Você está na página 1de 9

Applied Catalysis A: General 282 (2005) 163171 www.elsevier.

com/locate/apcata

Activity of CoMo/g-Al2O3 as a catalyst in hydrodesulfurization: effects of Co/Mo ratio and drying condition
Y.S. Al-Zeghayera, P. Sunderlandb, W. Al-Masrya, F. Al-Mubaddela, A.A. Ibrahima, B.K. Bhartiyac, B.Y. Jibrila,*
a

Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia b Department of Chemical Engineering, The University of Leeds, Leeds LS2 9JT, UK c Saudi Aramco, Riyadh, Saudi Arabia Received 1 July 2004; received in revised form 2 December 2004; accepted 9 December 2004 Available online 26 January 2005

Abstract A series of g-alumina-supported cobaltmolybdenum hydrodesulfurization catalysts have been prepared with different ratios of cobalt to molybdenum (0.01.0) and under different drying temperatures (323, 373 and 423 K). The active components ions were loaded on the alumina via an incipient wetness impregnation method. The catalysts were converted to active sulde form using dimethyldisulde. Hydrodesulfurization (HDS) of dibenzothiophene (DBT) was tested at 550683 K and 10 atm. Samples of the catalysts were characterized using SEM and XPS techniques. Changes in grain sizes and binding energy due to the cobaltmolybdenum interactions have been observed. The catalyst activities based on dibenzothiophene conversion rate constants and product (biphenyl BP and cyclohexylbenzene CHB) distributions, depended strongly on the Co/Mo ratios. CHB/BP ratio as high as 0.2 was obtained for Co/Mo of 0.20.4, while for higher Co/Mo ratios, no CHB was observed. The ratio of about 0.4 exhibited the highest rate constant. The drying temperature was found to improve the activities, with the lowest temperature showing the best performance in terms of dibenzothiophene degree of conversion. Such results suggest that cobalt and molybdenum are associated in a xed proportion and that their distribution on the support surface, which in turn improves catalyst performance, could be modied by the catalyst preparation conditions. # 2004 Elsevier B.V. All rights reserved.
Keywords: CoMo/g-Al2O3 catalysts; Cobalt/molybdenum ratios; Hydrodesulfurization; Dibenzothiophene; Dimethyldisulde; Co/Mo ratio; Drying condition; SEM; XPS

1. Introduction Sulfur reduction from gasoline using catalytic hydrodesulfurization (HDS) is a routine operation in petroleum rening processes. In this operation, sulfur-containing compounds such as thiophenes, dibenzothiophenes and mercaptans are removed. The removal is necessary in order to minimize the polluting effects of sulfur-containing gases such as SO2 that may be released to the atmosphere. In addition, many catalysts employed in the rening downstream operations could be poisoned by sulfur. This causes frequent shutdown and other operational difculties.
* Corresponding author. Tel.: +966 14676897; fax: +966 14678770. E-mail address: baba@ksu.edu.sa (B.Y. Jibril). 0926-860X/$ see front matter # 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.apcata.2004.12.021

Perhaps the most important challenge presented by sulfur is its environmental degradation. This has made it necessary for the US Environmental Protection Agency (EPA) to review the allowable sulfur content of fuels, ue gases and other petroleum products. Recently, new guidelines have been established that, for instance, restrict the amount of sulfur allowed in diesel fuels to 1.3 102 g/dl3 by the year 2010 [1]. This requires reduction of about 97% of the current level. Achieving this limit will require major improvement or redesign of the catalysts employed for HDS. Alumina-supported CoMo catalysts have been well established for hydrodesulfurization of oil fractions. The structures of both precursor and nal catalysts have been extensively studied and the nature of the active sites has been proposed [2]. Studies were done on increasing the dispersion

