You are on page 1of 9

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

Available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/he

Effect of N2O-mediated calcination on nickel species and the catalytic activity of nickel catalysts supported on g-Al2O3 in the steam reforming of glycerol
Youngbo Choi a, Nam Dong Kim a, Jayeon Baek a, Wooyoung Kim b, Hee Jong Lee b, Jongheop Yi a,*
World Class University (WCU) Program of Chemical Convergence for Energy & Environment (C2E2), School of Chemical and Biological Engineering, ICP, Seoul National University (SNU), Daehak-dong, Gwanak-gu, Seoul 151-742, Republic of Korea b GS-Caltex Corporation, R&D Center, Daejeon 305-380, Republic of Korea
a

article info
Article history: Received 5 August 2010 Received in revised form 2 December 2010 Accepted 18 December 2010 Available online 23 January 2011 Keywords: Glycerol Steam reforming N2O Ni/g-Al2O3 Calcination

abstract
The steam reforming of glycerol over supported nickel catalysts is a promising and costeffective method for producing hydrogen. The activity of nickel catalysts supported on g-Al2O3 is low, primarily due to the formation of inactive nickel species during high temperature calcination in air. In order to address this problem, a Ni/g-Al2O3 catalyst was prepared by calcination at 700  C in a nitrous oxide (N2O) environment. The N2O calcined catalyst showed an enhanced activity for the steam reforming of glycerol. A variety of characterization techniques (XRD, TPR, XPS and H2 Chemisorption) conrmed that the high temperature N2O calcination resulted in a signicant decrease in the levels of nickel aluminate. The N2O calcination also led to an enhancement in the amount of NiO as well as nickel ions present on the surface of the catalyst. Interestingly, compared to an air calcined catalyst, the N2O calcined catalyst contained larger nickel particles after reduction but the N2O calcined catalyst had a much larger nickel surface area and dispersion, which resulted in higher glycerol conversion and hydrogen yield. Copyright 2011, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.

1.

Introduction

Glycerol represents a promising renewable resource for producing hydrogen [1e4], because it is a major by-product of the production of biodiesel [5] and can also be produced by fermentation [6]. Hydrogen can be catalytically produced from glycerol with high selectivity by means of the steam reforming, a well established technology [3,7e9]. In addition, the steam reforming of glycerol has many merits, such as low toxicity, ease of handling and the expected growth in the availability of glycerol due to the rapid expansion of biodiesel production [5,10]. Among the various supported catalysts that have been investigated for the steam reforming of glycerol, nickel-based

catalysts have activities comparable to noble metal-based catalysts [7,11e14]. Moreover, the low cost and availability of nickel compared to noble metal have prompted investigations of nickel-based catalysts. One of the frequently encountered problems associated with supported nickel catalysts is the formation of catalytically inactive nickel species, produced by reactions between nickel and the support [15,16]. A good case in point is the development of nickel aluminate in Ni/g-Al2O3 catalysts [10,17e19]. The latter has been widely studied in the steam reforming of hydrocarbons (methane [20,21], ethanol [22,23], methanol [24] etc.) as well as glycerol [11,25]. It is well known that nickel aluminate is more difcult to reduce than nickel

* Corresponding author. Tel.: 82 2 880 7438; fax: 82 2 885 6670. E-mail address: jyi@snu.ac.kr (J. Yi). 0360-3199/$ e see front matter Copyright 2011, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.ijhydene.2010.12.081

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

3845

oxide (NiO), which reduces the number of active sites on the catalyst [26]. In this regard, the temperature used for calcination is an important factor that affects the types of nickel species produced in the catalysts, as well as catalytic activity. As the calcination temperature increases, the fraction of NiO decreases while the fraction of nickel aluminate increases due to the facile diffusion of nickel ions into the spinel structure of g-Al2O3 at high temperatures [27]. The objective of this study is to examine conditions favorable for suppressing the formation of inactive nickel species during high temperature calcination to enhance the catalytic activity. Our strategy is based on moderating the diffusion of nickel ions into g-Al2O3 in the calcination step. It has been reported that oxygen scavenging radicals (NO, N2O, CO and H2) can be used to prevent the sintering and redistribution of nickel nitrate supported on SBA-15 during calcination at 450  C, leading to the production of smaller NiO particles on the catalyst [28e30]. From these ndings, we hypothesized that if nickel precursors impregnated on g-Al2O3 resist the sintering and redistribution during the calcination, nickel ions would migrate into the structure of g-Al2O3 with difculty. Therefore, we have investigated the effects of nitrous oxide (N2O)-mediated calcination on Ni/g-Al2O3 catalysts for the steam reforming of glycerol, with particular attention to the formation of nickel species in the catalyst.

