO oxidation. J. Mol.
Catal. A Chem. 213, 235e240.
Konsolakis, M., Carabineiro, S.A., Tavares, P.B., Figueiredo, J.L., 2013. Redox properties and VOC oxidation activity of Cu catalysts supported on Ce 1 xSmxOd
mixed oxides. J. Hazard. Mater. 261, 512e521.
, J., Machovi
, L., Jira
tova
, K.,
Kovanda, F., Rojka, T., Dobesova
c, V., Bezdi
cka, P., Obalova
Grygar, T., 2006. Mixed oxides obtained from Co and Mn containing layered
double hydroxides: preparation, characterization, and catalytic properties.
J. Solid State Chem. 179, 812e823.
Krishnamoorthy, S., Baker, J.P., Amiridis, M.D., 1998. Catalytic oxidation of 1, 2dichlorobenzene over V2O5/TiO2-based catalysts. Catal. Today 40, 39e46.
Krishnamoorthy, S., Rivas, J.A., Amiridis, M.D., 2000. Catalytic oxidation of 1, 2dichlorobenzene over supported transition metal oxides. J. Catal. 193, 264e272.
Kujawa, J., Cerneaux, S., Kujawski, W., 2015. Removal of hazardous volatile organic
compounds from water by vacuum pervaporation with hydrophobic ceramic
membranes. J. Membr. Sci. 474, 11e19.
_ n
 ski, M., Van Ommen, J., Trawczyn
 ski, J., Walendziewski, J., 2002. Catalytic
Kuazy
combustion of trichloroethylene over TiO2-SiO2 supported catalysts. Appl. Catal.
B Environ. 36, 239e247.
Lahousse, C., Bernier, A., Grange, P., Delmon, B., Papaefthimiou, P., Ioannides, T.,
Verykios, X., 1998a. Evaluation of g-MnO2 as a VOC removal catalyst: comparison with a noble metal catalyst. J. Catal. 178, 214e225.
Lahousse, C., Bernier, A., Grange, P., Delmon, B., Papaefthimiou, P., Ioannides, T.,
Verykios, X., 1998b. Evaluation of g-MnO2 as a VOC removal catalyst: comparison with a noble metal catalyst. J. Catal. 178, 214e225.
thivier, C., Potvin, C., Louis, C., 2010.
Lakshmanan, P., Delannoy, L., Richard, V., Me
Total oxidation of propene over Au/xCeO2-Al2O3 catalysts: inuence of the CeO2
loading and the activation treatment. Appl. Catal. B Environ. 96, 117e125.
Lamaita, L., Peluso, M.A., Sambeth, J.E., Thomas, H., Mineli, G., Porta, P., 2005.
A theoretical and experimental study of manganese oxides used as catalysts for
VOCs emission reduction. Catal. Today 107, 133e138.
Larsson, P.-O., Andersson, A., 1998. Complete oxidation of CO, ethanol, and ethyl
acetate over copper oxide supported on titania and ceria modied titania.
J. Catal. 179, 72e89.