164

Y.S. Al-Zeghayer et al. / Applied Catalysis A: General 282 (2005) 163171

of the active material on a support matrix by change in preparation method [36]. Supports such as carbon [79], aluminosilicates [1012] and zeolites [13] were proposed. Recently, a direct relationship between the preparation parameters and the physicochemical and catalytic properties of alumina-supported CoMo in HDS has been demonstrated. The catalyst prepared using equilibrium deposition ltration, rather than conventional impregnation, was shown to have 3043% higher activity [14]. The increase in performance was on account of higher coverage of molybdenum phase. Similarly, higher coverage and better performance were observed when meso-structured g-alumina with lath-like framework morphology was employed as a support for CoMo catalysts [15]. Using the impregnation method, Inamura et al. have shown that the hydrodesulfurization activities could increase by up to 50% when chelating agent was added to precursor solutions [16]. The most frequently reported catalyst for hydrodesulfurization is sulded supported cobaltmolybdenum. This catalyst still needs further studies to reduce its rapid deactivation and wasteful hydrogen consumption by hydrocracking and hydrogenation reactions. This could be achieved through modication of catalyst preparation conditions, as demonstrated by others [1416]. The objective of this work, therefore, was to study the effects of the cobaltmolybdenum composition and drying condition in the catalyst preparation on the active components distribution on the support surface and hence on the hydrodesulfurization activities. Hydrodesulfurization of dibenzothiophene (DBT) was the reaction chosen to evaluate the activity of the catalysts. Dibenzothiophene can be regarded as a typical sulfur-containing hydrocarbon present in high-boiling oil fraction petroleum or coalderived liquids. Such study will lead to better understanding of catalyst structure and hydrodesulfurization performances.

de, (CH3)2S2. All chemicals were of analytical grade quality, supplied by Aldrich Chemical Company. A series of sulded cobalt molybdate catalysts was prepared by co-impregnation of g-Al2O3 with solutions containing appropriate amounts of cobalt nitrate and/or ammonium molybdate. Amount of salts enough to wet the support (2.2 ml); were dissolved in distilled water. The concentrations of the solutions were adjusted to give samples containing 15.4 wt.% of MoO3 and varying amounts of cobalt to produce samples with Co/Mo ratios of 0.01.0, but maintaining a xed amount of active components on the support. The support was dried at 373 K for 2 h. It was then co-impregnated with suitable solutions for 30 h and dried at 393 K in air for 9 h. Another sample of catalyst with Co/Mo ratio of 0.4 was prepared, but dried at different temperatures (323, 373 and 423 K) for 10 h. Hereafter, these catalysts are referred to as D323, D373 and D423, respectively. All dried catalyst precursors were calcined at 873 K for 24 h in dry air after reaching the calcination temperature at a heating rate of 1 K/min. The calcination decomposes the precursor salts to non-active oxidic form. The oxide form was sulded to catalytically active form using a solution of dimethyldisulde (2 wt.% sulfur equivalent) in decalin solvent at 323 K until H2S breakthrough was observed. 2.2. Catalyst testing The catalyst evaluation was carried out by passing solutions containing dibenzothiophene at a concentration of 1 or 2 wt.% in the hydrogen donor solvent tetralin through a xed bed containing about 1 g of the catalyst particles. The use of a hydrogen donor solvent ensured an ample supply of hydrogen readily available for the reaction without a need for adding gas phase hydrogen. The reaction was conducted over a temperature range of 550683 K, at a pressure of 10 atm and at ow rates required to give measurable conversions, which were between 0.5 and 2 g/min. The catalyst was initially sulded before testing. When sulding was completed, the ow was switched to the solution containing dibenzothiophene in tetralin that was pumped from a storage tank using a pulse-free HPLC pump. The mixture was then preheated to the required temperature before passing to the temperature-controlled reactor; a stainless steel tube 15 cm in length and 4.5 mm in internal diameter containing crushed sieved catalyst particles. Quenching then took place before pressure reduction and analysis using a capillary column gas chromatograph tted with a ame ionization detector. The degree of conversion of DBT for a given set of operating conditions was taken as a measure of catalyst activity. 2.3. Catalyst characterizations For a sample of the catalysts, particle grain sizes were observed by scanning electron microscopy (Camscan SEM)

