Sei sulla pagina 1di 7

Available online at www.sciencedirect.

com

Journal of Membrane Science 316 (2008) 4652

Metal doped silica membrane reactor: Operational effects of reaction and permeation for the water gas shift reaction
Scott Battersby, Mikel C. Duke, Shaomin Liu, Victor Rudolph, Jo ao C. Diniz da Costa
Films and Inorganic Membrane Laboratory (FIMLab), Division of Chemical Engineering, The University of Queensland, Brisbane, QLD 4072, Australia Received 26 June 2007; received in revised form 1 October 2007; accepted 5 November 2007 Available online 22 November 2007

Abstract In this work, we investigate the performance of metal (Cobalt) doped silica membranes in a membrane reactor (MR) conguration for the low temperature water gas shift (WGS) reaction. The membranes were hydrostable and showed activated transport even after 2 weeks exposure to steam. High CO conversions resulted in the H2 and CO partial pressures in the reaction chamber moving in opposite directions, thus favouring H2 /CO separation to treble (515) from 150 to 250 C. On the other hand, the separation of H2 /CO2 remained relatively low (24) as the driving force for diffusion or partial pressure of these gases remained equal in the reaction chamber irrespective of the extent of conversion. Below approximately 40% CO conversion, the MR is ineffective as the H2 driving force for permeation was so low that H2 /CO selectivity was below unity. Operating under equilibrium limited conversion (space velocities 7500 h1 ) conditions, very high conversions in excess of 95% were observed and there were no signicant advantages of the MR performance over the packed bed reactor (PBR). However, for higher throughputs (space velocities 38000 and 75000 h1 ) conversion is affected by the reaction rate, and relatively enough H2 is removed from the reactor through the membrane. Increasing temperature to 250 C as a function of the space velocity (75000 h1 ) allowed for the CO conversion in the MR to shift up to 12% as compared to the PBR. 2007 Elsevier B.V. All rights reserved.
Keywords: Metal doped silica; Membrane reactor; Water gas shift reaction; Gas separation and CO conversion

1. Introduction Hydrogen is widely seen as a clean energy carrier for future use in both transportation and electricity sectors. It is likely that, at least in the intermediate term, hydrogen will be derived from fossil fuels. Amongst the fossil fuels, coal is the most widely distributed and abundant. Many countries, based on foreign exchange and energy security issues, are considering coal gasication in association with carbon dioxide capture and storage. Large projects such as FutureGen (USA) and ZeroGen (Australia) represent early attempts to develop this technology into commercial readiness. The gasication of coal predominantly produces syngas (CO and H2 ) with some remaining hydrocarbons, CO2 and water [1]. The syngas requires further processing through the water gas shift (WGS) reaction, in order to maximise H2 production [2,3]. The reaction (Eq.

(1)) has been commercially used and developed over many decades. CO + H2 O CO2 + H2 H = 41.2 kJ mol1 (1)

Corresponding author. Tel.: +61 7 3365 6960; fax: +61 7 3365 4199. E-mail address: j.dacosta@eng.uq.edu.au (J.C.D.d. Costa).

This reaction is exothermic and the H2 product is thermodynamically favoured at low temperature, e.g. 175250 C and high steam ratios [4,5]. Two disadvantages arise: at low temperatures, the reaction kinetics slow, even using the preferred copper zinc alumina catalysts [5,6], and excess steam needs to be recycled. Typically, industrial reactors use a two-step shift to optimise conversion. A high temperature shift, operating between 300 and 450 C with a FeCr catalyst, provides most of the reaction, followed by a low temperature stage to maximise the conversion. While the two-step shift method provides both high reaction rates and high conversion, it requires large catalyst volumes and high steam recycling. An alternative strategy to optimise the WGS reaction is to employ a membrane reactor (MR) a unit integrating the reaction with product separation using a selective membrane. This operation can provide many benets over the conventional reactor

0376-7388/$ see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.memsci.2007.11.021