2.
2.1.

Experimental
Preparation of catalysts

The temperature of the furnace was increased from room temperature to 1000  C at a heating rate of 5  C min1. The signal for H2 consumption was continuously recorded by means of a thermal conductivity detector (TCD). The Ni surface area and dispersion of Ni on the reduced catalysts were measured by H2-pulse chemisorption using a Micromeritics Autochem II chemisorption analyzer. Prior to the analysis, about 0.1 g of the calcined sample was in situ reduced under a ow of 9.6 v/v% H2/Ar (50 mL min1) at 700  C for 3 h. The sample was then ushed with a stream of Ar (50 mL min1) at 700  C for 20 min, and cooled to 50  C. After this procedure, 36.2 mL of adsorbing gas (9.6% H2/Ar) was periodically injected onto the sample at intervals of 4 min using an injection loop. The amount of H2 uptake was monitored by a TCD until successive peaks on the chromatograph exhibited the same area. The Ni surface area was calculated from the following equation Am m2 =gNi Fs na Ag =M% 100, where Fs is an adsorption stoichiometry of H/Ni (1), na the number of moles of hydrogen adsorbed per gram of catalyst (mol/gcatalyst), Na Avogadros number (6.02 1023 mol1), Ag the cross-sectional area of nickel atom (6.49 1020 m2), and M% the weight percent of nickel in the catalyst (%). The surface chemical environment and composition of the samples were investigated by X-ray photoelectron spectroscopy (XPS). XPS spectra were recorded with a Kratos AXIS-HSi electron spectrometer equipped with a Mg-Ka X-ray source and a hemispherical electron energy analyzer. The charging potential of the samples was corrected by setting the binding energy of the adventitious carbon 1s to 284.6 eV. The carbon content of the catalysts, before and after the reaction, was determined by CHNS elemental analysis using a LECO CHNS932 instrument.

Ni catalysts supported on g-Al2O3 were prepared by incipient wetness impregnation. Ni(NO3)26H2O (Fluka) was dissolved in deionized water and the nickel precursor solution was added dropwise to g-Al2O3 (Degussa, specic surface area (BET) 100 m2/g) to give a 10 wt.% Ni loading. The impregnated sample was then dried overnight at 80  C. The dried samples were then calcined in both air and N2O/He (10 v/v%), respectively. The calcination was performed at 700  C for 4 h using a gas ow rate of 80 mL min1. The Ni/g-Al2O3 catalysts calcined in N2O/He and air are denoted as NiAleN2O and NiAleAir, respectively. A commercial NiO (II) powder (Junsei Chemical Co.) and nickel aluminate (NiAl2O4) were used as reference compounds. NiAl2O4 was prepared according to a previously reported method [31].

2.3.

Catalytic activity

2.2.

Characterizations

The BET surface area of the samples was determined by nitrogen adsorption at 196  C using a Micromeritics ASAP2010 system. The crystalline phase was examined by X-ray diffraction (XRD) using a Rigaku D-MAX2500-PC powder X-ray diffractometer with Cu Ka radiation (1.5406 A). The average particle size of metallic Ni was calculated from an XRD peak at 2q of 51.8 using the Scherrer equation. Hydrogen temperature programmed reduction (H2-TPR) measurements were carried out in a conventional apparatus. About 0.1 g of the sample was loaded into a U-shaped quartz microreactor, and a 10 v/v% H2/N2 mixture was allowed to ow at 50 mL min1.

The steam reforming of glycerol was carried out under atmospheric pressure and at a temperature of 600  C. A catalyst sample of 0.2 g was loaded into a quartz reactor, and the reactor was then placed in an electric furnace. The temperature of the catalyst bed was monitored by a k-type thermocouple located inside the reactor on top of the catalyst layer, and controlled by a PID controller. Prior to a reaction, the catalyst was reduced in situ at 700  C for 3 h in a mixture of H2 (3 mL min1) and N2 (30 mL min1), followed by cooling to the reaction temperature of 600  C in a stream of pure N2. A reactant mixture of glycerol (Acros, 99 %) and water, with a steam to carbon (S/C) ratio of 8 (molar ratio of glycerol/ water 1/24), was injected into a vaporizer using a Hamilton syringe mounted in a Cole-parmer syringe infusion pump at a rate of 1.2 mL h1. The reactant vapors were diluted with N2 carrier gas in the vaporizer, and then fed into the catalyst bed. The temperature of the vaporizer and a line between the vaporizer and the reactor was maintained at 300  C. The ow rate of the N2 carrier gas was adjusted so as to obtain gas 1 hourly space velocity (GHSV) values of 16,000 mL g1 catalyst h .  The reaction temperature of 600 C and the S/C ratio of 8 were selected, based on our and previously reported thermodynamic calculation results [32], which indicated that these were the optimum conditions for achieving maximum hydrogen production.