2. Experimental 2.1. Catalysts preparation The catalyst support was g-alumina supplied by Norton/ Hydronyl (UK). Its chemical composition was: Al2O3 > 99.85, SiO2 < 0.09, Na2O < 0.015 and Fe2O3 < 0.06. Its physical properties were: apparent porosity = 6575%, bulk (particle density) = 1.01.2 g/cm3, apparent specic gravity = 3.33.5, surface area = 200240 m2/g and total pore volume = 0.570.67 cm3 Hg/g. The support was ground and sieved to give a size range of 3672 mesh before being impregnated with catalyst precursor solution via incipient wetness method. The chemicals used in the catalyst preparation were ammonium molybdate(VI) tetrahydrate (NH4)6Mo7O244H2O and cobalt(II) nitrate hexahydrate Co(NO3)26H2O. The materials used in the catalyst testing were dibenzothiophene, tetralin, decalin and dimethyldisul-

Y.S. Al-Zeghayer et al. / Applied Catalysis A: General 282 (2005) 163171

165

interfaced to a digital image capture system. Particles of the calcined oxide from each of the catalysts dried at different temperatures were mounted in resin, ground smooth with emery paper, polished with diamond paste and then coated with carbon prior to examination using a Cam Scan Series-4 scanning electron microscope. When operating in backscattering electron image (BEI) mode, the instrument was able to provide information about the microstructure of the catalysts. The distributions of cobalt and molybdenum within the particles were measured using the energy dispersive X-ray analysis (EDAX) unit in the Cam Scan. Energy spectra were obtained by using spot mode analysis combined with the EDAX unit. The XPS spectra of two samples (Co/Mo of 0.0 or 0.4) were measured and recorded on a VG Escalab 200-C spectrometer using Mg Ka radiation (1253.6 eV). The base pressure in the analysis chamber was kept in the range 5 1010 to 1 109 mbar. Energy scales were referred to the C 1s peak, which was assumed to have a binding energy of 285.0 eV. The sample treatments were all carried out in a high-pressure gas cell directly connected to an ultra high vacuum system. The samples were pre-treated in synthetic air (1.1 bar, 773 K) in order to obtain comparable and reproducible surface conditions. All reduction treatments were carried out with 20% H2 in N2 (1.1 bar, 773 K).

Table 2 Analysis of Mo 3d and S 2s XPS signals for catalysts C1 (Co/Mo = 0.0) and C3 (Co/Mo = 0.4) C1 (Co/Mo = 0.0) Mo1 3d5/2 (eV) FWHM Intensity Area Mo2 3d5/2 (eV) FWHM Intensity Area Mo1 3d3/2 (eV) FWHM Intensity Area Mo2 3d3/2 (eV) FWHM Intensity Area S 2s (eV) FWHM Intensity Area 229.8 2.3 96.3 39.6 232.8 2.7 28.0 13.0 232.9 2.3 57.0 22.5 235.9 2.9 18.9 9.6 227.0 2.9 25.0 13.0 C3 (Co/Mo = 0.4) 229.4 1.8 98.0 39.0 232.4 2.6 22.0 12.1 232.6 2.0 58.0 25.6 235.6 2.8 15.0 9.2 226.5 2.5 24.0 13.2