S. Battersby et al. / Journal of Membrane Science 316 (2008) 4652

47

technology including: enhanced equilibrium; increased reaction rate; improved product yield; and a concentrated product stream at the membrane outlet [79]. The selective removal of the reaction products from the reactor shifts the equilibrium in the MR to favour the reaction products, resulting in a reduction in the amount of catalyst and driving the conversion [7,10]. Also, by providing the gas separation in the MR, there is a reduced need for downstream processing, which should result in a lower capital cost and smaller plant footprint [11]. Membrane reactors can be designed to operate as extractors, distributors and active contactors [7]. In the case of the WGS reaction, the extractor design has the advantages that it allows one of the products such as H2 to permeate through the membrane, thus shifting the reaction to high conversion [12]. Membrane reactors are more complex unit operations than simple packed beds and the individual characteristics of the catalyst and membranes as a combined unit operation are central to the overall efciency. The intrinsic relationships between these two parts determine whether the MR represents an improvement over a conventional packed bed reactor (PBR) or not. For instance, if the permeation rate is low, then the membrane contributes little to the performance and the MR behaviour is almost like a PBR. Using a membrane with high permeation and selectivity greatly enhances the reaction rates. An optimal conguration can be achieved through proper balancing of the Damkohler number with the Peclet number [13]. In addition, different zones in the reactor may be applied to different reaction or separation operations [14] providing additional exibility. Current work in membrane reactors has sought improvements to many aspects; including for example material limits and operational parameters, with good reviews of this topic elsewhere [7,12]. However, it is only recently that researchers have begun to investigate the effects of operational parameters on both conversion and separation within the MR unit with examples shown in Table 1. Barbieri et al. and Brunetti et al. [8,10] studied the effect of pressure on CO conversion within a silica membrane reactor, from both experimental and fundamental perspectives. They found that higher pressures increased separation leading to an enhanced equilibrium conversion and higher catalyst effectiveness. Brunetti et al. [10] established that there was a maximum pressure before selectivity issues arose within the MR system. However, Van Delft et al. [15] showed that higher pressures could still be utilised in silica membranes to achieve a high degree of H2 separation. Iyoha et al. [16] provided an analysis of a Pd/PdCu reactive membrane for the WGS at very high temperatures (1173 K) where the PdCu metal surface operated as both the catalyst for reaction and the H2 selective membrane. This work showed the advantages of Pd-based membranes for high temperature reaction and gas separation though equilibrium severely limited conversion and stability (i.e. pinhole formation) was highlighted as a serious issue at this temperature. The water to CO feed molar ratio affects the equilibrium conversion and has been investigated by several research groups [2,1618]. However, the effect of water was clearly focused on improving CO conversion, generally not addressing the effect of excess water on the membrane permeation and selectivity. In addition, literature reports are limited on the inuence of space

Table 1 Water gas shift MR Membrane types Silica Silica Silica PdAg PdCu H2O/CO ratio 1 0.752.5 1 11.5 12.5 Pressure (bar) 26 1 515 1 Temperature ( C) 220300 150325 250400 330 900 Ref. [10] [2] [15] [33] [16]

