Vous êtes sur la page 1sur 6

International Journal of Hydrogen Energy 31 (2006) 877 882 www.elsevier.

com/locate/ijhydene

Production of hydrogen by steam reforming of ethanol over a Ni/ZnO catalyst


Yu Yanga, b , Jianxin Mac , , Fei Wua
a Institute of Industrial Catalysis, East China University of Science and Technology, Shanghai 200237, China b Zhongyuan Institute of Technology, Zhengzhou 450007, China c Institute of Hydrogen Technologies, School of Automotive Studies, Tongji University, Shanghai 201804, China

Received in revised form 25 June 2005; accepted 25 June 2005 Available online 12 September 2005

Abstract In the present work a zinc oxide-supported nickel catalyst was investigated for the steam reforming of ethanol. Comparison was made to the nickel catalysts supported on La2 O3 , MgO and -Al2 O3 . Experimental results showed that Ni/ZnO is superior among the catalysts, especially in terms of selectivity and distribution of byproducts. Inuential factors, such as reaction temperature, water to ethanol molar ratio, liquid hourly space velocity (LHSV) and Ni loading, were investigated using a Ni/ZnO catalyst. The conversion was always complete at temperatures above 330 C, regardless of the changes in other reaction conditions. However, the selectivity to hydrogen was increased with the increase in temperature, H2 O/EtOH molar ratio and Ni loading. A selectivity of up to 95% was reached under conditions of T = 650 C, LHSV = 5.0 h1 , H2 O/EtOH ratio = 8, and Ni loading = 20 wt%. The effect of the inuential factors on the side reactions and the distribution of byproducts was also discussed. 2005 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
Keywords: Production of hydrogen; Ethanol; Ni/ZnO catalyst

1. Introduction Today, approximately 95% of hydrogen is produced from fossil-based materials and most of them are used as chemical feedstock in petrochemical, food, electronics and metallurgical processing industries. However, with the implementation of fuel cell systems and the growing demand for deep desulfurized fuels, production of hydrogen will need to increase remarkably. In the near future, the increased production can be fullled by conventional technologies, such as steam reforming of methane (SRM). In these processes, however, the carbon is converted to CO2 and released into the atmosphere, leading to global climate change. Sequestration
Corresponding author. Tel.: +86 21 69589461; fax: +86 21 69589978. E-mail address: jxma@fcv-sh.com (J. Ma).

of the CO2 has been proposed as a viable short-term solution. In addition, renewable energy-based processes like solar- or wind-driven electrolysis and photobiological water splitting are being investigated as long-term solutions for clean hydrogen production, while generation of hydrogen from biomass has been recognized as a more practical and viable option for the near- and mid-term solutions due to its renewable and carbon-neutral nature. In biomass technologies, ethanol is the most attractive feedstock because of its relatively high hydrogen content, availability, non-toxicity and storage and handling safety. More importantly, ethanol can be produced renewably from several biomass sources, including energy plants, waste materials from agroindustries or forestry residue materials [1]. One of the methods suggested processing of bio-ethanol to produce H2 to be used in fuel cell system, which is a catalytic steam reforming reaction, because this reaction is

0360-3199/$30.00 2005 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.ijhydene.2005.06.029

878

Y. Yang et al. / International Journal of Hydrogen Energy 31 (2006) 877 882

entirely feasible from a thermodynamic point of view [24]. A number of Ni-based catalysts were investigated by using different oxides as support, such as La2 O3 and YSZ [5], Al2 O3 [57], MgO [5,8,9], SiO2 [10] and CeO2 Zr 2 O3 [11], etc. The results of the investigation showed that the support of catalyst played an important role in the reaction. In the present work a Ni/ZnO catalyst, which has not been reported before, was investigated and compared with Ni/ -Al2 O3 , Ni/MgO and Ni/La2 O3 catalysts. The study was carried out by varying the operating conditions, such as reaction temperature, water to ethanol ratio and space velocity, and the results showed that the Ni/ZnO catalyst is very effective for steam reforming of ethanol in terms of ethanol conversion and selectivity toward hydrogen production. 2. Experimental 2.1. Preparation of catalyst Ni/ZnO, Ni/La2 O3 , Ni/MgO and Ni/ -Al2 O3 catalysts were prepared by the method of incipient wetness using nickel nitrate as a metal precursor. The support material ZnO was prepared by the decomposition of ZnCO3 . La2 O3 , MgO and -Al2 O3 were commercial products. After impregnation, the catalysts were dried at room temperature for 12 h, subsequently dried at 120 C for 12 h and nally calcined at 600 C for 6 h. The catalyst powder was tabletted at 5 MPa and crushed into particles with 2040 mesh size for use in a reactor. The specic surface area of the catalysts is listed in Table 1. 2.2. Evaluation of catalyst The steam reforming reaction was carried out at atmospheric pressure in a xed-bed ow reactor made of a stainless-steel tube (i.d. = 20 mm; L = 380 mm). In each run of the experiment, 3 ml of catalyst (2040 mesh) was used and before measurement the samples were reduced insitu in the reactor at 650 C and under a stream of 20% hydrogen/nitrogen for 2 h. Water and ethanol were pre-mixed

