You are on page 1of 8

ARTICLE IN PRESS

Chemical Engineering Science ( ) –


www.elsevier.com/locate/ces

Selective production of hydrogen via oxidative steam reforming


of methanol using Cu–Zn–Ce–Al oxide catalysts
Sanjay Patel a,b , K.K. Pant a,∗
a Department of Chemical Engineering, Indian Institute of Technology, Hauz Khas, New Delhi 110016, India
b Department of Chemical Engineering, Institute of Technology, Nirma University of Science and Technology, Ahmedabad 382481, India

Received 15 June 2006; received in revised form 19 January 2007; accepted 29 January 2007

Abstract
The oxidative steam reforming of methanol (OSRM) was carried out to produce the hydrogen selectively for polymer electrolyte membrane
(PEM) fuel cell applications over Cu–Zn–Ce–Al oxide and Cu–Zn–Al oxide catalysts of varying compositions prepared by co-precipitation
method. Catalyst performance was evaluated in a packed bed reactor over a wide range of operating conditions, and reaction parameters were
optimized in order to maximize the hydrogen production with minimum carbon monoxide formation. The incorporation of ceria in Cu–Zn–Al
oxide catalysts enhanced the activity greatly compared to without it. The Cu/Zn/Ce/Al:30/20/10/40 exhibited 100% methanol conversion and
244 mmol s−1 kg−1cat hydrogen rate at 553 K with carbon monoxide as low as 995 ppm, which reduces the load on preferential oxidation of CO
to CO2 significantly before feeding the hydrogen rich stream to the PEM fuel cell as a feed. Ceria had improved the dispersion and specific
surface area of copper in multi-component Cu–Zn–Ce–Al oxide catalysts which were confirmed by the physicochemical properties, X-ray
diffraction (XRD), temperature programmed reduction (TPR) and CO chemisorption studies. The chemisorption studies were performed at
193 K in order to hinder the spillover of carbon monoxide to ceria. The time-on-stream stability test had shown Cu–Zn–Ce–Al oxide catalysts
as more stable compared to Cu–Zn–Al oxide catalysts. The amount of carbon deposited onto the catalysts was determined using TG/DTA
thermogravimetric analyzer and the type of carbon species were identified using C1s X-ray photoelectron spectroscopy (XPS) spectra.
䉷 2007 Elsevier Ltd. All rights reserved.

Keywords: Hydrogen; OSRM; Ceria; PEM fuel cell

1. Introduction dimethyl ether (Semelsberger et al., 2006), etc. Methanol offers


several advantages for the hydrogen production compared to
Hydrogen is expected to play a major role in the future as other liquid organics (Velu et al., 2001; Patel and Pant, 2006a).
a carbon free energy carrier. Its use in the vehicles via poly- There are different routs available for the hydrogen production
mer electrolyte membrane (PEM) fuel cells can offer the non- from methanol as follows. One is partial oxidation of methanol
toxic tail-pipe emissions and high over all efficiency compared (POM),
to conventional internal combustion engines (Edwards et al.,
1998; Bowers et al., 2006). The on-board storage of hydro- CH3 OH + 0.5O2 ←→ 2H2 + CO2 , H 0 = −192 kJ mol−1 .
gen with high energy density is facing some technical prob- (1)
lems (Breen and Ross, 1999; Raimondi et al., 2002). One This is a highly exothermic reaction which leads to the problem
promising solution to this can be an on-board production of of heat removal and reactor temperature control, and also pro-
hydrogen using liquid hydrocarbons like methanol (Patel and duces significant amount of CO (Wang et al., 2003). Another
Pant, 2006c; Yan et al., 2006), ethanol (Sahoo et al., 2007), route is the steam reforming of methanol (SRM),

∗ Corresponding author. Tel.: +91 11 26596172; fax: +91 11 26581120. CH3 OH + H2 O ←→ 3H2 + CO2 , H 0 = 49.5 kJ mol−1 .
E-mail address: kkpant@chemical.iitd.ac.in (K.K. Pant). (2)
0009-2509/$ - see front matter 䉷 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2007.01.066

Please cite this article as: Patel, S., Pant, K.K., Selective production of hydrogen via oxidative steam reforming of methanol using Cu–Zn–Ce–Al oxide
catalysts. Chemical Engineering Science (2007), doi: 10.1016/j.ces.2007.01.066
ARTICLE IN PRESS
2 S. Patel, K.K. Pant / Chemical Engineering Science ( ) –

