Sie sind auf Seite 1von 6

Topics in Catalysis 6 (1998) 101106

101

New catalyst of SO2/Al2 O3ZrO2 for n-butane isomerization 4


Zi Gao , Yongde Xia, Weiming Hua and Changxi Miao
Department of Chemistry, Fudan University, Shanghai, 200433, P.R. China

The catalytic behavior of Al-promoted sulfated zirconia for n-butane isomerization at low temperature in the absence of H2 and at high temperature in the presence of H2 was studied. The addition of Al enhances the activity and stability of the catalysts for reaction at 250 C and in the presence of H2 signicantly. After on stream for 120 h, the n-butane conversion of the catalyst containing 3 mol% Al2 O3 keeps steadily at 88% of its equilibrium conversion and no observable trend of further deactivation has been observed. The difference in behavior of the promoted and unpromoted catalysts at low and high temperature is associated with a change of reaction mechanism from bimolecular to monomolecular. Experimental evidence is presented to show that the promoting effect of Al is different from that of the transition metals. Microcalorimetric measurements of NH3 adsorption on catalysts reveal that the remarkable activity and stability of the Al-promoted catalysts are caused by an enhancement in the number of acid sites effective for the isomerization reaction. Keywords: butane isomerization, sulfated zirconia, SO2 /Al2 O3 ZrO2 , acid strength distribution, promoting effect 4

1. Introduction The isomerization of light parafns to branched isomers is an important process in rening industry. n-Butane, which is undesirable for gasoline, can be converted to isobutane on strong acid catalysts. Isobutane is a valuable precursor for the production of MTBE and alkylated gasoline. As the equilibrium product distribution of the isomerization reaction is more favorable at lower temperature, the use of superacids as catalysts has been suggested [1]. Due to environmental concerns, researchers have shown more interest in solid superacid catalysts than in liquid superacids, such as HF + SbF5 and HCl + AlCl3 . A large number of papers have been devoted to the study of nbutane isomerization on sulfated zirconia and other oxides [214]. The SO2 /ZrO2 catalyst is very active for 4 n-butane isomerization, but a rapid deactivation of the catalyst has often been observed. The addition of a small amount of Pt to the catalyst and/or the presence of H2 have been recommended to increase the life time of SO2 /ZrO2 4 catalyst [7]. More recently, Hsu et al. [15] discovered that doping sulfated zirconia with 1.5 wt% Fe and 0.5 wt% Mn increased the rate of n-butane isomerization by 23 orders of magnitude. The promotion in activity has been conrmed by Jatia et al. [16], but other authors [17,18] have pointed out that the Fe,Mn-promoted catalyst also deactivates quickly at 60 C in the presence of N2 or at 250 C in the presence of H2 . Coelho et al. [19] reported that the addition of Ni to sulfated zirconia caused an activity enhancement comparable to that caused by the addition of Fe and Mn. They postulated that a bifunctional mechanism was responsible for this activity enhancement. Miao et al. [20,21] found that sulfated oxides of CrZr, FeCrZr and FeVZr were

23 times more active than sulfated FeMnZr for n-butane isomerization. In this work, n-butane isomerization on a series of Alpromoted sulfated zirconia catalysts was studied both at low and high temperatures and compared with that on Fe,Mnpromoted sulfated zirconia. A quantitative measurement of the number and acid strength distribution of the surface acid sites of the catalysts was accomplished by means of a microcalorimetric method. The reasons for the improvement on activity and stability of the catalysts in n-butane isomerization by the addition of Al were discussed.