3. Results and discussion 3.1. Catalyst characterizations For the XPS analysis, two catalysts were chosen to illustrate the changes in the surface composition due to addition of cobalt. The catalysts tested are: (i) spent sulded form of the catalyst that contained no cobalt and exhibited the lowest activity and (ii) spent sulded form of the catalyst (Co/Mo of 0.4), which exhibited the highest activity. The results are expressed as percent atoms in the surface layer as presented in Table 1. All binding energies (BE) are referred to C 1s (BE = 285.0 eV). The data for Mo 3d5/2 and Mo 3d3/2 are shown in Table 2 and Fig. 1. The elements detected in all samples were C, O, Mo, Al and S. The Mo spectra indicate
Table 1 XPS data for catalysts C1 (Co/Mo = 0.0) and C3 (Co/Mo = 0.4) C1 (Co/Mo = 0.0) Atom (%) Mo1 Mo2 O Al S1 S2 Co DE1 DE2 4.12 3.01 53.90 38.04 9.51 BE (eV) 229.8 232.8 532.5 75.5 163.2 616.8 C3 (Co/Mo = 0.4) Atom (%) 4.62 3.08 46.15 31.35 12.69 0.77 1.35 BE (eV) 229.4 232.4 532.3 75.3 162.8 171.3 779.6 66.6 66.6

the presence of more than one chemical state. Curve tting indicates two states, having BE value of 229.6 0.2 and 232.6 0.2 eV. In addition, observation of the Mo 3d envelope suggests that the presence of low levels of other

DE1 = Co 2pS 2p and DE2 = Mo1 3dS 2p.

Fig. 1. XPS pattern of Mo 3d envelopes: (a) catalyst C1 (Co/Mo = 0.0); (b) catalyst C3 (Co/Mo = 0.4).

166

Y.S. Al-Zeghayer et al. / Applied Catalysis A: General 282 (2005) 163171

chemical states may not be ruled out. The BE for the peaks tted could be assigned to Mo4+ and Mo6+, respectively. Also the gure suggests overlaps of XPS patterns of Mo 3d3/ 2 and Mo 3d5/2 with that of S 2s. The BE for the S 2p1 (S1) is 163.0 0.2 eV, which is similar to that for sulfate. The level of sulfur detected together with the binding energy measured, suggest Mo1 to be MoS2 while Mo2 is likely to be present as MoO2. The gure also suggests evidence of a second sulfur state (S2) in the second sample. It is difcult to distinguish among the oxide phases of cobalt and between CoMoS and Co9S8 from the values of the Co 2p3/2 energy alone. In the unsupported catalysts, the distinction can be made on the basis of the Co 2p3/2 peak shapes differences [1719]. In alumina-supported catalysts, the presence of Co dissolved in alumina complicated the analysis of peak shapes and so the distinction between Co MoS and Co9S8 is best made on the basis of the difference between the Co 2p3/2 binding energy and the S 2p3/2 binding energy [19]. The binding energy differences were used. Two binding energy differences DE1 and DE2 have been obtained by subtracting the binding energy of the S 2p3/2 peak from the binding energy of the Co 2p3/2 peak and from the binding energy of the Mo 3d5/2 peak, respectively. The values are shown in Table 1. The catalyst samples with DE1 = 617.0 eV were those where Co was present mainly in the CoMoS phases as observed by MES (Mossbauer emission spectroscopy) [19]. The catalysts where Co was present mainly as Co9S8, have DE1 = 616.3 eV. DE1 for catalyst (Co/Mo = 0.4) is equal to 616.8 eV, which is closer to 617.0 eV. It is observed that for catalysts C1 (Co/Mo = 0.0) and C3 (Co/ Mo = 0.4), which were prepared and analyzed under the same conditions, the second sulfur state was found only in catalyst C3, thus seeming to indicate a modied CoMoS site. It was also observed that Mo4+/Mo6+ of 1.5 for catalyst C3 is slightly higher than that for catalyst Cl (1.369), which indicates the ratio at which cobalt increases the reduction and/or sulding of molybdenum, in agreement with an earlier report [20]. X-ray maps were plotted to show the distribution of the elements in the catalyst dried at different temperatures 323, 373 and 423 K, referred to as D323, D373 and D423, respectively. Scanning electron microscopy and X-ray maps of cobalt and molybdenum are shown for the catalysts in Figs. 24. Energy spectra of the catalyst particles were obtained by using spot mode analysis combined with the (EDAX) units within the scanning electron microscope. These spectra give valuable information about the distribution. From the gure, it may be observed that: (1) Two regions can be recognized with a sharp interface between them: (i) a bright region inside the catalysts, which contains a higher concentration of cobalt and molybdenum, and (ii) an outer region of lower concentration of both components. The boundary between the two moved