SRM can produce H2 /CO2 in the molar ratio of 3/1. It pro- which was crushed to a fine powder and subsequently the
duces relatively small amount of carbon monoxide, usually less catalyst was produced by calcination in the presence of air at
than 1%, as a by-product over copper-based catalysts (Choi and 673 K for 4 h. The pellets of 3 mm size from fine catalyst pow-
Stenger, 2002; Lindstrom et al., 2003; Patel and Pant, 2006b). der were made in an automatic palletizing press Techno Search
But it is highly endothermic and has a slow rate of reaction. AP-15 and subsequently crushed and sieved to a size of 20/25
The limitations of partial oxidation and steam reforming of mesh. The same procedure was followed for the preparation
methanol for the hydrogen production for PEM fuel cell appli- of Cu–Zn–Al oxide catalysts. The final elemental composition
cations could be overcome by combining them as an oxidative was determined using atomic absorption spectroscopy (AAS)
steam reforming of methanol (OSRM) that can be performed Varian AA420FS. Surface area and pore volume of catalysts
auto-thermally with idealized reaction stoichiometry. were measured using ASAP 2010 Micromeritics micro-pore
surface area analyzer. The copper surface area, dispersion and
CH3 OH + (1 − 2a)H2 O + aO2 ←→ (3 − 2a)H2 + CO2 , particle size were measured by CO chemisorption at 193 K
H 0 = (49.5 − 241.8 ∗ 2a) kJ mol−1 , 0 a 0.5, (3) using Micromeritics Pulse Chemisorption ChemiSorb 2720
for Cu–Zn–Ce–Al oxide catalysts (Holmgren and Anderson,
where a is an oxygen to methanol (O/M) molar ratio. For 1998; Holmgren et al., 1999; Cheekatamarla et al., 2005). It
a = 0.125 OSRM can be operated auto-thermally at 573 K. The has been reported that the spillover of CO from metal onto the
limited work in the field of OSRM has been reported over the ceria support gives a higher dispersion of metal than the actual
different catalysts like CuO/ZnO/Al2 O3 (Murcia-Mascaros metal dispersion. Chemisorption of CO at 193 K alleviates this
et al., 2001; Turco et al., 2004a,b), Cu/Zn/Zr/Al oxide (Velu phenomenon. A 0.5 g catalyst was reduced in the stream of
et al., 2001), Ce/CuO (Shan et al., 2004), Pd/ZnO (Liu et al., 9.9% H2 /Ar with the flow of 30 cc min−1 at 533 K for 3 h with
2003), Pd–Zn/Cu–Zn–Al (Chen et al., 2007), etc. Turco a final temperature achieved by 10 K min−1 ramp rate. Then
et al. (2004b) observed the 100% methanol conversion at a the sample was flushed with N2 for 30 min and subsequently
temperature greater than 623 K. They observed the methane cooled to room temperature at a ramp rate of 10 K min−1 . Then
and other oxygenates in addition to H2 , CO2 and CO over 193 K temperature of sample was achieved by immersing the
CuO/ZnO/Al2 O3 . The copper–ceria based catalysts have been sample holder in the dewar of iso-propanol gel that was pre-
used widely for the preferential oxidation of carbon monoxide pared by quenching it with liquid nitrogen. At this temperature,
(PROX) (Wang et al., 2002; Papavasiliou et al., 2006). Also 0.2 ml pulses of CO were injected periodically in the stream
ceria is known to improve the stability of catalysts (Mattos of nitrogen in order to find out the total CO uptake. Tem-
et al., 2002). Therefore, in the present study we have devised perature programmed reduction (TPR) of calcined catalysts
the incorporation of ceria in Cu–Zn–Al oxide catalysts in order was carried out using Pulse Chemisorption ChemiSorb2720,
to improve the catalyst performance in OSRM process in terms Micromeritics. Different crystalline phases were identified by
of methanol conversion, hydrogen selectivity, suppression of means of Philips X’PERT PRO PW 3040/60 (PANalytical)
CO formation and deactivation compared to Cu–Zn–Al oxide powder diffractometer using monochromatic Cu-K 1.5418 Å
catalysts. The detailed characterization and activity study of radiation at a current of 30 mA with diffraction angle ranging
Cu–Zn–Ce–Al as well as Cu–Zn–Al oxide catalysts have been from 10–60◦ . The amount of carbon deposited on catalysts
carried out over the wide range of operating conditions of was determined by oxidizing it in the air by means of ther-
OSRM process. mogravimetric analyzer (TGA, Seiko TG/DTA 32 SSC 5100).
Surface analysis of post reaction catalysts was done by X-ray
2. Experimental photoelectron spectroscopy (XPS) using Perkin Elmer-1257.
Catalyst performance was evaluated in a fixed bed reactor
All the Cu–Zn–Ce–Al oxide and Cu–Zn–Al oxide catalysts, assembled with electrically heated furnace and PID con-
abbreviated as CZCeAi and CZAi , respectively, were prepared trollers. All the catalysts were reduced in situ in the stream
by co-precipitation method as follows. The 1.25 M solutions of 10% H2 /N2 mixture before OSRM reaction. The reduction
of Cu(NO3 )2 3H2 O, Zn(NO3 )2 6H2 O, Ce(NO3 )3 6H2 O and of Cu–Zn–Ce–Al oxide catalysts was carried out at a ramp of
Al(NO3 )3 9H2 O were prepared separately and then mixed in a 10 K min−1 with final temperature of 500 K and dwelled for
volume proportion corresponding to the final composition of 1 h; on the other hand Cu–Zn–Al oxide catalysts dwelled at
the catalysts to be obtained. The resulting solution was stirred 530 K for 1 h. A mixture of methanol and water vapor gener-
and heated to 343 K in a round-bottomed flask. The aqueous ated in an evaporator was passed through a preheating zone
0.5 M Na2 CO3 solution was added drop-wise to the nitrate along with oxygen and nitrogen prior to entering the reactor.
solution under vigorous stirring until pH 7 was attained at a The stainless steel reactor was used which consisted of a cylin-
temperature of 343 K. The precipitate was allowed to age for drical cup of 19 mm diameter and 30 mm length, perforated on
2 h at temperature 343 K with stirring. The excess solution was the bottom with orifices of 0.5 mm diameter to enable the flow
removed by filtration. The precipitate was washed three times of gases. The catalyst was retained by a plug of glass wool.
by double distilled water at a temperature of 343 K followed An equal amount of inert quartz particles of 20/25 mesh were
by several times by double distilled water at room temperature mixed with the catalyst to dilute the heat transfer effects. A
in order to remove the sodium salts. The drying was carried thermowell passed through the catalyst bed so that temperature
out at 383 K for 12 h. This material was the catalyst precursor, could be measured at different positions along the bed using
Please cite this article as: Patel, S., Pant, K.K., Selective production of hydrogen via oxidative steam reforming of methanol using Cu–Zn–Ce–Al oxide
catalysts. Chemical Engineering Science (2007), doi: 10.1016/j.ces.2007.01.066
ARTICLE IN PRESS
S. Patel, K.K. Pant / Chemical Engineering Science ( ) – 3