velocity (SV) and experiments generally designed for maximum separation of H2 using low SV (<2000 h1 ), thus allowing the reaction to exceed traditional equilibrium limits. Giessler et al. [2] investigated the effect of SV and found that at higher SV the MR conversion increased compared to a PBR. Low SV is used by researchers, which coupled with a small length MR [2,10], has particular advantages in reducing partial pressure difference across the membranes caused by the concentration gradient along the reactor length. However, these differences are likely to be accentuated for long tubes. Material stability plays a major role in the selection of the membranes and until recently the majority of literature on the WGS reaction in MR utilised Pd alloy dense membranes [1619]. With a theoretical separation of 100% possible, these membranes have major benets in clean syngas (i.e. H2 , CO, CO2 and water only) processing though often suffer from steam embitterment and H2 S poisoning. Recent improvements in porous silica membranes hydrostability, selectivity and H2 ux has created new interest in their use for MR in the WGS eld [2,10,15]. However, the chemically harsh nature of the WGS reaction in syngas conversion means that ongoing material developments are needed to achieve industrial relevance for this unit [20,21]. Despite recent work developing silica membranes for H2 separation [22,23] the hydrophilic nature of silica leads to pore blocking and degradation of the membranes selectivity [24]. De Vos et al. [25] demonstrated improvements in the hydrostability of silica membranes through the use of a ligand templating method to produce hydrophobic silica surfaces, while Giessler et al. [2] used surfactant as non-ligand templates in the solgel process to achieve similar properties. By reducing adsorption of water on the membrane surface these methods produced stable membranes with good structural characteristics but lower selectivities than typically desired. Duke et al. [24,26] developed a novel surfactant templating procedure to produce membranes with good hydrostability and selectivity. The surfactants were carbonised inside the silica membrane matrix, resulting in a hydrophilic yet hydrostable membrane. Doping silica with metals such as Ni and Co also resulted in improved structural support and hydrostability properties of silica derived membranes [27,32]. In this work we investigate the performance of metal (cobalt) doped silica membranes for the low temperature WGS reaction in both conventional PBR and also MR congurations for the effects of excess water, space velocity and temperature on the overall conversion efciencies. The same apparatus is used for both reactor architectures, permitting direct comparison. A PBR reactor conguration is simply achieved by directing the permeate ow through the retentate chamber, which eliminates the

48

S. Battersby et al. / Journal of Membrane Science 316 (2008) 4652

driving force for permeation. This method is discussed further below. 2. Experimental 2.1. Membrane preparation Commercial alumina platelet substrates having porosity of 30% and an average pore size of 0.51 m (Rojan Ceramics, Australia) were coated with Locron solution (Clariant, Germany) for the deposition of -alumina layer with a pore size of 4 nm. Metal doped solgel was prepared by mixing 42.26 g tetraethoxysilane (TEOS) in 600 g ethanol to form a solution. A second solution was then synthesised by the dissolution of 14.84 g Cobalt nitrate hexahydrate (Co(NO3 )2 6H2 O) in 51.77 g of 30% aqueous H2 O2 . Subsequently, both solutions were vigorous mixed together by stirring for 3 h in ice-cooled bath. The platelets were coated with the stable SiCoO solution using a controlled immersion time 1 min and withdrawal speed 2 cm min1 . Sintering was carried out at 600 C for 4 h at heating rate and cooling rate of 0.7 C min1 . A total of six metal doped silica layers were coated on the platelets to ensure the production of defect free membranes with an expected pore size in the region of (0.350.45 nm) for silica derived materials [25]. Subsequently, a sintering hydrogen rich atmosphere at 500 C was employed for 15 h to reduce cobalt oxides embedded in the silica matrix [27]. 2.2. Catalyst preparation Cu/Zn/Al2 O3 catalysts were synthesised using the Na2 CO3 co-precipitation method as described elsewhere [2]. In this method Na2 CO3 was added drop wise to the nitrate solution, maintaining a pH of 7. The precipitant was aged at 60 C for 1 h followed by ltration, washing and drying. The dried sample was pelletised, sintered and crushed to give a catalyst particle size of 0.51 mm size. Catalyst surface characterisation was carried out in a Quantachrome quadrasorb using N2 adsorption at 77 K to determine BET surface area and pore size. Powder X-ray diffraction patterns were obtained using a Rigaku miniex with CoK radiation. Crystallite sizes for CuO were calculated from the Sherrer equation. Table 2 lists the general properties of the catalysts that were used in this study. 2.3. Membrane and reactor The MR and PBR ow arrangements are shown in Fig. 1. A plate-type MR was used for this study as depicted in Fig. 1a.
Table 2 Catalyst properties Catalyst type Average pore size dp (nm) Total pore volume (cm3 /g) BET surface area (m2 /g) CuO crystal size (nm) Cu Zn Al2 O3 5.51 0.155 56.37 5.45

Fig. 1. MR and PBR schematic.

