You are on page 1of 6

Egyptian Journal of Petroleum (2012) 21, 119–124

Egyptian Petroleum Research Institute

Egyptian Journal of Petroleum


www.elsevier.com/locate/egyjp
www.sciencedirect.com

FULL LENGTH ARTICLE

Catalytic para-xylene maximization. Part X: Toluene


disproportionation on HF promoted H-ZSM-5 catalysts
Ahmed K. Aboul-Gheit a,*, Ateyya A. Aboul-Enein a, Salwa A. Ghoneim a,
Samia A. Hanafi b, Ahmed E. Awadallah a

a
Process Development Division, Egyptian Petroleum Research Institute, Nasr City, P.O. Box 9540, Cairo 11787, Egypt
b
Petroleum Refining Division, Egyptian Petroleum Research Institute, Nasr City, P.O. Box 9540, Cairo 11787, Egypt

Received 4 December 2011; accepted 28 February 2012


Available online 21 January 2013

KEYWORDS Abstract H-ZSM-5 zeolite catalysts were doped with 2%, 3% and 4%HF to be used for investi-
HZSM-5; gating their activities and selectivities for xylenes production and for para-xylene maximization at
Disproportionation; temperatures of 300–500 C via toluene disproportionation. This doping caused pore size modifica-
Toluene; tion of the H-ZSM-5 catalyst. The reaction was carried out in a fixed bed flow type reactor. The
Hydrofluorination; ratio of produced para-xylene relative to its thermodynamic composition reached as high as 3.29
Para-xylene at 300 C on the 4%HF doped H-ZSM-5 catalyst although this catalyst possessed the lowest
amount of the largest pores (3.0–5.7 nm) and the smallest pores (0.4–1.7 nm). The overall activities
of the catalysts were decreased with an increase in HF doping because of diffusion restriction. The
kinetics of the reaction were simply treated and found to give Ea and DS* values compatible with the
characterization data of the catalysts.
ª 2012 Egyptian Petroleum Research Institute. Production and hosting by Elsevier B.V.
Open access under CC BY-NC-ND license.

1. Introduction production of a wide range of materials such as polyesters


and polyethylene terephthalate (PET) [1]. Different types of
Toluene disproportionation is a very useful reaction for con- catalysts based on acidic zeolites have been reported to cata-
verting surplus toluene into the more valuable xylene, which lyze toluene conversion at high xylenes selectivity [2–12]. As-
is a key raw material in the petrochemical industry for the synthesized HZSM-5 zeolites always show low selectivity for
para-xylene. Post synthesis modification methods, described
as pore size engineering [13], are proposed to tailor the proper-
* Corresponding author. Tel.: +20 2 22729100/22808600; fax: +20 2
ties of zeolites to achieve high selectivity. Catalyst modifica-
22747433.
tions such as pore blocking and surface coating [14], which
E-mail address: aboulgheit2000@hotmail.com (A.K. Aboul-Gheit).
increase the intracrystalline diffusional resistance and inhibit
Peer review under responsibility of Egyptian Petroleum Research
Institute. the secondary isomerization of p-xylene on the external surface
of the crystallite, respectively, have been employed in indus-
trial practice [15] to enhance para-xylene selectivity. A typical
shape-selective toluene disproportionation reaction occurs in
Production and hosting by Elsevier the zeolitic pores of H-ZSM-5 to generate xylenes and benzene.
1110-0621 ª 2012 Egyptian Petroleum Research Institute. Production and hosting by Elsevier B.V. Open access under CC BY-NC-ND license.
http://dx.doi.org/10.1016/j.ejpe.2012.11.009
120 A.K. Aboul-Gheit et al.