in a molar ratio of 8 and fed to the vaporizer in the reactor at a rate of 15 ml h1 . The vaporization temperature was set at 200 C. The liquid components in the outlet stream were condensed in an ice-trap and the gaseous components were fed to on-line gas chromatography. Two sets of gas chromatography were used and equipped with a thermal conductivity detector (TDC) and ame ion detector (FID), respectively. The hydrogen content in the outlet was analyzed by TCD with a column packed with active carbon. The concentrations of CO, CO2 , CH4 and C2 H4 , etc. were analyzed by FID combined with a methanation converter and with a column packed with GDX. The composition of condensate was also determined by FID.

3. Results and discussion 3.1. Effect of support In Table 1 the conversion of ethanol and selectivity to hydrogen are compared for the steam reforming of ethanol over Ni/ZnO, Ni/La2 O3 , Ni/MgO and Ni/ -Al2 O3 catalysts at TR = 650 C. An almost complete conversion of ethanol was attained for all catalysts. However, selectivity to hydrogen was affected by the support used and occurred in the following order: Ni/ZnO Ni/La2 O3 > Ni/MgO > Ni/r -Al2 O3 . The data in Table 1 further show that less byproducts were formed over the Ni/ZnO catalyst than over the Ni/La2 O3 catalyst. Therefore, it is clear that the performance of a ZnOsupported nickel catalyst in steam reforming of ethanol is superior to other nickel catalysts reported so far. 3.2. Effect of reaction temperature Fig. 1 shows the behaviors of the Ni/ZnO catalyst at different temperatures. It can be seen from Fig.1A that the conversion of ethanol was almost complete in the range from 330 to 650 C, while selectivity to hydrogen increased with temperature and its increase could be roughly divided into three

Table 1 Catalytic performance of supported nickel catalysts in the ethanol steam-reforming reactiona
Catalystb BET surface area (m2 /g) Conversion of ethanol (%) Selectivity to hydrogen (%) Distribution of the products (%) CO Ni/r -Al2 O3 Ni/MgO Ni/La2 O3 Ni/ZnO
a Reaction b Ni

CH4 5.5 3.1 1.6 2.0

CO2 25.8 29.2 24.6 25.7

CH3 CHO 0.08

CH3 COCH3 0.7

113.0 24.0 7.2 7.8

100 100 100 100

78.2 82.2 89.3 89.1

11.0 4.4 6.1 5.5

conditions: T = 650 C, LSHV = 5 h1 , water/ethanol ratio = 8. loading = 10 wt%.

Y. Yang et al. / International Journal of Hydrogen Energy 31 (2006) 877 882

879

stages. In addition, the conversion of water increased with temperature and remained constant at temperatures above 530 C. Furthermore, as shown in Fig. 1B, the concentration of CO, CO2 and methane underwent a similar three-stage change. At temperatures below 380 C, the conversion of ethanol was already close to completion, the conversion of water was about 60%, and the selectivity to hydrogen was less than 60%. On the other hand, the concentration of CO and methane was high and decreased with the increase of temperature, while the concentration of CO2 increased. These phenomena indicate that at this stage the reforming reaction (1) and decomposition reaction (2) predominated; meanwhile, shift reaction (3), methanation reaction (4) and dehydrogenation reaction (5) were also involved: C2 H5 OH + 3H2 O 2CO2 + 6H2 , C2 H5 OH CH4 + H2 + CO, CO + H2 O CO2 + H2 , CO + 3H2 CH4 + H2 O, C2 H5 OH CH3 CHO + H2 . (1) (2) (3) (4) (5)

90 Selectivity to Hydrogen (%) 85

ethanol conversion
er si on

100
en

at er

dr og

80 75 70 65 60

co

nv

90 80 70 60 50

55 50 300 350 400 450 500 550 600 650 Temperature (C) 28 26 24 22 20 18 16 14 12 10 8 6 4 2 0 40

(A)

Product Distribution (%)

CO CH4 CO2 C2H4O

In the second stage, as the reaction temperature increased to 530 C, the consumption of water reached close to its stoichiometric value and the selectivity increased to over 80%. At the same time, the concentration of methane decreased further, the formation of CO2 passed through a maximum, and the concentration of CO underwent a minimum. These phenomena imply that besides reforming of ethanol (reaction 1) the reformation of methane (reactions (6)(8)) and the reverse shift reaction (9) were accelerated: CH4 + 2H2 O CO2 + 4H2 , CH4 + H2 O CO + 3H2 , CH4 + CO2 2CO + 2H2 , CO2 + 3H2 CO + H2 O. (6) (7) (8) (9)

(B)

300 350 400 450 500 550 600 650 700 Temperature (C )

Fig. 1. Effect of temperature on conversion and selectivity to hydrogen (A) and distribution of products (B).