131

Leson, G., Winer, A.M., 1991. Bioltration: an innovative air pollution control
technology for VOC emissions. J. Air Waste Manag. Assoc. 41, 1045e1054.
Li, H., Lu, G., Dai, Q., Wang, Y., Guo, Y., Guo, Y., 2010. Hierarchical organization and
catalytic activity of high-surface-area mesoporous ceria microspheres prepared
via hydrothermal routes. ACS Appl. Mater. Interfaces 2, 838e846.
Li, J., Zhao, P., Liu, S., 2014a. SnO xeMnO xeTiO 2 catalysts with high resistance to
chlorine poisoning for low-temperature chlorobenzene oxidation. Appl. Catal. A
Gen. 482, 363e369.
Li, J.W., Pan, K.L., Yu, S.J., Yan, S.Y., Chang, M.B., 2014b. Removal of formaldehyde
over MnxCe1 x O2 catalysts: thermal catalytic oxidation versus ozone catalytic
oxidation. J. Environ. Sci. 26, 2546e2553.
Li, S., Wang, H., Li, W., Wu, X., Tang, W., Chen, Y., 2015. Effect of Cu substitution on
promoted benzene oxidation over porous CuCo-based catalysts derived from
layered double hydroxide with resistance of water vapor. Appl. Catal. B Environ.
166e167, 260e269.
Li, T.-Y., Chiang, S.-J., Liaw, B.-J., Chen, Y.-Z., 2011. Catalytic oxidation of benzene over
CuO/Ce1 x MnxO2 catalysts. Appl. Catal. B Environ. 103, 143e148.
Li, W., Chu, W., Zhuang, M., Hua, J., 2004. Catalytic oxidation of toluene on
Mnecontaining mixed oxides prepared in reverse microemulsions. Catal. Today
93, 205e209.
Li, W., Wang, J., Gong, H., 2009. Catalytic combustion of VOCs on non-noble metal
catalysts. Catal. Today 148, 81e87.
Libanati, C., Ullenius, D., Pereira, C., 1998. Silica deactivation of bead VOC catalysts.
Appl. Catal. B Environ. 15, 21e28.
Lin, L.-Y., Wang, C., Bai, H., 2015. A comparative investigation on the lowtemperature catalytic oxidation of acetone over porous aluminosilicatesupported cerium oxides. Chem. Eng. J. 264, 835e844.
Lin, S.S., Chang, D.J., Wang, C.-H., Chen, C.C., 2003. Catalytic wet air oxidation of
phenol by CeO2 catalystdeffect of reaction conditions. Water Res. 37, 793e800.
Lin, S.S., Chen, C.L., Chang, D.J., Chen, C.C., 2002. Catalytic wet air oxidation of
phenol by various CeO2 catalysts. Water Res. 36, 3009e3014.
Liotta, L., 2010. Catalytic oxidation of volatile organic compounds on supported
noble metals. Appl. Catal. B Environ. 100, 403e412.
Liotta, L., Ousmane, M., Di Carlo, G., Pantaleo, G., Deganello, G., Marc, G.,
Retailleau, L., Giroir-Fendler, A., 2008. Total oxidation of propene at low temperature over Co3O4eCeO2 mixed oxides: role of surface oxygen vacancies and
bulk oxygen mobility in the catalytic activity. Appl. Catal. A General 347, 81e88.
Liotta, L.F., 2014. New trends in gold catalysts. Catalysts 4, 299e304.
Liotta, L.F., Wu, H., Pantaleo, G., Venezia, A.M., 2013. Co3O4 nanocrystals and
Co3O4eMOx binary oxides for CO, CH4 and VOC oxidation at low temperatures:
a review. Catal. Sci. Technol. 3, 3085e3102.
Liu, G., Li, J., Yang, K., Tang, W., Liu, H., Yang, J., Yue, R., Chen, Y., 2015. Effects of
cerium incorporation on the catalytic oxidation of benzene over ame-made
perovskite La1 xCe xMnO3 catalysts. Particuology 19, 60e68.
Liu, Q., Wang, L.-C., Chen, M., Cao, Y., He, H.-Y., Fan, K.-N., 2009. Dry citrateprecursor synthesized nanocrystalline cobalt oxide as highly active catalyst
for total oxidation of propane. J. Catal. 263, 104e113.
Liu, W., Flytzanistephanopoulos, M., 1995. Total oxidation of carbon-monoxide and
methane over transition metal uorite oxide composite catalysts II. Catalyst
characterization and reaction-kinetics. J. Catal. 153, 317e332.
Liu, W., Sarom, A.F., Flytzani-Stephanopoulos, M., 1994. Complete oxidation of
carbon monoxide and methane over metal-promoted uorite oxide catalysts.
Chem. Eng. Sci. 49, 4871e4888.
Luecken, D., Hutzell, W., Strum, M., Pouliot, G., 2012. Regional sources of atmospheric formaldehyde and acetaldehyde, and implications for atmospheric
modeling. Atmos. Environ. 47, 477e490.
Luo, J., Zhang, Q., Huang, A., Suib, S.L., 2000. Total oxidation of volatile organic
compounds with hydrophobic cryptomelane-type octahedral molecular sieves.
Microporous Mesoporous Mater. 35e36, 209e217.
Luo, M.-F., He, M., Xie, Y.-L., Fang, P., Jin, L.-Y., 2007. Toluene oxidation on Pd catalysts supported by CeO2eY2O3 washcoated cordierite honeycomb. Appl. Catal. B
Environ. 69, 213e218.
Main, D.M., Hogan, T.J., 1983. Health effects of low-level exposure to formaldehyde.
J. Occup. Environ. Med. 25, 896e900.
Makshina, E., Nesterenko, N., Siffert, S., Zhilinskaya, E., Aboukais, A., Romanovsky, B.,
2008. Methanol oxidation on LaCo mixed oxide supported onto MCM-41 molecular sieve. Catal. Today 131, 427e430.
Malaguarnera, G., Cataudella, E., Giordano, M., Nunnari, G., Chisari, G.,
Malaguarnera, M., 2012. Toxic hepatitis in occupational exposure to solvents.
World J. Gastroenterol. e WJG 18, 2756.
dar, F., Moreno-Castilla, C., Pe
rez-Cadenas, A., 2004. Catalytic comMaldonado-Ho
bustion of toluene on platinum-containing monolithic carbon aerogels. Appl.
Catal. B Environ. 54, 217e224.
cot, P., Fakche, A., Kellali, B., Mabilon, G., Prigent, P., Barbier, J., 1994. Propane
Mare
and propene oxidation over platinum and palladium on alumina: effects of
chloride and water. Appl. Catal. B Environ. 3, 283e294.
Masui, T., Imadzu, H., Matsuyama, N., Imanaka, N., 2010. Total oxidation of toluene
on Pt/CeO2-ZrO2-Bi2O3/gamma-Al2O3 catalysts prepared in the presence of
polyvinyl pyrrolidone. J. Hazard. Mater. 176, 1106e1109.
Masui, T., Imadzu, H., Terada, A., Imanaka, N., Complete oxidation of ethylene at
temperatures below 100 C over a Pt/Ce0.64Zr0.16Bi0.20O1.90/g-Al2O3 catalyst.
Milt, V., Lombardo, E., Ulla, M., 2002. Stability of cobalt supported on ZrO2 catalysts
for methane combustion. Appl. Catal. B Environ. 37, 63e73.
Miranda-Trevino, J.C., Coles, C.A., 2003. Kaolinite properties, structure and inuence
of metal retention on pH. Appl. Clay Sci. 23, 133e139.

