You are on page 1of 7

International Journal of Hydrogen Energy 29 (2004) 1075 1081

www.elsevier.com/locate/ijhydene

Hydrogen from steam reforming of ethanol in low and middle temperature range for fuel cell application
Jie Suna; b; c , Xinping Qiub , Feng Wua ; , Wentao Zhub , Wendong Wanga , Shaojun Haoa
a School

of Chemical and Environmental Science, Beijing Institute of Technology, Beijing 100081, China b Department of Chemistry, Tsinghua University, Beijing 100084, China c Institute of Chemical Defense of PLA, Beijing 102205, China

Abstract The catalyst, Ni nano-particles supported on Y2 O3 , which was prepared by three methods, was studied. The structural properties of the catalysts were tested through X-ray di raction and BET area. The catalyst of Ni= Y2 O3 exhibits high activity for ethanol steam reforming with conversion of ethanol of 98% and selectivity of hydrogen of 38% at 300 C, conversion of ethanol of 98% and selectivity of hydrogen of 55% at 380 C. With temperature increasing to and above 500 C, the conversion of ethanol increased to 100%, but the selectivity of hydrogen did not increase so much, it was 58% at 600 C. The catalyst has long-term stability for steam reforming of ethanol and is a good choice for ethanol processors for fuel cell applications. ? 2003 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
Keywords: Ethanol steam reforming; Hydrogen; Conversion; Selectivity

1. Introduction In recent years, the proton exchange membrane fuel cells (PEMFC) with hydrogen as fuel has attracted many attentions due to its potential application in electric vehicles and power station. The use of hydrogen as a fuel o ers an important reduction in NOx and COx emissions. There are four basic methods for hydrogen production at present: water electrolysis, gasication, partial oxidation reactions of heavy oil, and steam reforming reactions [1]. Ethanol as a source of hydrogen has several advantages compared to other primary fuel such as methanol, gasoline and hydrogen itself [2]. It can be manufactured by fermentation of crops and as such has an attractive potential because of its origin (agroproduct) and the consequent reduction in carbon dioxide emission. Its storage, transportation, and distribution are relatively safe and easy [13]. The reactions of ethanol over the surfaces of metal oxides and metals have been studied for more than one decade
Corresponding author. Tel.: +86-10-689-125-08; fax: +86-106845-1429. E-mail addresses: magnsun@bit.edu.cn (J. Sun), wufeng863@bit.edu.cn (F. Wu).

now. The reforming of ethanol can be found in Refs. [1,3 17]. Reactions of ethanol dehydrogenation over noble metal membrane [1820], and reactions of ethanol over M= CeO2 catalysts [2123] have been studied. The catalytic properties of the supported transition metal catalysts in one of the steam reforming reactions (reaction (2) in Table 1) show that the selectivity of H2 for the reforming reaction is in the order: Co much greater than Ni Rh Pt, Ru, Cu. The properties of the Co on di erent supports were also studied. Co= Al2 O3 exhibited the highest selectivity for steam reforming of ethanol by suppression of methanation of CO and decomposition of ethanol [811]. The studies of ethanol steam reforming in a molten carbonate fuel cell (MCFC) utilized the catalysts from the noble metals (Pt, Rh, etc.) to CuO= ZnO= Al2 O3 and NiO= CuO= SiO2 and they got very good results: the selectivity of hydrogen of 5% Rh= Al2 O3 can reach to 30.1 vol% [1,1214]. The steam reforming of ethanol at about 300 C over catalyst of Cu= Ni= K = -Al2 O3 has been studied by Marino et al. [15,16]. Compared to methanol, ethanol has one major complication if one considers its total decomposition: it contains a carboncarbon bond and as such requires a surface capable of breaking its bond. There are at least two further requirements for the surface. It must be capable of oxidizing

0360-3199/$ 30.00 ? 2003 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.ijhydene.2003.11.004

