Вы находитесь на странице: 1из 10

Environmental Science and Pollution Research

https://doi.org/10.1007/s11356-018-1305-y

REDUCING AIR AND SOIL POLLUTION IN CHINA: ISSUES IN ENVIRONMENTAL TECHNOLOGIES AND
MANAGEMENT

Photocatalytic removal of SO2 using natural zeolite modified by TiO2


and polyoxypropylene surfactant
Nasibeh Amini 1 & Mohsen Soleimani 1 & Nourollah Mirghaffari 1

Received: 25 October 2017 / Accepted: 15 January 2018


# Springer-Verlag GmbH Germany, part of Springer Nature 2018

Abstract
Air pollution due to emission of various hazardous gases such as SO2 into the atmosphere and its control is an important
environmental issue. Application of photocatalysts is considered as a suitable process to control the gaseous pollutants. In this
study, the efficiency of clinoptilolite as a natural zeolite (Ze) modified by TiO2 (Ze-Ti) and a polymeric surfactant
polyoxypropylene (Ze-Ti-POP) for removal of SO2 was investigated. The nanocomposites were characterized by SEM, EDX,
and BET analyses. The photocatalytic oxidation experiments of SO2 by the nanocomposites and natural zeolite were done under
UV irradiation with initial SO2 concentration of 500 ppm in a photoreactor. The effects of different factors including reaction
time, catalyst dose, UV irradiation intensity, humidity content, and calcination temperature and dose of TiO2 were studied. The
modification of clinoptilolite by TiO2 and POP increased considerably the BET specific surface area of the nanocomposites. The
results showed that maximum removal efficiencies of SO2 by Ze-Ti and Ze-Ti-POP under the optimum experimental conditions
were 82.1 and 87.4%, respectively. Adsorption kinetics data well fitted with the Langmuir–Hinshelwood model. Moreover,
reusing of nanocomposites after three regeneration cycles indicated that application of Ze-Ti and Ze-Ti-POP nanocomposites
could be a promising approach for SO2 removal.

Keywords Air pollution . Photocatalyst . Nanocomposites . Clinoptilolite . Sorption . TiO2

Introduction atmosphere and is a precursor to acid rain and stratospheric


ozone destruction and is harmful for human health through
Increase of fossil fuel combustion due to industrial activities asthma and respiratory diseases (Niu et al. 2011; Su et al.
and transportation has resulted in an excessive release of ni- 2013). Therefore, removal of SO2 from the atmosphere has
trogen oxides (NOx) and sulfur dioxide (SO2) into the atmo- been a significant challenge, particularly in developing
sphere in recent years. These pollutants lead to formation of countries.
secondary pollutants such as photochemical smog and ozone There are several technologies available for removal of
in the presence of sunlight and could adversely affect urban air SO2 emissions. Some of these technologies that are used for
quality. SO2 is well known as one of the noxious gases in the removal of SO2 require complex systems and high cost
implementing. Compared to other techniques, photocatalytic
Responsible editor: Suresh Pillai oxidation using the titanium dioxide (TiO2) nanoparticles has
been shown a successful approach in air pollution control
Electronic supplementary material The online version of this article
(Anpo and Kamat 2010). TiO2 photocatalysts have been in-
(https://doi.org/10.1007/s11356-018-1305-y) contains supplementary
material, which is available to authorized users. tensively studied and widely applied to environmental purifi-
cation in recent years. Having high specific surface area, fa-
* Mohsen Soleimani vorable chemical properties, high stability, low cost, and high
m.soleimani@cc.iut.ac.ir efficiency, they are effective for controlling air pollutants
(Anpo and Kamat 2010). Furthermore, mechanism of TiO2
1
Department of Natural Resources, Isfahan University of Technology, photoreaction only occur under ultraviolet (UV) irradiation,
Isfahan, Iran a drawback for any application where the UV region only
Environ Sci Pollut Res