Moreover, xylenes isomerize in the zeolitic pores simulta- 20 cm3 min1. The liquid product was collected and analyzed
neously and the products diffuse out of the pores at different by gas chromatography using a column packed with 5% Ben-
rates [14,16,17]. tone-34 plus 5% diisodecyl phthalate supported on Chromo-
The incorporation of fluorine in oxide catalysts enhances sorb-W (Resteck Corp., USA).
their activities for acid catalyzed hydrocarbon reactions [18–
20]. Fluoride ion is assumed to replace surface O or OH; be- 2.3. Temperature programmed desorption of ammonia (TPDA)
cause fluorine is more electronegative than these groups; it re-
places and increases the polarity of the lattice, thereby The TPDA procedure using differential scanning calorimetry
increasing the acidity and activity of the surface. Lok et al. (DSC) for detecting and evaluating the desorption of pre-
[21] have found that the treatment of zeolites with diluted fluo- sorbed ammonia from the catalysts via programmed tempera-
rine gas affected the samples in a manner dependent on the ture increase was carried out in an inert atmosphere [29,30].
severity of fluorination. Becker and Kowalak studied the mod- Accordingly, ammonia was adsorbed on the acid sites of the
ification of H-mordenite with ammonium fluoride solutions catalyst after previous heating in air flow at 500 C for 3 h in
and gaseous CHF3 [22]. The H-ZSM-5 zeolites exhibit en- a silica tube furnace. After cooling to 50 C, ammonia gas flow
hanced Brönsted acidity and therefore improved the catalytic of 50 cm3 min1 was applied over the evacuated catalyst. The
activity after treatment with various fluorine-containing com- DSC measurements were then carried out in a DSC-30 unit
pounds [23–25]. Ghosh and Kydd [26] have found that the of the TA-3000 Mettler system (Germany) using standard Al
treatment of mordenite catalysts with dilute HF (60.25 M) crucibles in a flow of 50 cm3 min1 oxygen-free nitrogen purge
does not damage the zeolitic structure of mordenite. Elemental gas. The DSC conditions were as follows: temperature, 50–
analysis shows that small amounts of fluorine are incorpo- 600 C; heating rate, 5 C/min; sample mass; 10 mg. The
rated, and aluminum is leached from the samples. Treatment desorption enthalpy obtained corresponds to the acid site den-
with more concentrated HF (1.5 M) causes silicon to be re- sity whereas the DSC peak temperature was used to compare
moved from the zeolitic structure and decreases slightly the the acid site strength. Two DSC endothermic peaks appear
size of the unit cell. With larger volumes of more concentrated in each thermogram; the lower temperature peak corresponds
HF, the surface area was found to be substantially reduced. to desorption of the weak acid sites and the higher temperature
The incorporation of small amounts of fluoride enhances both peak corresponds to desorption of the strong acid sites respon-
the Brönsted and Lewis acidity, but in different ways; it en- sible for catalyzing the reaction under study. The higher the
hances strong Lewis acidity and weak Brönsted acidity. desorption peak temperature, the higher the acid site strength.
Recently the authors have investigated the alkylation of toluene
with methanol using hydrofluorinated H-ZSM-5 zeolites [27] and 2.4. Surface analysis
hydrofluorinated Pt/H-ZSM-5 [28] catalysts to maximize para-xy-
lene. In the present communication, disproportionation of toluene
BET total surface area, and total pore volume were determined
on hydrofluorinated H-ZSM-5 zeolite catalysts is studied.
from nitrogen adsorption/desorption isotherms measured at
196 C, using Quatachrome NovaWin2 apparatus. The pore
2. Experimental
width distribution was calculated from these data using the
DFT method.
2.1. Catalysts preparation

Sodium ZSM-5 zeolite (Si/Al = 25) in powder form was


kindly supplied by Süd-Chemie AG, Germany. It was trans-
formed to the ammonium form by cation exchange using a
molar solution of NH4NO3 for five repetitions of 8 h each
using a fresh solution of NH4NO3 at 70 C under reflux and
stirring. The zeolite was then separated from the solution,
washed thoroughly with distilled water, dried overnight at
110 C, and calcined at 530 C for 4 h to obtain the H-ZSM-
5 zeolite. Three hydrofluorinated catalysts, H-ZSM-
5(2%HF), H-ZSM-5(3%HF), and H-ZSM-5(4%HF) were
prepared by doping H-ZSM-5 with 2%, 3%, and 4%HF,
respectively, followed by drying overnight at 110 C and calci-
nation at 450 C for 4 h.

2.2. Disproportionation procedure and apparatus

The catalytic runs were carried out at atmospheric pressure in


a fixed bed down flow silica tube reactor containing one gram
of the catalyst powder diluted with inert porcelain particles.
Toluene was fed into the reactor at a weight hourly space
velocity of 2.0 h1 via an electronic dosing pump. The dispro- Figure 1 Toluene conversion at different reaction temperatures
portionation runs were carried out at temperatures between using: (s) untreated H-ZSM-5; (4) H-ZSM-5(2%HF); (d) H-
300 and 500 C with a continuous hydrogen flow rate of ZSM-5(3%HF); (m) H-ZSM-5(4%HF) catalysts.
Catalytic para-xylene maximization. Part X: Toluene disproportionation on HF promoted H-ZSM-5 catalysts 121