2. Experimental Aqueous ammonia was added dropwise to a solution of ZrOCl2 8H2 O. After washing the hydroxide and drying at 110 C, it was immersed in a 0.5 M H2 SO4 solution for 30 min. The sulfated zirconia was then ltered, dried at 110 C and calcined at 650 C in air for 3 h. Al-promoted catalysts were prepared in the same way from a mixed solution of ZrOCl2 8H2 O and Al(NO3 )3 9H2 O. The unpromoted and promoted catalysts were labeled as SZ and SZA, respectively. SO2 /1.5%Fe/0.5%Mn/ZrO2 (SZFM) was pre4 pared according to the procedures in the literature [20]. X-ray powder diffraction measurements were performed on a Rigaku D/MAXIIA instrument with Cu K radiation, scan speed 16 /min and scan range 570 . Infrared spectra of the samples were recorded on a PerkinElmer 983 G spectrometer. The samples were pressed into thin disks with a density of 35 mg/cm2 and placed in a quartz cell with CaF2 windows. BET surface areas of the samples were acquired on a Micromeritics ASAP 2000 system. Microcalorimetric studies of the adsorption of NH3 were carried out at 150 C using a TianCalvet type heat-ux calorimeter. The catalysts were evacuated at 250 C for

To whom correspondence should be addressed.

J.C. Baltzer AG, Science Publishers

102

Z. Gao et al. / n-Butane isomerization over Al-promoted sulfated zirconia

3 h before measurements. The coke deposit on the catalysts was detected on a CarloErba 1106 elemental analysis instrument. Chemical method was used for the detection of sulfur content in the catalysts. Dehydrated Na2 CO3 and ZnO were used as fusing agents, and the sulfate was turned into BaSO4 and determined by gravimetric method. The isomerization of n-butane was performed both at low and high temperatures. At 35 C a closed reaction system was used. 0.5 g catalyst was placed in a glass cell, and 5 ml (S. T. P.) of n-butane of 99.9% purity was injected for each test. The reaction at 250 C was carried out in a ow-type xed bed reactor under ambient pressure. 1.0 g catalyst was loaded, and a mixture of butane (n-butane: isobutane = 4:1) and H2 (1:10 molar ratio) was fed at a rate of WHSV 0.3 h1 . The catalysts were preheated in situ in dry air at 450 C for 3 h. The reaction products were analyzed by a gas chromatograph equipped with FID.

3. Results and discussion 3.1. Catalyst characterization XRD patterns of the catalysts are shown in gure 1. After calcining at 650 C, a small portion of the monoclinic phase is present in SZ along with the tetragonal phase. For SZA catalysts with 0.510.0 mol% Al2 O3 , the transformation from the metastable tetragonal phase to the monoclinic phase is retarded. The tetragonal phase persists alone in these samples at 650 C. SZA catalysts containing 15 mol% or more Al2 O3 begin to crystallize above 700 C. The tetragonal phase alone is obtained at 750 C. The characteristic peaks of Al2 O3 are not observed in all the samples, implying that Al2 O3 is rather homogeneously mixed with zirconia. The sulfur content and the surface area of the catalysts with different Al2 O3 contents are listed in table 1. BET surface areas of the SZA series catalysts are slightly higher than that of SZ, but the sulfate concentration on the catalysts increases signicantly with Al2 O3 content. The amount of sulfate on SZ corresponds to approximately half a monolayer coverage, assuming each sulfate group covers 0.25 nm2 [22]. On SZA-5 catalyst the surface coverage of sulfate groups is close to unity, showing that the incorporation of Al into ZrO2 helps to stabilize the surface sulfate complexes remarkably. Data of SZFM are also listed in table 1 for comparison. Infrared spectra of the catalysts after evacuation at 350 C for 3 h display a strong band at 1390 cm1 , characteristic of the surface sulfate species having covalent S=O bonds [23]. When water is adsorbed on the surface sulfate, a red shift of this IR band indicating a strong interaction between the adsorbed water molecules and the surface sulfate species is observed. This frequency shift corresponding to a decrease in the bond order of S=O covalent bond and an increase in the partial charge on oxygen atom is associated with the acid strength of the catalyst [24]. The S=O