accounts for ~4% of the total solar energy. However, mecha- et al. (2012) synthesized montmorillonite-TiO2 composites in
nism of TiO2 photoreaction occurs at the presence of UV the presence of POP and evaluated photocatalytic oxidation of
irradiation with large band-gap energy of 3.2 eV (λ methylene blue in water. According to their results, presence of
≥365 nm) (Anpo and Kamat 2010). Generally, photocatalytic POP into the structure of montmorillonite-TiO2 composites
oxidation of pollutants in the gas phase is more efficient than could significantly increase the porosity and surface area of
the liquid phase (Shang et al. 2002). However, it has also been the composites (Chen et al. 2012).
reported that the homogeneous and heterogeneous photooxi- Zeolite-TiO2 (Ze-Ti) composites have been used for degra-
dation reactions of SO2 gas occur upon UV irradiation in the dation of volatile organic compounds (Jansson et al. 2015),
absence and presence of TiO2, respectively (Shang et al. NOx removal (Hashimoto et al. 2001), and H2S removal and
2002). A lot of studies were conducted on using of TiO2 methane production as well (Hao et al. 2016). The photocat-
photocatalysts to reduce air pollution in 1990 (Zhao et al. alytic behavior of the composite materials consisted of miner-
2009). Afterwards, Japanese researchers investigated photo- al substrate and TiO2 was investigated for NOx oxidation
catalytic oxidation of NOx using TiO2 photocatalysts in the which showed different oxidation rates depending on the type
ambient atmosphere (Ibusuki and Takeuchi 1994). of substrates (Todorova et al. 2014). However, it seems that
Photocatalytic oxidation of SO2 highly depends on some im- there is a lack of information about photocatalytic removal of
portant factors such as UV irradiation intensity, type of TiO2 SO2 using Ze-Ti nanocomposites. Sulfur dioxide is one of the
supporting materials, and ambient humidity (Ao et al. 2004; main air pollutants in metropolitan cities in Iran because of
Kachina 2008). It has been reported that the presence of H2O industrial activities and diesel fuel consumption by vehicles.
molecules on the surface of TiO2 plays an important role on Among the natural zeolites, clinoptilolite are abundant and
photocatalytic oxidation of pollutants. So that, photocatalytic low-cost resources which are widely distributed in different
activity of TiO2 could be properly maintained (Ibusuki and regions of Iran and have a high potential for water, soil, and air
Takeuchi 1994; Shang et al. 2002). pollution control (Hamidpour et al. 2010; Malekian et al.
In recent years, many researchers reported that the effective- 2011). Therefore, the aim of this study was modification of
ness of supported photocatalysts could improve photocatalytic clinoptilolite obtained from the Semnan Mine in Iran as a
activity of pure TiO2 (Kachina 2008). However, some studies natural zeolite with TiO2 nanoparticles in the presence of a
have focused on various methods for increasing TiO2 efficien- polymeric surfactant (i.e., polyoxypropylene) for removal of
cy in removal of environmental pollutants. One of these impor- SO2 in a laboratory experiment. Besides, the factors affecting
tant approaches is incorporation of TiO2 nanoparticles with the removal efficiency of SO2 were investigated to find the
porous materials such as silica gel, activated carbon, and clay optimal conditions for photocatalytic removal of the pollutant.
minerals, which are often used for adsorbing target contami-
nants (Carriazo et al. 2010; Kibanova et al. 2009; Nieto-Suárez
et al. 2009; Papoulis et al. 2010). Among these different porous Experiment
materials, clay minerals such as zeolites (Ze) have been widely
studied, because they are inexpensive and have a useful prop- Photocatalyst preparation
erties such as large specific surface area and excellent adsorp-
tion capacity (Hao et al. 2016; Nasonova and Kim 2013). In this research, clinoptilolite as a natural zeolite from Semnan
Additionally, natural zeolites are an interesting kind of alumi- Mine (Iran), titanium tetraisopropoxide (Aldrich, 97%) as the
nosilicate minerals with unique structures, uniform pores, and precursor of TiO2 photocatalyst, and polyoxypropylene as a
channels that they are commonly used as adsorbent (Weitkamp polymeric surfactant were used. Preparation of zeolite samples
2000). As a result, a synergetic effect arises between TiO2 was implemented through two processes. At first, zeolite sus-
nanoparticles and nanoclays can significantly influence the pension was obtained by dispersion of 2 g raw zeolite into
adsorption and photocatalytic degradation of pollutants. 100 mL distilled water and stirring for 4 h. In the second
Torimoto et al. (1996) studied the role of porous solids with process, 2 g of raw zeolite was dispersed into the 100 mL
various adsorption capacities such as zeolite, activated carbon, distilled water and 0.5 g POP was added into the zeolite sus-
silica, and alumina for the photocatalytic degradation of pension; afterwards, reaction mixture was stirred for 4 h at
propyzamide which revealed the role of valuable physicochem- 80 °C. TiO2 sol was prepared by the hydrolysis of titanium
ical properties (e.g., cation exchange and molecular sieves) of tetraisopropoxide in distilled water. Therefore, 6 mL
zeolite over other adsorbents on the pollutant removal (Liu isopropanol was added dropwise into 31 mL titanium
et al. 2014). Besides, several reports have published about tetraisopropoxide for a period of 10 min with vigorous stir-
using of polymeric surfactant with high molecular weight such ring. Then, 200 mL distilled water and 2 mL nitric acid were
as polyoxypropylene (POP) in synthesis of clay-TiO2 nano- slowly added and subsequently mixed for 2 h at 80 °C with
composites that can increase pore volume and pore diameter vigorous stirring. Afterwards, TiO2 sol was mixed dropwise
of nanocomposites (Chou et al. 2003; Lin et al. 2001). Chen with zeolite suspension for 30 min while was vigorously
Environ Sci Pollut Res