132

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

n
~ ez, S., Vega, A., Dez, F.V., 2007. Oxidation of triMiranda, B., Daz, E., Ordo
chloroethene over metal oxide catalysts: kinetic studies and correlation with
adsorption properties. Chemosphere 66, 1706e1715.
Mitsui, T., Tsutsui, K., Matsui, T., Kikuchi, R., Eguchi, K., 2008. Catalytic abatement of
acetaldehyde over oxide-supported precious metal catalysts. Appl. Catal. B
Environ. 78, 158e165.
Molina, M.J., Rowland, F.S., 1974. Stratospheric sink for chlorouoromethanes chlorine atom catalyzed destruction of ozone. B. Am. Meteorol. Soc. 55, 491.
Morales, M.R., Agero, F.N., Cadus, L.E., 2013. Catalytic combustion of n-hexane over
alumina supported MneCueCe catalysts. Catal. Lett. 143, 1003e1011.
Morales, M.R., Barbero, B.P., Cads, L.E., 2006. Total oxidation of ethanol and propane over Mn-Cu mixed oxide catalysts. Appl. Catal. B Environ. 67, 229e236.
Morales, M.R., Barbero, B.P., Cads, L.E., 2008. Evaluation and characterization of
MneCu mixed oxide catalysts for ethanol total oxidation: inuence of copper
content. Fuel 87, 1177e1186.
Moretti, E.C., 2002. Reduce VOC and HAP emissions. Chem. Eng. Prog. 98, 30e40.
Musialik-Piotrowska, A., 2007. Destruction of trichloroethylene (TCE) and trichloromethane (TCM) in the presence of selected VOCs over Pt-Pd-based
catalyst. Catal. Today 119, 301e304.
Musialik-Piotrowska, A., Landmesser, H., 2008. Noble metal-doped perovskites for
the oxidation of organic air pollutants. Catal. Today 137, 357e361.
Musialik-Piotrowska, A., Syczewska, K., 2000. Combustion of volatile organic
compounds in two-component mixtures over monolithic perovskite catalysts.
Catal. Today 59, 269e278.
Neyestanaki, A.K., Klingstedt, F., Salmi, T., Murzin, D.Y., 2004. Deactivation of
postcombustion catalysts, a review. Fuel 83, 395e408.
aho, S., Laitinen, T., Koivikko, N.N., Brahmi, R., Gaa
lova
, J., Matejova, L.,
Ojala, S., Pitka
Kucherov, A., P
aiv
arinta, S., Hirschmann, C., 2011. Catalysis in VOC abatement.
Top. Catal. 54, 1224e1256.
Okumura, K., Kobayashi, T., Tanaka, H., Niwa, M., 2003. Toluene combustion over
palladium supported on various metal oxide supports. Appl. Catal. B Environ.
44, 325e331.
Oliveira, L., Lago, R.M., Fabris, J., Sapag, K., 2008. Catalytic oxidation of aromatic
VOCs with Cr or Pd-impregnated Al-pillared bentonite: byproduct formation
and deactivation studies. Appl. Clay Sci. 39, 218e222.
Olsen, E., Nielsen, F., 2001. Predicting vapour pressures of organic compounds from
their chemical structure for classication according to the VOCDirective and
risk assessment in general. Molecules 6, 370e389.
n
~ ez, S., Bello, L., Sastre, H., Rosal, R., Dez, F.V., 2002. Kinetics of the deep
Ordo
oxidation of benzene, toluene, n-hexane and their binary mixtures over a
platinum on g-alumina catalyst. Appl. Catal. B Environ. 38, 139e149.

_ 2009. Catalytic combustion of toluene over Mn, Fe


Ozelik,
Z., Soylu, G.S.P., Boz, I.,
and Co-exchanged clinoptilolite support. Chem. Eng. J. 155, 94e100.
Ozturk, B., Yilmaz, D., 2006. Absorptive removal of volatile organic compounds from
ue gas streams. Process Saf. Environ. Prot. 84, 391e398.
Padilla, A.M., Corella, J., Toledo, J.M., 1999. Total oxidation of some chlorinated hydrocarbons with commercial chromia based catalysts. Appl. Catal. B Environ. 22,
107e121.
Palacio, L., Silva, E., Catal~
ao, R., Silva, J., Hoyos, D., Ribeiro, F., Ribeiro, M., 2008.
Performance of supported catalysts based on a new copper vanadate-type
precursor for catalytic oxidation of toluene. J. Hazard. Mater. 153, 628e634.
Papaefthimiou, P., Ioannides, T., Verykios, X., 1998a. Catalytic incineration of volatile
organic compounds present in industrial waste streams. Appl. Therm. Eng. 18,
1005e1012.
Papaefthimiou, P., Ioannides, T., Verykios, X.E., 1997. Combustion of nonhalogenated volatile organic compounds over group VIII metal catalysts. Appl.
Catal. B Environ. 13, 175e184.
Papaefthimiou, P., Ioannides, T., Verykios, X.E., 1998b. Performance of doped Pt/TiO2
(W6) catalysts for combustion of volatile organic compounds (VOCs). Appl.
Catal. B Environ. 15, 75e92.
Papaefthimiou, P., Ioannides, T., Verykios, X.E., 1998c. Performance of doped Pt/TiO2
(W6) catalysts for combustion of volatile organic compounds (VOCs). Appl.
Catal. B Environ. 15, 75e92.
Papaefthimiou, P., Ioannides, T., Verykios, X.E., 1999. VOC removal: investigation of
ethylacetate oxidation over supported Pt catalysts. Catal. Today 54, 81e92.
Parida, K., Samal, A., 1999. Catalytic combustion of volatile organic compounds on
Indian Ocean manganese nodules. Appl. Catal. A Gen. 182, 249e256.
Patdhanagul, N., Rangsriwatananon, K., Siriwong, K., Hengrasmee, S., 2012. Combined modication of zeolite NaY by phenyl trimethyl ammonium bromide and
potassium for ethylene gas adsorption. Microporous Mesoporous Mater. 153,
30e34.
Patdhanagul, N., Srithanratana, T., Rangsriwatananon, K., Hengrasmee, S., 2010.
Ethylene adsorption on cationic surfactant modied zeolite NaY. Microporous
Mesoporous Mater. 131, 97e102.
Paulis, M.a., Peyrard, H., Montes, M., 2001. Inuence of chlorine on the activity and
stability of Pt/Al2O3Catalysts in the complete oxidation of toluene. J. Catal. 199,
30e40.
Pei, J., Han, X., Lu, Y., 2015. Performance and kinetics of catalytic oxidation of
formaldehyde over copper manganese oxide catalyst. Build. Environ. 84,
134e141.
Peluso, M.A., Pronsato, E., Sambeth, J.E., Thomas, H.J., Busca, G., 2008. Catalytic
combustion of ethanol on pure and alumina supported K-Mn oxides: an IR and
ow reactor study. Appl. Catal. B Environ. 78, 73e79.
Pena, M., Fierro, J., 2001. Chemical structures and performance of perovskite oxides.
Chem. Rev. 101, 1981e2018.

