You are on page 1of 12

INTERNATIONAL JOURNAL OF ENERGY RESEARCH

Int. J. Energy Res. 2013; 37:1896–1907


Published online 10 January 2013 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/er.3009

Hydrogen-rich gas production from ethanol steam-


reforming reaction using NiZr-loaded MCM-48 catalysts
at mild temperature
Jun Su Lee1, Dongjin Kim1, Byung-Hyun Choi2 and Misook Kang1,*,†
1
Department of Chemistry, College of Science, Yeungnam University, Gyeongsan, Gyeongbuk, 712-749, Republic of Korea
2
Korean Institutes of Ceramic Engineering and Technology (KICET), Geumcheon-gu, Seoul 153-801, Republic of Korea

SUMMARY
The rich-hydrogen generation from ethanol steam reforming over NiZr, which is used as an anode material in solid oxide
fuel cells, -loaded MCM-48 (NiZr/MCM-48) catalyst was investigated in this study. We used an impregnation approach to
synthesize an MCM-48 (70.0 wt-%) support loaded with bimetallic NiZr (30.0-wt%, Ni:Zr atomic ratio = 4:6, 5:5, and 6:4),
and the prepared catalysts were applied to the steam-reforming reactions of ethanol. These three bimetallic NiZr/MCM-48
catalysts exhibited significantly higher reforming reactivity than the mono-metal, Ni-loaded MCM-48 (Ni/MCM-48)
catalyst. The hydrogen production was started from 350 C over the three NiZr/MCM-48 catalysts, compared to above
550 C over the Ni/MCM-48 catalyst. The catalytic performance was affected by the Zr content. The H2 production
and ethanol conversion were maximized at 85% and 95%, respectively, over Ni4Zr6/MCM-48 at 750 C for 1 h at
CH3CH2OH:H2O = 1:1 and a gas hourly space velocity of 4000 h-1. This high performance was maintained for up to
60 h. Copyright © 2013 John Wiley & Sons, Ltd.
KEY WORDS
hydrogen production; ethanol steam-reforming reaction; NiZr/MCM-48

Correspondence
*Misook Kang, Department of Chemistry, College of Science, Yeungnam University, Gyeongsan, Gyeongbuk, 712–749, Republic of Korea.

E-mail: mskang@ynu.ac.kr

Received 10 September 2012; Revised 28 November 2012; Accepted 30 November 2012

1. INTRODUCTION higher than that of methanol or dimethyl ether, making


its decomposition difficult. This method offers the
The processes for converting petroleum fuels into hydrogen- advantage of producing more hydrogen per mole of
rich gas products include steam reforming [1], partial ethanol that is reformed because hydrogen can also be
oxidation [2], auto thermal reforming [3], dry reforming extracted from the steam. On the other hand, Ni-based
[4], and a combination of two or more [5]. Steam reform- Al2O3 support catalysts [10,11] have been used for ordi-
ing has the highest efficiency for hydrogen production. nary steam-reforming processes because of their accept-
Gaseous hydrocarbons can reduce the catalyst life and ably high activity and significantly lower cost than the
degrade the performance of the hydrogen production alternative precious metal-based catalysts [12]. However,
because of the catalytic poison that results from the CO Ni-based Al2O3 support catalysts are susceptible to deacti-
generation during the oxidation reaction. To solve these vation resulting from the deposition of carbon, even
problems, methanol [6], acetic acid [7], and dimethyl when operating at steam-to-carbon ratios predicted to be
ether [8], which including rich-oxygen, have been used thermodynamically outside of the carbon-forming regime.
as raw materials in the manufacture of hydrogen. These Additionally, an especially serious problem in NiAlO4
are completely oxidized to CO2 during the reforming catalysts is the abrupt catalytic deactivation that occurs at
reaction, and the reformation of ethanol in the presence high temperatures above 650 C, due to the formation of a
of steam has been a general practice for hydrogen NiAlO3 spinal structure resulting from the strong sintering
production. between Ni and Al, and this deactivation leads to reactor
Recently, ethanol has been used for steam reforming [9] shutdown and the reversal of the feed gases [10,11]. To
because of its non-toxicity and ease of transport and overcome these problems, many researchers have used the
storage. However, the oxidation temperature of ethanol is non-nickel-based and non-alumina supported catalysts for

1896 Copyright © 2013 John Wiley & Sons, Ltd.


Hydrogen-rich gas production from ethanol steam reforming J. S. Lee et al.