Figure 1. XRD patterns of various catalysts with different Al2 O3 content (mol%) calcined at 650 C and SO2 /15%Al2 O3 ZrO2 catalyst calcined 4 at different temperatures. (a) 0; (b) 0.5%; (c) 1.5%; (d) 3.0%; (e) 6.0%; (f) 10.0%; (g) 15.0%; (h) 15.0% (700 C); (i) 15.0% (750 C). ( ) Tetragonal phase; () monoclinic phase. Table 1 Surface area and sulfur content of various catalysts. Catalyst SZ SZA-1 SZA-2 SZA-3 SZA-4 SZA-5 SZA-6a SZFM
a

Al2 O3 (mol%) 0 0.5 1.5 3.0 6.0 10.0 15.0 0

Surface area (m2 /g) 113 124 126 140 125 133 104 105

SO3 content (mol/g) 412 587 612 799 874 974 487 524

The calcination temperature for this sample is 750 C, and for all the others 650 C.

Z. Gao et al. / n-Butane isomerization over Al-promoted sulfated zirconia Table 2 Stretching frequency, bond order and partial charge on oxygen of S=O before and after water adsorptiona . Catalyst SO stretching frequency (cm1 ) B SZ SZA-2 SZA-3 SZA-4 SZFM
a

103

Bond order B 1.87 1.85 1.85 1.87 1.85 A 1.80 1.79 1.79 1.80 1.78

Partial charge on oxygen B 0.13 0.15 0.15 0.13 0.15 A 0.20 0.21 0.21 0.20 0.22

A 1352 1342 1342 1351 1336

shift 40 41 42 41 49

1392 1383 1384 1392 1385

B: before water adsorption; A: after water adsorption. Table 4 Activity for n-butane isomerization at 250 C. Catalyst k1 103 (h1 ) 12.8 13.1 18.3 17.7 14.4 13.4 3.0 43.1 SZ SZA-1 SZA-2 SZA-3 SZA-4 SZA-5 SZA-6a SZFM
a

Table 3 Activity for n-butane isomerization at 35 C. Catalyst k1 103 (h1 ) SZ SZA-1 SZA-2 SZA-3 SZA-4 SZA-5 SZA-6a SZFM
a

Activity

Conversion (%) 2 min 38.6 27.5 26.0 28.0 27.5 26.2 20.7 26.1 10 min 30.0 24.0 24.0 27.0 24.6 25.1 18.8 13.3 60 min 15.4 22.5 21.0 26.5 19.9 20.1 15.2 1.5 120 min 12.1 21.0 20.0 25.0 19.5 19.9 13.7 180 min 11.8 19.5 19.0 25.0 19.1 19.5 13.2 360 min 11.5 16.0 19.0 25.0 19.1 18.5 13.0

40.1 41.0 57.2 55.3 45.2 41.8 9.4 134.8

The calcination temperature for this sample is 750 C, and for all the others 650 C.

The calcination temperature for this sample is 750 C, and for all the others 650 C.

stretching frequency and the bond order and partial charge on oxygen atom calculated according to equations in the literature [23,25] for the SZ, SZA and SZFM catalysts are listed in table 2. The data show that the acid strengths of the SZA catalysts are almost identical with that of the SZ catalyst, whereas the acid strength of SZFM is slightly higher than those of the others. 3.2. Isomerization reaction The n-butane isomerization activities of the SZ and SZA catalysts running on stream for about 2 d at 35 C are listed in table 3. Since the isomerization of n-butane on the catalysts at 35 C follows the rate law of a rst-order reversible reaction [10], the activities are given in terms of the forward and backward rate constants, k1 and k1 . The Al-promoted catalysts with 0.510.0 mol% Al2 O3 are more active than the unpromoted SZ catalyst. The isomerization activity of the Al-promoted catalysts increases with the Al2 O3 content up to 1.53.0 mol% and then decreases as the Al2 O3 content is further increased. The isomerization of n-butane at low temperature can be used as a test reaction for superacidity, comprising the effect of acid strength and density of acid sites [10]. In the case of SZA catalysts, the promotion in isomerization activity is probably caused by an increase in the number of surface acid sites rather than in acid strength, if taking the above IR results and the sulfate contents of the catalysts into consideration. The isomerization activity of SZFM at 35 C is 23 times higher than those of the others. Such a great difference in activity cannot be explained by the slight increase in acid strength of SZFM. The pro-