stirred. The Ti/clay ratio (mmol of Ti per gram of zeolite) was each sample, the experiment was repeated for three times.
3, 7, and 13 mmol/g, which corresponded to about 0.24, 0.56 Investigation of regeneration and reuse ability of the Ze-Ti
and 1.04 g of TiO2 per gram of zeolite, respectively. The and Ze-Ti-POP nanocomposites was carried out at 3 cycles
suspensions were stirred to be homogenized using a magnetic after washing the samples with a solution of H2O2 (0.03% w/w
stirrer for 6 h and hydrothermally treated at 150 °C for 4 h. in H2O) for 15 min after each cycle at the presence of UVA
Finally, the obtained suspensions were centrifuged and lamp with irradiation intensity of 1.6 mW cm−2. Afterwards,
washed with distilled water and dried at the 80 °C and then samples were washed with distilled water and dried at the
calcined at different temperatures including 300, 400, and 80 °C. SO2 removal efficiency under UV light irradiation
500 °C for 4 h (Chen et al. 2012; Daniel et al. 2007; was calculated according to the following equation:
Petkowicz et al. 2010) to find the influence of calcination  
C inlet −C outlet
temperatures on SO2 removal efficiency of nanocomposites. SO2 removal efficiency ð%Þ ¼  100 ð1Þ
Details of experiments have been described in the C inlet
BExperimental setup and procedure^ section.
where Cinlet and Coutlet are the concentrations of SO2 in the
photoreactor inlet and outlet streams, respectively. The Cinlet
Physicochemical characterization of modified concentration tests were performed under simulated flue gas
photocatalysts condition and without the presence of photocatalyst.
The nanocomposites and natural zeolite were characterized
Kinetic model simulation
via a series of complementary analytical techniques. The spe-
cific surface area of the samples was determined by nitrogen
Langmuir–Hinshelwood (L-H) model of monomolecular re-
adsorption-desorption isotherms (Belsorp mini II Japan). The
action kinetics has been most commonly used for the expres-
results obtained from the specific surface area and pore struc-
sion to explain heterogeneously photocatalytic reactions
ture were characterized by Brunauer–Emmett–Teller and
(Kachina 2008). Experimental data were also analyzed with
Barrett–Joyner–Halenda methods. Surface morphology and
the L-H model which can be expressed as the following equa-
semiquantitative elemental analysis were characterized using
tions:
scanning electron microscopy (SEM) model PHILIPS XL30-
FEG combined with an energy-dispersive X-ray (EDX) spec- kKC
r¼ ð2Þ
trometer and operated at 20 kV. 1 þ KC
1 1 1 1
Experimental setup and procedure ¼ þ ð3Þ
r k kK C
The photocatalytic oxidation was evaluated in a fixed bed where r is the reaction rate (mol m−3 s), k is the L-H rate
continuous flow photoreactor. The photoreactor was made of constant (mol m−3 s), K is the Langmuir adsorption coefficient
a Pyrex glass with capacity of 50 mL and equipped with air (m3 mol), and C is the concentration of the reactant (mol m−3).
inlet, air outlet, UVA-light lamp and its holder, temperature A linear plot of 1/r versus 1/C is often obtained, that the
controller, mass flow controllers, cooler, and a protected con- value k as the L-H rate constant and K as the Langmuir ad-
tainer (Fig. 1). The photoreactor was sealed to avoid the leak sorption constant can be derived from the combination of the
of gas. The UV-light source was a 64-W m-2 tube with a intercept and the slope of the linear line from the SO2 in the
spectral peak centered at 365–380 nm. The nanocomposites photocatalytic degradation reaction.
and raw zeolite as well as TiO2 as blank (0.5 g) were placed on
a sample board in the photoreactor and treated at room tem-
perature (25 °C) in an N2 stream to attain an equilibrium Results and discussion
adsorption of SO2. Thus, SO2 and N2 were mixed in a gas
blender and were adjusted by mass flow controllers and intro- Structure characteristics of nanocomposites
duced into the photoreactor under the pressure of 1 atm and
flow rate of 60 mL min-1. The inlet concentration of SO2 BET analysis
monitored before the experiment was up to 500 ppm.
Analyses of the SO2 concentration at inlet and outlet were Generally, the photocatalytic properties of TiO2 are influ-
continuously measured using the flue gas analyzer online enced by physicochemical variable such as pore volume,
(MRU VARIO plus SE, Germany). Details of the experimen- particle size, and surface area. Table 1 represents the spe-
tal apparatus are shown in Fig. 1. cific surface area and pore structure information that ob-
The removal efficiencies were calculated by comparing the tained from the nitrogen adsorption–desorption isotherm
concentrations of SO2 after and before the experiments. For data. From these results, it can be concluded that specific
Environ Sci Pollut Res

Fig. 1 Schematic diagram of 4


experimental setup for SO2
removal. 1 SO2 cylinder, 2 N2
cylinder, 3 mass flow controller, 4 3 3 6 8
valve, 5 water bath, 6 barometer,
7 gas blender, 8 holder, 9 UV 9
lamp, 10 photoreactor, 11 gas 7 Mix
10
GAS
analyzer, 12 computer 4 3 4
4

3
SO 2 N2
H2O
4 11
12

5
1 2

surface area of Ze-Ti-POP nanocomposites enhanced sig- variation in surface morphology with intercalation TiO2 nano-
nificantly compared to the zeolite particles. However, particles and polyoxypropylene surfactant. When TiO2 nano-
when TiO2 nanoparticles and POP surfactant were intro- particles are intercalated into the zeolite, small particles be-
duced, the porosity of the synthesized nanocomposites come dispersing on the surface and interlayer spacing of zeo-
was greatly increased. According to IUPAC classification, lite, and this can be the reason of disordered structures of the
nitrogen adsorption–desorption isotherms for Ze, Ze-Ti, samples. In the image of Ze-Ti-POP nanocomposites, the larg-
and Ze-Ti-POP nanocomposites were a typical type IV er platelets with smooth surfaces can be observed compared to
isotherms and pore size distributions were in the mesopo- natural zeolite and Ze-Ti nanocomposites (Fig. 2). More struc-
rous region (S1). tural information of the samples was investigated using the
elemental analysis through EDX. By comparison of EDX
SEM-EDX analysis analyses of natural zeolite, Ze-Ti, and Ze-Ti-POP nanocom-
posites, it was revealed that with increasing Ti at nanocom-
Figure 2 shows the SEM and EDX results of Ze, as well posites, the amounts of Si, Al, and other metals were signifi-
as Ze-Ti, and Ze-Ti-POP nanocomposites, which indicates a cantly decreased (Fig. 2).