application to ethanol steam reforming [13,14]. However, at 550 C for 2 h in air. Finally, the following four
some problems remain in terms of the catalytic activity experimental catalysts were prepared for comparison:
and lifetime. Ni(30.0 wt-%)/MCM-48(70.0 wt-%), Ni4Zr6/MCM-48,
We herein have tried the introduction of the Zr Ni5Zr5/MCM-48, and Ni6Zr4/MCM-48 with various Ni
component with the capability for hydrocarbon cracking, and Zr contents.
hydrogen attraction from hydrocarbons, and depressing
the sintering between Ni and the support added between 2.2. Characterizations of the NiZr/MCM-48
them, as a catalytic active site in ethanol steam-reforming catalysts
reaction. Ni-loaded yittrium stabilized zirconia has been
widely used as an anode materials in solid oxide fuel The four prepared experimental catalysts were identified
cells. The material has a high conductivity at high temper- through powder XRD (MPD model, PANalytical, Yeungnam
ature and endurances for hydrogen reduction and CO University Instrumental Analysis Center, Korea) analysis
poison [15]; therefore, the hydrogen production will be with nickel-filtered Cu Ka radiation (30 kV, 30 mA) at 2-theta
improved by introduction of Zr component into an active angles of 10–100o. The scan speed was 10o/min, and the
site of the catalyst for ethanol steam reforming. Moreover, time constant was 1 s. The shapes and surface atomic
we use MCM-48 with high meso-porosities and consisting compositions of the particles of the four experimental
of a silica framework with no Al component as a support catalysts were determined by field emission scanning elec-
to eliminate the catalytic deactivation resulting from tron microscopy (S-4100, Hitachi, Yeungnam University
the strong sintering between Ni and Al. Therefore, in this Instrumental Analysis Center, Korea)/energy dispersive
study, we prepare the NiZr-loaded MCM-48 catalysts spectroscopy (EDS-EX-250, Horiba) operated at 120 kV.
for application into ethanol steam reforming. The ethanol The pore sizes and shapes of the MCM-48 particles were also
steam-reforming reactions are conducted in the tempera- measured by TEM (H-7600, Hitachi) operated at 120 kV.
ture range of 300–800 C at intervals of 50 C. The physico- The Brunauer, Emmett, and Teller (BET) surface area was
chemical properties of the catalysts were determined by measured using a Micrometrics ASAP 2000 instrument.
X-ray diffraction (XRD), transmission electron microscopy Before BET surface measurement, the three bimetallic
(TEM), H2-temperature-programmed reduction (TPR), and Ni-Zr-loaded catalysts were degassed under vacuum at
X-ray photon spectroscopy (XPS) analyses. 120 C for 12 h, and then thermally treated at 300 C for 30
min. The BET surface areas of the catalysts were measured
by nitrogen gas adsorption using a continuous flow method
2. EXPERIMENTAL SECTION with a mixture of nitrogen and helium as the carrier gas.
XPS measurements of Ni2p, Zr3d, Si2p, C1s, and O1s were
2.1. Preparation of MCM-48 support and recorded with a model ESCALAB250 XPS system (Thermo
NiZr/MCM-48 catalysts Fisher Scientific (U.K), Busan Center, Korea Basic Science
Institute, Korea), equipped with a non-monochromatic Al
For the preparation of MCM-48 (70.0 wt-%) catalysts Ka (1486.6 eV) X-ray source. The powders were pelletized
with various loadings of NiZr (30.0 wt-%, Ni:Zr atomic at 1.2  104 kPa for 1 min, and the 1.0-mm pellets were
ratios =4:6, 5:5, and 6:4), MCM-48 mesoporous material then maintained overnight in a vacuum (1.0  10–7 Pa) to
was hydrothermally synthesized from the assembly of cetyl- remove any water molecules from the surface prior to the
trimethylammonium bromide (CTAB) and Ludox HS40 measurement. The base pressure of the system was
with NaOH. The general preparation method of MCM-48 below 1  10–9 Pa. The experiments were conducted with a
is reported elsewhere [16]. Briefly, solution A is a mixture 15kV and 150-W source power and an angular acceptance
containing Ludox HS40 (SiO2 sol 40.0 wt-%, DuPont Co., of 5º. The analyzer axis made an angle of 90º with the
USA) as a Si precursor, NaOH, and distilled H2O. The mix- specimen surface. A Shirley function was used to subtract
ture was stirred for 2 h at 80 C. Solution B consisted of a the background in the XPS data analysis. The Ni2p, Zr2p,
mixture of CTAB (Junsei Co., Japan), distilled H2O, and Si2p, C1s, and O1s signals were fitted using mixed
ethanol. Solution B was added to solution A slowly. After Lorentzian– Gaussian curves. H2-TPR was conducted as
continuous stirring for 2 h at room temperature, the final follows. About 0.3 g of the four experimental catalysts was
gel solution was transferred to an autoclave and heated at pre-treated under helium flow (30 ml/min) at 700 C for 2 h
140 C for 16 h. The resulting as-synthesized solid product and then cooled to room temperature. The analysis was
was obtained by filtration, dried in air, and calcined at carried out by flowing 30 ml/min of H2 (10 vol. %)/N2
550 C for 3 h to obtain the porosity. The three NiZr/ and raising the temperature of the catalyst from room temper-
MCM-48 catalysts were obtained using an impregnation ature to 700 C at 5  C/min. The change in the hydrogen
method. Nickel and zirconium nitrate (Ni(NO3)2 and Zr concentration was measured with a gas chromatograph (GC
(NO3)4, 99.9%, Junsei Co., Japan), as precursors of the Ni series 580, GOW-MAC) equipped with a thermal conductiv-
and Zr components, were added at an atomic ratio and ity detector (TCD). In order to study the formation of carbon
were simultaneously dropped into 25.0 ml of ethanol species on the catalyst surface, temperature-programmed
including MCM-48 powder. The slurry was stirred oxidation (TPO) was performed using the TGA N-1000
for 2 h, evaporated at 50 C for 3 h, and thermally treated instrument by introducing high purity oxygen gas into the

Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd. 1897
DOI: 10.1002/er
J. S. Lee et al. Hydrogen-rich gas production from ethanol steam reforming

system after purging it with N2. A 7.0 mg sample was placed 10 and 30 ml/min for ethanol and steam, respectively. Argon
in a sample pan and heated from 50 to 900 C at a constant gas was used to transport the vaporized mixture to the reac-
heating rate of 10 C/min. High purity O2 gas was flowed at tor. The reaction products were measured using on-line GC
20 ~ 40 ml/min into the TPO system to combust the carbon on a Donam DS6200 (Donam company, Korea) that was
accumulated on the catalyst after the reaction. The profiles equipped with a TCD and a flame ionizing detector (FID).
were obtained in the same manner as that described for The TCD was used to detect H2, CH3CHO, C2H5OH, CO,
TPR, and the coke contents were calculated from the weight and CO2, and the FID to detect CH4, C2–C5 hydrocarbons,
loss in the temperature range from 50 to 900 C. and other products. The ethanol conversion (XEtOH) and
the selectivity of the C-containing products (SC) of the vari-
2.3. Ethanol steam reforming over NiZr/MCM- ous samples were calculated using the following equations:
48 catalysts
XEtOH ¼ ðmol EtOHin  mol EtOHout Þ=mol EtOHin x 100%
Figure 1 shows the reactor for the ethanol steam reforming, SCH2 ¼ mol H2 =½ðmol EtOHin  mol EtOHout Þ  ðmol H2 Oin
and the catalytic activities of the four experimental catalysts
mol H2 Oout Þ x 100%
were measured in the temperature range for 300–800 C for
SCothers ¼ mol othersout =ðmol EtOHin  mol EtOHout Þ x 100%
1 h reaction time intervals at steam-to-ethanol ratios of 1:1,
1:2, and 1:3 with gas hourly space velocities (GHSV) of In this study, GHSV was calculated based on the total flow
3000, 4000, and 6000 h-1. The three bimetallic NiZr-loaded rate of the feed mixture in the gas phase.
catalysts (0.4 g) were pelletized to 20–24 mesh and then
packed with a small amount of quartz wool to prevent their
movement in the fixed bed quartz reactor, which was verti- 3. RESULTS AND DISCUSSION
cally mounted inside the furnace. An ethanol/water solution
(mol/mol) was then introduced into the vaporizer. Typically, 3.1. Characteristics of the four
according to most articles [16,17], the ethanol and water experimental catalysts
were mixed at a desired weight percentage and vaporized
immediately at a boiling point temperature before the The four experimental catalysts were characterized by XRD
mixture was sent to the reactor. However, in this study, the before the reaction. The XRD patterns are shown in Figure 2.
amount of steam was adjusted by regulating the temperature, The diffraction lines of the NiO phase at 2-theta angles of
according to the following partial pressure law, as described around 37.27o, 43.29o, 62.89o, 75.44o, and 79.43o could be
in detail in a previous paper [17]. Here, the vaporization seen in the XRD patterns of all the catalysts, corresponding
temperatures were affected by the ethanol and steam to the (d111), (d200), (d220), (d311), and (d222) spaces, respec-
concentrations. The flow rate was held constant at a rate of tively. They were ascribed to the cubic structure and assigned

Figure 1. Apparatus of the catalytic reactor used for ethanol steam reforming.

1898 Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd.
DOI: 10.1002/er
Hydrogen-rich gas production from ethanol steam reforming J. S. Lee et al.