The MR had an effective permeation area of 1.25 cm2 and reactive length of 20 mm with the membrane placed in direct contact with the pelletised catalyst. An initial pre-reaction zone was used to provide partial H2 conversion before the membrane reactor main reaction chamber with 1/3 of the catalyst placed in this pre reaction chamber. In some literature reports researchers have carried out the PBR testing in a separate apparatus from the MR. In this work, we use the same apparatus for both congurations. In the case of the PBR (Fig. 1b), the retentate stream is passed through the permeate side of the membrane, thus becoming a product stream only. The only partial pressure difference across the membrane in this system will then be caused by the concentration gradient across the reactor length. This effect is minimised with the use of a pre-reaction zone and excess water. In addition, the effective membrane reactor length is only 2 cm, which reduced the partial pressure difference to approximately 2500 Pa (2.5% concentration difference at 1atm). As gas to gas diffusion is at least four orders of magnitude higher then diffusion through the membrane, the effect of permeation on the reaction within this PBR set up will be <1%. Hence, the concentration difference of gases across the membrane is minimised and gas diffusion becomes negligible for the PBR. The catalyst was activated prior to experimentation at 250 C for 3 h using a gas composition of 20 vol% H2 and 20 vol% H2 O (4.65 103 mol l1 ) and 60 vol% N2 (1.39 103 mol l1 ) at 20 ml min1 total ow rate. After activation the H2 ow was switched off and the reactant ow set at the desired ow rate. A N2 sweep gas was maintained on the permeate side to carry permeate away. All gas ows were controlled using high precision Cole Palmer rotameters while reactant pressure was maintained using a back pressure valve. The operating conditions are given in Table 3. All reactant and permeate gases were tested in a Shimadzu GC-2014 with TCD and FID detectors to determine gas concentrations. The mechanisms that govern gas transport in membranes are Knudsen diffusion, Poiseuille ow and surface diffusion. Uhlhorn and Burggraaf [28] have shown that transport rate of

S. Battersby et al. / Journal of Membrane Science 316 (2008) 4652 Table 3 Operating conditions Temperature range ( C) Feed pressure (bar) H2 O/CO molar feed ratio Space velocity (h1 ) Mass of catalyst (g) 150250 2 0.54 75005000 0.28

49

gas molecules through mesoporous (2 < dp < 50 nm) membranes decreases with increasing temperature. On the other hand, de Lange et al. [29] have shown that the transport of diffusing molecules through microporous (dp < 2 nm) membranes is activated in which the ux J increases as a function of temperature according to J J0 exp Eact RT (2)

where J is the ux (mol m2 s1 ) through the membrane, Eact (kJ mol1 ) is an apparent activation energy, R the gas constant and T the absolute temperature (K). This model has been derived from Barrer [30] proposed model of transport through microporous crystal membrane. The permeation experiments in this work were conducted for several gases (He, H2 , CO2 , CO, and N2 ) at temperatures 100 < T < 250 C and 1bar feed pressure. At these conditions, adsorption is generally in the low coverage, including metal doped silica, and surface diffusion can be considered as negligible for He, H2 and N2 [29,31] except CO2 which shows negative activation energy [22]. Hence, using Ficks Law, Eq. (2) can be further derived to give a temperature dependency ux Jx = D0 K0 exp Eact RT dp dx (3)
Fig. 2. (a) Single gas permeation after completion of MR testing, and (b) MR binary mixture selectivity during reaction.

where D0 and K0 are temperature independent proportionalities for the Arrhenius and Vant Hoff equations, respectively. This model assumes that microporous ux is rate determining. In other words, the membrane top layer regiments the total ow rate of gases, whilst the effects of sub layers (i.e. mesoporous) or substrates (i.e. macroporous) are negligible. Predominantly, gas uxes increase with temperature, except in the case of molecules showing a good adsorption afnity to the membrane material (i.e. sorbent). In this work, the membrane has the capability to selectively separate H2 in the WGS reaction. The selectivity for binary gas mixtures (1/2 ) of the membranes can be calculated as a function of the concentration of the gas components in the permeate (cp,i ) over the retentate (cr,i ) 1/2 = cp,1 cr,2 cr,1 cp,2 (4)