thermocouple. Product effluent was passed through the con- Pt/CeO2 catalyst. However, they observed that the spillover
denser and separator in order to separate the unconverted of CO to ceria support could be hindered by performing CO
methanol and water from the product gases. Operating tem- chemisorption at 195 K. The catalysts containing Ce exhibited
perature, contact time (W/F) and oxygen to methanol (O/M) higher copper surface area and better dispersion, and hence
molar ratio varied from 473 to 573 K, 3 to 15 kgcat s mol−1 the smaller copper particle size compared to those containing
and 0.1 to 0.5, respectively, with steam to methanol (S/M) only Cu–Zn–Al oxides. However, optimum loading of Ce was
molar ratio = 1.5 and pressure P = 1 atm being kept con- required as shown in Table 1. The alumina and ceria enhanced
stant. Reaction products were analyzed by Nucon-5700 Gas the dispersion of copper oxide species by stabilization of iso-
chromatograph, equipped with thermal conductivity detector lated Cu2+ ions in their matrix and moderately by formation
having carbosphere column for gaseous product concentration of spinal like CuAl2 O4 . This is in agreement with the work of
measurement and flame ionization detector with silica–alumina Fernandez-Garcia et al. (1997) who confirmed that the ceria
fused capillary column for unconverted liquid reactants. The increases the copper dispersion by reducing the particle size of
very low concentration of CO was determined using molec- copper in the CuO/CeO2 /Al2 O3 catalysts compared to only
ular sieve-5A followed by methanizer installed with flame CuO/Al2 O3 catalysts. XRD spectra of catalyst precursors dis-
ionization detector. played in Fig. 1(a) exhibit mainly hydrotalcite (HT)-like lay-
ered double hydroxide (LDH; (CuZn)6 Al2 (OH)16 CO3 · 4H2 O;
3. Results and discussion JCPDS file No. 38-487) with other phases such as aurichal-
cite (AC; (CuZn)5 (CO3 )2 (OH)6 ; JSPDS file No. 7-743) and
3.1. Catalyst characterization bayerite (AH; Al(OH)3 ; JCPDS file No. 20-11). At pH of
precipitates higher than 7 during catalyst preparation, aurichal-
The specific surface area and pore volume of various cat- cite was the dominated phase of catalyst precursors, and on
alyst samples shown in Table 1 were measured according to the other hand at pH lower than 7, the hydroxynitrates were
the Brunauer–Emmett–Teller (BET) method by nitrogen ad- dominated. At pH 7 HT-like LDH phase was dominant which
sorption at 77 K after degassing the samples at least for 12 h exhibited the higher catalytic activity. The calcination of these
at 523 K. There is no significant difference in the BET sur- precursors resulted in the catalyst comprising copper, zinc,
face area was observed among the catalysts. The surface area, cerium and aluminum oxide. When catalysts were reduced in
dispersion and particle size of copper in the catalyst samples the stream of hydrogen at the beginning of activity test, CuO
shown in Table 1 were calculated assuming atomic adsorption converted into copper where as zinc, cerium and aluminum
of CO on a copper site and the site density of copper to be remained in oxide form as shown in Fig. 1(b). A small peak of
0.069 nm2 atm−1 . All the copper particles were assumed to spinal CuAl2 O4 was also observed in the XRD spectra. As per
be spherical for calculating the mean particle size of copper. the Scherrer formula the crystal size is inversely proportional to
Stoichiometrically one atom of CO is adsorbed on each copper the peak width at constant diffraction angle and monochromatic
atom site. The CO chemisorption was carried out at 193 K wavelength. The peak width of copper for a CZCeA2 catalyst
in order to hinder the spillover of CO to the ceria that was is broader than that for the CZA2 catalyst (Fig. 1(b)), which re-
confirmed by comparing the chemisorption results of catalysts sulted in a smaller copper crystallite size. The smaller the crys-
which were obtained at room temperature 302 K. Holmgren tallite sizes, the better the copper dispersion and the more the
and Anderson (1998) have confirmed the CO spillover to copper surface area for CZCeA2 catalyst compared to CZA2
the ceria support by performing CO chemisorption at 298 K catalyst. The reducibility of copper species in catalysts was
followed by temperature programme desorption (TPD) over investigated by TPR experiments, and profiles are displayed in

Table 1
Physiochemical properties of catalysts

CZA1 CZA2 CZCeA1 CZCeA2 CZCeA3


Cu/Zn/Al Cu/Zn/Al Cu/Zn/Ce/Al Cu/Zn/Ce/Al Cu/Zn/Ce/Al

Preparative composition (wt%) 30/30/40 30/20/50 30/25/5/40 30/20/10/40 30/10/20/40


Final composition (wt%) 29/26/45 29/19/52 30/24/6/40 29/18/10/43 29/10/19/42
BET surface area (m2 g−1 ) 92 106 96 108 101
Pore volume (cm3 g−1 ) 0.26 0.32 0.28 0.34 0.29
Cu dispersion (%) 9.4 12.8 10.2 19.6 14.8
Cu surface area (m2 g−1 ) 18.3 25.1 20.2 38.6 29.3
Cu particle size (Å) 108 80 101 52 69
Methanol conversiona (%) 60 77 69 100 90
H2 ratea (mmol s−1 kg−1
cat ) 132 180 160 244 217
CO formationa (ppm) 9400 3400 1400 995 1240
a Results at T = 553 K, S/M = 1.5 M, O/M = 0.15 M and W/F = 11 kgcat mol−1 s.
Please cite this article as: Patel, S., Pant, K.K., Selective production of hydrogen via oxidative steam reforming of methanol using Cu–Zn–Ce–Al oxide
catalysts. Chemical Engineering Science (2007), doi: 10.1016/j.ces.2007.01.066
ARTICLE IN PRESS
4 S. Patel, K.K. Pant / Chemical Engineering Science ( ) –

LDH
AH
AC
(003) (012) (015)
(006)
Intensity (a.u.)