crystalline domain size. The full width at half maximum


(FWHM) height of the peaks at 2 theta = 43.29o (d200) and
30.27o (d101) of NiO and ZrO2, respectively, were measured.
The Scherrer equation, t = 0.9l/bcosθ, where l_ is the wave-
length of the incident X-rays, b_ the FWHM in radians, and θ_
the diffraction angle, was used to determine the crystalline
domain size. The calculated crystalline domain sizes based
on a special peak of 43.29o (d200) of NiO were 23.24,
18.08, 21.67, and 29.59 for Ni/MCM-48, Ni4Zr6/MCM-48,
Ni5Zr5/MCM-48, and Ni6Zr4/MCM-48, respectively. In the
same way, the crystalline domain sizes at 30.27o (d101) over
the NiZr/MCM-48 samples were 4.68, 3.13, and 7.84 nm,
referred to as b), c), and d), respectively. The smaller grain
sizes indicated that the Ni and Zr ingredients were well
dispersed in the MCM-48 support.
In Table I, the EDS analysis (atomic composition)
revealed variation in loaded metal concentrations, with
the Ni and Zr surface concentrations being relatively
Figure 2. XRD patterns of the four experimental catalysts
before ethanol steam reforming.
higher and lower, respectively. The atomic ratios of each
metal component in Ni:Zr were 4.42:6.21, 6.71:5.70, and
7.63:5.31 in Ni4Zr6/MCM-48, Ni5Zr5/MCM-48, and
to JCPDS card No. 78–0643. The special peaks decreased Ni6Zr4/MCM-48, respectively. However, the amounts of
with increasing concentration of loaded Zr. The diffraction Ni and Zr presented on the surface were not greatly
lines of the ZrO2 phase at 2-theta angles of around 30.27o affected by the amount inserted during synthesis, which
(d101), 35.24o (d110), 50.37o (d112), and 60.19o (d211) also indicated high dispersion.
could be also exhibited in the XRD patterns of the three The pore size and shapes of the four experimental
bimetallic NiZr-loaded catalysts, which have indicated the catalysts were characterized using the TEM images shown
tetragonal structure assigned to JCPDS card No. 88–1007. in Figure 3. The observed mesopores with regular hexago-
The special peaks were broad with increasing concentration nal form were not clear because the NiZr/MCM-48 cata-
of loaded Ni. However, the peak corresponding to NiZr alloy lysts were covered by Ni and Zr components compared
was not observed in this study. Conversely, the line widths of to the unique mesoporous MCM-48. The black spots of
the peaks were broad, which generally indicates a smaller cubic and spherical shapes, which were considered to be

Table I. Atomic compositions calculated from the EDS analysis for the four experimental catalysts.

Catalysts/elements O Ni Zr Si

Ni/MCM48 76.04 6.77 - 15.80


Ni4Zr6/McM48 61.69 4.42 6.21 27.68
Ni5Zr5/McM48 61.84 6.71 5.70 25.69
Ni6Zr4/McM48 65.60 7.63 5.31 21.45

Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd. 1899
DOI: 10.1002/er
J. S. Lee et al. Hydrogen-rich gas production from ethanol steam reforming

Figure 3. TEM images of the four experimental catalysts before ethanol steam reforming.

aggregated NiO particles, were observed on the external slope was observed at intermediate and high relative pres-
surface of the mesopores in MCM- 48, and the size of these sures in all four catalysts, which is indicative of the pres-
particles increased from 50 to 70 nm with increasing Ni ence of large mesopores.
mole fraction. In Table II, the average pore diameter value, Dp, ranged
The pore size distribution is an important characteristic from 3.47 to 6.85 nm, and the pore size decreased with
for porous materials. Among these methods, Barrett-Joyner- increasing Zr content. The adsorption and desorption lines
Halenda plots are a suitable method for the range of for the four experimental catalysts overlapped completely
mesopores [18]. The relative pressure at which pore filling in the low relative pressure range, while the hysteresis loop
takes place by capillary condensation can be calculated from was in the high relative pressure region (P/P0 0.49 ~ 0.85),
Kelvin’s equation. By using Kelvin’s equation, the pore mainly due to the presence of bottle types of pore. These
radius in which the capillary condensation occurs actively bottle pore types had a larger pore size in the bottle body,
can be determined as a function of the relative pressure which induced hysteresis in the high relative pressure
(P/P0). The mean pore diameter, Dp, was calculated from region. The BET surface areas in the three NiZr/MCM-
Dp = 4VT/S, where VT is the total volume of pores, and S 48 catalysts were not greatly decreased compared to the
the BET surface area. Figure 4 shows the adsorption– value of 560 m2/g for pure MCM-48, and their surface areas
desorption isotherms of N2 at 77 K for the four mesopor- were in the range of 403–463 m2/g; however, it was dramat-
ous catalysts. They illustrate the shape and behavior of ically decreased in Ni/MCM-48 to 80.97 m2/g, which attrib-
the N2 adsorption isotherms for the mesoporous materials. uted to the sintering of Ni.
All the isotherms belonged to IV type in the IUPAC The H2–TPR profiles of the four experimental catalysts
classification [19] in all samples. The isotherms were are shown in Figure 5. The changes corresponding to the
wide, without any clear plateau, and a certain hysteresis reduction of the NiO/Ni [20] components in the four

1900 Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd.
DOI: 10.1002/er
Hydrogen-rich gas production from ethanol steam reforming J. S. Lee et al.

Figure 4. Adsorption-desorption isotherm curves of the four experimental catalysts.

Table II. BET surface areas and pore diameters determined by adsorption-desorption isotherm curves for the four experimental catalysts.

Catalysts BET specific surfacearea(m2/g) Total pore volume(p/p0 = 0.009) Pore diameter(nm)

Ni/MCM48 80.97 0.14 6.85


Ni4Zr6/McM48 403.63 0.37 3.62
Ni5Zr5/McM48 463.76 0.40 3.47
Ni6Zr4/McM48 449.78 0.40 3.56

experimental catalysts were observed in the H2–TPR


profiles. In general, the H2–TPR results indicated that
the peak area corresponds to the hydrogen uptake and the
peak at high temperatures corresponds to the catalytic
reaction involved in the reduction mechanism. A reduction
type for the Ni component was seen in all four catalysts at
300–400 C, which was considered to be NiO/Ni. The
reduction peak of the Ni oxides was gradually shifted to
lower temperatures with increasing Zr content, and the great-
est reduction was observed in the cases of Ni4Zr6/MCM-48
and Ni6Zr4/MCM-48. Otherwise, the reduced peak of ZrO2
to ZrO was not shown in this study.