3. Results and discussion Fig. 2 shows the effect of temperature on permeation and selectivity. After the MR work was completed, the catalyst was removed from the reactor chamber and the membrane was tested for single gas permeation. Even after 2 weeks of MR testing,

the metal doped silica membrane showed activated transport for the smaller molecules (He and H2 ) while activation was almost neutral or slightly negative for the larger molecules (CO2 and N2 ) as observed in Fig. 2a. Hence, as the temperature increases for the single gas permeation, the difference in H2 and the larger molecular gases permeation also increases. This is reected in the separation of gas mixtures as shown in Fig. 2b during the MR testing. For instance, the H2 /CO separation more than doubled from 150 to 250 C providing a two-fold improvement through higher H2 uxes and better separation. Comparatively, the separation of H2 /CO2 remains relatively low even with increasing temperature. The results in Fig. 2b are intrinsically linked with the concentrations in the reaction chamber as depicted in Fig. 3a, and explained as follows. Selectivities of binary mixtures are calculated based on the ratio of gases in permeate and retentate streams. Temperature affects not only gas separation, as discussed above, but also the extent of reaction: high temperatures lead to high CO conversions in the WGS reaction. As this reaction proceeds, each mole of CO converted produces 1 mole of H2 and 1 mole of CO2 , i.e. the concentrations of CO2 and H2 remain equal as a function of conversion. Hence, the partial pressure the driving force for diffusion of these gases is equivalent (see retentate line concen-

50

S. Battersby et al. / Journal of Membrane Science 316 (2008) 4652

Fig. 4. CO conversion as a function of space velocity at varying temperature (error bars showed for the highest experimental variation).

Fig. 3. (a) Permeate concentrations in a WGS membrane reactor at 300 C and space velocity of 7500 h1 and (b) MR conversion as a function of temperature.

trations in Fig. 3a), irrespective of the extent of conversion. In this case, the membrane slightly favours H2 over CO2 as shown by the concentration of these gases in the permeate stream, also Fig. 3a. On the other hand, as H2 (or CO2 ) is produced by the WGS reaction, CO is consumed, so the partial pressures move in opposite directions. At high conversions, which in this case are also correlated to high temperatures (see Fig. 3b), the driving force for H2 permeation is high and for CO is low. Taken together with the H2 selectivity preference of the membrane, Fig. 3a shows that the concentrations in the permeate stream reach values above 45 vol/vol% for H2 and below 2 vol/vol% for CO, respectively. Fig. 3a also reveals a performance limitation for this set-up: for reaction conversions below approximately 40% (crossing H2 and CO permeate lines), the H2 driving force for permeation is so low that H2 /CO selectivity is below unity. At these low conversions, the CO concentration in the permeate stream becomes larger than the H2 concentration and the MR retards rather than promotes the WGS reaction. This is important as it dictates the minimum conversion where the MR becomes effective for H2 /CO separation. Fig. 4 shows the MR and PBR conversion as a function of space velocities at different temperatures. Changing the space velocity allows for investigating both systems operating under