CZCeA2

CZA2

10 15 20 25 30 35 40 45 50 55 60

Cu
CuO
ZnO
Al2O3
CeO2
Intensity (a.u.)

CuAl2O4
Fig. 3. Catalytic activity in terms of methanol conversion for different catalysts
as a function of temperature (W/F = 11 kgcat s mol−1 methanol , O/M = 0.15 M,
CZCeA2
S/M = 1.5 M, P = 1 atm).

attributed to the reduction of CuO into Cu whereas the peak


CZA2
at 459 K might be due to spinal CuAl2 O4 . But it is expected
that the limited amount of surface CuAl2 O4 species contribute
10 15 20 25 30 35 40 45 50 55 60
to the TPR signals; therefore, the peak at 459 K might be a

composite peak of CuAl2 O4 and Cu2 O reduction. The forma-
Fig. 1. (a) X-ray diffraction patterns of catalysts precursors. Layered dou-
tion of Cu2 O is also confirmed by Lindstrom et al. (2002) and
ble hydroxide (LDH) (CuZn)6 Al2 (OH)16 CO3 · 4H2 O; aurichalcite (AC): Reitz et al. (2001) in which they found two step reduction of
(CuZn)5 (CO3 )2 (OH)6 ; bayerite (AH): Al(OH)3 . (b) X-ray diffraction pat- copper: Cu(II) → Cu(I) → Cu(0). For CZCeA2 catalyst three
terns of reduced catalysts. composite peaks were observed with shoulders at 443, 463 and
488 K. The lower valent Cu species cause structural defects of
ceria lattice. However, cerium sublattice is not strongly per-
turbed by these defects, which instead cause the formation of
a defective oxygen sublattice (Shan et al., 2004). In this way
many oxygen vacancies are generated with the formation of
solid solutions. Therefore, oxygen vacancies of CZCeA2 cata-
lyst were reduced at a low temperature of 443 K during TPR to
give first peak. The peak at temperature 463 K was attributed
to spinal CuAl2 O4 and Cu2 O, and a peak at 488 K was due to
CuO reduction to Cu. Breen and Ross (1999) and Agrell et al.
(2001) reported that the highly dispersed CuO gave the TPR
signals at lower temperature than the bulk CuO. Dow et al.
(2000) and Liu and Flyzani-Stephanopoulos (1996) found that
strong interaction occurs at the interface between highly dis-
persed copper oxide and ceria (interfacial metal oxide–support
interaction (IMOSI)) due to their close contact with each other
that leads to the low temperature reduction of copper oxide.
Fig. 2. TPR profiles of calcined catalysts. This is also observed in the present study with the catalyst
CZCeA2 for which CuO was reduced at lower temperature than
Fig. 2. All the catalysts exhibited composite peaks, with two that of CZA2 due to better dispersion of copper in CZCeA2 .
or three unresolved signals. These might be due to reduction
of different Cu(II) species such as, CuO, CuAl2 O4 and Cu2+ 3.2. Effect of ceria over catalysts performance
ions incorporated in octahedral sites of the alumina phase. The
pure CuO sample had shown a single large peak at 613 K. For Figs. 3 and 4 clearly depict that the addition of ceria in
CZA2 catalyst two peaks were observed, the peak at 513 K is CZAi catalysts improved the performance greatly in terms
Please cite this article as: Patel, S., Pant, K.K., Selective production of hydrogen via oxidative steam reforming of methanol using Cu–Zn–Ce–Al oxide
catalysts. Chemical Engineering Science (2007), doi: 10.1016/j.ces.2007.01.066
ARTICLE IN PRESS
S. Patel, K.K. Pant / Chemical Engineering Science ( ) – 5

Fig. 4. Catalytic activity in terms of hydrogen production rate for differ- Fig. 5. Comparison of CO formation for different catalysts as a function
ent catalysts as a function of temperature (W/F = 11 kgcat s mol−1methanol , of temperature (W/F = 11 kgcat s mol−1
methanol , O/M = 0.15 M, S/M = 1.5 M,
O/M = 0.15 M, S/M = 1.5 M, P = 1 atm). P = 1 atm).