3.2. Ethanol steam-reforming reaction over


the four experimental catalysts

Ethanol steam reforming was carried out with 0.5 g of


Figure 5. H2-TPR curves of the four experimental catalysts. each of the four experimental catalysts under the reaction

Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd. 1901
DOI: 10.1002/er
J. S. Lee et al. Hydrogen-rich gas production from ethanol steam reforming

conditions of temperature = 300–800 C, GHSV = 4,000 h1, the introduction of Zr between the Ni and the MCM-48
and H2O/EtOH = 1.0. Figure 6 compares the time-on-stream support on the catalytic performance and the decreased
activity of the four experimental catalysts. In (b), in which sintering phenomenon and the consequently reduced
only the ethanol conversion is compared, the conversion catalytic deactivation.
over the Ni/MCM-48 catalyst was higher than those of Figure 7a and b shows the catalytic performances for
the three NiZr/MCM-48 catalysts, which was attributed to the ethanol steam-reforming reaction over the Ni4Zr6/
the methanation of carbon dioxide and carbon monoxide MCM- 48 catalyst at various GHSV values and with
on the Ni, as previously reported [10,11]. Particularly, the various ethanol/steam concentrations, diluted in argon gas
ethanol conversions over Ni/MCM-48 and Ni4Zr6/MCM- of 30 vol-percent. The H2 production was positively corre-
48 catalysts were maximized at 98% at 800 C. On the lated with GHSV as the reaction temperature exceeded
other hand, the hydrogen production gradually increased 350 C, with an ethanol conversion exceeding 85% over
over the three NiZr/MCM-48 catalysts as the temperature the entire GHSV range, as shown in figure a. The efficien-
was increased from 350 to 750 C, to a maximum of 85% cies of the ethanol conversion and hydrogen production
at 750 C over Ni4Zr6/MCM-48. Otherwise, the starting point were both reduced with decreasing GHSV below 3000
for hydrogen evolution was shifted to the higher temperature and increasing GHSV above 6000 over the complete
of 500 C over the Ni/MCM-48 catalyst in spite of the higher temperature range. Therefore, the optimal reaction condi-
ethanol conversion. The Ni4Zr6/MCM-48 catalyst provided a tions according to the active response for the production
significantly higher reforming reactivity than the conven- of H2 over the Ni4Zr6/MCM-48 catalyst were a GHSV of
tional Ni-based MCM-48 catalyst. This suggested that the 4000 h1, an ethanol steam concentration ratio of 1:1
Zr particles prevented sintering between the Ni particles (30 vol-percent), and a reaction temperature of 750 C. In
and was further attributed to a synergy effect between Ni figure b, the optimal conditions corresponded to an EtOH:
and Zr. Here, five species were present in the product H2O ratio of 1:1, and 85% of the hydrogen was emitted at
distribution: H2, CO2, CO, CH4, and feed CH3CH2OH. 750 C, with an ethanol conversion of 95%. However, the
Possibly, the conversion of ethanol to hydrocarbons like ethanol conversion was dramatically decreased at higher
C3–C5 prevailed over the hydrogen synthesis at the lower water concentrations at an EtOH: H2O ratio of 1:3, in spite
temperature of 200–350 C, and the reaction products were of the high ethanol conversion of 99%. FID-GC investigation
transferred to CH4, H2, and CO2 at the medium temperature revealed that methane and other light hydrocarbons (~C4)
of 400–550 C. Finally, the synthesis of hydrogen from were produced with high water molecules, possibly because
ethanol steam reforming prevailed over the hydrocarbon the H ions in H2O act as Brønsted acidic sites.
synthesis at the higher temperature of 600–800 C, and the re-
action products were distributed in the ratio of 40–85% H2: 3.3. Characteristics of catalysts after
below 10–15% CH4 and CO2. Most importantly, the CO ethanol steam reforming
selectivity approached zero percent over the whole tempera-
ture range, and the CH4 production was slightly increased Figure 8 compares the XRD patterns of the four experimen-
compared to that over the catalysts loaded only with Ni. tal catalysts after 10 h of reaction at 750 C. The diffraction
Therefore, our results demonstrated the synergistic effect of lines of the Ni metal phase shown at 2-theta angles of

Figure 6. (a) H2 production and (b) ethanol conversion over the four experimental catalysts according to the reaction temperature.

1902 Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd.
DOI: 10.1002/er
Hydrogen-rich gas production from ethanol steam reforming J. S. Lee et al.

Figure 7. Ethanol conversion and H2 production according to the (a) GHSV and (b) ethanol/steam ratio over Ni4Zr6/MCM-48.