3846

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

The product stream was separated into a liquid phase and a gas phase in a condenser connected to the reactor outlet. Seven gaseous products were analyzed on-line by gas chromatography using a Younglin ACME 6100 model chromatograph equipped with two TCDs and two automated injection valves. H2 was separated on a 13 molecular sieve-packed column using Ar as the carrier gas. N2, CO, CO2, CH4, C2H2, C2H4 and C2H6 were separated in a Carboxen-1000 column using He as the carrier gas. In the steam reforming of glycerol, hydrogen can be produced via the following overall reaction: C3 H8 O3 g 3H2 Og43CO2 g 7H2 g (1)

The overall reaction (1) may be considered as the combination of glycerol decomposition (2) and water-gas-shift reaction (3): C3 H8 O3 g 4 3COg 4H2 g COg H2 Og 4 CO2 g 4H2 g (2) (3) Fig. 1 e XRD patterns of NiO, g-Al2O3, fresh NiAleN2O sample, fresh NiAleAir sample and NiAl2O4.

Other possible reactions can be described as follows [33,34]: C3 H8 O3 g 4 C2 H4 g COg 2H2 Og C3 H8 O3 g 2H2 g 4 2CH4 g COg 2H2 Og COg 3H2 g 4 CH4 g H2 Og CO2 g 4H2 g 4 CH4 g 2H2 Og CO2 g CH4 g 4 2COg 2H2 g C2 H4 g H2 g 4 C2 H6 g (4) (5) (6) (7) (8) (9)

The conversion of glycerol to gaseous products (Xg) was calculated according to the following equation (10) on the basis of the carbon balance from the gaseous products. Xg % moles of C atoms in the gaseous products 100 3 moles of glycerol in the feed (10)

The gaseous products included CO, CO2, CH4, C2H2, C2H4 and C2H6. The yield of hydrogen YH2 was calculated using equation (2) on the assumption that 7 moles of hydrogen are produced per mole of glycerol. YH2 % moles of H2 products 100 7 moles of glycerol in the feed (11)

A close examination of the NiAleN2O sample, however, revealed a small diffraction peak at 2q of 43.3 , suggesting the presence of NiO in the NiAleN2O sample. It should be noted that both samples showed a (440) diffraction peak at 2q between 65.5 (for NiAl2O4) and 67.3 (for g-Al2O3). In addition, the (440) diffraction peak for the NiAleAir sample was located at a lower diffraction angle compared to that for the NiAleN2O sample. Because nickel ions have a larger ionic radius than aluminum, as nickel ions are incorporated into the lattice of g-Al2O3, the lattice parameters of the Ni/g-Al2O3 catalyst are expanded and the (440) diffraction peak shifts to a lower diffraction angle [36e38]. Therefore, these results suggest that nickel aluminate is formed in both samples, and that more nickel ions are strongly incorporated into the lattice of alumina in the case of the NiAleAir sample.

3.2.

Crystalline structure of reduced catalysts

3.
3.1.

Results and discussion


Crystalline structure of fresh samples

The XRD patterns of the fresh NiAleN2O and NiAleAir samples are shown in Fig. 1. For comparison, the XRD patterns of NiO, g-Al2O3, and NiAl2O4 are also included. For the NiAleAir sample, no characteristic diffraction peaks corresponding to crystalline NiO were observed, which is consistent with the XRD data reported in the literature [35,36].

After the samples had been reduced at 700  C for 3 h, some additional differences were observed in the XRD patterns. The XRD patterns for the reduced NiAleN2O and NiAleAir catalysts, as well as NiAl2O4 and g-Al2O3 are shown in Fig. 2. Characteristic diffraction peaks indicating the presence of metallic nickel (Nio) were observed for both catalysts. It has been reported that the N2O calcination assists in forming smaller NiO particles than the air calcination [29]. Therefore, it would be reasonably expected that the reduction of the NiAleN2O sample would lead to the formation of smaller Nio particles than that of the NiAleAir sample. Compared to the NiAleAir catalyst, however, the diffraction peaks for Nio in the NiAleN2O catalyst were more intense and sharper, indicating the presence of larger Nio particles. Indeed, the average particle size of Nio particles, as estimated by the Scherrer

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

3847

Fig. 2 e XRD patterns of g-Al2O3, reduced NiAleN2O catalyst, reduced NiAleAir catalyst and NiAl2O4 (O : metallic Ni). Fig. 3 e TPR proles of NiAleN2O and NiAleAir samples.