moting effect of SZFM is more reasonably ascribed to an enhanced surface concentration of C8 intermediate caused by the presence of transition metal oxides since the isomerization proceeds via a bimolecular mechanism at low temperature as suggested in the literature [17,19]. The major reaction product of n-butane isomerization at 250 C is isobutane, and the main by-products are propane and isopentane. The selectivity to isobutane for all the catalysts is above 95%. The variation of the conversion of n-butane at 250 C with time on stream for the SZ, SZA and SZFM catalysts is given in table 4. Although the initial conversion of SZ is the highest, it decreases rapidly with time on stream. The steady state conversion of SZ after running for 6 h is reduced by more than 3 times. Under the same reaction conditions, the initial conversion of SZFM is lower than that of SZ and it deactivates even more rapidly than SZ. Previous reaction studies indicate that at low temperature and in the absence of hydrogen butane conversion on SZ and SZFM catalysts occurs through a complex bimolecular process involving C8 intermediates [26,27], whereas at high temperature and in the presence of hydrogen a monomolecular mechanism predominates [28]. The extraordinary change in catalytic behavior of SZFM under the two different reaction conditions seems to be consistent with these observations. A bimolecular mechanism operates over the catalysts at 35 C, but at 250 C the classical monomolecular mechanism is still valid. The disappearance of the positive effect of Fe and Mn on the reaction at 250 C can be interpreted as that the main function of the transition metal oxides promotors is to enhance the local concentration of olens near the acid sites rather than the

104

Z. Gao et al. / n-Butane isomerization over Al-promoted sulfated zirconia Table 5 Coke deposition on various catalysts. Catalyst SZ SZA-1 SZA-2 SZA-3 SZA-3 SZA-4 SZA-5 SZFM Reaction time (h) 6 6 6 6 120 6 6 6 Coke (wt%) 0.42 1.30 1.12 1.18 1.20 1.18 1.19 2.34

Figure 2. Steady state isomerization activity as a function of Al2 O3 content.

coke deposited on SZA catalysts after on stream for 6 h is higher than that on SZ. However, the amount of coke deposited on SZA-3 remained almost unchanged from 6 h to 120 h, showing that the initial drop in activity during 02 h was probably caused by catalyst coking and after that coking and deactivation slowed down. The high steady state activity of SZA-3 catalyst is probably associated with an higher amount of active acid sites left on the surface after the initial period. Coking on SZFM catalyst is more serious than that on others under the same conditions, which explains the rapid deactivation of the catalyst in reaction as shown in table 4. The increased surface concentration of olens on SZFM is probably responsible for the acceleration of coke formation. 3.3. Microcalorimetric measurement

Figure 3. Long-term test of SZA-3 catalyst.

acid strength of the acid sites [19]. The initial conversions of SZA catalysts are almost identical to SZFM but they deactivate much more slowly than SZ and SZFM, which is rather extraordinary for sulfated oxide catalysts. The effect of varying Al2 O3 content on the steady state activity of the SZA catalysts after on stream for 6 h is shown in gure 2. A maximum steady state activity is observed at a loading of 3 mol% Al2 O3 (SZA-3). As compared with SZ, SZA-3 catalyst is 2.2 times more active at steady state. This is indeed very promising from a practical point of view. To investigate the stability of SZA-3 catalyst for longer terms, the reaction has been run at 250 C continuously for 120 h. As illustrated in gure 3, the initial conversion is 28% and it drops to 25% after 1 h, and from then on it keeps at 25% steadily without any observable trend of deactivation. Since our original reactant contains 20% isobutane the actual concentration of isobutane in the reaction product is 45%, which means that isomerization of n-butane has proceeded on SZA-3 at a level of 88% of its equilibrium conversion stably. Compared to other sulfated zirconia based catalysts, the advantage of SZA-3 is obvious. In view of its high isomerization activity and stability, SZA-3 can be considered as an excellent candidate for a commercial-scale n-butane isomerization catalyst. The coke deposited on the catalysts after reaction was analyzed. The results are given in table 5. The amount of