3. Results and discussion Table 2 Effect of HF concentration on specific surface area


and total pore volume of H-ZSM-5 zeolite.
3.1. Toluene conversion and total xylene production
Catalysts BET surface Total pore
area, m2/g volume, ml/g
The effect of reaction temperature on the conversion of tolu-
Untreated H-ZSM-5 345.4 0.220
ene (Fig. 1) and total xylene production (Fig. 2) during the H-ZSM-5(2%HF) 371.0 0.231
process of toluene disproportionation using the current H- H-ZSM-5(3%HF) 324.3 0.203
ZSM-5 catalysts, either promoted or unprompted with HF, H-ZSM-5(4%HF) 311.6 0.197
have been investigated. Both Figures show that successive in-
crease of the HF concentration in the H-ZSM-5 catalyst, re-
sults in a successive decrease of both the reactions in the The first addition of HF in H-ZSM-5 increases the acid site
following order: H-ZSM-5 > H-ZSM-5(2%HF) > H-ZSM- density from 105.1 to 107.5 J g1, whereas a further increase of
5(3%HF) > H-ZSM-5(4%HF). the HF concentration results in a continuous decrease of acid
Toluene conversion (Fig. 1) and total xylenes production site density with a further increase of HF up to 4% (Table 1).
(Fig. 2) are slowly increased with an increase in the reaction On the other hand, the acid site strength of catalysts continu-
temperature from 300 to 450 C on all catalysts except for ally increases as a function of the HF concentration.
using the H-ZSM-5(4%HF) catalyst. However, from 450 to Moreover, the surface area and total pore volume are found
500 C, the activity and total xylene production increase at to increase at the first treatment with HF, beyond which these
higher rates than from 300 to 450 C, indicating no diffusion parameters decrease with a further increase of the HF concen-
limitation in these catalysts. On the other hand, doping the tration (Table 2). Nevertheless, the disproportionation activity
zeolite with 4%HF does not show the increased rate observed that produces the xylenes (Fig. 2) is decreased continually with
during the increase of temperature from 450 to 500 C as it oc- increasing the HF concentration, which implies that increased
curs with the lower HF doping, which implies the presence of acid strength in the catalysts (Table 1) is unfavorable, most
relatively more significant diffusion limitation. probably due to the shifting of the disproportionation equilib-
rium to the reverse direction. Another cause of this decrease
can be attributed to the increase in the diffusion restriction
in the zeolitic channels of the hydrofluorinated H-ZSM-5
catalysts.
The pore volume distribution in the catalysts has been eval-
uated from N2 adsorption isotherms at 196 C applying the
DFT differential method through plotting the total pore vol-
ume vs. pore width of the catalyst (Fig. 3). The largest pores
in the zeolite have widths ranging between 3.0 and 5.7 nm,
whereas the pores of intermediate width range between 1.7
and 3.0 nm, and the smallest pores in the catalyst range be-
tween 0.4 and 1.7 nm. The amount of the smallest pores
(0.4–1.7 nm) is in the order: H-ZSM-5 > H-ZSM-
5(2%HF) > H-ZSM-5(3%HF) > H-ZSM-5(4%HF). The
overall behavior of the sequential increase of the HF concen-
tration in the zeolite catalyst is due to the leaching of Al and
Si and formation of minute particles of fluoro-alumino- and
fluoro-silico-species (debris) [31] which are deposited in the
pores of the zeolite. As a summation, the deposition of debris
occurs sequentially via HF increase in the widest (3.0–5.7 nm)
and in the narrowest pores (0.4–1.7 nm). This means that
increasing of the HF concentration causes pore size modifica-
Figure 2 Total xylenes production during toluene dispropor- tion of the H-ZSM-5 catalyst.
tionation using: (s) untreated H-ZSM-5; (4) H-ZSM-5(2%HF);
(d) H-ZSM-5(3%HF); (m)H-ZSM-5(4%HF) catalysts. 3.2. Para-xylene selectivity of the catalysts