Table 1 Structural properties of


the natural zeolite (clinoptilolite) Sorbent BJH plot BET plot
and nanocomposites synthe-
sized at various calcination tem- Vp Sp P/P0 Vm dpore Sa
peratures (300, 400 and 500°C) (cm3 g−1) (m2 g−1) (cm3 g−1) (cm3 g−1) (nm) (m2 g−1)

Ze 0.116 30.18 0.120 6.06 18.17 26.41


TiO2 0.125 56.90 0.125 9.62 11.97 42.01
Ze-Ti-(3 mmol Ti/g 0.233 122.98 0.244 26.89 8.34 117.04
zeolite)
Ze-Ti-(7 mmol Ti/g 0.228 134.46 0.234 28.30 7.60 123.21
zeolite)
Ze-Ti-(13 mmol Ti/g 0.223 174.41 0.186 28.01 7.24 167.04
zeolite)
Ze-Ti-300 0.231 116.97 0.239 27.28 8.07 118.75
Ze-Ti-400 0.223 174.41 0.186 28.01 7.24 167.04
Ze-Ti-500 0.256 137.88 0.253 26.63 8.76 115.94
Ze-Ti-POP-300 0.237 198.07 0.237 44.68 4.88 194.49
Ze-Ti-POP-400 0.250 115.4 0.258 27.00 8.78 117.5

Vp primary mesopore volume, Sp specific surface area of primary mesopores, P/P0 total pore volume, Vm mono-
layer volume, dpore mean pore diameter, Sa BET specific surface area, Ze zeolite, Ti TiO2, POP polyoxypropylene
Environ Sci Pollut Res

Fig. 2 SEM images with EDX Ze


spectrum results of zeolite (a), Ze-
Ti (b), and Ze-Ti-POP (c) (Ze ze-
olite, Ti TiO2, POP
polyoxypropylene)

Ze-Ti

Ze-Ti-POP

Effect of different content of TiO2 in nanocomposites Ti nanocomposites decreased with increasing the calcination
on SO2 removal efficiency temperature from 400 to 500 °C (Table 1). Generally, in-
creasing the calcination temperatures could lead to phase
The photocatalytic removal of SO2 using Ze-Ti nanocompos- transformation of anatase to rutile which affects specific sur-
ites having different TiO2 contents is shown in Fig. 3. The SO2 face area (Hussain et al. 2011). Figure 4 shows the removal
removal efficiency was enhanced by increasing TiO2 content efficiency of SO2 by Ze-Ti nanocomposites which enhanced
of nanocomposites probably because of increasing specific by increasing calcination temperature from 300 to 400 °C
surface area (Table 1) which could provide more surface ac-
tive sites and improved the adsorption ability of the nanocom-
posites (Kachina 2008). So, Ze-Ti nanocomposites synthe-
sized with 13 mmol Ti/g clay ratio showed the highest surface
area and SO2 removal efficiency. Furthermore, pore size dis-
tribution was also investigated using BJH plot and results
showed that the maximum distributed pore size of nanocom-
posites was up to 2.10 nm (S2).

Effect of different calcination temperature


of nanocomposites on SO2 removal efficiency

Based on the obtained results, increasing the calcination tem-


peratures from 300 to 400 °C led to increasing and decreas-
ing the specific surface area of Ze-Ti and Ze-Ti-POP nano-
Fig. 3 Effect of different content of TiO2 in Ze-Ti nanocomposites on
composites, respectively. It revealed that Ze-Ti composites SO2 removal efficiency (conditions: SO2 concentration, 500 ppm;
had a good thermal stability in comparison to Ze-Ti-POP photocatalyst dose 1.3 g m−3, gas flow, 60 mL min−1, run time 25 min,
nanocomposites. However, the specific surface area of Ze- UVA irradiation intensity, 6.4 mW cm−2). Ze zeolite, Ti TiO2
Environ Sci Pollut Res

Fig. 5 SO2 removal efficiency of sorbents in various times (conditions:


Fig. 4 Effect of different calcination temperatures (300, 400 and 500 °C) SO2 concentration, 500 ppm; photocatalyst dose 1.3 g m−3 gas flow,
of Ze-Ti and Ze-Ti-POP nanocomposites on SO2 removal efficiency 60 mL min−1, UVA irradiation intensity, 6.4 mW cm−2). Ze zeolite, Ti
(conditions: SO2 concentration, 500 ppm; photocatalyst dose 1.3 g m−3 TiO2, POP polyoxypropylene
gas flow, 60 mL min−1, run time 25 min, UVA irradiation intensity,
6.4 mW cm−2). Ze zeolite, Ti TiO2, POP polyoxypropylene
which led to further oxidation of SO2. The key process in the
photocatalytic reaction rate is the interaction between ac-
and then decreased from 400 to 500 °C. However, the re- tive oxygen species, such as hydroxyl radical and superox-
moval efficiency of all samples was more than 80%. In the ide anion radicals and SO2 that occurs on the surface of the
case of Ze-Ti-POP nanocomposites, increasing calcination TiO2 (Campostrini et al. 1994). The removal of SO2 is
temperature from 300 to 400 °C led to decreasing removal usually done through physical adsorption leading to the
efficiency of SO2 from 87.4 to 76.5%. It has been reported formation of strong bonds between SO2 molecules and
that increasing the calcination temperatures more than TiO2 surface, but chemical absorption may occur through
500 °C can decrease surface area of composites and conse- weak bonds between SO2 molecules and oxidizing species
quently has an adverse effect on photocatalytic activity (Sun on the TiO2 surface. As a result of this chemical reaction,
et al. 2015; Turki et al. 2014). In this study, Ze-Ti-POP nano- SO3 molecules that remain on the TiO2 surface subsequent-
composites calcined at temperature of 300 °C showed the ly react with water from humid air and convert into H2SO4
highest specific surface area (up to 194 m2 g−1) compared on the TiO2 surface (Zhao et al. 2009). According to the
to the other nanocomposites (Table 1). It seems that increas- results achieved from the study of Shang et al. (2002)) in
ing temperature from 300 to 400 °C could reduce thermal the SO2–O2–N2 system under UV irradiation, SO2 mole-
stability of Ze-Ti-POP nanocomposites and therefore de- cules produced singlet-state (1SO2) and triplet-state (3SO2)
creased their specific surface area and consequently reduced species, and this demonstrated that 3SO2 molecules was
the SO2 removal efficiency (Fig. 4). mainly dominant species in the photochemical reactions.
The reaction mechanism of SO 2 removal by TiO 2
Kinetic model of SO2 removal by nanocomposites photocatalysts according to results of previous investiga-
tions is as follows (Sanchez and Augustynski 1979; Shang
To assess the photocatalytic performance of natural zeolite et al. 2002; Zhao et al. 2009):
and prepared nanocomposites, the removal rate of SO2 was
calculated over the time. Figure 5 shows the removal percent- TiO2 þ hv→TiO2 þ e− ðCBÞ þ hþ ðVBÞ ð5Þ
age of SO2 by Ze, Ze-Ti-POP, and Ze-Ti nanocomposites un- O2 ðgÞ →O2 ðadsÞ →2OðadsÞ ð6Þ
der UV irradiation (UV intensity; 6.4 mW cm-2). The results
showed that the removal efficiencies of SO2 by Ze, Ze-Ti, and OðadsÞ þ e− →O−ðadsÞ þ hþ →O*ðadsÞ ð7Þ
Ze-Ti-POP nanocomposites were 17.5, 82.1, and 87.4%, re-
O* →O−ðadsÞ þ hþ →O*ðadsÞ ð8Þ
spectively, whereas the removal efficiency of blank (i.e., TiO2)
was about 46.3%. During the whole process (25 min), SO2 O* þ H2 OðadsÞ →2OH* ðadsÞ ð9Þ
removal rate indicated only a slight difference between Ze-Ti
and Ze-Ti-POP nanocomposites, which demonstrated the high
adsorption capacity of both sorbents. In fact, the synergic ef- Moreover, homogeneous photooxidation reactions of SO2
fects between zeolite adsorption and TiO2 photocatalysts was without nanocomposites and zeolite in the presence of UV
the main reason for much more efficiency of the Ze-Ti and Ze- irradiation may occur. The SO2 removal efficiency under ho-
Ti-POP nanocomposites in comparison to the natural zeolite mogeneous photooxidation reactions was about 7%.
Environ Sci Pollut Res