Peng, J., Wang, S., 2007. Performance and characterization of supported metal
catalysts for complete oxidation of formaldehyde at low temperatures. Appl.
Catal. B Environ. 73, 282e291.
Periyat, P., Baiju, K., Mukundan, P., Pillai, P., Warrier, K., 2008. High temperature
stable mesoporous anatase TiO 2 photocatalyst achieved by silica addition.
Appl. Catal. A Gen. 349, 13e19.
Petrosius, S.C., Drago, R.S., Young, V., Grunewald, G.C., 1993. Low-temperature
decomposition of some halogenated hydrocarbons using metal oxide/porous
carbon catalysts. J. Am. Chem. Soc. 115, 6131e6137.
rrez, M., Pina, M., Herguido, J., 2007. Preparation and characterPicasso, G., Gutie
ization of Ce-Zr and Ce-Mn based oxides for n-hexane combustion: application
to catalytic membrane reactors. Chem. Eng. J. 126, 119e130.
Pinna, F., 1998. Supported metal catalysts preparation. Catal. Today 41, 129e137.
Pope, D., Walker, D., Moss, R., 1978. Evaluation of platinum-honeycomb catalysts for
the destructive oxidation of low concentrations of odorous compounds in air.
Atmos. Environ. (1967) 12, 1921e1927.
Puertolas, B., Solsona, B., Agouram, S., Murillo, R., Mastral, A.M., Aranda, A.,
Taylor, S.H., Garcia, T., 2010. The catalytic performance of mesoporous cerium
oxides prepared through a nanocasting route for the total oxidation of naphthalene. Appl. Catal. B Environ. 93, 395e405.
Qi, G., Yang, R.T., Chang, R., 2004. MnO x-CeO2 mixed oxides prepared by coprecipitation for selective catalytic reduction of NO with NH 3 at low temperatures. Appl. Catal. B Environ. 51, 93e106.
Qu, Z., Gao, K., Fu, Q., Qin, Y., 2014. Low-temperature catalytic oxidation of toluene
over nanocrystal-like MneCo oxides prepared by two-step hydrothermal
method. Catal. Commun. 52, 31e35.
Radic, N., Grbic, B., Terlecki-Baricevic, A., 2004. Kinetics of deep oxidation of nhexane and toluene over Pt/Al2O3 catalysts: platinum crystallite size effect.
Appl. Catal. B Environ. 50, 153e159.
Ramesh, K., Chen, L., Chen, F., Liu, Y., Wang, Z., Han, Y.-F., 2008. Re-investigating the
CO oxidation mechanism over unsupported MnO, Mn2O3 and MnO2 catalysts.
Catal. Today 131, 477e482.
Rao, G.R., Sahu, H.R., Mishra, B.G., 2003. Surface and catalytic properties of
CueCeeO composite oxides prepared by combustion method. Colloid Surf. A
220, 261e269.
Rao, T., Shen, M., Jia, L., Hao, J., Wang, J., 2007. Oxidation of ethanol over MneCeeO
and MneCeeZreO complex compounds synthesized by solegel method. Catal.
Commun. 8, 1743e1747.
Rezlescu, N., Rezlescu, E., Popa, P.D., Doroftei, C., Ignat, M., 2015. Some nanograined
ferrites and perovskites for catalytic combustion of acetone at low temperature.
Ceram. Int. 41, 4430e4437.
Rodhe, H., 1990. A comparison of the contribution of various gases to the greenhouse effect. Science 1217e1219.
Rooke, J.C., Barakat, T., Brunet, J., Li, Y., Finol, M.F., Lamonier, J.-F., Giraudon, J.-M.,
Cousin, R., Siffert, S., Su, B.L., 2015. Hierarchically nanostructured porous group
V b metal oxides from alkoxide precursors and their role in the catalytic
remediation of VOCs. Appl. Catal. B Environ. 162, 300e309.
Rossner, A., Knappe, D.R., 2008. MTBE adsorption on alternative adsorbents and
packed bed adsorber performance. Water Res. 42, 2287e2299.
Rostrup-Nielsen, J., Trimm, D.L., 1977. Mechanisms of carbon formation on nickelcontaining catalysts. J. Catal. 48, 155e165.
Roth, W.J., Nachtigall, P., Morris, R.E., Cejka, J.I., 2014. Two-dimensional zeolites:
current status and perspectives. Chem. Rev. 114, 4807e4837.
Rotter, H., Landau, M., Carrera, M., Goldfarb, D., Herskowitz, M., 2004. High surface
area chromia aerogel efcient catalyst and catalyst support for ethylacetate
combustion. Appl. Catal. B Environ. 47, 111e126.
Rotter, H., Landau, M., Herskowitz, M., 2005. Combustion of chlorinated VOC on
nanostructured chromia aerogel as catalyst and catalyst support. Environ. Sci.
Technol. 39, 6845e6850.
Rui, Z., Chen, C., Lu, Y., Ji, H., 2014. Anodic alumina supported Pt catalyst for total
oxidation of trace toluene. Chin. J. Chem. Eng. 22, 882e887.
Salagre, P., Fierro, J., Medina, F., Sueiras, J., 1996. Characterization of nickel species
on several g-alumina supported nickel samples. J. Mol. Catal. A Chem. 106,
125e134.
 ~
Santos, V., Pereira, M., Orf
ao, J., Figueiredo, J., 2011. Mixture effects during the
oxidation of toluene, ethyl acetate and ethanol over a cryptomelane catalyst.
J. Hazard. Mater. 185, 1236e1240.
 a
~o, J.J., Figueiredo, J.L.,
Santos, V.P., Carabineiro, S.A., Tavares, P.B., Pereira, M.F., Orf
2010. Oxidation of CO, ethanol and toluene over TiO2 supported noble metal
catalysts. Appl. Catal. B Environ. 99, 198e205.
Saqer, S.M., Kondarides, D.I., Verykios, X.E., 2011. Catalytic oxidation of toluene over
binary mixtures of copper, manganese and cerium oxides supported on gAl2O3. Appl. Catal. B Environ. 103, 275e286.
, S., Liotta, L.F., 2012. Supported gold catalysts for the total oxidation of volatile
Scire
organic compounds. Appl. Catal. B Environ. 125, 222e246.
, S., Minico
, S., Crisafulli, C., Galvagno, S., 2001a. Catalytic combustion of volatile
Scire
organic compounds over group IB metal catalysts on Fe2O3. Catal. Commun. 2,