of methanol conversion as well as hydrogen production. The corresponding H2 rate compared to 10% Ce catalyst as shown in
CZCeA2 exhibited 100% methanol and 244 mmol s−1 kg−1 cat hy-
Figs. 3 and 4. The carbon monoxide formation was raised dras-
drogen rate at 553 K whereas 50% and 112 mmol s kg−1 −1
cat
tically at a temperature greater than 513 K for all the Cu–Zn–Al
given by CZA1 . The methanol conversion was increased as a oxide catalysts, whereas for the 10% and 15% ceria-based cat-
function of temperature for all the catalysts. The enhanced ac- alysts it was increased gradually as a function of temperature
tivity of CZCeA2 could be due to higher copper surface area and as shown in Fig. 5. The CO concentration of 10% and 15%
better copper dispersion (Table 1 and XRD spectra of Fig. 1(b)). Cu–Zn–Ce–Al oxide catalysts was in the range of 0.05–0.1%,
Cheekatamarla et al. (2005) observed the better copper disper- which is quite low. Incorporation of ceria leads to the inhibition
sion in ceria promoted copper–alumina catalyst than catalyst of CO formation envisaged due to the oxygen storage–release
without it. The addition of zinc oxide and ceria caused the syn- ability of ceria; this was expected because copper–ceria based
ergistic effect to give the activity enhancement. Ceria has been catalysts have been used for the CO oxidation and PROX. Pillai
widely used as a promoter in three-way catalysts for removal of and Deevi (2006) reported that catalytic activity of CuO/ZnO
carbon monoxide, nitrogen oxides and hydrocarbons under con- could be improved significantly by incorporating CeO2 for the
ditions with rich-lean air/fuel in automotive exhaust (Golunski carbon monoxide oxidation. The CuO/ZnO/CeO2 based cata-
et al., 1995). It has also been widely studied for PROX to clean lysts have also been quite active for the PROX in hydrogen rich
the hydrogen rich feed stream for PEM fuel cell applications stream for the PEM fuel cell applications (Jung et al., 2004;
(Wang et al., 2002; Pillai and Deevi, 2006) and methane par- Cheekatamarla et al., 2005).
tial oxidation (Pantu and Gavalas, 2002). This is because ceria
has a fluorite-like cubic structure in which each cerium site is 3.3. Effect of contact time and oxygen to methanol molar
surrounded by eight oxygen sites in fcc arrangement and each ratio on catalyst performance
oxygen site has a tetrahedron of cerium sites. Under various
redox conditions, the oxidation state of cerium may vary be- The effect of contact time (W/F) on the methanol conver-
tween 3+ and 4+. The better redox properties than those CeO2 sion and CO formation at a temperature of 553 K is shown
alone can be obtained by the incorporation of metal ions into in Fig. 6. The methanol conversion reached 100% at contact
the CeO2 lattice forming Ce1−x Mx Oy solid solutions (Shan time more than 11 kgcat s mol−1 . The CO formation increased
et al., 2004). Surface oxygen and oxygen vaccines are involved from 300 ppm at contact time 3 to 2400 ppm at contact time
in the catalytic activity. This enhanced oxygen-mobility facili- 15 kgcat s mol−1 . The contact time 11 kgcat s mol−1 is the opti-
tates the occurrence of redox process like OSRM at lower tem- mum because the 100% methanol conversion with CO forma-
perature. Therefore, enhanced performance of Cu–Zn–Ce–Al tion as low as 995 ppm was obtained with it.
oxide catalysts in terms of methanol conversion, high hydro- It was found that as contact time was increased the hydro-
gen production rate and low CO formation was observed in the gen selectivity marginally decreased. The concentration of car-
present study. It can be seen that optimum loading of ceria was bon monoxide for all the catalysts was always below 1% at all
required among the Cu–Zn–Ce–Al oxide catalysts. The 15% the operating conditions, which is well below the equilibrium
and 5% Ce catalysts had given lower methanol conversion and concentration of water gas shift reaction. At a temperature less
Please cite this article as: Patel, S., Pant, K.K., Selective production of hydrogen via oxidative steam reforming of methanol using Cu–Zn–Ce–Al oxide
catalysts. Chemical Engineering Science (2007), doi: 10.1016/j.ces.2007.01.066
ARTICLE IN PRESS
6 S. Patel, K.K. Pant / Chemical Engineering Science ( ) –

Fig. 7. Effect of oxygen to methanol molar ratio over methanol conversion


and hydrogen production for CZCeA2 catalyst (W/F = 11 kgcat s mol−1
methanol ,
T = 553 K, S/M = 1.5 M, P = 1 atm).
Fig. 6. Effect of contact time over methanol conversion and CO formation
for CZCeA2 catalyst (T = 553 K, O/M = 0.15 M, S/M = 1.5 M, P = 1 atm).

analysis as well as kinetic studies. They confirmed that the CO


than 473 K and contact time 3 kgcat s mol−1 there was no CO formation over CuO/ZnO/ZrO2 /Al2 O3 catalyst for steam re-
detected; therefore it can be postulated that the carbon monox- forming of methanol occurs via rWGS (Eq. (4)). Recently, many
ide is not a primary product. All these suggest that the CO researchers have also proposed the CO formation via rWGS
formation takes place through reverse water gas shift reaction reaction (Eq. (4)) (Agrell et al., 2001; Purnama et al., 2004;
(rWGS: Eq. (4)), which uses the hydrogen and carbon dioxide (Reuse et al., 2004; Horny et al., 2007).
produced by reforming reaction (Eq. (3)). The O/M molar ratio is expected to have a strong influence
on the catalytic performance because the OSRM is a combi-
H2 + CO2 ←→ CO + H2 O. (4) nation of POM and SRM reactions. The data were collected
It can be seen in from Fig. 4 that at 573 K temperature H2 rate by keeping the steam to methanol molar ratio at 1.5 as shown
is lower than that at 553 K even if methanol conversion was in Fig. 7. The methanol conversion and hydrogen production
100% at both the temperatures. This was because of conversion rate increased with increasing O/M ratio and attained maximum
of hydrogen into carbon monoxide via WGS reaction as it was corresponding values at O/M = 0.15. With further increase in
accelerated at higher temperature. Various routes have been O/M ratio the methanol conversion was 100% but the rate of
proposed for the formation of CO in the literature of methanol H2 production declined drastically up to O/M = 0.25 due to the
reforming. Santacesaria and Carra (1983) have suggested the increase in CO formation rate by rWGS and also due to some
decomposition/water gas shift reaction sequence: extent of methanol partial oxidation (Eq. (1)). At O/M ratio
greater than 0.25 both methanol conversion and hydrogen pro-
CH3 OH → CO + 2H2 , (5) duction decreased steeply because methanol partial oxidation
(Eq. (1)) was the dominant reaction. The high exothermicity
CO + H2 O ←→ H2 + CO2 . (6) of POM resulted in inefficient heat removal and consequently
decline of methanol conversion. The S/M molar ratio 1.5 was
Jiang et al. (1993) proposed the kinetic model for steam re-
kept constant in the present study based on our previous study
forming of methanol in which they suggested CO formation
(Agarwal et al., 2005). The excess steam retarded the methanol
via decomposition of methyl formate:
decomposition by providing sufficient amount of steam for the
2CH3 OH → CH3 OCHO + 2H2 , (7) methanol reforming reaction (Eq. (2)), suppressed the CO for-
mation by inhibiting the rWGS reaction (Eq. (4)) to proceed
CH3 OCHO + H2 O → HCOOH + CH3 OH, (8) toward the right side and also reduced the catalyst deactivation.
HCOOH → CO2 + H2 . (9)
3.4. Reaction pathway
Peppley et al. (1999) reported that the methanol decomposi-
tion is much slower than the steam reforming methanol reaction The OSRM involves the following reactions in addition to
but it must be included in the overall reaction scheme compris- the OSRM reaction Eq. (3). The POM (Eq. (11)) and SRM
ing methanol steam reforming, methanol decomposition and (Eq. (12)) are expected to occur simultaneously at the entry
water gas shift reactions. Horng et al. (2006) reported the CO of reactor, while at the downstream side of reactor the major
formation via WGS (Eq. (6)) for the auto-thermal reforming reaction may be the SRM.
of methanol. Breen et al. (1999) have done extensive study
to understand the CO formation mechanism through DRIFT CH3 OH + 0.25O2 ←→ 2H2 + 0.5CO2 + 0.5CO, (10)
Please cite this article as: Patel, S., Pant, K.K., Selective production of hydrogen via oxidative steam reforming of methanol using Cu–Zn–Ce–Al oxide
catalysts. Chemical Engineering Science (2007), doi: 10.1016/j.ces.2007.01.066
ARTICLE IN PRESS
S. Patel, K.K. Pant / Chemical Engineering Science ( ) – 7