The catalysts were evacuated at 250 C before microcalorimetric measurements. Figure 4 shows the microcalorimetric results of NH3 adsorption at 150 C on SZ and SZA-3 catalysts. The differential heat of adsorption decreases with increasing NH3 coverage, indicating a distribution of acid site strengths in the catalysts [29]. The initial heats of NH3 adsorption on the strong acid sites of SZ and SZA-3 are all in the range of 140170 kJ/mol, but SZA-3 contains more acid sites with intermediate acid strengths, having differential heats of NH3 adsorption in the range of 125140 kJ/mol.

Figure 4. Differential heat of NH3 adsorption versus adsorbate coverage at 150 C for () SZ and () SZA-3 catalysts.

Z. Gao et al. / n-Butane isomerization over Al-promoted sulfated zirconia

105

the promoting effect of aluminum on sulfated zirconia is very different from that of the transition metals, such as Fe, Mn and Ni. The high isomerization activities at low reaction temperature of the catalysts promoted by the addition of transition metals can be explained in terms of a bifunctional mechanism in which the metal promoters are responsible for an enhancement in the surface concentration of olens [19]. The decrease in activity of these catalysts at higher temperatures and in the presence of hydrogen can be interpreted as that hydrogen suppresses the formation of unsaturated intermediate species [31] or a monomolecular mechanism predominates under such conditions [28]. In contrast, the addition of Al to sulfated zirconia helps to stabilize the surface sulfate complex on the oxide and increases the number of effective acid sites on the catalyst for n-butane isomerization. The remarkable activity and stability of the Al-promoted catalyst under H2 at higher temperatures are caused by an appropriate distribution of acid site strengths and an enhanced number of acid sites with intermediate acid strengths. Acknowledgement
Figure 5. Histograms of acid strength distributions for (a) SZ and (b) SZA-3 catalysts.

We thank Professor Jianyi Shen and Dr. Mai Tu for assisting in microcalorimetric studies of the catalysts. References

Table 6 Microcalorimetric results of the distribution of acid site strengths. Catalyst Total SZ SZA-3 327.9 510.2 Acid sites (mol/g) 40125 kJ/mol 255.8 350.5 125140 kJ/mol 26.4 113.0 140170 kJ/mol 45.7 46.7

Histograms and data of the distribution of acid site strengths are shown in gure 5 and table 6. The total acid sites with differential adsorption heat of NH3 above 40 kJ/mol for SZ and SZA-3 are 327.9 mol/g and 510.2 mol/g, which amount to 80% and 64% of the surface sulfate in the catalysts (as seen in table 1), respectively. The acid strengths of the acid sites on SZ are more evenly distributed, whereas SZA-3 possesses a greater number of acid sites with differential heats between 125 and 140 kJ/mol. The number of acid sites with differential heats in the range of 125 to 140 kJ/mol for SZA-3 is 4.3 times greater than that for SZ. Earlier studies on sulfated zirconia catalysts [30] have shown that the acid sites with intermediate acid strengths, namely with differential heats of NH3 adsorption between 125 and 140 kJ/mol, are active for n-butane isomerization. Hence, the abundance of these acid sites on SZA-3 explains its extraordinarily high catalytic activity and stability for n-butane isomerization reaction. On the other hand, the strong acid sites on the catalysts with differential heats above 140 kJ/mol must be also involved in the n-butane isomerization, but they are probably deactivated rapidly in the initial period of the reaction. From the results of this study it can be concluded that