1076

J. Sun et al. / International Journal of Hydrogen Energy 29 (2004) 1075 1081

Table 1 Relative reactions of ethanol steam reforming in this work Independent reactions which involve ethanol, water and hydrogen Selectivity (mole fraction) CH3 CHO (1) C2 H5 OH + 3H2 O 6H2 + 2CO2 (2) C2 H5 OH + H2 O 4H2 + 2CO (3) C2 H5 OH H2 + CH3 CHO (4) C2 H5 OH H2 + CO + CH4 (5) C2 H5 OH 3H2 + CO + C Dependent reactions (6) 2CO CO2 + C (7) CO + H2 O CO2 + H2 (8) CH4 + 2H2 O CO2 + 4H2 (9) CO2 + CH4 2CO + 2H2 (10) C2 H5 OH + H2 O 2H2 + CO2 + CH4 (11) 2C2 H5 OH + 3H2 O 7H2 + 2CO2 + CO + CH4 (12) 2C2 H5 OH + H2 O 3H2 + CO2 + CO + 2CH4 50 CO2 25 100 50 20 25 18.2 14.3 CH4 33.3 25 9.1 28.6 CO 33.3 33.3 25 50 9.1 14.3 H2 75 66.7 50 33.3 75 50 80 50 50 63.6 42.9

both carbon atoms to CO2 , and in the case of hydrogen production, it must not be active for the oxidation of H2 . Consequently, the choice of the support is very crucial. An ideal support would not favor dehydration reactions (such as -Al2 O3 ) and would have no (or mild) capability for other CC bond formation reactions [24]. Based on the above studies, we chose Ni as the supported metal and Y2 O3 as the support. Ni has been widely used as catalyst in hydrogenation and dehydrogenation reactions because of its high activity and low cost. It is well known that rare earth metal oxides are the highly alkaline and are favorable for dehydrogenation of alcohols. Y2 O3 is more active for dehydrogenation than oxides of other elements in lanthanides. In this work, the catalyst of Ni= Y2 O3 for ethanol steam reforming for hydrogen production is studied.

solution of K2 C2 O4 under continuous stirring and NiC2 O4 precipitated. During the precipitation of NiC2 O4 ; Y2 O3 powder was put into the mixture at the same time. After 24 h continuous stirring, the mixture was centrifuged and washed to eliminate SO2 and K + ions, then the product 4 was dried at 100 C for 24 h. The dry product was then ground, sieved, and heated at 500 C for 2 h in N2 ow. NiC2 O4 would decompose to Ni and CO. X-ray powder di raction (XRD) was carried out on a Rigaku X-ray di raction equipment, using the Cu K radiation, at 40 kV and 20 mA. The BET surface area of each catalyst was determined by means of nitrogen physisorption, using a Quantachrome NOVA automated gas sorption instrument. 2.2. Ethanol steam reforming

2. Experimental 2.1. Catalyst preparation Three simple methods were employed to prepare the Ni= Y2 O3 catalysts. In method (1), Y2 O3 powder was put into the solution of Ni(NO3 )2 under continuous stirring for no less than 6 h at 90 C. After water being dried out, the dry product was heated at 120 C for 12 h to release NO2 , then ground, sieved and heated at 500 C for 2 h in H2 ow to reduce the NiO to Ni. In method (2), NiSO4 ; NaBH4 and Y2 O3 were used as starting materials. Y2 O3 powder was put into the solution of NiSO4 under continuous stirring, and then the solution of NaBH4 was dropped into the mixture of Y2 O3 and the solution of NiSO4 , and black loose Ni was produced. Then the solid was ltrated, washed and dried out, and then heated at 500 C for 2 h in H2 ow. In method (3), the solution of NiSO4 was dropped into the

The ethanol steam reforming experiments were conducted with a xed-bed reactor (inner diameter 16 mm) tted in a programmable oven with three heat sections (Fig. 1). The nether heat section with an operating range up to 300 C was used as evaporator, and we chose 280 C as the vaporizing temperature to vaporize the in uent instantly. The two up sections were used for the reforming reactions with an operating range up to 700 C and 1000 C, respectively. In this work, the catalyst prepared by method (1) or method (2) was previously in situ decomposed in nitrogen at 500 C for 2 h and then reduced in hydrogen at 500 C for 2 h before the reforming reactions. While the catalyst prepared by method (3) only need previously in situ decomposed in nitrogen at 500 C for 2 h without hydrogen reduction before the reforming reactions. The ow rate of ethanolwater solution was controlled as 0:05 cm3 min1 . The reforming reactions take place through the catalyst bed. The components of e uents were out through a condenser and a dryer,

J. Sun et al. / International Journal of Hydrogen Energy 29 (2004) 1075 1081

1077

Products

GC reactor

(222)

700-1000C
Intensity ( a.u.)
(211) (411) (111) (200) (400) (541) (631) (600) (620) (622) (220)

catalyst bed 300-700C quartz powder

(332)

(521)