different reaction conditions, namely under (i) equilibrium limit conversion (high conversion) and (ii) reaction rate limited conversion (low conversion). At equilibrium limited conversion (space velocities 7,500 h1 ) the membrane does not show any signicant advantage. Under these conditions, very high conversions in excess of 95% are observed in the PBR, so there is little or no opportunity to improve this further. The reactor is oversized for the throughput. For higher throughputs (space velocities 38,000 and 75,000 h1 ) conversion is affected by the reaction rate, under which condition the membrane system may be expected to show an improved conversion. As H2 is removed from the reactor through the membrane, the concentration in the reaction chamber is reduced and the reaction pushed to higher conversions. Up to 12% shifting is observed for space velocity 75,000 h1 at 250 C. Increasing temperature also results in better MR performance as compared to PBR. These improvements stem from the fact that higher temperature increases CO conversion, leading to higher H2 partial pressure in the MRs reaction chamber which in turn increases the driving force for H2 diffusion through the membrane. In addition, the higher temperature also improves the H2 permeation through the activated transport properties of the metal doped silica membranes. These outcomes are consistent with the results showed in Figs. 2 and 3. H2 O as a reactant in the WGS is of particular importance and in industrial systems a large excess of steam, e.g. 34 stoichiometric, is used to drive the equilibrium towards higher H2 . In a MR design, however, the adverse effects water has on the membrane permeation and stability require a more delicate consideration to H2 O concentration. Many silica derived membranes are hydrophilic and break down in the presence of water [2]. In this work, metal doping was used to hydrostabilise the membrane. For the membranes used here, no condensate was found in the permeate stream even after 12 h operation, demonstrating that the membrane blocks water diffusion. Water reduced H2 permeation, consequently reducing both H2 /CO and H2 /CO2 selectivity of gas mixture permeation as shown in Fig. 5a. This effect was found to be reversible.

S. Battersby et al. / Journal of Membrane Science 316 (2008) 4652

51

Fig. 5. (a) The effect of water molar ratio on the MR selectivity during reaction and (b) water to CO molar ratio effect on conversion.

Excess water allows for higher CO conversion (see Fig. 5b), which reduces the concentration of CO and increases the concentration of H2 , CO2 , and excess water in the MR reactor chamber. The reasonable explanation is that water adsorbs on the surface of the membrane due the hydrophilic nature of silica, even in hydrostable silica membranes as reported, elsewhere, [26]. Assuming an ideal homogeneous gas mixture in the reaction chamber, there is a high probability that excess water may adsorb onto the membrane surface, blocking pore sizes lower then 3.4 A diffusion. which would normally be available for H2 (dk = 2.9 A) This adsorption preferentially occurs in the smaller pores, leaving larger pores unblocked by water, thus allowing the diffusion and CO2 (dk = 3.3 A). Nevertheless, the of H2 , CO (dk = 3.76 A) reduction in selectivity, Fig. 5a, is puzzling and requires further research to provide an explanation. 4. Conclusion Hydrostable metal (Cobalt) doped silica membranes were tested for the low temperature WGS reaction. The H2 /CO separation (515) increased by threefold from 150 to 250 C though the separation of H2 /CO2 (24) remained relatively low. Temperature affects not only gas separation due to H2 activated

transport, but also the extent of reaction: high temperatures lead to high CO conversions in the WGS reaction. As consumption of CO leads to production of H2 , higher conversion leads to an improved driving force of H2 across the membrane resulting in increased H2 /CO separation. On the other hand, H2 and CO2 are produced in equal volumes by reaction, so the partial pressures remain equal and all selectivity improvement is due solely to temperature activation. In the case of low reaction (conversions below approximately 40% for this membrane) the H2 driving force for permeation is so low that H2 /CO selectivity is below unity, dictating the minimum conversion where the MR becomes effective for H2 /CO separation. The metal doped silica membranes proved to be hydrostable, which is also consistent with Ni doped silica membranes results reported in the literature. However, water had an effect on the operational selectivity of the membrane. This could mainly be attributed to water adsorption on the silica membrane surface. Water may have resulted in microporous pore blocking via the hydrophilic silanol groups, thus hindering the diffusion of H2 and decreasing selectivities. However, water adsorption was reversible and in equilibrium with the water content in the gas phase. Operating under equilibrium limited conversion (space velocities 7,500 h1 ) conditions, very high conversions in excess of 95% are observed and there are no signicant advantages of the MR performance over the PBR. The reactor is oversized for the throughput. For higher throughputs (space velocities 38,000 and 75,000 h1 ) conversion is affected by the reaction rate, resulting in lower conversions. In this case, H2 is removed from the reactor through the membrane, the concentration in the reaction chamber is reduced and the reaction pushed to higher conversions. Increasing temperature as a function of the space velocity also results in better MR performance as compared to PBR. CO conversions shifted up to 12% at space velocity 75,000 h1 and at 250 C. Acknowledgment Scott Battersby acknowledges scholarship support from the CCSDCooperative Research Centre (CRC) for Coal in Sustainable Development. Shaomin Liu and Jo ao C. Diniz da Costa acknowledge nancial support trough the Innovation Funds given by the Queensland Government (Australia). References
[1] E. Shoko, B. McLellan, A.L. Dicks, J.C.D. da Costa, Hydrogen from coal: generation and utilisation technologies, Int. J. Coal Geol. 65 (2006) 213222. [2] S. Giessler, L. Jordan, J.C. Diniz da Costa, G.Q.M. Lu, Performance of hydrophobic and hydrophilic silica membrane reactors for the water gas shift reaction, Separ. Purif. Technol. 32 (2003) 255264. [3] C. Wheeler, A. Jhalani, E.J. Klein, S. Tummala, L.D. Schmidt, The watergas-shift reaction at short contact times, J. Catal. 223 (2004) 191199. [4] T. Utaka, K. Sekizawa, K. Eguchi, CO removal by oxygen-assisted water gas shift reaction over supported Cu catalysts, Appl. Catal. A: Gen. 194 (2000) 2126. [5] M.J.L. Gines, N. Amadeo, M. Laborde, C.R. Apesteguia, Activity and structure-sensitivity of the watergas shift reaction over CuZnAl mixed oxide catalysts, Appl. Catal. A:Gen. 131 (1995) 283.