The effect of reaction time on SO2 removal efficiency Table 2 Langmuir–Hinshelwood parameters obtained from SO2
photocatalytic removal experiments using various sorbents
is shown in Fig. 4. SO2 removal by the natural zeolite
exhibited different trend and was the highest during the Sorbents L-H kinetic model
first 5 min of reaction time. Then, it was reduced because
of zeolite surface saturation by sulfate. Compared to zeo- k (mol m−3 s−1) K (m3 mol−1) R2
lite, Ze-Ti and Ze-Ti-POP nanocomposites showed more Ze 154,548 3.8 × 10−9 0.97
than 97% SO2 removal efficiency after 5 min and then it −6
Ze-Ti 3 × 10 5.2 × 10−12 0.92
showed a constant trend. This revealed the significant ef-
Ze-Ti-POP 5 × 10−6 1.8 × 10−12 0.97
fect of zeolite modification by TiO2 and POP on the gas
removal. k reaction rate constant, K Langmuir adsorption coefficient, R2 correlation
However, Langmuir-Hinshelwood kinetic model was coefficient, Ze zeolite, Ti TiO2, POP polyoxypropylene
used to explain the heterogeneous photocatalytic mecha-
nisms of SO2 oxidation using Ze-Ti, Ze-Ti-POP nanocom- Effect of humidity on SO2 removal efficiency
posites, and clinoptilolite. Table 2 shows the k, K, and R2
values for SO2 oxidation via the L-H simulation. The R2 The effect of humidity on SO2 removal efficiency was studied
values obtained from the L-H kinetic model ranged be- at different humidity contents (Fig. 7). The SO2 removal effi-
tween 0.92 and 0.97. The comparing values obtained from ciency of Ze-Ti and Ze-Ti-POP nanocomposites increased
the L-H kinetic model indicated that the adsorption con- with increasing the humidity content from 0 to 2%, and a
stant (K) for Ze-Ti and Ze-Ti-POP nanocomposites was further increase in the humidity content to 5% resulted in a
approximately the same and it was more than the adsorp- decrease of SO2 removal. It is known that the increasing ad-
tion constant of the natural zeolite. But the reaction rate sorption of water on the irradiated TiO2 surface leads to re-
(k) value of the zeolite was more than that of Ze-Ti and combination of the photogenerated electrons and holes and
Ze-Ti-POP nanocomposites. This explains that higher ad- decreases photocatalytic reactivity (Park et al. 2001). The re-
sorption constant cannot indicate a higher reaction rate sults showed that the maximum SO2 removal efficiency of
(Boulamanti et al. 2008). Chiu et al. (2015) explained that 71.6 and 77.6% were achieved for Ze-Ti and Ze-Ti-POP nano-
higher R2 values obtained from the L-H kinetic model composites with 2% humidity, respectively. This could be in
could indicate impregnated surface of photocatalyst. By account of SO2 oxidizing by active radicals such as ·OH, ·O2−,
the way, results of the current study revealed that the and .O on TiO2 surface. The overall mechanism of photocat-
experimental data had a good agreement with Langmuir- alytic oxidation in the presence of H2O molecules can be
Hinshelwood kinetic model. described by the following reactions (Zhao et al. 2009):