Fig. 4 depicts the effect of the extent of HF doping on the selec-


Table 1 Ammonia desorption enthalpy (DHdes) and peak
tivity of H-ZSM-5 concerning the reaction of toluene dispro-
temperature for untreated and hydrofluorinated H-ZSM-5
portionation to produce para-xylene. Effectively, the
catalysts.
selectivity of this reaction is the highest on the most diffusion-
Catalysts TPD of ammonia ally restricted catalyst (H-ZSM-5(4%HF)) and selectivity grad-
Peak temp., C DHdes, J g1 ually decreases with decreasing the HF doping; i.e., the para-
Untreated H-ZSM-5 380.0 105.1
xylene selectivity decreases in the order: H-ZSM-
H-ZSM-5(2%HF) 385.0 107.5 5(4%HF) > H-ZSM-5(3%HF) > H-ZSM-5(2%HF) > H-
H-ZSM-5(3%HF) 387.0 101.9 ZSM-5. Using the 4%HF doped catalyst, the selectivity ranges
H-ZSM-5(4%HF) 388.0 98.3 between 78.6% and 42.4% at temperatures between 300 and
500 C, respectively, whereas using the H-ZSM-5 catalyst, the
122 A.K. Aboul-Gheit et al.

benzene diffuses out of the pores rapidly and xylenes isomerize


also rapidly within the pores. The small particles of the depos-
ited debris can cause partial blocking in the pore entrance and
hence increase the difference in diffusivity between para-xylene
and both meta- and ortho-xylenes. The para-xylene diffuses
out moderately fast while the ortho- and meta-xylenes move
within the pores relatively slowly and so further convert to
the para-isomer before escaping from the channel system. In
addition, this debris may be enriched on the external surface
of the fluorinated H-ZSM-5 zeolite and causes masking of
the unselective acid sites and decrease the secondary isomeriza-
tion of p-xylene to the other isomers.

3.3. Para-xylene in product relative to thermodynamic


equilibrium

Fig. 5 is a graphical representation of the behavior of individ-


ual xylenes relative to their thermodynamic equilibrium values,
i.e. X/Xe (where X is the percentage of the individual xylenes in
the total xylenes produced on a given catalyst and at a given
Figure 3 Pore volume distribution by the DFT differential temperature; Xe is the thermodynamic equilibrium value of
method for the catalysts under study: (––) untreated H-ZSM-5; the xylene isomer at the same temperature [34].
( Æ Æ) H-ZSM-5(2%HF); (  ) H-ZSM-5(3%HF); (ÆÆÆÆÆÆÆÆ) Using the untreated H-ZSM-5 catalyst, the X/Xe ratio of
H-ZSM-5(4%HF) catalysts. the three xylenes in product at the whole range of temperature
(300–500 C) investigated does not significantly differ from
unity, i.e., each one of the three xylenes is obtained in the
para-selectivity is 25.0% all over the temperature range inves-
xylenes mixture yields almost equal to their thermodynamic
tigated. This persisting low selectivity of the as-synthesized
values at all temperatures. Using the HF doped catalysts, H-
HZSM-5 zeolite can be attributed to the presence of the
ZSM-5(2%HF), H-ZSM-5(3%HF) and H-ZSM-5(4%HF),
unselective acid sites [15] which cause a secondary
para-xylene shows a superior production over the other two
isomerization of p-xylene to take place on the catalyst’s surface
xylenes, with highest X/Xe values at 300 C and lowest values
whereby the small difference in the diffusivities between the
at 500 C.
para-xylene and its isomers leads to a decrease in the para-selec-
On the other hand, the more voluminous xylenes (ortho-
tivity [14,32].
and meta-xylenes) are not significantly different and yielded
The para-selectivity can be classified under controlling
lowest X/Xe values at 300 C and highest values at 500 C.
selectivity or/and transition state shape selectivity [33]. Dispro-
Using the H-ZSM-5(4%HF) catalyst, the X/Xe values show
portionation of toluene produces benzene and xylenes;
significant production of para-xylene at 3.22 times the equilib-
rium value at 300 C, and 1.8 times at 500 C. Ortho- and
meta-xylenes each gives X/Xe  0.3 times the equilibrium value
at 300 C, whereas at 500 C, ortho-xylene gives X/Xe = 0.60
and meta-xylene gives X/Xe = 0.82. In other words, the de-
crease below the unity of both ortho- and meta-xylenes is com-
pensated by an increase of the para-isomer.