equation, was 15.8 and 10.0 nm for the reduced NiAleN2O and NiAleAir samples, respectively. This phenomenon can be explained in terms of the different states of nickel species present in the fresh samples. In the NiAleN2O sample, NiO as well as nickel aluminate can be formed during the N2O calcination. NiO is more easily reduced and is also more susceptible to sintering during the reduction step than nickel aluminate [38]. Accordingly, both NiO and nickel aluminate can form Nio particles after reduction and the Nio particles that are produced from NiO can account for the large size particles observed in the NiAleN2O catalyst. On the other hand, the fresh NiAleAir sample is mainly composed of nickel aluminate as conrmed by the data shown in Fig. 1, so the Nio particles can originate from the reduction of nickel aluminate and these particles appear to be more resistant to sintering. It should also be noted that the (440) diffraction peak of the reduced catalysts shifted back to a high diffraction angle close to that of g-Al2O3. However, their position was still slightly lower, especially in the case of the NiAleAir catalyst. This implies that some amount of nickel ions in both catalysts were not completely reduced, and that more nickel ions remained unreduced in the NiAleAir catalyst.

3.3.

Reducibility and nickel species

TPR analyses were performed to investigate the reducibility and nickel species present in the fresh samples and the results are shown in Fig. 3. The NiAleAir sample showed a peak maximum at 790  C, while the NiAleN2O sample exhibited a peak maximum at 761  C. This indicates that the NiAleN2O sample is more easily reduced.

Compared to the NiAleAir sample, the TPR prole of the NiAleN2O sample was relatively asymmetric with tailing to low temperature, suggesting a more heterogeneous distribution of nickel species. The difference in nickel species was also investigated through deconvolution of the TPR prole. For the NiAleN2O sample, three reduction peaks with maxima at 558, 706 and 769  C were observed. In the case of the NiAleAir sample, however, best tting result was achieved with only two reduction peaks with maxima at 740 and 796  C. Generally, NiO, when bonded to alumina, is reduced at temperature of 400e690  C, while nickel aluminate is reduced at temperatures in excess of 700  C [27,39,40]. It is also known that nickel aluminate can exist in two forms: (i) bulk nickel aluminate and (ii) surface nickel aluminate. The reducibility of surface nickel aluminate is between bulk nickel aluminate and NiO [38,41,42]. In a previous study, Guo et al. reported that nickel aluminate on the surface of the support is a nickelsaturated compound (stoichiometric nickel aluminate, NiAl2O4), while that in the inner parts of the support is a nickel-unsaturated compound (non-stoichiometric nickel aluminate, NiAlxOy, x > 2) [43]. Thus, it can be reasonably concluded that the surface and bulk nickel aluminate are NiAl2O4 and NiAlxOy species, respectively. Based on these literature reports, it can be concluded that the NiAleN2O sample contained readily reducible NiO, reducible NiAl2O4 and NiAlxOy, which is reduced only with difculty, while the NiAleAir sample contained NiAl2O4 and NiAlxOy. This conclusion is consistent with the XRD results. Table 1 lists the areas for each reduction peak and their sum. Although the total peak area was nearly same for both samples, the NiAleN2O sample had smaller peak area for

3848

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

Table 1 e Area of reduction peaks in TPR proles of NiAleN2O and NiAleAir samples. Total peak area (a.u.) NiO peak area (a.u.) Nickel aluminate peak area (a.u.) Stoichiometric NonNiAl2O4 stoichiometric NiAlxOy
174.9 205.9 191.9 240.6

Table 2 e Binding energies for Ni 2p3/2 and nominal and surface Ni/Al atomic ratios, as determined by XPS. Ni 2p3/2 (eV) Ni/Al atomic ratio Nominal
NiO NiAl2O4 NiAleN2O NiAleAir 855.4 855.6 855.5 855.6 853.6 e 854.1 854.0 e 0.500 0.097 0.097

Surface
e 0.513 0.236 0.183

NiAleN2O NiAleAir

447.9 446.5

81.1 e

nickel aluminate, especially for non-stoichiometric NiAlxOy, than the NiAleAir sample. This result indicates that the NiAleN2O sample retained a lower amount of nickel aluminate, particularly NiAlxOy, than the NiAleAir sample.

3.4.