(440)

(3)

**

**

(2)

0-300C
*
20 30 40 50 60

(1)
70

evaporator nitrogen or hydrogen ethanol solution

2 (degree)
Fig. 2. XRD of the Ni= Y2 O3 catalysts prepared by three methods. The peaks with symbol and vertical labels of Miller indexes are corresponding to di raction peaks of NiO, and the peaks with horizontal labels of Miller indexes are corresponding to that of Y2 O3 . (a) XRD patterns of catalyst prepared by method (1), (b) XRD patterns of catalyst prepared by method (2), (c) XRD patterns of catalyst prepared by method (3). Table 2 Crystal grain sizes of catalysts calculated in this work Method of catalyst (1) (2) (3) Y2 O3 (nm) 59 71 81 NiO (nm) 67 69

Fig. 1. Scheme of the reforming reactor and oven.

and then introduced to a gas chromatograph (GC). The GC was equipped with two packed columns (Porapak, 80 100 mesh, 1:5 m long; Tdx-01, 60 80 mesh, 1:5 m long) and one thermal conductive detector (TCD). The Porapak column was used for the separation of C2 H5 OH; H2 O and CH3 CHO with H2 as carrier gas, the column temperature being 100 C. The Tdx-01 column was used for the separation of CO; CO2 ; CH4 ; C2 H4 , and C2 H6 with H2 as carrier gas, the column temperature being 100 C. 3. Results and discussion It has been shown, from a thermodynamic point of view, that the steam reforming of ethanol is entirely feasible [4,5,17]. The system of steam reforming of ethanol involves a set of response reactions which have di erent relative fractions at di erent operative condition. The yield of produced hydrogen may be a ected by each response reaction [5]. Therefore, the yield of hydrogen depends largely on the operative conditions of the reforming reaction, such as pressure, temperature, H2 O= EtOH mole ratio, etc. In order to increase the yield of hydrogen, it is necessary to know the e ect of these operative conditions on the product components. 3.1. Structural characteristics of Ni=Y2 O3 3.1.1. XRD Fig. 2 shows the XRD patterns of Ni= Y2 O3 catalysts prepared by three di erent methods. The peaks labeled with the symbol are corresponding to di raction peaks of NiO, and the other peaks are corresponding to that of Y2 O3 . No new

peaks appeared, meaning that the catalysts have only the phases of NiO and Y2 O3 and no new compound was formed. The crystal size of the heterogeneous catalysts, calculated by Scherrer formula, was listed in Table 2. The crystal size of Y2 O3 was calculated based on his peaks with Miller indexes of (4 0 0) and (5 4 1). The crystal size of NiO was calculated based on his peak with Miller indexes of (2 0 0). It can be seen that the crystal sizes of Y2 O3 changed in different methods: method (1) 59 nm, method (2) 71 nm, and method (3) 81 nm, respectively. The crystal size of NiO in the catalyst prepared by method (1) was hard to calculate because of its too weak di raction. The crystal sizes of NiO in the catalyst prepared by method (2) and method (3) are 67 and 69 nm, respectively. 3.1.2. BET The results of BET measurement (Table 3) indicated that the specic surface area of catalyst prepared by method (3) was 52:2 m2 g1 , larger than that of the catalysts prepared by method (1) 26:3 m2 g1 and by method (2) 17:0 m2 g1 .

1078

J. Sun et al. / International Journal of Hydrogen Energy 29 (2004) 1075 1081

Table 3 Catalyst BET area tested in this work Method of catalyst (1) (2) (3) Description Y2 O3 Ni(NO3 )2 = Y2 O3 NiO= Y2 O3 NiSO4 = Y2 O3 = NaBH4 NiO= Y2 O3 NiC2 O4 = Y2 O3 NiO= Y2 O3 BET area (m2 = g)

60

50

SH (mol%)