Figure 8. Comparison of XRD patterns for the four experimental


catalysts before and after ethanol steam-reforming reaction. Figure 9. Comparison of XPS curves for Si2p, Zr3d, Ni2p, and
O1s orbital of two catalysts, (a) Ni/MCM-48 and (b) Ni4Zr6/
MCM-48, before and after ethanol reforming reaction.
44.49o (d111) and 51.85o (d420) [JCPDS#87-0712] were
clearly different after the reaction in all four catalysts,
and the peaks assigned to Ni oxides were smaller than catalyst. Additionally, the peaks of the Zr metal phases
before reaction. This indicated that the Ni components appeared after the reaction at 2-theta angles of 25.44o
acted as active sites in the ethanol steam-reforming reaction. (d110) and 35.78o (d110) [JCPDS #34-0657]. This
The ZrO2 peak decreased after the reaction, which automat- result confirmed that the Zr component also acted as an
ically implied a comparison with before the reaction, and the active site in the ethanol steam-reforming reaction, as
trend was particularly pronounced in the Ni4Zr6/MCM-48 with Ni.

Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd. 1903
DOI: 10.1002/er
J. S. Lee et al. Hydrogen-rich gas production from ethanol steam reforming

Figure 9 compares the typical survey and high-resolution both samples before and after reaction. However, the peak
spectra obtained from the quantitative XPS analyses of the intensities were largely decreased with no change of peak
two catalysts, Ni/MCM-48 and Ni4Zr6/MCM-48, before position after reaction, particularly in Ni4Zr6/MCM-48. This
and after ethanol steam reforming. The survey spectra of indicated that the Ni components acted as a strong catalyst
the particles contained Zr3d, Ni2p, Si2p, and O1s peaks, which for the ethanol steam reforming and then deactivated. These
were analyzed based on an XPS handbook [21]. The Si2p3/2 results indicated that the Zr oxidation state was changed after
spin–orbital photoelectron, which was assigned to the Si the ethanol reforming and, consequently, confirmed that the
component in the MCM-48 support, was located at a binding Zr ions were involved in the reduction of ethanol or other
energy of 103.3 eV. The peak in Ni/MCM-48 was not hydrocarbons such as Ni ions that were evolved during the
changed after the reaction compared to that before reaction; ethanol-reforming reaction to afford the production of CH4
however, the binding energy curve for the Si2p3/2 exhibited from CO or CO2 molecules.
a lower energy of 102.0 eV in Ni4Zr6/MCM-48, and the peak To determine the amount of carbon deposited on the
was shifted to a higher binding energy after reaction. In catalysts, we carried out TPO measurement (a) and C1s
general, a large binding energy indicates a more oxidized XPS analysis (b), as shown in Figure 10. The deposited
state. The same phenomenon occurred for O1s (at 531.5 eV amount (peak area) and species of carbons (decomposed
(hydroxides) for Ni/MCM-48), which was distributed at temperature) are closely related to the catalytic deactivation.
529.0 eV for the Ni4Zr6/MCM-48 assigned to metal oxides, Generally, the extent of catalytic deactivation is reduced
and it also shifted to a higher energy after reaction. These when a smaller amount of carbon is deposited. When the
results indicated that the Si and O ions were oxidized to Ni-only component was exposed to the outside surface of
higher oxidation states after the ethanol reforming and the Ni/MCM-48 sample, the deposited carbons underwent
confirmed the involvement of the Si and O ions in the reduc- greater oxidation at high temperatures of 530 C, indicat-
tion of ethanol or other hydrocarbons that were evolved ing the deposition of heavy carbons. Similar amounts
during the ethanol-reforming reaction to afford the produc- of carbon were deposited in the Ni6Zr4/MCM-48 and
tion of H2 or hydrocarbons in Ni4Zr6/MCM-48. On the other Ni5Zr5/MCM-48 catalysts, but the peak areas were
hand, the Zr3d region in Ni4Zr6/MCM-48 was decomposed smaller in the latter, whereas the deposited carbon amount
into two contributions, which were assigned to Zr3d5/2 and was the smallest in Ni4Zr6/MCM-48. We therefore
Zr3d3/2 at 181.5 and 183.0 eV in ZrO2, respectively [21]. concluded that the Zr addition helped to retain the stability
The oxidation state was divided into three types after of the Ni crystallites and avoided their conglomeration in
reaction, and the values were smaller after the reaction. the ethanol steam-reforming reaction, which corresponds
Generally, the same phenomenon occurred for Ni2p3/2 and to the better stability demonstrated by the three NiZr/
Ni2p1/2 at binding energies of 853.8 and 871.29 eV, which MCM-48 catalysts compared with Ni/MCM-48. On
were assigned to NiO [21]. In this study, Ni2p3/2 and Ni2p1/2 the other hand, although one C1s orbital is generally seen
were also expressed at 853.5 and 871.0 eV, respectively, in at 284.5 eV for bulky carbons [21], the peak presented

Figure 10. (a) TPO profiles and (b) XPS curve of C1s of the four experimental catalysts after ethanol reforming reaction.