229e232.
, S., Minico
, S., Crisafulli, C., Galvagno, S., 2001b. Catalytic combustion of volScire
atile organic compounds over group IB metal catalysts on Fe2O3. Catal. Commun. 2, 229e232.
Scire, S., Minico, S., Crisafulli, C., Satriano, C., Pistone, A., 2003. Catalytic combustion
of volatile organic compounds on gold/cerium oxide catalysts. Appl. Catal. B
Environ. 40, 43e49.
Sedjame, H.-J., Fontaine, C., Lafaye, G., Barbier Jr., J., 2014. On the promoting effect of

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134


the addition of ceria to platinum based alumina catalysts for VOCs oxidation.
Appl. Catal. B Environ. 144, 233e242.
Shah, R., Thonon, B., Benforado, D., 2000. Opportunities for heat exchanger applications in environmental systems. Appl. Therm. Eng. 20, 631e650.
Shahna, F.G., Golbabaei, F., Hamedi, J., Mahjub, H., Darabi, H.R., Shahtaheri, S.J., 2010.
Treatment of benzene, toluene and xylene contaminated air in a bioactive foam
emulsion reactor. Chin. J. Chem. Eng. 18, 113e121.
Shen, Y., Yang, X., Wang, Y., Zhang, Y., Zhu, H., Gao, L., Jia, M., 2008. The states of gold
species in CeO2 supported gold catalyst for formaldehyde oxidation. Appl. Catal.
B Environ. 79, 142e148.
Silva, A.M., Marques, R.R., Quinta-Ferreira, R.M., 2004. Catalysts based in cerium
oxide for wet oxidation of acrylic acid in the prevention of environmental risks.
Appl. Catal. B Environ. 47, 269e279.
Sinha, A.K., Suzuki, K., 2007. Novel mesoporous chromium oxide for VOCs elimination. Appl. Catal. B Environ. 70, 417e422.
Sleiman, M., Conchon, P., Ferronato, C., Chovelon, J.-M., 2009. Photocatalytic
oxidation of toluene at indoor air levels (ppbv): towards a better assessment of
conversion, reaction intermediates and mineralization. Appl. Catal. B Environ.
86, 159e165.
 n, P., Herna
ndez, S., Demicol, B., Nieto, J.M.L., 2012. Oxidative
Solsona, B., Concepcio
dehydrogenation of ethane over NiOeCeO2 mixed oxides catalysts. Catal. Today
180, 51e58.
Solsona, B., Davies, T.E., Garcia, T., V
azquez, I., Dejoz, A., Taylor, S.H., 2008. Total
oxidation of propane using nanocrystalline cobalt oxide and supported cobalt
oxide catalysts. Appl. Catal. B Environ. 84, 176e184.
 n, E., Dejoz, A.M., Va
zquez, I., Agouram, S., Davies, T.E.,
Solsona, B., Garcia, T., Aylo
Taylor, S.H., 2011. Promoting the activity and selectivity of high surface area
NieCeeO mixed oxides by gold deposition for VOC catalytic combustion. Chem.
Eng. J. 175, 271e278.
Song, K.S., Klvana, D., Kirchnerova, J., 2001. Kinetics of propane combustion over La
0.66 Sr 0.34 Ni 0.3 Co 0.7 O 3 perovskite. Appl. Catal. A Gen. 213, 113e121.