H2 O↔Hþ þ OH− ð10Þ


Effect of UV irradiation intensity on SO2 removal
efficiency OH− þ hþ ðVBÞ→∙OH ð11Þ
þ þ
H2 O þ h ðVBÞ↔H þ ∙OH ð12Þ
As it is seen in Fig. 6, when UV irradiation intensity
increased from 3.2 to 6.4 mW cm−2, SO2 removal effi-
ciency enhanced significantly from 45.2 to 72.6% and
from 60.9 to 77.6% for Ze-Ti and Ze-Ti-POP, respective-
ly. Besides, SO2 removal efficiency without UV irradia-
tion was only 14.9 and 25.1% for Ze-Ti and Ze-Ti-POP,
respectively. However, comparison of the results indicated
that UV irradiation intensity played an important role on
the photocatalytic reaction. The effects of the UV irradia-
tion intensity on SO2 removal efficiency can be explained
by two main reasons. On the one hand, after activation,
the surface of TiO2 by UV irradiation could lead to the
transfer of electrons from the valence band to the conduc-
tion band and therefore electron–hole pairs on TiO2 would
be generated (Sun et al. 2015). Moreover, increasing the
UV irradiation intensity could enhance production of pho-
Fig. 6 Influence of the UVA irradiance on the SO2 removal efficiency
toexcited active species on the surface of TiO2 and sub- (conditions: SO2 concentration, 500 ppm; photocatalyst dose 0.6 g m−3
sequently the photocatalytic activity would increase (Yuan gas flow, 60 mL min−1, run time 25 min). Ze zeolite, Ti TiO2, POP
et al. 2012). polyoxypropylene
Environ Sci Pollut Res

Fig. 7 Influence of the humidity content on the SO2 removal efficiency


(conditions SO2 concentration, 500 ppm; photocatalyst dose 0.6 g m−3, Fig. 8 Influence of the photocatalyst dose on SO2 removal efficiency
gas flow 60 mL min−1, run time 25 min, UVA irradiation intensity, (conditions: SO2 concentration, 500 ppm; photocatalyst dose 0.6 g m−3,
6.4 mW cm−2). Ze zeolite, Ti TiO2, POP polyoxypropylene gas flow 60 mL min−1, run time 25 min, UVA irradiation intensity,
6.4 mW cm−2). Ze zeolite, Ti TiO2, POP polyoxypropylene

∙OHðadsÞ þ SO2ðadsÞ →HOSO2ðadsÞ ð13Þ Regeneration ability and reuse of the photocatalysts

SO3ðadsÞ →SO3ðgÞ ð14Þ During the photocatalytic processes, sulfur dioxide is convert-
ed into sulfur trioxide, which accumulated on the surface of
TiO2 catalyst and leads to gradual disappearance of catalytic
activity (Shang et al. 2002). Therefore, the adsorbent stability
The results also showed that with increaseing the hu-
and regeneration of Ze-Ti and Ze-Ti-POP nanocomposites is
midity content from 0 to 5%, the SO2 removal efficiency
very important factor for SO2 removal by adsorption and pho-
by zeolite decreased from 10.5 to 4.4%. It was reported
tocatalytic oxidation. The regeneration results of Ze-Ti and
that SO2 removal efficiency decreased due to the inhibi-
Ze-Ti-POP after 3 cycles are indicated in Fig. 9. It was seen
tory effect of H2O molecules into the zeolite pores (Rouf
that the regenerated Ze-Ti and Ze-Ti-POP showed an excellent
and Eić 1998). Furthermore, excess of H 2O molecules
adsorption and high photocatalytic capacity as the fresh sor-
which occupy the active sites on the surface of TiO2 com-
bents after three regeneration cycles with only a slight de-
pete with the gas molecules and reduce photocatalytic
crease in efficiency. The SO2 removal efficiency of the regen-
reaction rate and removal efficiency (Kachina 2008).
erated Ze-Ti and Ze-Ti-POP was about 66.1 and 76.3%,
However, it should be considered that a certain amount
of humidity is necessary to maintain hydroxylation and to
prevent saturation of TiO2 surface by partially oxidized
products (Maira et al. 2001).

Effect of photocatalyst dose on SO2 removal


efficiency

The amount of removed SO 2 increased monotonously


with increasing photocatalyst dose (Fig. 8). Maximum re-
moval efficiency of SO 2 achieved to 17.5, 82.1, and
87.2% for Ze, Ze-Ti, and Ze-Ti-POP, respectively. One
main reason for increasing the removal rate of SO2 is
generation of more photocatalytic reactive sites for the
oxidation of SO2 into SO2− 4 (Kamegawa et al. 2013).
However, it should be considered that increasing the
photocatalyst dose may also lead to particle aggregation Fig. 9 SO2 removal efficiency of Ze-Ti and Ze-Ti-POP nanocomposites
in three reuse cycles (conditions: SO 2 concentration, 500 ppm;
and formation of multilayers which are not accessible to
photocatalyst dose 1.3 g m−3, gas flow 60 mL min−1, run time 25 min,
the reaction and decreases the exposed surface area of UVA irradiation intensity, 6.4 mW cm−2). Ze zeolite, Ti TiO2, POP
sorbents (Fida et al. 2015). polyoxypropylene
Environ Sci Pollut Res