3.4. Kinetics of toluene disproportionation reaction using


hydrofluorinated H-ZSM-5 catalysts

From the data obtained for the total xylenes production in


Fig. 2, the apparent reaction rate constant (k) has been calcu-
lated according to the simple first order flow reactor Eq. (1):
F 1 WHSV 1
k¼ ln ¼ ln ð1Þ
W 1x 3600 1x

k ¼ A  eEa =RT ð2Þ


where W is the catalyst weight (1 gm); F is the rate of the feed
injection (cm3/s); WHSV is the weight hourly space velocity of
the feed (h1) and x is the mole fraction of the total xylenes.
The apparent activation energy (Ea) is calculated via apply-
Figure 4 Para-xylene selectivity during toluene disproportion- ing the Arrhenius plot (Fig. 6) of ln k vs. 1/T, where k is the
ation using: (s) untreated H-ZSM-5; (4) H-ZSM-5(2%HF); (d) reaction rate constant and T is the absolute reaction tempera-
H-ZSM-5(3%HF); (m) H-ZSM-5(4%HF) catalysts. ture. The Ea values calculated accordingly are given in Table 3.
Catalytic para-xylene maximization. Part X: Toluene disproportionation on HF promoted H-ZSM-5 catalysts 123

Figure 5 X/Xe values during toluene disproportionation for: (4) para-xylene; (h) meta-xylene; (¤) ortho-xylene; using: (a) untreated H-
ZSM-5; (b) H-ZSM-5(2%HF); (c) H-ZSM-5(3%HF); (d) H-ZSM-5(4%HF) catalysts.

Applying absolute reaction rate theory, the activation enthalpy


(DH*), and the activation entropy (DS*) are calculated from
the following Eq. (3):

kb  T DH  DS
k¼ e RT  e R ð3Þ
h
The Ea values obtained for catalysts doped with 2%, 3%
and 4%HF amount to 17.54, 12.63 and 9.73 kJ mol1, respec-
tively. This sequential decrease of Ea is attributed to a sequen-
tial increase of diffusion restriction in the channels of the
zeolite. The markedly lower value of Ea using the unloaded
H-ZSM-5 catalyst is attributed to the higher value of rate con-
stant of the current reaction both at lower and higher temper-
atures, such that at the higher temperatures the kinetics are
limited by thermodynamic equilibrium (Fig. 6).
Indeed, the Ea values in Table 3 are relatively low, indicat-
ing that the temperature has a modest effect in accelerating the
current reaction. The difference between the catalytic re-
sponses of these three catalysts toward HF doping is very
slight but exhibiting a decreasing trend with increasing HF
concentration in the catalyst. This indicates that the increase Figure 6 Arrhenius plot for disproportionation of toluene using:
of amorphous debris formed via leaching Al and Si in the zeo- (s) untreated H-ZSM-5; (4) H-ZSM-5(2%HF); (d) H-ZSM-
lite composition is slightly effective on the physical structure of 5(3%HF); (m) H-ZSM-5(4%HF) catalysts.
the porous system of these catalysts. The activation entropy
(DS*) values calculated for xylenes production using the
current catalysts show insignificantly different values. In significant molecular restriction during adsorption and reac-
general, these values reflect that this reaction encounters a tion of the activated state in the zeolitic channels.
124 A.K. Aboul-Gheit et al.