Surface chemical environment and composition

In order to investigate the surface chemistry of the prepared samples, XPS analyses were conducted. Fig. 4 shows the baseline subtracted XPS spectra of the samples for the Ni 2p3/2 region, along with corresponding spectra for the reference NiAl2O4 and NiO compounds. The binding energies of the Ni 2p3/2 spectra were summarized in Table 2. The Ni 2p3/2 spectrum of NiO showed a clear multiplet splitting at 855.4 and 853.6 eV with a shake-up satellite peak at 860.9 eV, which is in good agreement with data from literature reports [44,45]. The reference NiAl2O4 exhibited a characteristic peak at 855.6 eV and a shake-up satellite peak at 861.8 eV, similar to reported values [31,46]. For the NiAleN2O sample, the Ni 2p3/2 peak was slightly asymmetric and consisted of a main peak centered at 855.5 eV

and a shoulder peak centered at 854.1 eV. This indicates the presence of both NiO and nickel aluminate on the surface of the NiAleN2O sample. For the NiAleAir sample, however, a shoulder peak centered at 854.0 eV was signicantly small and the shape of the Ni 2p3/2 peak was close to that of NiAl2O4. This clearly indicates that nickel aluminate was predominantly formed in the NiAleAir sample. The surface Ni/Al atomic ratios of the samples were also determined by XPS, and these ndings are listed along with the nominal Ni/Al atomic ratios in Table 2. The surface Ni/Al atomic ratio for the NiAleN2O sample was quite high compared to the NiAleAir sample. Therefore, it can be concluded that more nickel ions were located on the surface of the support in the NiAleN2O sample, which is well consistent with the TPR results.

3.5.

Nickel surface area of reduced catalysts

BET surface areas for fresh samples and H2 chemisorption results for the reduced catalysts are summarized in Table 3. A larger amount of hydrogen was chemisorbed on the NiAleN2O catalyst compared to the NiAleAir catalyst, although both samples showed almost the same BET surface area. Consequently, the NiAleN2O catalyst showed a larger nickel surface area and dispersion. The large nickel surface area in the NiAleN2O catalyst can be attributed to the presence of easily reducible NiO and a relatively low amount of NiAlxOy. On the other hand, the formation of larger amounts of NiAlxOy explains the small nickel surface area for the reduced NiAleAir catalyst.

3.6.

Catalytic activity

The activity of the NiAleN2O and NiAleAir catalysts for the steam reforming of glycerol was measured at 600  C. Reaction results are shown in Figs. 5 and 6. The NiAleN2O catalyst

Table 3 e BET surface area and H2 chemisorption results for NiAleN2O and NiAleAir samples. Hydrogen Nickel Nickel BET uptake surface dispersion surface area (%)a area (m2/g) (mmol/gcatalyst) (m2/gNi)a
NiAleN2O NiAleAir 92.8 95.3 24.6 19.8 9.6 7.7 1.4 1.1

Fig. 4 e XPS spectra of NiAl2O4, NiAleAir sample, NiAleN2O sample and NiO for the Ni 2p3/2 region after baseline subtraction.

a Calculated by assuming the adsorption stoichiometry of one hydrogen atom per nickel atom (H/Ni 1).

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

3849

showed a 13e17% higher glycerol conversion and hydrogen yield than the NiAleAir catalyst during all time on stream. The glycerol conversion and hydrogen yield over the NiAleN2O catalyst were also close to the thermodynamically calculated values. For the NiAleAir and NiAleN2O catalysts, the gas products were mainly H2, CO2, CO and CH4. C2 compounds (C2H2, C2H4 and C2H6), however, were not detected and the CH4 production was quietly low for both catalysts. This can be explained by the high S/C ratio of 8, which favors the products in the thermodynamic equilibrium and inhibits the formation of CH4 [12,32]. Compared to the NiAleAir catalyst, the NiAleN2O catalyst exhibited higher H2 and CO2 production and slightly lower CO production, indicating that the NiAleN2O catalyst has better performance for the steam reforming of glycerol. A decrease in activity was observed with time on stream in both catalysts, and this appears to be due to coke formation. Table 4

Fig. 6 e Molar amount of gaseous products per mole of glycerol: (a) NiAleAir and (b) NiAleN2O.

presents the carbon contents of the catalysts. Both fresh catalysts showed a small amount of carbon contents, which is probably due to the carbon contamination. The amount of carbon contents of both catalysts were increased after the reaction. However, the carbon content was slightly lower in the case of the NiAleN2O catalyst than the NiAleAir catalyst. The enhanced activity of the NiAleN2O catalyst can be attributed to its favorable properties. Metallic nickel is known to be the active component for the steam reforming of glycerol

Fig. 5 e Catalytic activity of NiAleAir and NiAleN2O catalysts for the steam reforming of glycerol (reaction conditions: T [ 600  C, S/C ratio [ 8 and L1 ): (a) glycerol conversion and GHSV [ 16,000 mL gL1 catalyst h (b) hydrogen yield. Thermodynamic conversions and yields were calculated with Rgibbs reactor module of the Aspen plus software program, version 7.0, assuming ten components (C3H8O3, H2O, H2, CO2, CO, CH4, C, C2H2, C2H4 and C2H6).