catalytic activity might be due to some transient transforma-


tion to a stable form under the reaction condition that occurred
in the catalyst. Catalyst CZCeA2 containing ceria was found
to be quite durable because ceria promoter has high oxygen
storage–release capacity, which inhibits the coke deposition
significantly by coke gasification. The higher amount of lat-
tice oxygen near the metal particles promotes the mechanism
of carbon removal from the metallic surface (Fernandez-Garcia
et al., 1997; Mattos et al., 2002, Patel and Pant, 2006a). They
also reported that ceria improves the stability against sinter-
ing of metallic copper phase formed during the reaction. The
amount of coke deposited on the catalysts obtained using TGA
was 0.18 and 1.55 wt% for the CZCeA2 and CZA2 catalysts,
Fig. 8. Time on stream (TOS) stability test of catalysts
respectively. The C1s XPS spectra of spent catalysts revealed
(W/F = 11 kgcat s mol−1
methanol , T = 553 K, O/M = 0.15 M, S/M = 1.5 M,
that the graphite carbon –C–C– (284 eV, BE) and oxidized car-
P = 1 atm). bon CO3 (288 eV, BE) were present on the catalyst surface.

4. Conclusion
CH3 OH + 0.5O2 ←→ 2H2 + CO2 , (11)
In the present study to investigate the effect of Ce loading
CH3 OH + H2 O ←→ 3H2 + CO2 , (12) on the performance of Cu–Zn–Al oxide catalysts for OSRM,
the optimum 10 wt% doping of cerium improved the catalyst
CH3 OH ←→ 2H2 + CO, (13)
performance greatly in terms of methanol conversion, hydro-
gen production rate and CO suppression. The 100% methanol
CO + 0.5O2 ←→ CO2 , (14)
conversion with CO formation as low as 995 ppm was obtained
H2 + 0.5O2 ←→ H2 O, (15) that reduce the load on preferential oxidation of CO to CO2
(PROX) before feeding the hydrogen rich stream as a feed for
CH3 OH + 0.25O2 ←→ HCHO + 0.5H2 O + 0.5H2 , (16) the PEM fuel cells. The optimum operating parameters could
be suggested as reaction temperature 553 K, contact time (W/F)
The experimental hydrogen selectivity (moles of hydro- 11 kgcat s mol−1 , oxygen to methanol (O/M) molar ratio 0.15
gen produced per mole of methanol converted) was observed and steam to methanol molar (S/M) ratio 1.5. The spillover of
slightly lower than that stoichiometrically given by Eq. (3). carbon monoxide to ceria can be hindered by performing the
The selectivity of carbon dioxide decreased with tempera- CO chemisorption over ceria-based catalysts at low tempera-
ture, whereas that of carbon monoxide was increased. These ture of 193 K. The CO chemisorption, TPR and XRD studies
above all suggest the occurrence of rWGS reaction (Eq. (4)), have revealed that the incorporation of Ce in Cu–Zn–Al oxide
and also the possibility of some hydrogen produced via Eqs. catalysts improved the copper dispersion by reducing the par-
(10) and (13). For the OSRM process, benefit in reduction ticle size of copper, and also facilitated catalyst reduction at
of CO concentration might be obtained due to conversion of lower temperature. The incorporation of ceria greatly improved
CO to CO2 via oxidation reaction (Eq. (14)). The formation the stability of Cu–Zn–Al oxide catalysts.
of formaldehyde via Eq. (16) was prevented by operating the
OSRM at higher space velocities (Velu et al., 2001). The ceria-
based Cu–Zn–Ce–Al oxide catalysts enhanced the activity References
that facilitated the complete methanol conversion at relatively
Agarwal, V., Patel, S., Pant, K.K., 2005. H2 production by steam reforming
lower temperature of 553 K so that there was no formation of of methanol over Cu/ZnO/Al2 O3 catalysts: transient deactivation kinetics
methane and other oxygenates which were observed by Turco modeling. Applied Catalysis A: General 279, 155–164.
(et al. (2004b). Agrell, J., Hasselbo, K., Jansson, K., Jaras, S.G., Boutonnet, M., 2001.
Production of hydrogen from methanol over Cu/ZnO catalysts promoted
by ZrO2 and Al2 O3 . Applied Catalysis A: General 211, 239–250.
3.5. Stability of catalysts Bowers, B.J., Zhao J.L., Ruffo, M., Khan, R., Dattatraya, D., Dushman, N.,
Beziat, J., Boudjemaa, F., 2006. Onboard fuel processor for PEM fuel cell
The Cu/Zn/Al2 O3 catalysts face the great problem of deac- vehicles. International Journal of Hydrogen Energy, in press. Available
tivation (Agarwal et al., 2005). The CZA2 catalyst deactivated online at sciencedirect.com.
greatly by a magnitude of 26% methanol conversion over a Breen, J.P., Ross, J.R.H., 1999. Methanol reforming for fuel-cell applications:
development of zirconia-containing Cu–Zn–Al catalysts. Catalysis Today
72 h TOS stability test as shown in Fig. 8. On the other hand
51, 521–533.
the activity of CZCeA2 catalyst declined only by a magni- Breen, J.P., Meunier, F.C., Ross, J.R.H., 1999. Mechnistic aspects of the steam
tude of 8%. For a CZCeA2 catalyst activity declined initially reforming of methanol over a CuO/ZnO/ZrO2 /Al2 O3 catalyst. Chemical
and then became nearly constant. This initial decline in the Communications, 2247–2248.