towards the center of the particle with increasing drying temperature; (2) For samples with low drying temperature (323 K), the results indicate a uniform dispersion of cobalt and molybdenum. A much less uniform distribution of cobalt and molybdenum is observed on the particle with high drying temperature (423 K). 3.2. Catalyst testing Catalysts of different compositions were tested to study the interaction between cobalt and molybdenum in determining the performance of the catalysts. The degree of conversion of dibenzothiophene at a particular temperature and ow rate (0.6 0.2 ml/min) was used as a measure of activity. The reaction was conducted at 633, 653 and 683 K. The range is limited due to the wide variation in the activities of the catalysts. Fig. 5 shows the changes in conversion with the cobaltmolybdenum compositions. Biphenyl BP and cyclohexylbenzene CHB are the two products identied. The conversions have been scaled to compare the conversions of different compositions relative to the most active sample. The promoting effect of cobalt could be observed from the gure. The activity passed through a maximum at Co/Mo values in 0.20.5 range. In another set of experiments, conversion rates were measured as functions of residence time at given temperatures. From these, pseudo rst-order rate constants were obtained by measuring the gradient corresponding to initial rate conditions on conversion contact time (W/F) plot. The results, as shown in Fig. 6, indicate that optimum performance will occur at a cobaltto-molybdenum ratio of about 0.4. The products distribution is also affected by the composition. Table 3 shows the amounts of BP and CHB produced at 633 K for the compositions. It could be observed that the promotion of the hydrogenolysis reaction was greater than that of the hydrogenation reaction. For the Co/Mo ratios of 0.2 or 0.4 the CHB/BP ratio as high as 0.2 was obtained. For the higher Co/Mo ratio, which exhibited lower conversion, no CHB was observed. As shown by the XPS results, two catalysts: C1 (Co/ Mo = 0.0) and C3 (Co/Mo = 0.4), exhibited the same DE2 (Mo1 3d5/2S1 2p3/2) values. This indicates that the nature of the MoS2 phase on C1 and C3 is the same. Comparing the molybdenum lines (Fig. 1 and Table 2) showed that adding cobalt has an effect on the line widths. In the catalyst with cobalt, the line widths were narrower than in the catalyst with only molybdenum, especially for Mo1 and S 2s signals, indicating less degree of dispersion. The presence of a second sulfur state in the second sample (Co/Mo = 0.4) could be associated with the more active phase. The oxide state of molybdenum is still present in both samples but the Mo4+/Mo6+ ratio is higher by about 9% in the catalyst containing cobalt. This indicates that cobalt increased the reduction or sulding of catalyst. The increase may be

Y.S. Al-Zeghayer et al. / Applied Catalysis A: General 282 (2005) 163171

167

Fig. 2. BEI micrograph showing phases in catalyst D323 (dried at 323 K) (top) and X-ray map distribution for molybdenum and cobalt elements in the catalyst (bottom).

responsible for the high activity exhibited by C3 (Co/ Mo = 0.4). During impregnation of a solid adsorbent by metal ions from solution and subsequent drying, metal species are adsorbed and deposited on the surface. The conditions under which these steps take place inuence the distribution of metal ions throughout the internal structure of the support [2124]. Therefore, we consider it of interest to study the effect of the distribution on the performance of the catalyst. In order to demonstrate the effect of drying conditions on the performances, catalysts (with Co/Mo = 0.4) were prepared using different drying temperatures (323, 373 and 423 K). This may shed light on the effect of drying conditions on the distributions of the two types of cations within the catalysts and how this, in turn, may affect the activity and selectivity.