3.3. Effect of water to ethanol ratio It was found that the ratio of water to ethanol plays an important role in ethanol steam reforming, especially in the selectivity towards hydrogen [3]. The effect of water to ethanol ratio was investigated for the Zi/ZnO catalyst and the results are given in Fig. 2. The experiments were carried out at 650 C. It was observed that whatever the amount of water was initially introduced, the conversion of ethanol was complete and no intermediate products, such as acetaldehyde and ethylene, could be detected. It can be seen in Fig. 2 that the selectivity to hydrogen was about 81% at a stoichiometirc ratio of 3 and increased signicantly with the increase in the water to ethanol ratio. At a ratio of 9, the selectivity was about 91%. On further increasing the ratio to 12, the selectivity increased only by about 1%. The concentration of CO2 in the outlet gas stream also increased with the increase of water to ethanol ratio, while the formation of the byproducts, CO and methane, decreased with the in-

Furthermore, at temperatures above 430 C the formation of acetaldehyde could no longer be detected, indicating that acetaldehyde was decomposed completely at this stage: CH3 CHO CH4 + CO. (10)

In the third stage (from 530 to 650 C), the reaction between ethanol and water was stoichiometrically complete and selectivity to hydrogen increased further with temperature to over 90% at 650 C. The concentration of methane was decreased further to below 2%. However, the content of CO2 did not recover and correspondingly, the formation of CO increased. Therefore, it can be recognized that the dry reforming reaction (8) and the reverse shift reaction (9) were accelerated in this stage, although the steam reforming of methane was also promoted at a high temperature.

Conversion of Reactant (%)

se

le

ct iv

ity

to

hy

880

Y. Yang et al. / International Journal of Hydrogen Energy 31 (2006) 877 882

30 25 Product Distribution (%) 20 15 10 5 0 2 4 6 8 10 Water: Ethanol (molar ratio) 12 CO CH4 CO2

92 25 Selectivity to Hydrogen (%) 90 88 86 84 82 80 CO CH4 CO2

92 91 90 89 88 10 87 5 86 85 25 Selectivity to Hydrogen (%)

Product Distribution (%)

20

15

0 5 10 15 20 Liquid Hourly Space Velocity (h-1)

Fig. 2. Effect of water to ethanol molar ratio on selectivity to hydrogen and distribution of byproducts.

Fig. 3. Effect of LHSV on selectivity to hydrogen and byproduct distribution.

crease of H2 O/EtOH molar ratio. Obviously, in accordance with the mass action law, at higher water to ethanol ratios, not only were the reforming reactions (1) and (6) and the shift reaction (3) favored but also the reverse shift reaction (9) could be inhibited. Similar observations were also reported for other catalyst systems [3,12]. On the other hand, it would also be benecial for the suppression of carbon deposit on the catalyst to conduct steam reforming of ethanol at a molar ratio of higher than the stoichiometric ones [12]. 3.4. Effect of liquid space velocity The effect of Liquid Hourly Space Velocity (LHSV) on the catalytic performance of the 10 wt% Ni/ZnO at reaction temperatures of 650 C is illustrated in Fig. 3. The conversion of ethanol was complete and no intermediate byproducts (acetaldehyde and ethylene) were detected in the whole range of LHSV investigated. The selectivity decreased with the increase in LHSV, but it was still over 88% at LHSV = 24.0 h1 . At the same time, no signicant change in the concentration of CO and CO2 occurred, while the concentration of methane increased slightly. According to the reactions mentioned in Section 3.2, the decrease in selectivity and the increase of methane formation may be caused by the competition between the reformation reactions (i.e. Eqs. (1), (6) and (8)) and the decomposition reaction (2), because the reformation of byproduct methane, whether it was formed by the decomposition of ethanol or by methanation, will be more difcult when its resident time is shortened, and on the other hand, the increase in the formation of methane by one mole would lead to a loss of hydrogen by 4 mol. 3.5. Effect of Ni loading Nickel loading on zinc oxide was changed in the range between 5 and 20 wt% and the catalysts were evaluated at