9.1 26.3 17.0 52.2

40

30

20

3.2. Catalyst selectivity and conversion Generally, the denitions of catalyst selectivity are different according to the convenience for description catalytic reaction [2528]. For the ethanol steam reforming reactions, selectivity of hydrogen was usually dened as the mole ratio of produced hydrogen to the consumed ethanol [16]. While, for the convenience of tests, we dened the catalyst selectivity as the mole fraction of each product: molP SP = 100%: (1) molsp Here, molp represents the number of moles of each product; molsp represents the number of sum moles of products, but the number of moles of solid products (such as small amount of coke) is not included. Catalyst activity is evaluated in terms of ethanol conversion. We dened ethanol conversion as molconv CE = 100%: (2) molsEt Here, molconv represents the number of moles of converted ethanol; molsEt represents the number of sum moles of ethanol feed into reactor. 3.3. E ect of the prepared methods on the selectivity of hydrogen Six grams of catalyst was used to test its catalysis properties for the ethanol reforming reactions. Fig. 3 shows the selectivity of hydrogen of catalyst Ni= Y2 O3 prepared by three methods for ethanol steam reforming at 600 C. Catalysts selectivity of hydrogen prepared by methods (1), (2), and (3) is 53.5%, 53.7% and 59.6%, respectively. The selectivity of hydrogen of the catalyst prepared by method (3) was the highest and this maybe due to its high specic surface area. In the following sections, we used catalyst prepared by method (3) for further experiments. 3.4. E ect of the mole ratio of H2 O=EtOH on the selectivity of hydrogen Fig. 4 shows the e ect of the mole ratio of water and ethanol (H2 O= EtOH) on the selectivity of hydrogen at temperature 500 C. At mole ratios of 3:1 and 8:1, the selectivity of H2 reaches higher values of 45% and 46.5%,
10

(1)

(2)

(3)

Methods of Catalysts
Fig. 3. Selectivity of hydrogen, over the catalysts prepared by three methods. Experimental conditions: mass of catalyst 6 g; pellets size 0.5 1 mm; H2 O= EtOH mol ratio 3:1, ow rate 0:05 cm3 min1 (W= F = 7200 g s cm3 ), P = 1 atm; T = 600 C. (a) Selectivity of hydrogen, over the catalyst prepared by method (1) (b) Selectivity of hydrogen, over the catalyst prepared by method (2) (c) Selectivity of hydrogen, over the catalyst prepared by method (3).

respectively. It is interesting that there are two peak values in Fig. 4. From reaction (1) (in Table 1), it can be seen that when the H2 O= EtOH mols ratio is 3:1, the mole fraction of hydrogen is 75%, very high. So the appearance of a peak near mole ratio of 3:1 is reasonable. However, when mole ratio reached 8:1, the extent of watergas shift reaction (as reaction (7) in Table 1) increases when there is much excess water. The shift reaction contributes to the form of the second peak, so we chose the H2 O= C2 H5 OH mol ratio of 3:1 in the further experiments considering the selectivity of hydrogen and the reaction rate. 3.5. E ect of temperature on selectivity and conversion Typical experimental results obtained are presented in Figs. 5 and 6, in which the selectivity of each product and the conversion of ethanol are shown as a function of reaction temperature in the low-temperature range of 300 400 C and middle temperature range of 400 650 C. At temperature 300 C, steam reforming of ethanol has occurred appreciably and the selectivity of hydrogen, methane, carbon dioxide and carbon oxide are 38.2%, 34.3%, 18.3% and 9.2%, respectively. A slight amount of acetaldehyde was also detected with the selectivity of 0.02%. When the temperature increased from 300 C to 380 C, the selectivity of hydrogen increased from 38.2% to 55.2%, while the selectivity of methane decreased from 34.3% to 27.8%. The selectivity of carbon dioxide and carbon monoxide kept decreasing from 18.3% to 15% and from 9.2% to 2.0%, respectively. The

J. Sun et al. / International Journal of Hydrogen Energy 29 (2004) 1075 1081

1079

60

50

40

SH2 (mol%)

30

20

10

0 0 2 4 6 8 10

H2O/EtOH (mol ratio)


Fig. 4. E ect of mol ratios of H2 O= EtOH on the selectivity of hydrogen, over the Ni= Y2 O3 catalyst. Experimental conditions: mass of catalyst 6 g, pellets size 0.5 1 mm, ow rate 0:05 cm3 min1 (W= F = 7200 g s cm3 ); P = 1 atm; T = 500 C.

100 90 80 C Ethanol

100

CEthanol

80

C (%), S (mol%)

70 60 50 40 30 20 10 0

C (%), S (mol%)

SH2 SCH SCO SCH


300

60

SH

40

SCH SCO
2

20

SCO
320 340 360 380 400
0 400 450 500

3 CHO

SCO
550 600 650

Temperature(oC)
Fig. 5. E ect of low reaction temperature range (300 400 C) on conversion of ethanol (CEtOH ) and on selectivity of hydrogen (SH2 ), carbon monoxide (SCO ), carbon dioxide (SCO2 ), methane (SCH4 ) and acetaldehyde (SCH3 CHO ), obtained over the Ni= Y2 O3 catalyst. Experimental conditions: mass of catalyst 15 g, particle size 0.5 1 mm; H2 O= EtOH mol ratio 3:1, ow rate 0:4 cm3 min1 (W= F = 2250 g s cm3 ); P = 1 atm.