respectively, while the efficiency of original samples was 82.1 Chen D, Zhu Q, Zhou F, Deng X, Li F (2012) Synthesis and photocata-
lytic performances of the TiO2 pillared montmorillonite. J Hazard
and 87.4%, respectively. The regeneration could be due to
Mater 235:186–193
formation of •OH radicals from H2O2 aqueous solutions under Chiu CH, Hsi HC, Lin HP (2015) Multipollutant control of Hg/SO /NO
UVA irradiation which react with the pollutants on from coal-combustion flue gases using transition metal oxide-im-
photocatalyst surface and play the most relevant role for the pregnated SCR catalysts. Catal Today 245:2–9
Chou C-C, Shieu F-S, Lin J-J (2003) Preparation, organophilicity, and self-
successful regeneration (Miranda-García et al. 2014).
assembly of poly (oxypropylene) amine-clay hybrids. Macromolecules
However, H2O2 in the presence of UVA irradiation might 36(7):2187–2189. https://doi.org/10.1021/ma025773h
affect the structure and surface of the nanocomposites. Daniel LM, Frost RL, Zhu HY (2007) Synthesis and characterisation of
clay-supported titania photocatalysts. J Colloid Interface Sci 316(1):
72–79. https://doi.org/10.1016/j.jcis.2007.08.023
Fida H, Guo S, Zhang G (2015) Preparation and characterization of bi-
Conclusions functional Ti–Fe kaolinite composite for Cr (VI) removal. J Colloid
Interface Sci 442:30–38
This study showed that modification of clinoptilolite as a Hamidpour M, Kalbasi M, Afyuni M, Shariatmadari H, Holm PE,
Hansen HCB (2010) Sorption hysteresis of Cd (II) and Pb (II) on
natural zeolite with TiO2 and POP surfactant could signifi-
natural zeolite and bentonite. J Hazard Mater 181:686–691
cantly affect SO2 removal efficiency through the changing the Hao X, Hou G, Zheng P, Liu R, Liu C (2016) H2S in-situ removal from
sorbent structure properties (specific surface area and pore biogas using a tubular zeolite/TiO2 photocatalytic reactor and the
volume) as well as improving photocatalytic oxidation. improvement on methane production. Chem Eng J 294:105–110
Furthermore, the factors affecting the gas removal efficiency Hashimoto K, Wasada K, Osaki M, Shono E, Adachi K, Toukai N,
Kominami H, Kera Y (2001) Photocatalytic oxidation of nitrogen
including calcination temperature, reaction time, dose of oxide over titania–zeolite composite catalyst to remove nitrogen
photocatalysts, water content, and UV irradiation intensity oxides in the atmosphere. Appl Catal B Environ 30(3-4):429–436.
were optimized. The L-H kinetic model was successfully https://doi.org/10.1016/S0926-3373(00)00258-7
fitted with the obtained data from the experiments. Hussain M, Russo N, Saracco G (2011) Photocatalytic abatement of VOCs
by novel optimized TiO2 nanoparticles. Chem Eng J 166:138–149
Regeneration process of the sorbents revealed that they could Ibusuki T, Takeuchi K (1994) Removal of low concentration nitrogen
be reused as promising photocatalysts for SO2 oxidation. By oxides through photoassisted heterogeneous catalysis. J Mol Catal
the way, in practical point of view, the economic aspects of 88(1):93–102. https://doi.org/10.1016/0304-5102(93)E0247-E
using these modified photocatalysts, interactions of other pol- Jansson I, Suárez S, Garcia-Garcia FJ, Sánchez B (2015) Zeolite–TiO2 hy-
brid composites for pollutant degradation in gas phase. Appl Catal B
lutants in real conditions, and management of the by-products Environ 178:100–107. https://doi.org/10.1016/j.apcatb.2014.10.022
of the process should be considered. Kachina A (2008) Gas-phase photocatalytic oxidation of volatile organic
compounds. Lappeenranta University of Technology, Lappeenranta
Acknowledgements The current work was financially supported by the Kamegawa T, Kido R, Yamahana D, Yamashita H (2013) Design of TiO2-
Iran Nanotechnology Initiative Council (INIC-84100) under the project zeolite composites with enhanced photocatalytic performances under
entitled BApplication of TiO2-clay nanocomposites for removal of SO2 as irradiation of UV and visible light. Microporous Mesoporous Mater
a pollutant gas,^ which is appreciated. The authors gratefully acknowl- 165:142–147. https://doi.org/10.1016/j.micromeso.2012.08.013
edge Dr. Ali Ashrafi for his great help to set up the photoreactor equip- Kibanova D, Trejo M, Destaillats H, Cervini-Silva J (2009) Synthesis of
ment for laboratory experiments. hectorite–TiO2 and kaolinite–TiO2 nanocomposites with photocata-
lytic activity for the degradation of model air pollutants. Appl Clay
Sci 42(3-4):563–568. https://doi.org/10.1016/j.clay.2008.03.009
Lin J-J, Cheng I-J, Wang R, Lee R-J (2001) Tailoring basal spacings of
References montmorillonite by poly (oxyalkylene) diamine intercalation.
Macromolecules 34(26):8832–8834. https://doi.org/10.1021/ma011169f
Anpo M, Kamat PV (2010) Environmentally benign photocatalysts: ap- Liu S, Lim M, Amal R (2014) TiO2-coated natural zeolite: rapid humic
plications of titanium oxide-based materials. Springer Science & acid adsorption and effective photocatalytic regeneration. Chem Eng
Business Media, Berlin. https://doi.org/10.1007/978-0-387-48444-0 Sci 105:46–52
Maira A, Yeung KL, Soria J, Coronado J, Belver C, Lee C, Augugliaro V
Ao C, Lee S, Zou S, Mak C (2004) Inhibition effect of SO2 on NOx and
(2001) Gas-phase photo-oxidation of toluene using nanometer-size
VOCs during the photodegradation of synchronous indoor air pol-
TiO2 catalysts. Appl Catal B Environ 29(4):327–336. https://doi.
lutants at parts per billion (ppb) level by TiO2. Appl Catal B Environ
org/10.1016/S0926-3373(00)00211-3
49(3):187–193. https://doi.org/10.1016/j.apcatb.2003.12.011
Malekian R, Abedi-Koupai J, Eslamian SS, Mousavi SF, Abbaspour KC,
Boulamanti AK, Korologos CA, Philippopoulos CJ (2008) The rate of Afyuni M (2011) Ion-exchange process for ammonium removal and
photocatalytic oxidation of aromatic volatile organic compounds in release using natural Iranian zeolite. Appl Clay Sci 51(3):323–329.
the gas-phase. Atmos Environ 42(34):7844–7850. https://doi.org/ https://doi.org/10.1016/j.clay.2010.12.020
10.1016/j.atmosenv.2008.07.016 Miranda-García N, Suárez S, Maldonado MI, Malato S, Sánchez B
Campostrini R, Carturan G, Palmisano L, Schiavello M, Sclafani A (2014) Regeneration approaches for TiO 2 immobilized
(1994) Sol-gel derived anatase TiO2: morphology and photoactivity. photocatalyst used in the elimination of emerging contaminants in
Mater Chem Phys 38:277–283 water. Catal Today 230:27–34
Carriazo J, Moreno-Forero M, Molina R, Moreno S (2010) Incorporation Nasonova A, Kim K-S (2013) Effects of TiO2 coating on zeolite particles
of titanium and titanium–iron species inside a smectite-type mineral for NO and SO2 removal by dielectric barrier discharge process.
for photocatalysis. Appl Clay Sci 50(3):401–408. https://doi.org/10. Catal Today 211:90–95. https://doi.org/10.1016/j.cattod.2013.03.
1016/j.clay.2010.09.007 006
Environ Sci Pollut Res