[10] M.A. Uguina, J.L. Sotelo, D.P. Serrano, Ind. Eng. Chem. Res.
Table 3 Activation parameters for the disproportionation of 32 (1993) 49–55.
toluene reaction using the current. [11] J. Das, Y.S. Bhat, A.B. Halgeri, Ind. Eng. Chem. Res. 33 (1994)
Catalysts Ea (kJ mol1) DS* (eu) 246–250.
[12] T. Kunieda, J. Kim, M. Niwa, J. Catal. 188 (1999) 431–433.
Untreated H-ZSM-5 4.5 79.86
[13] E.F. Vansant, Pore Size Engineering in Zeolite, Wiley,
H-ZSM-5(2%HF) 17.54 75.45
Chichester, UK, 1990.
H-ZSM-5(3%HF) 12.63 77.59
[14] D.H. Olson, W.O. Haag, ACS Symp. Ser. 248 (1984) 275–
H-ZSM-5(4%HF) 10.37 80.53
307.
[15] W.W. Kaeding, G.C. Basile, M.M. Wu, Cat. Rev. Sci. Eng. 26
(1984) 597–612.
[16] W.W. Kaeding, C. Chu, L.B. Young, S.A. Butter, J. Catal. 69
4. Conclusion
(69) (1981) 392–398.
[17] P.B. Weisz, Pure Appl. Chem. 52 (1980) 2091–2103.
Although the activities of toluene disproportionation reaction [18] A.K. Ghosh, P. Harvey, U.S. Patent, 7, 105, 713 B2 (2006).
decrease with increasing HF doping concentration, the selec- [19] A.K. Ghosh, R.A. Kydd, Catal. Rev. Sci. Eng. 27 (1985) 539–
tivity for para-xylene production increases. HF leaches Al 589.
and Si forming fine debris which is deposited in the wide cat- [20] V.R. Choudhury, Ind. Eng. Chem. Prod. Res. Dev. 16 (1977)
12–22.
alytic pores (3.0–5.7 nm) leading to decrease in the number
[21] B.M. Lok, F.P. Gortsema, C.A. Messina, H. Rastelli, T.P.J.
of wider pores and transferring them to intermediate pores
Izod, ACS Symp. Ser. 218 (1983) 41–58.
of 1.7–3.0 nm width. This pore size modification leads to max- [22] K.A. Becker, S. Kowalak, J. Chem. Soc. Faraday Trans. I (81)
imize para-xylene selectivity via toluene disproportionation. (1985) 1161–1166.
The 4%HF doped catalyst produces 3.29, 0.27 and 0.31 times [23] R. Le Van Mao, T.S. Le, M. Fairbairn, A. Muntasar, S. Xiao,
of para-, meta- and ortho-xylenes, respectively, relative to their G. Denes, Appl. Catal. A 185 (1999) 41–52.
thermodynamic equilibrium values at 300 C. The process ap- [24] J.N. Miale, C.D. Chang, U.S. Patents (a) 4,427,786 (1984), (b)
pears in general to be grouped under pore engineering effec- 4,427,787 (1984), (c) 4,427,788 (1984), (d) 4,427,789 (1984), (e)
tiveness. The entropy of activation values are highly negative 4,427,790 (1984).
indicating that all the current catalysts suffer significant diffu- [25] G.T. Kokotailo, A.C. Rohrman, S. Sawruk, U.S. Patents, (a)
4,414,189 (1983), (b) 4;415,544 (1983).
sion restriction.
[26] A.K. Ghosh, R.A. Kydd, J. Catal. 103 (1987) 399–406.
[27] A.K. Aboul-Gheit, S.A. Hanafi, S.A. Ghoneim, A.A. Aboul-
References Enein, Egypt, J. Petrol. 19 (2010) 73–83.
[28] A.K. Aboul-Gheit, A.A. Aboul-Enein, A.E. Awadallah, S.A.
[1] T.C. Tsai, S.B. Liu, I. Wang, Appl. Catal. A 181 (1999) 355–398. Ghoneim, E.A. Emam, Chin. J. Catal. 31 (2010) 1209–1216.
[2] I. Kolev, V. Mavrodinova, M.R. Mihályi, M. Kollár, [29] A.K. Aboul-Gheit, J. Catal. 113 (1988) 490–496.
Microporous Mesoporous Mater. 118 (2009) 258–266. [30] A.K. Aboul-Gheit, Thermochem. Acta 191 (191) (1991) 233–
[3] S.B. Hong, H.K. Min, C.H. Shin, P.A. Cox, S.J. Warrender, 240.
P.A. Wright, J. Am. Chem. Soc. 129 (2007) 10870–10885. [31] Y. Wang, X. Guo, C. Zhanga, F. Song, X. Wang, H. Liu, X. Xu,
[4] K. Wang, X. Wang, G. Li, Catal. Commun. 8 (2007) 324–328. C. Song, W. Zhang, X. Liu, X. Han, X. Bao, Catal. Lett. 107
[5] X. Ren, J. Liang, J. Wang, J. Porous Mater. 13 (2006) 353–357. (2006) 209–214.
[6] A.G. Bhavani, D. Karthekayen, A.S. Rao, N. Lingappan, Catal. [32] J. Wei, J. Catal. 76 (1982) 433–439.
Lett. 103 (2005) 89–100. [33] L.B. Young, S.A. Butter, W.W. Kaeding, J. Catal. 76 (1982)
[7] V.P. Mavrodinova, M.D. Papova, Catal. Commun. 6 (2005) 418–432.
247–252. [34] W.J. Taylor, D.D. Wagman, M.G. Williams, K.S. Pitzer, F.D.
[8] Q. Chen, X. Chen, L. Mao, W. Cheng, Catal. Today 51 (1999) Rossini, J. Res. Natl. Bur. Stand. 37 (1946) 95–122.
141–146.
[9] M. Yang, I. Nakamura, K. Fujimoto, Appl. Catal. A 127 (1995)
115–124.

You might also like