The extent of conversion of dibenzothiophene at a particular temperature and ow rate was measured. The results, as shown in Fig. 7, demonstrate that the activity decreased appreciably with increasing drying temperature. The results show that at 553 K the conversion obtained using D323 (dried at 323 K) was 12.6% whilst with D373 (dried at 373 K) only 0.61% conversion was achieved and with D423 (dried at 423 K) there was no conversion at all. D423 activity was very low over the whole temperature range studied compared with the other two. For instance at 613 K, conversion was 0.4%, while the value was more than 95% when D323 was used. The activity decreased with increasing drying temperature. The drying temperature, as shown in Table 4, also affects the product distributions. At similar levels of DBT conversions, the selectivity to cyclohex-

168

Y.S. Al-Zeghayer et al. / Applied Catalysis A: General 282 (2005) 163171

Fig. 3. BEI micrograph showing phases in catalyst D373 (dried at 373 K) (top) and X-ray map distribution for molybdenum and cobalt elements in the catalyst (bottom).

ylbenzene decrease as the drying temperature decreased. This suggests that the low drying temperature led to a catalytic phase to favor path 1 rather than 10 and/or 2 in the reaction networks as suggested in Scheme 1. The product ratios (CHB/BP) at 590 K and with conversion of 32.5 1.1% were found to be 0.07 with the catalyst D323 and to be 0.119 with the catalyst D373. The catalyst D323 has the highest activity and selectivity. The high CHB/ BP value of 0.222 for D373 may be due to experimental error. The D423 was found to be inactive when tested at the same conditions as used for D323 and D373. As indicated earlier, the effect of the drying rate depends on the adsorptivity of the species being adsorbed from solution. Few works appear to have been done on the adsorption on molybdenum and the inuence of drying

conditions on the molybdenum distribution [25,26]. In one report, it was concluded that the adsorption process was slow [25]. It was also shown that the mildest drying conditions gave better macroscale distribution and activity [26]. The results we have obtained show that the drying rate has a strong effect on both the distribution of the cobalt and molybdenum and the activity of the catalyst. Analysis of the three catalysts shows that a uniform distribution of both metals could be achieved with slow drying rate (323 K). Increasing the drying rate led to an increase in concentration of adsorbates towards the center of the pellet (Figs. 24). In the literature, inhomogeneous distribution of molybdenum in the alumina pellets in the absence of cobalt was obtained with fast drying [27]. The maximum activity was obtained using the catalyst dried at low temperature that coincided

Y.S. Al-Zeghayer et al. / Applied Catalysis A: General 282 (2005) 163171

169

Fig. 4. BEI micrograph showing phases in catalyst D423 (dried at 423 K) (top) and X-ray map distribution for molybdenum and cobalt elements in the catalyst (bottom).

with a uniform distribution of the active metals. From the point of view of the physical properties of the catalysts, it was concluded that for the supported catalysts, the active surface area per unit weight of catalyst drops rapidly as the distribution becomes more inhomogeneous [28]. The color of the catalyst (D423) dried at 423 K was graygreen instead of the gray-purple coloration seen with the other two catalysts, thus indicating some change in the nature of the phases present. The green color has been associated with the formation of CoMoO4, which is hardly suldable, while a gray-purple color is indicative of high dispersion of Co and Mo in the solid. The formation of CoMoO4 was shown to lead to lower activity in HDS of thiophene [29]. The optimum composition of MoO3 alone for HDS of DBT has been shown to be about 15% [26]. The optimum concentration of CoO when mixed with molyb-

dena may be different. We observed that the distribution of the active metals salts towards the center resulted in the concentration of the active components being higher than the optimum near the center, while the concentration was lower than the optimum near the external surface. High concentrations of MoO3 and CoO result in the formation of inactive phases such as Al2(MoO4)3 in the subsequent calcination stage [30] and Co3O4 [21,31,32], whereas at low concentrations the amounts of active phases formed are small. Both effects may cause a decrease in activity. Few experimental observations on adsorbate redistribution during drying have been reported and there are inconclusive opinions about the signicance of such redistribution [25 28,33]. Our results, whereby the cobalt to molybdenum ratio was constant throughout and in which uniform distribution was obtained by drying at low temperatures and egg yolk