_ 2010. Total oxidation of toluene over metal oxides


Soylu, G.S.P., Ozelik,
Z., Boz, I.,
supported on a natural clinoptilolite-type zeolite. Chem. Eng. J. 162, 380e387.
Spivey, J.J., 1987. Complete catalytic oxidation of volatile organics. Ind. Eng. Chem.
Res. 26, 2165e2180.
Sue-aok, N., Srithanratana, T., Rangsriwatananon, K., Hengrasmee, S., 2010. Study of
ethylene adsorption on zeolite NaY modied with group I metal ions. Appl. Surf.
Sci. 256, 3997e4002.
Suib, S.L., 2013. New and Future Developments in Catalysis: Catalysis for Remediation and Environmental Concerns. Newnes.
Sultana, S., Vandenbroucke, A.M., Leys, C., De Geyter, N., Morent, R., 2015. Abatement of VOCs with alternate adsorption and plasma-assisted regeneration: a
review. Catalysts 5, 718e746.
Sun, H., Liu, Z., Chen, S., Quan, X., 2015. The role of lattice oxygen on the activity and
selectivity of the OMS-2 catalyst for the total oxidation of toluene. Chem. Eng. J.
270, 58e65.
Tabakova, T., Ilieva, L., Petrova, P., Venezia, A., Avdeev, G., Zanella, R., Karakirova, Y.,
2015. Complete benzene oxidation over mono and bimetallic AuePd catalysts
supported on Fe-modied ceria. Chem. Eng. J. 260, 133e141.
Tang, W., Wu, X., Li, D., Wang, Z., Liu, G., Liu, H., Chen, Y., 2014a. Oxalate route for
promoting activity of manganese oxide catalysts in total VOCs oxidation: effect
of calcination temperature and preparation method. J. Mater. Chem. A 2,
2544e2554.
Tang, W., Wu, X., Li, S., Li, W., Chen, Y., 2014b. Porous MneCo mixed oxide nanorod
as a novel catalyst with enhanced catalytic activity for removal of VOCs. Catal.
Commun. 56, 134e138.
Tang, X., Li, Y., Huang, X., Xu, Y., Zhu, H., Wang, J., Shen, W., 2006. MnO xeCeO2
mixed oxide catalysts for complete oxidation of formaldehyde: effect of preparation method and calcination temperature. Appl. Catal. B Environ. 62,
265e273.
Teh, L., Triwahyono, S., Jalil, A., Mukti, R., Aziz, M., Shishido, T., 2015. Mesoporous
ZSM5 having both intrinsic acidic and basic sites for cracking and methanation.
Chem. Eng. J. 270, 196e204.
Thevenet, F., Guillard, C., Rousseau, A., 2014. Acetylene photocatalytic oxidation
using continuous ow reactor: gas phase and adsorbed phase investigation,
assessment of the photocatalyst deactivation. Chem. Eng. J. 244, 50e58.
Tichenor, B.A., Palazzolo, M.A., 1987. Destruction of volatile organic compounds via
catalytic incineration. Environ. Prog. 6, 172e176.
Tidahy, H., Hosseni, M., Siffert, S., Cousin, R., Lamonier, J.-F., Aboukas, A., Su, B.-L.,
Giraudon, J.-M., Leclercq, G., 2008. Nanostructured macro-mesoporous zirconia
impregnated by noble metal for catalytic total oxidation of toluene. Catal. Today
137, 335e339.
Tidahy, H., Siffert, S., Wyrwalski, F., Lamonier, J.-F., Aboukas, A., 2007. Catalytic
activity of copper and palladium based catalysts for toluene total oxidation.
Catal. Today 119, 317e320.
Trimm, D., 1983. Catalyst design for reduced coking (review). Appl. Catal. 5,
263e290.
Trimm, D., Stanislaus, A., 1986. The control of pore size in alumina catalyst supports:
a review. Appl. Catal. 21, 215e238.
Trinh, Q.H., Lee, S.B., Mok, Y.S., 2015. Removal of ethylene from air stream by
adsorption and plasma-catalytic oxidation using silver-based bimetallic catalysts supported on zeolite. J. Hazard. Mater. 285, 525e534.
Trovarelli, A., 1996. Catalytic properties of ceria and CeO2-containing materials.
Catal. Rev. 38, 439e520.
Trovarelli, A., de Leitenburg, C., Boaro, M., Dolcetti, G., 1999. The utilization of ceria

133

in industrial catalysis. Catal. Today 50, 353e367.