1904 Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd.
DOI: 10.1002/er
Hydrogen-rich gas production from ethanol steam reforming J. S. Lee et al.

in this study could be split into three peaks at around oxides, in the absence of any kind of metal or other catalyst
281.0, 283.2-284.8, and 288.2 eV, due to the mixture of capable of functioning. Finally, the carbons react with H2
carbide, carbon, and carboxyl species in the coke deposi- and are transferred into C = C molecules, which can be a
tion over the catalyst. Compared to over Ni/MCM-48, precursor of carbon nanofilaments or tubes [22], as we may
the carbon and carboxyl species were strongly seen in infer from some CH4 evolution during the ethanol cracking.
Ni4Zr6/MCM-48. Interestingly, many carbon lumps were generated in the
In an interesting result shown in Figure 11, a carbon lump catalysts with high Ni content; otherwise, the carbon
was found after ethanol steam reforming over the four exper- nanotubes increased with increasing Zr content, particularly
imental catalysts and was observed by TEM. At a certain in Ni4Zr6/MCM-48.
time after the beginning of the reaction, e.g. 10 h at 750 C, Finally, the catalytic deactivation was tested for the
the carbons were deposited on the surface of the catalyst. In Ni4Zr6/MCM-48 catalyst, and the results are shown in
general, CO2 and ethanol could be converted into carbon Figure 12. In a dramatic result, the H2 production exceeded
nanotubes (or filaments) on the carbon deposited on the 76–64% mol with above 95% ethanol conversion depend-
catalyst surface. As we have suggested above, a molecule ing on the presence of Zr. The catalyst lifetime was greatly
of ethanol was dissolved into two molecules of CO and H2 improved in the Ni4Zr6/MCM-48 catalyst compared to that
as intermediates, and subsequently converted at higher of the catalyst without Zr. The catalytic deactivation was
temperatures to pure carbon. The hydrogen and oxygen retarded until 60 h over the former, in contrast to the
may have disappeared due to reaction with the main metal rapid deactivation over the latter, indicating that the initial

Figure 11. TEM images of deposited carbon over the four experimental catalysts after ethanol steam reforming.

Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd. 1905
DOI: 10.1002/er
J. S. Lee et al. Hydrogen-rich gas production from ethanol steam reforming

Figure 12. Test of catalytic deactivation over the Ni4Zr6/MCM-48 catalyst.

catalyst deactivation may have resulted from a combina- REFERENCES


tion of steam-induced nickel sintering and carbon deposi-
tion. However, the apparent deactivation rate was far 1. Rossia CCRS, Alonso CG, Antunes OAC, Guirardello R,
lower for the Zr-promoted catalyst, and its higher ethanol Cardozo-Filho L. Thermodynamic analysis of steam
conversion was maintained for up to 60 h. After 60 h, the
reforming of ethanol and glycerine for hydrogen produc-
used catalyst was very black and was very dense. Thus,
tion. International Journal of Hydrogen Energy 2009;
the improved stability at the slower deactivation rate
achieved with the NiZr sample could not be unequivocally 34:323–332.
attributed to carbon formation, as shown in Figure 11. 2. Kim SC, Chun YN. Experimental study on partial
oxidation of methane to produce hydrogen using
low-temperature plasma in AC Glidarc discharge.
4. CONCLUSIONS International Journal of Energy Research 2008;
32:1185–1193.
The two key study findings are the effects of the Zr addi- 3. Youn MH, Seo JG, Jung JC, Park SU, Song IK.
tive in the Ni-loaded MCM-48 catalyst on the variation Hydrogen production by auto-thermal reforming of
in H2 production and the retardation of the catalytic deacti- ethanol over nickel catalyst supported on mesoporous
vation. On the basis of the performance results and a vari-
yttria-stabilized zirconia. International Journal of
ety of physical measurements, we proposed that both the
Hydrogen Energy 2009; 24:5390–5397.
Ni and Zr components played a role in the oxidation of
the feed gases during ethanol reforming. The simultaneous 4. Yasyerli S, Filizgok S, Arbag H, Yasyerli N, Dogu G.
addition of Ni and Zr may have depressed the sintering Ru incorporated Ni–MCM-41 mesoporous catalysts
between the Ni particles, and retarded the catalytic deacti- for dry reforming of methane: Effects of Mg addition,
vation. Consequently, their synergistic effect increased the feed composition and temperature. International
ethanol conversion and H2 production over the Ni4Zr6/ Journal of Hydrogen Energy 2011; 36:4863–4874.
MCM-48 catalysts to 95% and 85%, respectively, over 5. Sania M de L, Ivna O da C, Jacobs G, Burtron HD,
a period up to 60 h. The optimum operation conditions, Lisiane VM, Fabio BN. Steam reforming, partial
as identified by a parametric study, were a reaction temper- oxidation, and oxidative steam reforming of ethanol
ature of 750 C, GHSV of 4000 h1, and feed ratio of over Pt/CeZrO2 catalyst. Journal of Catalysis 2008;
ethanol: H2O of 1:1.
257:356–368.
6. Chein RY, Chen YC, Lin YS, Chung JN. Hydrogen
production using integrated ethanol-steam eforming
ACKNOWLEDGEMENT reactor with various reformer designs or PEM fuel
cells. International Journal of Energy Research
This work was supported by The New and Renewable En-
ergy of the Korea Institute of Energy Technology Evalua- 2012; 36:466–476.
tion and Planning (KETEP) grant funded by the Korea 7. Thaicharoensutcharittham S, Meeyoo V, Kitiyanan B,
Ministry of Knowledge Economy (No. 2010T100100622). Rangsunvigit P, Rirksomboon T. Hydrogen production