Nieto-Suárez M, Palmisano G, Ferrer ML, Gutiérrez MC, Yurdakal S, Su C, Ran X, Hu J, Shao C (2013) Photocatalytic process of simultaneous
Augugliaro V, Pagliaro M, del Monte F (2009) Self-assembled tita- desulfurization and denitrification of flue gas by TiO2–polyacrylo-
nia–silica–sepiolite based nanocomposites for water decontamina- nitrile nanofibers. Environ Sci Technol 47(20):11562–11568.
tion. J Mater Chem 19(14):2070–2075. https://doi.org/10.1039/ https://doi.org/10.1021/es4025595
b813864h Sun Q, Hu X, Zheng S, Sun Z, Liu S, Li H (2015) Influence of calcination
Niu S, Han K, Lu C (2011) Release of sulfur dioxide and nitric oxide and temperature on the structural, adsorption and photocatalytic proper-
characteristic of coal combustion under the effect of calcium based ties of TiO2 nanoparticles supported on natural zeolite. Powder
organic compounds. Chem Eng J 168(1):255–261. https://doi.org/ Technol 274:88–97. https://doi.org/10.1016/j.powtec.2014.12.052
10.1016/j.cej.2010.10.082 Todorova N, Giannakopoulou T, Karapati S, Petridis D, Vaimakis T,
Papoulis D, Komarneni S, Nikolopoulou A, Tsolis-Katagas P, Trapalis C (2014) Composite TiO2/clays materials for photocatalytic
Panagiotaras D, Kacandes H, Zhang P, Yin S, Sato T, Katsuki H NOx oxidation. Appl Surf Sci 319:113–120. https://doi.org/10.1016/
(2010) Palygorskite-and halloysite-TiO2 nanocomposites: synthesis j.apsusc.2014.07.020
and photocatalytic activity. Appl Clay Sci 50(1):118–124. https:// Torimoto T, Ito S, Kuwabata S, Yoneyama H (1996) Effects of adsorbents
doi.org/10.1016/j.clay.2010.07.013 used as supports for titanium dioxide loading on photocatalytic deg-
Park D-R, Ahn B-J, Park H-S, Yamashita H, Anpo M (2001) radation of propyzamide. Environ Sci Technol 30(4):1275–1281.
Photocatalytic oxidation of ethylene to CO2 and H2O on ultrafine https://doi.org/10.1021/es950483k
powdered TiO2 photocatalysts: effect of the presence of O2 and H2O Turki A, Guillard C, Dappozze F, Berhault G, Ksibi Z, Kochkar H (2014)
and the addition of Pt. Korean J Chem Eng 18(6):930–934. https:// Design of TiO2 nanomaterials for the photodegradation of formic
doi.org/10.1007/BF02705621 acid—adsorption isotherms and kinetics study. J Photochem
Petkowicz DI, Pergher SB, Da Silva CDS, Da Rocha ZN, Dos Santos JH Photobiol A Chem 279:8–16. https://doi.org/10.1016/j.
(2010) Catalytic photodegradation of dyes by in situ zeolite- jphotochem.2014.01.008
supported titania. Chem Eng J 158(3):505–512. https://doi.org/10.
Weitkamp J (2000) Zeolites and catalysis. Solid State Ionics 131(1-2):
1016/j.cej.2010.01.039
175–188. https://doi.org/10.1016/S0167-2738(00)00632-9
Rouf SA, Eić M (1998) Adsorption of SO2 from wet mixtures on hydro-
phobic zeolites. Adsorption 4(1):25–33. https://doi.org/10.1023/A: Yuan Y, Zhang J, Li H, Li Y, Zhao Y, Zheng C (2012) Simultaneous
1008883219403 removal of SO2, NO and mercury using TiO2-aluminum silicate
Sanchez J, Augustynski J (1979) X-ray photoelectron spectroscopic study fiber by photocatalysis. Chem Eng J 192:21–28. https://doi.org/10.
of the interaction of various anions with the oxide-covered titanium 1016/j.cej.2012.03.043
metal. J Electroanal Chem Interfacial Electrochem 103(3):423–426. Zhao Y, Han J, Shao Y, Feng Y (2009) Simultaneous SO2 and NO re-
https://doi.org/10.1016/S0022-0728(79)80366-6 moval from flue gas based on TiO2 photocatalytic oxidation.
Shang J, Zhu Y, Du Y, Xu Z (2002) Comparative studies on the deacti- Environ Technol 30(14):1555–1563. https://doi.org/10.1080/
vation and regeneration of TiO2 nanoparticles in three photocatalytic 09593330903313786
oxidation systems: C7H16, SO2, and C7H16–SO2. J Solid State Chem
166(2):395–399. https://doi.org/10.1006/jssc.2002.9613

Вам также может понравиться