Temperature(oC)
Fig. 6. E ect of middle reaction temperature range (400 650 C) on conversion of ethanol (CEtOH ) and on selectivity of hydrogen (SH2 ), carbon monoxide (SCO ), carbon dioxide (SCO2 ), methane (SCH4 ) and acetaldehyde (SCH3 CHO ), obtained over the Ni= Y2 O3 catalyst. Experimental conditions: mass of catalyst 15 g, particle size 0.5 1 mm; H2 O= EtOH mol ratio 3:1, ow rate 0:4 cm3 min1 (W= F = 2250 g s cm3 ); P = 1 atm.

selectivity of acetaldehyde reached the maximum of 0.04% at 360 C, down to the bottom value of 0.02% at 380 C. No acetaldehyde was detected measurably above 380 C. When the temperature increased from 380 C to 400 C, the

selectivity of hydrogen decreased to 42.7%, while the selectivity of methane and carbon dioxide increased to 38.1% and 18.3%, respectively. The selectivity of carbon oxide kept decreasing to 0.9%.

1080

J. Sun et al. / International Journal of Hydrogen Energy 29 (2004) 1075 1081

This distribution of selectivity of each product should follow relative reactions in Table 1, where the selectivity of each product in each reaction at equilibrium was listed, but the contribution of each reaction to the selectivity of each product in the reactions system changed with di erent operative conditions. In the temperature range of 300 330 C, reactions (4), (10) and (12) seem to contribute more than the other reactions considering the selectivity (mole fraction) of the products. In this temperature range, the selectivity of hydrogen increased from 38.2% to 46.9%, and the selectivity of methane, carbon dioxide and carbon oxide decreased from 34.3% to 31.9%, from 18.3% to 17.7% and from 9.2% to 3.5%, respectively, which was closely consistent with co-reactions of (4), (10) and (12). In the temperature range of 330 380 C, reactions (1) and (4) should be the dominate reactions, since the selectivity of methane, carbon dioxide and carbon oxide kept decreasing, from 31.9% to 27.8%, 17.7% to 15% and 3.5% to 2.0%, respectively. Only the selectivity of hydrogen kept increasing, from 46.9% to 55.2%. In the temperature range of 380 400 C, the decrease of selectivity of hydrogen and carbon oxide contributed to the increase of that of methane and carbon dioxide. This means that reaction (10) became the dominated reaction. The conversion of ethanol reached to 98% at the low temperature range of 300 400 C. This high result means very good activity of the catalyst. The e ects of temperatures between 400 C and 650 C on the selectivity and conversion were presented in Fig. 6. With the increase of the temperature, the conversion of ethanol increased. The conversion of ethanol reached 100% when the temperature was 500 C and above. However, the hydrogen selectivity did not increase much compared with the selectivity at the low temperature range (300 400 C). In the temperature range of 400 600 C, the selectivity of hydrogen, carbon dioxide and carbon oxide kept increasing and reached a peak value of 58%, 20.2% and 4.9% at 600 C, respectively. While the selectivity of methane kept decreasing and down to the minimum value of 16.9% at 600 C also, which indicated the reformation of methane with H2 O had occurred (as reaction (8) in Table 1). Here, the dependent reaction (11) (in Table 1) seems to contribute more. With temperature increasing to 650 C, the selectivity of hydrogen and carbon dioxide decreased while the selectivity of methane and carbon oxide increased, which indicated that inverse watergas shift reaction took place with temperature increasing. This means that the Ni= Y2 O3 catalyst is also active for the inverse watergas shift reaction. In the temperature ranges we studied, no C2 H2 ; C2 H4 and C2 H6 were not detected measurably in the reaction products, indicating that no dehydration of ethanol was taking place, as might be expected. This is due to the fact this particular catalyst, which uses Y2 O3 as the support material, dose not possess dominating acidic sites, which are required for the dehydration route [3]. In the temperature range of 300 380 C, there had occurred the reforming reactions evidently, which proved the fact that the catalyst of Ni= Y2 O3 is active