52

S. Battersby et al. / Journal of Membrane Science 316 (2008) 4652 [20] A. Criscuoli, A. Basile, E. Drioli, O. Loiacono, Economic feasibility study for water gas shift membrane reactor, J. Membr. Sci. 181 (2001) 2127. [21] A. Criscuoli, A. Basile, E. Drioli, An analysis of the performance of membrane reactors for the watergas shift reaction using gas feed mixtures, Catal. Today 56 (2000) 5364. [22] R.M. de Vos, H. Verweij, Improved performance of silica membranes for gas separation, J. Membr. Sci. 143 (1998) 3751. [23] C.J. Brinker, G.W. Scherer, SolGel Science: The Physics and Chemistry of SolGel Processing, Academic press, Boston, 1990. [24] M.C. Duke, J.C. Diniz da Costa, G.Q. Max Lu, M. Petch, P. Gray, Carbonised template molecular sieve silica membranes in fuel processing systems: permeation, hydrostability and regeneration, J. Membr. Sci. 241 (2004) 325333. [25] R.M. de Vos, W.F. Maier, H. Verweij, Hydrophobic silica membranes for gas separation, J. Membr. Sci. 158 (1999) 277288. [26] M.C. Duke, J.C. Diniz Da Costa, D.D. Do, P.G. Gray, G.Q. Lu, Hydrothermally robust molecular sieve silica for wet gas separation, Adv. Funct. Mater. 16 (2006) 12151220. [27] H. Mori, S. Fujisaki, T. Saito, T. Sumino, T. Iwamoto, Characterisation and hydrogen interaction studies on cobalt-doped amorphous silica composite materials for high-temperature hydrogen separation membranes, in: R. Bredesen, H. Raeder (Eds.), Proceedings of the International Conference on Inorganic Membranes, Lillehammer, Norway, June 2529, 2006, pp. 382385. [28] R.J.R. Uhlhorn, A.J. Burggraaf, Gas separation with inorganic membranes, in: R.R. Bhave (Ed.), Inorganic Membranes Synthesis, Characteristics and Application, Chapman and Hall, New York, 1991. [29] R.S.A. De Lange, J.H.A. Hekkink, K. Keizer, A.J. Burggraaf, Formation and characterization of supported microporous ceramic membranes prepared by solgel modication techniques, J. Membr. Sci. 99 (1995) 57. [30] R.M. Barrer, Porous crystal membranes, J. Chem. Soc. Faraday Trans. 86 (1990) 11231130. [31] J.C. Diniz da Costa, G.Q. Lu, V. Rudolph, Y.S. Lin, Novel molecular sieve silica (MSS) membranes: characterisation and permeation of single-step and two-step solgel membranes, J. Membr. Sci. 198 (2002) 921. [32] M. Kanezashi, M. Asaeda, Hydrogen permeation characteristics and stability of Ni-doped silica membranes in steam at high temperature, J. Membr. Sci. 271 (2006) 8693. [33] S. Tosti, A. Basile, G. Chiappetta, C. Rizzello, V. Violante, PdAg membrane reactors for water gas shift reaction, Chem. Eng. J. 93 (2003) 2330.