Tseng, T.K., Wang, L., Ho, C.T., Chu, H., 2010. The destruction of dichloroethane over
a g-alumina supported manganese oxide catalyst. J. Hazard. Mater. 178,
1035e1040.
Tsou, J., Magnoux, P., Guisnet, M., Orfao, J., Figueiredo, J.L., 2005. Catalytic oxidation
of volatile organic compounds: oxidation of methyl-isobutyl-ketone over Pt/
zeolite catalysts. Appl. Catal. B Environ. 57, 117e123.
U.S. Department of Health and Human Services, 1993. Hazardous Substances Data
Bank (HSDB, Online Database). National Toxicology Information Program. National Library of Medicine, Bethesda, MD.
U.S. Environmental Protection Agency, 1990. Updated Health Effects Assessment for
Methyl Ethyl Ketone. EPA/600/8e89/093. Environmental Criteria and Assessment Ofce, Ofce of Health and Environmental Assessment, Ofce of Research
and Development, Cincinnati, OH.
Van der Vaart, D., Vatvuk, W., Wehe, A., 1991. Thermal and catalytic incinerators for
the control of VOCs. J. Air Waste Manag. Assoc. 41, 92e98.
 Lozano-Castello
 , D., Cazorla-Amoro
s, D.,
Varela-Ganda, F.J., Berenguer-Murcia, A.,
Sellick, D.R., Taylor, S.H., 2013. Total oxidation of naphthalene using palladium
nanoparticles supported on BETA, ZSM-5, SAPO-5 and alumina powders. Appl.
Catal. B Environ. 129, 98e105.
Vasile, A., Bratan, V., Hornoiu, C., Caldararu, M., Ionescu, N.I., Yuzhakova, T.,
 2013. Electrical and catalytic properties of ceriumetin mixed oxides
dey, A.,
Re
in CO depollution reaction. Appl. Catal. B Environ. 140, 25e31.
Wang, C.-H., Lin, S.-S., 2004. Preparing an active cerium oxide catalyst for the catalytic incineration of aromatic hydrocarbons. Appl. Catal. A Gen. 268, 227e233.
Wang, L., Sakurai, M., Kameyama, H., 2008. Catalytic oxidation of dichloromethane
and toluene over platinum alumite catalyst. J. Hazard. Mater. 154, 390e395.
Wang, X., Zheng, Y., Lin, J., 2013a. Highly dispersed MneCe mixed oxides supported
on carbon nanotubes for low-temperature NO reduction with NH 3. Catal.
Commun. 37, 96e99.
Wang, Y., Zhang, C., Liu, F., He, H., 2013b. Well-dispersed palladium supported on
ordered mesoporous Co3O4 for catalytic oxidation of o-xylene. Appl. Catal. B
Environ. 142, 72e79.
Wang, Y., Zhang, C., Yu, Y., Yue, R., He, H., 2015. Ordered mesoporous and bulk Co3O4
supported Pd catalysts for catalytic oxidation of o-xylene. Catal. Today 242,
294e299.
Waters, R., Weimer, J., Smith, J., 1994. An investigation of the activity of coprecipitated gold catalysts for methane oxidation. Catal. Lett. 30, 181e188.
Wenxiang, T., Xiaofeng, W., Gang, L., Shuangde, L., Dongyan, L., Wenhui, L., Yunfa, C.,
2015. Preparation of hierarchical layer-stacking Mn-Ce composite oxide for
catalytic total oxidation of VOCs. J. Rare Earths 33, 62e69.
Wu, J.C.-S., Chang, T.-Y., 1998. VOC deep oxidation over Pt catalysts using hydrophobic supports. Catal. Today 44, 111e118.
Wu, J.C.-S., Lin, Z.-A., Tsai, F.-M., Pan, J.-W., 2000. Low-temperature complete
oxidation of BTX on Pt/activated carbon catalysts. Catal. Today 63, 419e426.
Wyrwalski, F., Lamonier, J.-F., Perez-Zurita, M., Siffert, S., Aboukas, A., 2006. Inuence of the ethylenediamine addition on the activity, dispersion and reducibility of cobalt oxide catalysts supported over ZrO2 for complete VOC oxidation.
Catal. Lett. 108, 87e95.
Wyrwalski, F., Lamonier, J.-F., Siffert, S., Aboukas, A., 2007. Additional effects of
cobalt precursor and zirconia support modications for the design of efcient
VOC oxidation catalysts. Appl. Catal. B Environ. 70, 393e399.
Xia, Y., Dai, H., Jiang, H., Deng, J., He, H., Au, C.T., 2009. Mesoporous chromia with
ordered three-dimensional structures for the complete oxidation of toluene
and ethyl acetate. Environ. Sci. Technol. 43, 8355e8360.
Xia, Y., Dai, H., Jiang, H., Zhang, L., 2010. Three-dimensional ordered mesoporous
cobalt oxides: highly active catalysts for the oxidation of toluene and methanol.
Catal. Commun. 11, 1171e1175.
Xiaoyuan, J., Guanglie, L., Renxian, Z., Jianxin, M., Yu, C., Xiaoming, Z., 2001. Studies
of pore structure, temperature-programmed reduction performance, and
micro-structure of CuO/CeO2 catalysts. Appl. Surf. Sci. 173, 208e220.
Xin, Y., Yang, B., Wang, H., Anderson, S., Law, C., 2015. Kinetics of catalytic oxidation
of ethylene over palladium oxide. Proc. Combust. Inst. 35, 2233e2240.
Yan, Y., Wang, L., Zhang, H., 2014. Catalytic combustion of volatile organic compounds over Co/ZSM-5 coated on stainless steel bers. Chem. Eng. J. 255,
195e204.
Yang, P., Shi, Z., Yang, S., Zhou, R., 2015a. High catalytic performances of CeO2eCrOx
catalysts for chlorinated VOCs elimination. Chem. Eng. Sci. 126, 361e369.
Yang, P., Yang, S., Shi, Z., Meng, Z., Zhou, R., 2015b. Deep oxidation of chlorinated
VOCs over CeO2-based transition metal mixed oxide catalysts. Appl. Catal. B
Environ. 162, 227e235.
Yazawa, Y., Yoshida, H., Takagi, N., Komai, S.-I., Satsuma, A., Hattori, T., 1998.
Oxidation state of palladium as a factor controlling catalytic activity of Pd/
SiO2eAl2O 3 in propane combustion. Appl. Catal. B Environ. 19, 261e266.
Yim, S.D., Chang, K.-H., Nam, I.-S., 2001. Deactivation of chromium oxide catalyst for
the removal of perchloroethylene (PCE). Stud. Surf. Sci. Catal. 139, 173e180.
Yose, L., Haghighi, M., Allahyari, S., Shokrani, R., Ashkriz, S., 2015. Abatement of
toluene from polluted air over Mn/ClinoptiloliteeCeO2 nanopowder: impregnation vs. ultrasound assisted synthesis with various Mn-loading. Adv. Powder
Technol. 26, 602e611.
Yu, C., Crump, D., 1998. A review of the emission of VOCs from polymeric materials
used in buildings. Build. Environ. 33, 357e374.
Yu, D., Liu, Y., Wu, Z., 2010. Low-temperature catalytic oxidation of toluene over
mesoporous MnOxeCeO2/TiO2 prepared by solegel method. Catal. Commun. 11,
788e791.