for carboncarbon bond breaking. On the other hand, there was a slight amount of acetaldehyde (the highest selectivity is 0.04%) to be detected, which indicated two important possibilities: the particular catalyst Ni= Y2 O3 has a mild capability for dehydrogenation of ethanol and the catalyst has very strong ability to break the CC bond of acetaldehyde which is just an in-process product. In the temperature range of 500 650 C, the conversion of ethanol reached 100%, at the temperature of 600 C, the selectivity of hydrogen exceeded 58%. Under these conditions, one of the undesirable products was methane, which consumed hydrogen atoms. There appeared a small amount of coke also at the middle temperature range of 400 650 C. In our early work, Raman spectrum of the deposition had proved that it was carbon. The TG analysis used pure ethanol as in uent introduced by nitrogen indicated that coke deposition began at about 380 C. If the reforming temperature increased above 380 C, the coke deposition would start very soon. Although the product of coke may increase the selectivity of hydrogen in gaseous products (as reaction (5) in Table 1, the selectivity of hydrogen is 75%), and did not a ect the activity of the catalyst much, coke may cover part of the surface of catalyst. 3.6. Stability The stability of the catalyst with time-on-stream was examined in the middle temperature range at 550 C for 66 h. Before the experiment, the catalyst of NiC2 O4 = Y2 O3 was mixed with water and glycerol to form pellets of 0.5 1 mm and then decomposed in nitrogen at 500 C for 2 h. The relationships of selectivity of each product, conversion of ethanol with time-on-stream were shown in Fig. 7. Under the experimental conditions, the conversion of ethanol was 100%. The selectivity of hydrogen obviously increased
100 90

CETOH

C(mol%),S(mol%)

80 70 60 50 40 30 20 10 0 30 35 40

SH

SCH

SCO

SCO
45 50 55 60 65

Time (h)
Fig. 7. Conversion of ethanol, and selectivity of hydrogen as function of time-on-stream, obtained over the Ni= Y2 O3 catalyst. Experimental conditions: mass of catalyst 10 g, particle size 0.5 1 mm; H2 O= EtOH mol ratio 3:1, ow rate 0:4 cm3 min1 (W= F= 1500 g s cm3 ); T = 550 C; P = 1 atm.

J. Sun et al. / International Journal of Hydrogen Energy 29 (2004) 1075 1081

1081

during the rst 59 h from 49.6% to 61.4% and decreased during the last 7 h from 61.4% to 53.8%. And, if the stability of the catalyst was tested at a lower temperature, longer time of stability will be found. 4. Conclusion To sum up, the present results clarify that the novel Ni= Y2 O3 catalyst produced hydrogen-rich gas mixture in low and middle temperature range, possessed high activity, and good stability for steam reforming of ethanol to hydrogen production. The catalyst of Ni= Y2 O3 is a good choice to be used in ethanol steam reforming processors for fuel cell applications. Acknowledgements The funding provided for this work by 973 Project (2002CB211800) of China, the help in directions of Prof. C.M. Hong, the cooperation of Ms. X.L. Liu in XRD measurements and the technical assistance of Dr. J.T. Ma in BET measurements are gratefully acknowledged. References
[1] Freni S. Rh based catalysts for indirect internal reforming ethanol applications in molten carbonate fuel cells. J. Power Sources 2001;94:149. [2] Brown LF. A comparative study of fuels for on-board hydrogen production for fuel-cell-powered automobiles. Int J Hydrogen Energy 2001;26:38197. [3] Fatsikoatas AN, Kondarides DI, Verykios XE. Steam reforming of biomass-derived ethanol for the production of hydrogen for fuel cell applications. Chem Commun 2001;8512. [4] Vasudeva K, Mitra N, Umasankar P, Dhingra SC. Steam reforming of ethanol for hydrogen production: thermodynamic analysis. Int J Hydrogen Energy 1996;21:138. [5] Fishtik I, Alexander A, Datta R, Geana D. A thermodynamic analysis of hydrogen production by steam reforming of ethanol via response reactions. Int J Hydrogen Energy 2000;25: 3145. [6] Llorca J, de la Piscina PR, Sales J, Homes N. Direct production of hydrogen from ethanolic aqueous solutions over oxide catalysts. Chem Commun 2001; 6412. [7] Cavallaro S. Ethanol steam reforming on Rh= Al2 O3 catalysts. Energy Fuels 2000;14:11959. [8] Haga F, Nakajima T, Yamashita K, Mishina S. Catalytic properties of supported transition metal catalysts for conversion of ethanol in the presence water vapor. Nippon Kagaku Kishi 1997;1:336. [9] Haga F, Nakajima T, Miya H, Mishima S. Catalytic properties of supported cobalt catalysts for steam reforming of ethanol. Catal Lett 1997;48:2237. [10] Haga F, Nakajima T, Yamashita K, Mishina S. E ect of particle size on steam reforming of ethanol over alumina-supported cobalt catalyst. Nippon Kagaku Kaishi 1997;11:75862. [11] Haga F, Nakajima T, Yamashita K, Mishina S. E ect of crystallite size on the catalysis of alumina-supported cobalt