temperatures ranging from 450 to 650 C. It was observed that the conversion of ethanol was complete over all the catalysts and at the experimental temperature range, with the exception of the catalyst 5 wt% Ni/ZnO, on which the conversion was about 99% at 450 C. Fig. 4A shows the effect of nickel loading on selectivity to hydrogen. It is evident that increasing Ni loading resulted in an increase in the selectivity. Furthermore, the increase of temperature led monotonously to an increase in selectivity, except for the catalyst with the highest nickel loading (20 wt% Ni/ZnO), on which the selectivity was over 90% at temperatures above ca. 520 C and reached a maximum at 550 C. The effect of nickel loading on the distribution of byproduct is depicted in Fig. 4BD. It can be found that in accordance with the change in selectivity the formation of intermediate byproducts was decreased with the increase in Ni loading and especially, the formation of methane was lowered by increasing Ni loading. On the catalyst 20 wt% Ni/ZnO the concentration of methane became minimal at 550 C. 4. Conclusions The following conclusions can be drawn from the present investigation: (1) Zinc oxide is a promising support for preparing a nickelbased reforming catalyst. Over Ni/ZnO catalyst a complete conversion of ethanol and over 90% of selectivity to hydrogen could be obtained at temperatures above 520 C. (2) Increasing the molar ratio of water to ethanol in the range between 3 and 12 and the Ni-loading in the range between

Y. Yang et al. / International Journal of Hydrogen Energy 31 (2006) 877 882

881

95 Product Distribution (%)

90

Product Distribution (%)

20% 15% 10% 5%

27 26 25 24 23 22 21 20

20% 15% 10% 5%

85

80

75

70 450 (A) 25 20% 15% 10% 5% 9 8 Product Distribution (%) 7 6 5 4 3 2 1 0 450 (B) 500 550 Temperature (C) 600 650 (D) 0 450 500 550 600 650 Temperature (C) 20% 15% 10% 5% 500 550 600 Temperature (C) 650 (C) 450 500 550 600 Temperature (C) 650

20 Product Distribution (%)

15

10

Fig. 4. Effect of nickel loading on selectivity to hydrogen (A) and formation of ethane (B), carbon monoxide (C) and carbon dioxide (D).

5 and 20 wt% increased the selectivity to hydrogen and suppressed the formation of methane. (3) The increase in LHSV within the range from 5.0 to 24.0 h1 resulted in a decrease in the selectivity to hydrogen, but the inuence was insignicant. References
[1] Milne TA, Elam CC, Evans RJ. Hydrogen from biomass. Golden, CO, USA: National Renewable Energy Laboratory. [2] Garcia EY, Lordorde MA. Hydrogen production by steam reforming of ethanol: thermodynamic analysis. Int J Hydrogen Energy 1991;16(5):30712.

[3] 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(1): 3145. [4] Freni S, Maggio G, Cavallaro S. Ethanol steam reforming in a molten fuel cell: a thermodynamic approach. J Power Sources 1996;62(1):6773. [5] Athanasios NF, Dimitris IK, Xenophon EV. Production of hydrogen for fuel cells by reformation of biomass-derived ethanol. Catal Today 2002;75(14):14555. [6] Marino FJ, Boveri M, Baronetti G, Laborde M. Hydrogen production from steam reforming of bioethanol using Cu/Ni/K / -Al2 O3 catalysts. Effect of Ni. Int J Hydrogen Energy 2001;26(7):6658.

882

Y. Yang et al. / International Journal of Hydrogen Energy 31 (2006) 877 882 [10] Freni S, Mondello N, Cavallaro S, Cacciola G, ParmonVN, Sobyamin VA. Hydrogen production by steam reforming of ethanol: a two step process. React Kinet Catal Lett 2000;71(1):14352. [11] Breen JP, Burch R, Coleman HM. Metal-catalysed steam reforming of ethanol in the production of hydrogen for fuel cell applications. Appl Catal B: Environmental 2002;39(1): 6574. [12] Goula MA, Kontou SK, Tsiakaras PE. Hydrogen production by ethanol steam reforming over a commercial Pd/ -Al2 O3 catalyst. Appl Catal B: Environmental 2004;49(2):13544.

[7] Comas J, Marino F, Laborde M, Amadeo N. Bio-ethanol steam reforming on Ni/Al2 O3 catalyst. J Chem Eng 2004;98(12):618. [8] Freni S, Cavallaro S, Mondello N. Steam reforming of ethanol on Ni/MgO catalysts: H2 production for MCFC. J Power Sources 2002;108(12):537. [9] Freni S, Cavallaro S, Mondello N, Spadaro L, Frusteri F. Production of hydrogen for MC fuel cell by steam reforming of ethanol over MgO supported Ni and Co catalysts. Catal Commun 2003;4(6):25968.

Vous aimerez peut-être aussi