170

Y.S. Al-Zeghayer et al. / Applied Catalysis A: General 282 (2005) 163171 Table 3 DBT conversions and product distributions for different Co/Mo ratios compositions at 633 K and reaction concentration of 0.542 104 mol DBT/g feed Co/Mo ratio DBT reacted (%) 0.2 39.5 0.4 53.9 0.5 18.2 0.078 0.000 0.000 1.0 13.3 0.063 0.000 0.000

Distribution, concentration 104 mol/g feed BP 0.155 0.211 CHB 0.026 0.032 CHB/BP 0.167 0.151

Table 4 DBT conversions and products distributions for 323 and 372 K drying temperatures (TD) at 590 K, reaction concentration of 0.542 104 mol DBT/g feed Fig. 5. DBT conversions with Co/Mo ratios at different temperatures. TD = 323 K DBT reacted (%) Flow rate (g/min) 94.75 0.810 90.02 1.070 67.23 1.710 48.16 2.480 0.181 0.023 0.127 10.21 1.780 0.039 0.009 0.222 31.39 3.410 0.134 0.009 0.070 5.17 2.920 0.013 0.000 0.000

Distribution, concentration 104 mol/g feed BP 0.369 0.349 0.281 CHB 0.066 0.062 0.033 CHB/BP 0.179 0.178 0.117 TD = 373 K DBT reacted (%) Flow rate (g/min) 33.58 0.586 23.74 1.017 15.13 1.145

Distribution, concentration 104 mol/g feed BP 0.134 0.091 0.059 CHB 0.016 0.009 0.000 CHB/BP 0.119 0.099 0.000

Fig. 6. Rate constants for conversion of DBT at 633 K with Co/Mo ratios.

Scheme 1.

distribution was obtained for more rapid rates of solvent evaporation, are in agreement with an earlier report [26]. This was evidenced by the X-ray map obtained by scanning electron microscopy and energy dispersive X-ray analyses. In line with this observation, the catalyst produced by drying at 323 K, which produced a uniform distribution of the active components, turned out to have an optimum performance among the samples tested.

4. Conclusion
Fig. 7. DBT conversions with reaction temperatures for catalysts dried at different temperatures.

The performances of alumina-supported cobaltmolybdenum catalysts in hydrodesulfurization have been shown to

Y.S. Al-Zeghayer et al. / Applied Catalysis A: General 282 (2005) 163171

171

depend on the distribution of the active components in the catalyst. Such variation in distribution could be as a result of either difference in the amounts of the active components in the precursor or different drying temperatures during preparation. The selectivities to biphenyl and cyclohexylbenzene also varied with distribution of the active components. As shown by XPS analysis, the presence of cobalt increased the rate of reduction and sulding of molybdenum in the catalyst. The performance obtained on the cobalt-containing sample depended on the Co/Mo ratio in the catalyst. Furthermore, as evidenced by X-ray map, the catalyst with uniform distribution of active components exhibited the best performance.