Table 4 e Carbon content of fresh and used NiAleN2O and NiAleAir catalysts. C content (wt.%) Fresh
NiAleN2O NiAleAir 0.95 0.12

Useda
9.14 9.86

a Measured after a 360 min reaction.

3850

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

Fig. 7 e Schematic representation of nickel species in NiAleAir and NiAleN2O samples.

due to its ability to catalyze the cleavage of CeC bonds [12,13,47]. The enhanced reducibility as well as high nickel surface area and dispersion of the NiAleN2O catalysts provides more active sites for the steam reforming of glycerol, resulting in a higher activity.

4.

Conclusions

3.7.

Comparison of N2O and air calcination

Because nickel species in the Ni/g-Al2O3 catalyst can affect the reducibility and activity of the catalyst, the preparation of the appropriate nickel species is an important issue, especially when the samples are treated at high temperatures, which promotes the formation of inactive nickel species. The characterization results revealed that, for the NiAleAir sample, nickel nitrate impregnated on g-Al2O3 was converted into poorly reducible nickel aluminate. In particular, the data indicate that relatively large amounts of NiAlxOy are formed in the support. In contrast, for the NiAleN2O sample, NiO and NiAl2O4 were formed on the surface of the support. The amount of nickel ions present on the surface of the support was also enhanced in the case of the NiAleN2O sample (Fig. 7). Owing to the ability of N2O to prevent the sintering and redistribution of supported nickel nitrate [29], N2O appears to inhibit the diffusion of nickel ions into the structure of g-Al2O3 during the calcination at 700  C. This leads to the formation smaller amounts of NiAlxOy in the inner part of the support. After reduction, for the case of the NiAleAir catalyst, Nio particles were not grown to a large size but an active Ni surface was not sufciently formed, which results in low catalytic activity for the steam reforming of glycerol. The reason for this may be because Ni ions in the inner parts of the support are more resistant to sintering as well as more difcult to reduce [38], and hydrogen molecules are able to diffuse deeper into the support during the reduction step only with difculty [43]. On the other hand, the NiAleN2O catalyst was sufciently reduced to provide a high Ni surface area and dispersion even with larger Nio particles present, resulting in a higher catalytic activity (Fig. 7).

The Ni/g-Al2O3 catalyst prepared by calcination at 700  C in a N2O environment showed improved catalytic activities for the steam reforming of glycerol, as compared to a catalyst that was subjected to the calcination in air. The N2O calcined sample retained readily reducible NiO, reducible NiAl2O4 and barely reducible NiAlxOy. On the other hand, the air calcined sample mainly consisted of NiAl2O4 and NiAlxOy. In addition, smaller amounts of nickel aluminate, especially NiAlxOy, were formed in the N2O calcined sample and more nickel ions were enriched on the surface of the N2O calcined sample. After the reduction treatment, although larger nickel particles were produced in the N2O calcined catalyst, probably due to the sintering of NiO, the N2O calcined catalyst had a larger nickel surface area and dispersion. This resulted in an enhancement in glycerol conversion and hydrogen yield.

Acknowledgments
This research was partially supported by WCU (World Class University) program through the National Research Foundation of Korea funded by the Ministry of Education, Science and Technology (R31-10013) and Korea Ministry of Environment as The Eco-technopia 21 project.

references

[1] Behr A, Eilting J, Irawadi K, Leschinski J, Lindner F. Improved utilisation of renewable resources: new important derivatives of glycerol. Green Chem 2008;10:13e30. [2] Cortright RD, Davda RR, Dumesic JA. Hydrogen from catalytic reforming of biomass-derived hydrocarbons in liquid water. Nature 2002;418:964e7.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