1906 Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd.
DOI: 10.1002/er
Hydrogen-rich gas production from ethanol steam reforming J. S. Lee et al.

by steam reforming of acetic acid over Ni-based Ba1xCe0.4Zr0.4Y0.2O3-d. International Journal of


catalysts. Catalysis Today 2011; 164:257–261. Hydrogen Energy 2011; 36:8450–8460.
8. Wang X, Pan X, Lin R, Kou Y, Zou W, Ma J-X. Steam 16. Wang L, Shao Y, Zhang J, Anpo M. Synthesis of
reforming of dimethyl ether over Cu–Ni/g-Al2O3 bi- MCM-48 mesoporous molecular sieve with thermal
functional catalyst prepared by deposition–precipitation and hydrothermal stability with the aid of promoter
method. International Journal of Hydrogen Energy anions. Microporous and Mesoporous Materials
2010; 35:4060–4068. 2006; 95:17–25.
9. Wang W, Wang Y, Liu Y. Production of hydrogen by 17. Kwak BS, Lee JS, Lee JS, Choi B-H, Ji MJ,
ethanol steam reforming over nickel–metal oxide cata- Kang M. Hydrogen-rich gas production from etha-
lysts prepared via urea–nitrate combustion method. Inter- nol steam reforming over Ni/Ga/Mg/Zeolite Y cata-
national Journal of Energy Research 2011; 35:501–506. lysts at mild temperature. Applied Energy 2011;
10. Alberton Andre L, Souza Mariana MVM, Schmal M. 88:4366–4375.
Carbon formation and its influence on ethanol steam 18. Zhao Q, Zhou X, Ji M, Wu D, Jiang T, Yin H.
reforming over Ni/Al2O3. Catalysis Today 2007; Synthesis of ordered cubic mesoporous ZrMCM-48
123:257–264. molecular sieve with aid of fluoride ions. Colloid
11. Zhang L, Liu J, Li W, Guo C, Zhang J. Ethanol steam Surface A 2011; 384:513–518.
reforming over Ni-Cu/Al2O3-MyOz (M = Si, La, Mg, 19. Jiang T, Wu D, Song J, Zhou X, Zhao Q, Ji M, Yin H.
and Zn) catalysts. Journal of Natural Gas Chemistry Synthesis and characterization of mesoporous
2009; 18:55–65. ZrMCM-48 molecular sieves with good thermal and
12. Santucci A, Annesini MC, Borgognoni F, Marrelli L, hydrothermal stability. Power Technology 2011;
Rega M, Tosti S. Oxidative steam reforming of ethanol 207:422–427.
over a Pt/Al2O3 catalysts in a Pd-based membrane 20. Pirouzmand M, Amini M.M, Safari N. Immobilization
reactor. International Journal of Hydrogen Energy of iron tetrasulfophthalocyanine on functionalized
2011; 36:1503–1511. MCM-48 and MCM-41 mesoporous silicas: Catalysts
13. Eswaramoorthi I, Dalai AK. A comparative study on for oxidation of styrene. Journal of Colloid Interface
the performance of mesoporous SBA-15 supported Science 2008; 319:199–205.
Pd-Zn catalysts in partial oxidation and steam reforming 21. Gies H, Grabowski S, Bandyopadhyay M, Grünert W,
of methanol for hydrogen production. International Tkachenko OP, Klementiev KV, Birkner A. Synthesis
Journal of Hydrogen Energy 2009; 34:2580–2590. and characterization of silica MCM-48 as carrier of
14. Wang F, Cai W, Provendier H, Schuurman Y, size-confined nanocrystalline metal oxides particles
Descorme C, Mirodatos C, Shen W. Hydrogen produc- inside the pore system. Microporous and Mesoporous
tion from ethanol steam reforming over Ir/CeO2 catalysts: Materials 2003; 60:31–42.
Enhanced stability by PrOx promotion. International 22. Jankhah S, Abatzoglou N, Gitzhofer F, Blanchard J,
Journal of Hydrogen Energy 2011; 36:13566–13574. Oudghiri-Hassani H. Catalytic properties of carbon
15. Guo Y, Ran R, Shao Z, Liu S. Effect of Ba nano-filaments produced by iron-catalysed reforming
nonstoichiometry on the phase structure, sintering, of ethanol. Chemical Engineering Journal 2008;
electrical conductivity and phase stability of 139:532–539.

Int. J. Energy Res. 2013; 37:1896–1907 © 2013 John Wiley & Sons, Ltd. 1907
DOI: 10.1002/er

You might also like