References
[1] Environmental Protection Agency, Federal Register 65 (2000) 35430. [2] H. Topsoe, B.S. Clausen, Catal. Rev.-Sci. Eng. 26 (1984) 395. [3] R. Prins, V.H.J. de Beer, G.A. Somorjai, Catal. Rev.-Sci. Eng. 31 (1989) 1. [4] A.N. Startsev, Catal Rev.-Sci. Eng. 37 (1995) 353. [5] H. Topsoe, B.S. Clausen, F.E. Massoth, in: J.R. Anderson, M. Boudart (Eds.), Hydrotreating Catalysis, Series in Catalysis Science and Technology, vol. 11, 1996. [6] B. Delmon, G.F. Froment, Catal. Rev.-Sci. Eng. 38 (1996) 69. [7] E.J.M. Hensen, H.J.A. Brans, G.M.H.J. Lardinois, V.H.J. de Beer, J.A.R. Van Veen, R.A. Van Santen, J. Catal. 192 (2000) 98. [8] H. Farag, D.D. Whitehurst, K. Skanishi, I. Mochida, Catal. Today 50 (1999) 9. [9] M.J. Ledoux, O. Michaux, G. Agostini, P.J. Panissod, Catalysts 102 (1986) 275. [10] Y. Yue, Y. Sun, Z. Gao, Catal. Lett. 47 (1997) 167. [11] C. Song, K.M. Reddy, Appl. Catal. A 176 (1999) 1.

[12] J. Ramirez, R. Contreras, P. Castillo, T. Klimova, R. Zarate, R. Luna, Appl. Catal. A 197 (2000) 69. [13] Y. Okamoto, Catal. Today 39 (1997) 45. [14] Ch. Papadopoulou, J. Vakros, H.K. Matralis, Ch. Kordulis, A. Lycourghiotis, J. Colloid Interface Sci. 261 (2003) 146. [15] R.W. Hicks, N.B. Castagnola, Z. Zhang, T.J. Pinnavaia, C.L. Marshall, Appl. Catal. A 254 (2003) 311. [16] K. Inamura, K. Uchikawa, S. Matsuda, Y. Akai, Appl. Surf. Sci. 121 122 (1997) 468. [17] R.A. Walton, J. Catal. 44 (1976) 335. [18] Y. Okamoto, H. Nakamo, T. Shimokawa, T. Imanaka, S. Teranishu, J. Catal. 50 (1977) 447. [19] I. Alstrup, I. Chorkendoref, R. Candia, B.S. Clausen, H. Topsoe, J. Catal. 77 (1982) 397. [20] P. Ratnasamy, A.V. Ramaswamy, K. Banerjee, D.K. Sharma, N. Ray, J. Catal. 38 (1975) 19. [21] B. Delmon, A. Houalla, Preparation of Catalysts II, Elsevier, Amsterdam, 1979. [22] M. Komiyama, Cat. Rev.-Sci. Eng. 27 (1985) 341. [23] S.Y. Lee, R. Aris, Cat. Rev.-Sci. Eng. 27 (1985) 207. [24] A.V. Neimark, L.I. Keifez, V.B. Fenelonov, Ind. Eng. Chem. Prod. Res. Dev. 20 (1981) 439. [25] J. Sonnemans, P. Mars, J. Catal. 31 (1973) 209. [26] O. Ochoa, R. Galiasso, P. Andreu, Preparation of Catalysts II, Elsevier, Amsterdam, 1979. [27] R. Galiasso, Preparation of Catalysts II, Elsevier, Amsterdam, 1979. [28] J.W. Geus, in: G. Poncelet, P. Grange, P.A. Jacobs (Eds.), Preparation of Catalysts III, Elsevier, Amsterdam, 1983. [29] J.A. Rob van Veen, E. Gerkema, A.M. van Der Kraan, P.A.J.M. Hendriks, H. Beens, J. Catal. 133 (1992) 112. [30] N. Giordano, J.C.J. Bart, A. Vaghi, A. Castellan, G. Matinohi, J. Catal. 36 (1975) 81. [31] F.E. Massoth, Adv. Catal. 27 (1978) 265. [32] C. Wivel, B.S. Clausen, R. Candia, S. Morup, H. Topsoe, J. Catal. 87 (1984) 497. [33] R.W. Maatman, C.D. Prater, Ind. Eng. Chem. 49 (1957) 253.

Você também pode gostar