[12] [13] [14] [15]

[16] [17] [18] [19] [20] [21]

[22]

[23]

[24]

[25] [26]

[27]

[28]

catalyst for steam reforming of ethanol. React Kinetics and Catal Lett 1998;63:2539. Cavallaro S, Freni S. Ethanol steam reforming in a molten carbonate fuel cell. A preliminary kinetic investigation. Int J Hydrogen Energy 1996;21:4659. Maggio G, Freni S, Cavallaro S. Light alcohols/methane fuelled molten carbonate fuel cells: a comparative study. J Power Sources 1998;74:1723. Freni S, Chiodo V, Carvallaro S. Ethanol application for multifuel molten carbonate fuel cell. Abstracts of Papers of the American Chemical Society Part 1 2001;222:141. Marino FJ, Cerrella EG, Duhalde S. Hydrogen from steam reforming of ethanol. Characterization and performance of coppernickel supported catalysts. Int J Hydrogen Energy 1998;23:1095101. Marino FJ, Boveri M, Baronetti G. Hydrogen production from steam reforming of bioethanol using Cu= Ni= Al2 O3 catalysts. E ect of Ni. Int J Hydrogen Energy 2001;26:6658. Garcia EY, Laborde MA. Hydrogen production by the steam reforming of ethanol: thermodynamic analysis. Int J Hydrogen Energy 1991;16:30712. Amandusson H, Ekedahl LG, Dannetun H. Isotopic study of ethanol dehydrogenation over a palladium membrane. J Catal 2000;195:37682. Amandusson H, Ekedahl LG, Dannetun H. Alcohol dehydrogenation over Pd versus Pd/Ag membranes. Appl Catal A 2001;217:15764. Amandusson H, Ekedahl LG, Dannetun H. Hydrogen permeation through surface modied Pd and PdAg membranes. J Membrane Sci 2001;193:3547. Yee A, Morrison SJ, Idriss H. A study of ethanol reactions over Pt = CeO2 by temperature programmed desorption and in situ FT-IR spectroscopy: evidence of benzene formation. J Catal 2000;191:3045. Yee A, Morrison SJ, Idriss H. The reactions of ethanol over M= CeO2 catalysts. Evidence of carboncarbon bond dissociation at low temperatures over Rh= CeO2 . Catal Today 2000;63:32735. Yee A, Morrison SJ, Idriss H. A study of the reactions of ethanol on CeO2 and Pd = CeO2 by steady state reactions, temperature programmed desorption, and in situ FT-IR. J Catal 1999;186:27995. Sheng PY, Yee A, Bowmaker GA, Idriss H. H2 production from ethanol over Rh-Pt = CeO2 catalyst: the role of Rh for the e cient dissociation of the carboncarbon bond. J Catal 2002;208:393403. Bukur DB, Lang Xi, Ding Y. Pretreatment e ect studies with a precipitated iron FischerTropsch catalyst in a slurry reactor. Appl Catal A 1999;186:25575. Wallenstein D, Haas BKA. In uence of coke deactivation and vanadium and nickel contamination, on the performance of low ZSM-5 levels in FCC catalysts. Appl Catal A 2000;92: 10523. Hedrick SA, Chuang SSC, Pant A, Dastidar AG. Activity and selectivity of Group VIII, alkali-promoted MnNi, and Mo-based catalysts for C2+ oxygenate synthesis from the CO hydrogenation and CO= H2 = C2 H4 reactions. Catal Today 2000;55:24757. Englisch M, Ranade VS, Lercher TA. Hydrogenation of crotonaldehyde over Pt based bimetallic catalysts. J Mol Catal A Chem 1997;121:6980.

You might also like