3851

[3] Slinn M, Kendall K, Mallon C, Andrews J. Steam reforming of biodiesel by-product to make renewable hydrogen. Bioresour Technol 2008;99:5851e8. [4] Soares RR, Simonetti DA, Dumesic JA. Glycerol as a source for fuels and chemicals by low-temperature catalytic processing. Angew Chem Int Ed 2006;45:3982e5. [5] Johnson DT, Taconi KA. The glycerin glut: options for the value-added conversion of crude glycerol resulting from biodiesel production. Environ Prog 2007;26:338e48. [6] Remize F, Roustan JL, Sablayrolles JM, Barre P, Dequin S. Glycerol overproduction by engineered Saccharomyces cerevisiae wine yeast strains leads to substantial changes in byproduct formation and to a stimulation of fermentation rate in stationary phase. Appl Environ Microbiol 1999;65:143e9. [7] Hirai T, Ikenaga N, Miyake T, Suzuki T. Production of hydrogen by steam reforming of glycerin on ruthenium catalyst. Energy Fuels 2005;19:1761e2. [8] Pompeo F, Santori G, Nichio NN. Hydrogen and/or syngas from steam reforming of glycerol. Study of platinum catalysts. Int J Hydrogen Energy 2010;35:8912e20. [9] iriondo A, Barrio VL, Cambra JF, Arias PL, Guemez MB, Sanchez-Sanchez MC, et al. Glycerol steam reforming over Ni catalysts supported on ceria and ceria-promoted alumina. Int J Hydrogen Energy 2010;35:11622e33. [10] Profeti LPR, Ticianelli EA, Assaf EM. Production of hydrogen via steam reforming of biofuels on Ni/CeO2eAl2O3 catalysts promoted by noble metals. Int J Hydrogen Energy 2009;34: 5049e60. [11] Adhikari S, Fernando S, Haryanto A. Production of hydrogen by steam reforming of glycerin over alumina-supported metal catalysts. Catal Today 2007;129:355e64. [12] Cui Y, Galvita V, Rihko-Struckmann L, Lorenz H, Sundmacher K. Steam reforming of glycerol: the experimental activity of La1xCexNiO3 catalyst in comparison to the thermodynamic reaction equilibrium. Appl Catal B 2009;90:29e37. [13] Zhang B, Tang X, Li Y, Xu Y, Shen W. Hydrogen production from steam reforming of ethanol and glycerol over ceriasupported metal catalysts. Int J Hydrogen Energy 2007;32: 2367e73. [14] Iriondo A, Barrio VL, Cambra JF, Arias PL, Guemez MB, Navarro RM, et al. Inuence of La2O3 modied support and Ni and Pt active phases on glycerol steam reforming to produce hydrogen. Catal Commun 2009;10:1275e8. [15] Choudhary VR, Uphade BS, Mamman AS. Large enhancement in methane-to-syngas conversion activity of supported Ni catalysts due to precoating of catalyst supports with MgO, CaO or rare-earth oxide. Catal Lett 1995;32:387e90. [16] Gil A, Daz A, Ganda LM, Montes M. Inuence of the preparation method and the nature of the support on the stability of nickel catalysts. Appl Catal A 1994;109:167e79. [17] Souza MdMVM, Clav L, Dubois V, Perez CAC, Schmal M. Activation of supported nickel catalysts for carbon dioxide reforming of methane. Appl Catal A 2004;272:133e9. [18] Ruckenstein E, Hu XD. The effect of steam on supported metal catalysts. J Catal 1986;100:1e16. [19] Profeti LPR, Dias JAC, Assaf JM, Assaf EM. Hydrogen production by steam reforming of ethanol over Ni-based catalysts promoted with noble metals. J Power Sources 2009; 190:525e33. [20] Navarro RM, Pena MA, Fierro JLG. Hydrogen production reactions from carbon feedstocks: fossil fuels and biomass. Chem Rev 2007;107:3952e91. [21] Yokota O, Oku Y, Sano T, Hasegawa N, Matsunami J, Tsuji M, et al. Stoichiometric consideration of steam reforming of methane on Ni/Al2O3 catalyst at 650  C by using a solar furnace simulator. Int J Hydrogen Energy 2000;25:81e6.