[1] G.A. Olah, C.S.K. Prakash and J. Sommer, in: Superacids (Wiley, New York, 1985). [2] M. Hino, S. Kobayashi and K. Arata, J. Am. Chem. Soc. 101 (1979) 6439. [3] M. Hino and K. Arata, J. Chem. Soc. Chem. Commun. (1979) 1148. [4] M. Hino and K. Arata, J. Chem. Soc. Chem. Commun. (1980) 851. [5] H. Matsuhashi, M. Hino and K. Arata, Chem. Lett. (1988) 1027. [6] J.C. Yori, J.C. Luy and J.M. Parera, Appl. Catal. 46 (1989) 103. [7] F. Garin, D. Andriamasinoro, A. Abdulsamad and J. Sommer, J. Catal. 131 (1991) 199. [8] F.R. Chen, G. Coudurier, J.-F. Joly and J.C. Vedrine, J. Catal. 143 (1993) 616. [9] J.H. Lunsford, H. Sang, S.M. Campbell, C.H. Liang and R.G. Anthony, Catal. Lett. 27 (1994) 305. [10] Z. Gao, J.M. Chen, W.M. Hua and Y. Tang, Stud. Surf. Sci. Catal. 90 (1994) 507. [11] C. Morterra, G. Cerrato, F. Pinna, M. Signoretto and G. Strukul, J. Catal. 149 (1994) 181. [12] F.T.T. Ng and N. Hovat, Appl. Catal. A: General 123 (1995) L197. [13] M.R. Gonzalez, J.M. Kobe, K.B. Fogash and J.A. Dumesic, J. Catal. 160 (1996) 290. [14] C.X. Miao, W.M. Hua and Z. Gao, Chinese J. Catal. 18 (1997) 13. [15] C.Y. Hsu, C.R. Heimbuch, C.T. Armes and B.C. Gates, J. Chem. Soc. Chem. Commun. (1992) 1645. [16] A. Jatia, C. Chang, J.D. MacLeod, T. Okubo and M.E. Davis, Catal. Lett. 25 (1994) 21. [17] A. Adeeva, J.W. de Haan, J. Janchen, G.D. Lei, V. Schunemann, L.J.M. van de Ven, W.M.H. Sachtler and R.A. van Santen, J. Catal. 151 (1995) 364. [18] C.X. Miao and Z. Gao, Chem. J. Chinese Univ. 18 (1997) 424. [19] M.A. Coelho, D.E. Resasco, E.C. Sikabwe and R.L. White, Catal. Lett. 32 (1995) 253. [20] C.X. Miao, W.M. Hua, J.M. Chen and Z. Gao, Catal. Lett. 37 (1996) 187.

106

Z. Gao et al. / n-Butane isomerization over Al-promoted sulfated zirconia [27] V. Adeeva, G.D. Lei and W.M.H. Sachtler, Appl. Catal. A: General 118 (1994) L11. [28] F. Garin, L. Seyfried, P. Girard, G. Maire, A. Abdulsamad and J. Sommer, J. Catal. 151 (1995) 26. [29] B.E. Spiewak, B.E. Handy, S.B. Sharma and J.A. Dumesic, Catal. Lett. 23 (1994) 207. [30] K.B. Fogash, G. Yaluris, M.R. Gonzalez, P. Ouraipryvan, D.A. Ward, E.I. Ko and J.A. Dumesic, Catal. Lett. 32 (1995) 241. [31] X.M. Song and A. Sayari, Catal. Rev.-Sci. Eng. 38 (1996) 329.

[21] C.X. Miao, W.M. Hua, J.M. Chen and Z. Gao, Science in China (Series B) 39 (1996) 406. [22] C. Morterra, G. Cerrato, C. Emanuel and V. Bolis, J. Catal. 142 (1993) 349. [23] T. Jin, T. Yamaguchi and K. Tanabe, J. Phys. Chem. 90 (1986) 4794. [24] Z. Gao, J.M. Chen and Y. Tang, Chem. J. Chinese Univ. 14 (1993) 658. [25] R.J. Gillespie and E.A. Robinson, Can. J. Chem. 41 (1963) 2074. [26] M.R. Guisnet, Acc. Chem. Res. 23 (1990) 392.

Das könnte Ihnen auch gefallen