Please cite this article as: Patel, S., Pant, K.K., Selective production of hydrogen via oxidative steam reforming of methanol using Cu–Zn–Ce–Al oxide
catalysts. Chemical Engineering Science (2007), doi: 10.1016/j.ces.2007.01.066
ARTICLE IN PRESS
8 S. Patel, K.K. Pant / Chemical Engineering Science ( ) –

Cheekatamarla, P.K., Epling, W.S., Lane, A.M., 2005. Selective low- Patel, S., Pant, K.K., 2006a. Activity and stability enhancement of
temperature removal of carbon monoxide from hydrogen-rich fuels over copper–alumina catalysts using cerium and zinc promoters for the selective
Cu–Ce–Al catalysts. Journal of Power Sources 147, 178–183. production of hydrogen via steam reforming of methanol. Journal of Power
Chen, G., Li, S., Yuan, Q., 2007. Pd −Zn/Cu −Zn −Al catalysts prepared for Sources 159, 139–143.
methanol oxidation reforming in micro channel reactors. Catalysis Today Patel, S., Pant, K.K., 2006b. Influence of preparation method on performance
120, 63–70. of Cu(Zn)(Zr)–alumina catalysts for the hydrogen production via steam
Choi, Y., Stenger, H.G., 2002. Fuel cell grade hydrogen from methanol on a reforming of methanol. Journal of Porous Materials 13, 381–386.
commercial Cu/ZnO/Al2 O3 catalyst. Applied Catalysis B: Environmental Patel, S., Pant, K.K., 2006c. Production of hydrogen with low carbon
38, 259–269. monoxide formation via catalytic steam reforming of methanol. ASME
Dow, W., Wang, Y., Huang, T., 2000. TPR and XRD studies of yttria- Journal of Fuel Cell Science and Technology 3, 369–374.
doped ceria/g-alumina-supported copper oxide catalyst. Applied Catalysis Peppley, B.A., Amphlett, J.C., Kearns, L.M., Mann, R.F., 1999.
A: General 190, 25–34. Methanol–steam reforming on Cu/ZnO/Al2 O3 catalysts. Part 2. A
Edwards, N., Ellis, S.R., Frost, J.C., Golunski, S.E., Keulen, A.N.J., comprehensive kinetic model. Applied Catalysis A: General 179, 31–49.
Lindewald, N.G., Reinkingh, J.G., 1998. On-board hydrogen generation Pillai, U.R., Deevi, S., 2006. Copper–zinc oxide and ceria promoted
for transportation applications: the HotSpot methanol processor. Journal copper–zinc oxide as highly active catalysts for low temperature oxidation
of Power Sources 71, 123–128. of carbon monoxide. Applied Catalysis B: Environmental 65, 110–117.
Fernandez-Garcia, M., Rebollo, E.G., Ruiz, A.G., Conesa, J.C., Soria, Purnama, H., Ressler, T., Jentoft, R.E., Soerijanto, H., Schlogl, R.,
J., 1997. Influence of ceria on the dispersion and reduction/oxidation Schomacker, R., 2004. CO formation/selectivity for steam reforming of
behavior of alumina-supported copper catalysts. Journal of Catalysis 172, methanol with a commercial CuO/ZnO/Al2 O3 catalyst. Applied Catalysis
146–159. A: General 259, 83–94.
Golunski, S.E., Hatcher, H.A., Rajaram, R.R., Truex, T.J., 1995. Origins Raimondi, A., Geissler, K., Wambach, J., Wokaun, A., 2002. Hydrogen
of low-temperature three-way activity in Pt/CeO2 . Applied Catalysis B: production by methanol reforming: post-reaction characterisation of a
Environmental 3, 367–376. Cu/ZnO/Al2 O3 catalyst by XPS and TPD. Applied Surface Science 189,
Holmgren, A., Anderson, B., 1998. Oxygen storage dynamics in 59–71.
Pt/CeO2 /Al2 O3 catalysts. Journal of Catalysis 178, 14–25. Reitz, T.L., Lee, P.L., Czaplewski, K.F., Lang, J.C., Popp, K.E., Kung, H.H.,
Holmgren, A., Anderson, B., Duprez, D., 1999. Interactions of CO 2001. Time-resolved XANES investigation of CuO/ZnO in the oxidative
with Pt/ceria catalysts. Applied Catalysis B: Environmental 22, methanol reforming reaction. Journal of Catalysis 199, 193–201.
215–230. Reuse, P., Renken, A., Hass-Santo, K., Gorke, O., Schubert, K., 2004.
Horng, R., Chou, H., Lee, C., Tsai, H., 2006. Characterisation of hydrogen Hydrogen production for fuel cell application in an autothermal micro-
produced by partial oxidation and auto-thermal reforming in a small channel reactor. Chemical Engineering Journal 101, 133–141.