[22] Haryanto A, Fernando S, Murali N, Adhikari S. Current status of hydrogen production techniques by steam reforming of ethanol: a review. Energy Fuels 2005;19:2098e106. [23] Vizcano AJ, Arena P, Baronetti G, Carrero A, Calles JA, Laborde MA, et al. Ethanol steam reforming on Ni/Al2O3 catalysts: effect of Mg addition. Int J Hydrogen Energy 2008; 33:3489e92. [24] Palo DR, Dagle RA, Holladay JD. Methanol steam reforming for hydrogen production. Chem Rev 2007;107:3992e4021. [25] Sanchez EA, DAngelo MA, Comelli RA. Hydrogen production from glycerol on Ni/Al2O3 catalyst. Int J Hydrogen Energy;35: 5902e7. [26] Roh H-S, Jun K-W, Park S-E. Methane-reforming reactions over Ni/Ce-ZrO2/q-Al2O3 catalysts. Appl Catal A 2003;251:275e83. [27] Rynkowski JM, Paryjczak T, Lenik M. On the nature of oxidic nickel phases in NiO/g-Al2O3 catalysts. Appl Catal A 1993; 106:73e82. [28] Sietsma JRA, Meeldijk JD, Breejen JPd, Versluijs-Helder M, Dillen AJv, Jongh PEd, et al. The Preparation of supported NiO and Co3O4 nanoparticles by the nitric oxide controlled thermal decomposition of nitrates. Angew Chem Int Ed 2007; 46:4547e9. [29] Sietsma JRA, Friedrich H, Broersma A, Versluijs-Helder M, van Dillen A Jos, de Jongh PE, et al. How nitric oxide affects the decomposition of supported nickel nitrate to arrive at highly dispersed catalysts. J Catal 2008;260:227e35. [30] Sietsma JRA, Van Dillen AJ, de Jongh PE, de Jong KP. Metal nitrate conversion method, patent application WO2008029177; 2008. [31] Heracleous E, Lee AF, Wilson K, Lemonidou AA. Investigation of Ni-based alumina-supported catalysts for the oxidative dehydrogenation of ethane to ethylene: structural characterization and reactivity studies. J Catal 2005;231:159e71. [32] Adhikari S, Fernando S, Gwaltney SR, Filip To SD, Mark Bricka R, Steele PH, et al. A thermodynamic analysis of hydrogen production by steam reforming of glycerol. Int J Hydrogen Energy 2007;32:2875e80. [33] Li Y, Wang W, Chen B, Cao Y. Thermodynamic anaylsis of hydrogen production via glycerol steam reforming with CO2 adsorption. Int J Hydrogen Energy 2010;35:7768e77. [34] Cheng CK, Foo SY, Adesina AA. Glycerol steam reforming over bimetallic Co-Ni/Al2O3. Ind Eng Chem Res 2010;49: 10804e17. [35] Alberton AL, Souza MMVM, Schmal M. Carbon formation and its inuence on ethanol steam reforming over Ni/Al2O3 catalysts. Catal Today 2007;123:257e64. [36] Kim P, Kim Y, Kim H, Song IK, Yi J. Synthesis and characterization of mesoporous alumina with nickel incorporated for use in the partial oxidation of methane into synthesis gas. Appl Catal A 2004;272:157e66. [37] Seo JG, Youn MH, Jung JC, Song IK. Effect of preparation method of mesoporous NieAl2O3 catalysts on their catalytic activity for hydrogen production by steam reforming of liqueed natural gas (LNG). Int J Hydrogen Energy 2009;34: 5409e16. [38] Li G, Hu L, Hill JM. Comparison of reducibility and stability of alumina-supported Ni catalysts prepared by impregnation and co-precipitation. Appl Catal A 2006;301:16e24. [39] Zhu X, Huo P, Zhang Y-P, Cheng D-G, Liu C-J. Structure and reactivity of plasma treated Ni/Al2O3 catalyst for CO2 reforming of methane. Appl Catal B 2008;81:132e40. [40] Li C, Chen Y-W. Temperature-programmed-reduction studies of nickel oxide/alumina catalysts: effects of the preparation method. Thermochim Acta 1995;256:457e65. [41] Cimino A, Lo Jacono M, Schiavello M. Structural, magnetic, and optical properties of nickel oxide supported on h- and g-aluminas. J Phys Chem 1971;75:1044e50.

3852

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 6 ( 2 0 1 1 ) 3 8 4 4 e3 8 5 2

[42] Kharat AN, Pendleton P, Badalyan A, Abedini M, Amini MM. Decomposition of nickel formate on solegel alumina and characterization of product by X-ray photoelectron and TOF-SIMS spectroscopies. J Catal 2002;205:7e15. [43] Yang R, Li X, Wu J, Zhang X, Zhang Z, Cheng Y, et al. Hydrotreating of crude 2-ethylhexanol over Ni/Al2O3 catalysts: surface Ni species-catalytic activity correlation. Appl Catal A 2009;368:105e12. [44] Gavalas GR, Phichitkul C, Voecks GE. Structure and activity of NiO/a-Al2O3 and NiO/ZrO2 calcined at high temperatures: I.Structure. J Catal 1984;88:54e64.

[45] Loechel BP, Strehblow HH. Breakdown of passivity of nickel by uoride. II. Surface analytical studies. J Electrochem Soc 1984;131:713e23. [46] Shalvoy RB, Davis BH, Reucroft PJ. Studies of the metal-support interaction in coprecipitated nickel on alumina methanation catalysts using X-ray photoelectron spectroscopy (XPS). Surf Interface Anal 1980; 2:11e6. [47] Buffoni IN, Pompeo F, Santori GF, Nichio NN. Nickel catalysts applied in steam reforming of glycerol for hydrogen production. Catal Commun 2009;10:1656e60.

You might also like