134

M.S. Kamal et al. / Atmospheric Environment 140 (2016) 117e134

Yuan, H.-K., Ren, J., Ma, X.-H., Xu, Z.-L., 2011. Dehydration of ethyl acetate aqueous
solution by pervaporation using PVA/PAN hollow ber composite membrane.
Desalination 280, 252e258.
Yuranov, I., Dunand, N., Kiwi-Minsker, L., Renken, A., 2002. Metal grids with highporous surface as structured catalysts: preparation, characterization and activity in propane total oxidation. Appl. Catal. B Environ. 36, 183e191.
Zadaka-Amir, D., Nasser, A., Nir, S., Mishael, Y.G., 2012. Removal of methyl tertiarybutyl ether (MTBE) from water by polymerezeolite composites. Microporous
Mesoporous Mater. 151, 216e222.
Zhang, Q., Zhao, L., Teng, B., Xie, Y., Yue, L., 2008. Pd/Ce0.8Zr0.2O2Substrate monolithic catalyst for toluene catalytic combustion. Chin. J. Catal. 29, 373e378.
Zhang, X.H., Liu, Q.L., Xiong, Y., Zhu, A.M., Chen, Y., Zhang, Q.G., 2009. Pervaporation
dehydration of ethyl acetate/ethanol/water azeotrope using chitosan/poly (vinyl pyrrolidone) blend membranes. J. Membr. Sci. 327, 274e280.
Zheng, X.-C., Wu, S.-H., Wang, S.-P., Wang, S.-R., Zhang, S.-M., Huang, W.-P., 2005.
The preparation and catalytic behavior of copperecerium oxide catalysts for
low-temperature carbon monoxide oxidation. Appl. Catal. A Gen. 283, 217e223.
Zhou, G., He, X., Liu, S., Xie, H., Fu, M., 2015. Phenyl VOCs catalytic combustion on
supported CoMn/AC oxide catalyst. J. Ind. Eng. Chem. 21, 932e941.

Zhou, K., Zhang, Q.G., Han, G.L., Zhu, A.M., Liu, Q.L., 2013. Pervaporation of watereethanol and methanoleMTBE mixtures using poly (vinyl alcohol)/cellulose
acetate blended membranes. J. Membr. Sci. 448, 93e101.
Zhu, B., Li, X.-S., Liu, J.-L., Liu, J.-B., Zhu, X., Zhu, A.-M., 2015. In-situ regeneration of
Au nanocatalysts by atmospheric-pressure air plasma: signicant contribution
of water vapor. Appl. Catal. B Environ. 179, 69e77.
Zhu, Z., Wu, R.-J., 2015. The degradation of formaldehyde using a Pt@ TiO2 nanoparticles in presence of visible light irradiation at room temperature. J. Taiwan
Inst. Chem. Eng. 50, 276e281.
pe, A., Birringer, R., 2002. Temperature-programmed reaction
Zimmer, P., Tscho
spectroscopy of ceria-and Cu/ceria-supported oxide catalyst. J. Catal. 205,
339e345.
Zimowska, M., Michalik-Zym, A., Janik, R., Machej, T., Gurgul, J., Socha, R.,
 ski, J., Serwicka, E., 2007. Catalytic combustion of toluene over mixed
Podobin
CueMn oxides. Catal. Today 119, 321e326.
Zou, L., Luo, Y., Hooper, M., Hu, E., 2006. Removal of VOCs by photocatalysis process
using adsorption enhanced TiO2eSiO2 catalyst. Chem. Eng. Process. Process
Intensif. 45, 959e964.

You might also like