[6] E. Xue, M. OKeeffe, J.R.H. Ross, Watergas shift conversion using a feed with a low steam to carbon monoxide ratio and containing sulphur, Catal. Today 30 (1996) 107. [7] A. Julbe, D. Farrusseng, C. Guizard, Porous ceramic membranes for catalytic reactorsoverview and new ideas, J. Membr. Sci. 181 (2001) 320. [8] G. Barbieri, A. Brunetti, T. Granato, P. Bernardo, E. Drioli, Engineering evaluations of a catalytic membrane reactor for the water gas shift reaction, Ind. Eng. Chem. Res. 44 (2005) 76767683. [9] J.N. Armor, Applications of catalytic inorganic membrane reactors to renery products, J. Membr. Sci. 147 (1998) 217233. [10] A. Brunetti, G. Barbieri, E. Drioli, K.-H. Lee, B. Sea, D.-W. Lee, WGS reaction in a membrane reactor using a porous stainless steel supported silica membrane, Chem. Eng. Process. 46 (2007) 119126. [11] S. Battersby, D. Miller, M. Zed, J. Patch, V. Rudolph, M.C. Duke, J.C.d. Costa, Silica membrane reactors for hydrogen processing, Adv. Appl. Ceram. 106 (2007) 2934. [12] A.G. Dixon, Recent research in catalytic inorganic membrane reactors, Int. J. Chem. React. Eng. 1 (2003). [13] S. Battersby, P.W. Teixeira, J. Beltramini, M.C. Duke, V. Rudolph, J.C. Diniz da Costa, An analysis of the Peclet and Damkohler numbers for dehydrogenation reactions using molecular sieve silica (MSS) membrane reactors, Catal. Today 116 (2006) 1217. [14] W.S. Moon, S.B. Park, Design guide of a membrane for a membrane reactor in terms of permeability and selectivity, J. Membr. Sci. 170 (2000) 4351. [15] Y.C. Van Delft, L.A. Correia, J.P. Overbeek, D.F. Meyer, P.P.A.C. Pex, J.W. Dijkstra, D. Jansen. Hydrogen membrane reactor for industrial hydrogen production and power generation, in: Sustainable (Bio) Chemical Process Technology, Sixh Intenational Conference on Process Intensication, Sep 2729, 2005 (Eds.) Delft, Netherlands pp. 397402 BN-1855980681. [16] O. Iyoha, R. Enick, R. Killmeyer, B. Howard, B. Morreale, M. Ciocco, Wall-catalyzed watergas shift reaction in multi-tubular Pd and 80 wt% Pd-20 wt% Cu membrane reactors at 1173 K, J. Membr. Sci. 298 (2007) 1423. [17] A. Basile, E. Drioli, F. Santella, V. Violante, G. Capannelli, G. Vitulli, Study on catalytic membrane reactors for water gas shift reaction, Gas Sep. Purif. 10 (1996) 5361. [18] A. Basile, A. Criscuoli, F. Santella, E. Drioli, Membrane reactor for water gas shift reaction, Gas Sep. Purif. 10 (1996) 243254. [19] A. Basile, G. Chiappetta, S. Tosti, V. Violante, Experimental and simulation of both Pd and Pd/Ag for a water gas shift membrane reactor, Sep. Purif. Technol. 25 (2001) 549571.

Potrebbero piacerti anche