methanol reformer. Journal of Power Sources 161, 1225–1233. Sahoo, D.R., Vajpai, S., Patel, S., Pant, K.K., 2007. Kinetic modeling of
Horny, C., Renken, A., Kiwi-Minsker, L., 2007. Compact string reactor for steam reforming of ethanol for the production of hydrogen over Co/Al2 O3
autothermal hydrogen production. Catalysis Today 120, 45–53. catalyst. Chemical Engineering Journal 125, 139–147.
Jiang, C.J., Trimm, D.L., Wainwright, M.S., Cant, N.W., 1993. Kinetic Santacesaria, E., Carra, S., 1983. Kinetics of catalytic steam reforming of
mechanism for the reaction between methanol and water over a methanol in a CSTR reactor. Applied Catalysis 5, 345–358.
Cu–ZnO–Al2 O3 catalyst. Applied Catalysis A: General 97, 145–158. Semelsberger, T.A., Ott, K.C., Borup, R.L., Greene, H.L., 2006. Generating
Jung, C.R., Han, J., Nam, S.W., Lim, T.H., Hong, S.A., Lee, H.I., 2004. hydrogen-rich fuel-cell feeds from dimethyl ether (DME) using Cu/Zn
Selective oxidation of CO over CuO–CeO2 catalyst: effect of calcination supported on various solid–acid substrates. Applied Catalysis A: General
temperature. Catalysis Today 93–95, 183–190. 309, 210–223.
Lindstrom, B., Pettersson, L.J., Menon, P.G., 2002. Activity and Shan, W., Feng, Z., Li, Z., Zhang, J., Shen, W., Li, C., 2004. Oxidative
characterization of Cu/Zn, Cu/Cr and Cu/Zr on -alumina for methanol steam reforming of methanol on Ce0.9Cu0.1Oy catalysts prepared by
reforming for fuel cell vehicles. Applied Catalysis A: General 234, deposition–precipitation, coprecipitation, and complexation–combustion
111–125. methods. Journal of Catalysis 228, 206–217.
Lindstrom, B., Agrell, J., Pettersson, L.J., 2003. Combined methanol Turco, M., Bagnasco, G., Costantino, U., Marmottini, F., Montanari, T.,
reforming for hydrogen generation over monolithic catalysts. Chemical Ramis, G., Busca, G., 2004a. Production of hydrogen from oxidative
Engineering Journal 93, 91–101. steam reforming of methanol I. Preparation and characterization of
Liu, S., Takahashi, K., Ayabe, M., 2003. Hydrogen production by oxidative Cu/ZnO/Al2 O3 catalysts from a hydrotalcite-like LDH precursor. Journal
steam reforming of methanol on Pd/ZnO catalyst: effects of Pd loading. of Catalysis 228, 43–55.
Catalysis Today 87, 247–253. Turco, M., Bagnasco, G., Costantino, U., Marmottini, F., Montanari, T.,
Liu, W., Flyzani-Stephanopoulos, M., 1996. Transition metal-promoted Ramis, G., Busca, G., 2004b. Production of hydrogen from oxidative steam
oxidation catalysis by fluorite oxides: a study of CO oxidation over reforming of methanol II. Catalytic activity and reaction mechanism on
Cu–CeO2 . The Chemical Engineering Journal and the Biochemical Cu/ZnO/Al2 O3 hydrotalcite-derived catalysts. Journal of Catalysis 228,
Engineering Journal 64, 283–294. 56–65.
Mattos, L.V., Oliveira, E.R., Resende, P.D., Noronha, F.B., Passos, F.B., 2002. Velu, S., Suzuki, K., Kapoor, M.P., Ohashi, F., Osaki, T., 2001. Selective
Partial oxidation of methane on Pt/Ce–ZrO2 catalysts. Catalysis Today production of hydrogen for fuel cells via oxidative steam reforming of
77, 245–256. methanol over CuZnAl(Zr)-oxide catalysts. Applied Catalysis A: General
Murcia-Mascaros, S., Navarro, R.M., Gomez-Sainero, L., Costantino, U., 213, 47–63.
Nocchetti, M., Fierro, J.L.G., 2001. Oxidative methanol reforming reactions Wang, J.B., Tsai, D., Huang, T., 2002. Synergistic catalysts of carbon
on CuZnAl catalysts derived from hydrotalcite-like precursors. Journal of monoxide oxidation over copper oxide supported on samaria-doped ceria.
Catalysis 198, 338–347. Journal of Catalysis 208, 370–380.
Pantu, P., Gavalas, G.R., 2002. Methane partial oxidation on Pt/CeO2 and Wang, Z., Xi, J., Wang, W., Lu, G., 2003. Selective production of hydrogen by
Pt/Al2 O3 catalysts. Applied Catalysis A: General 223, 253–260. partial oxidation of methanol over Cu/Cr catalysts. Journal of Molecular
Papavasiliou, J., Avgouropoulos, G., Ioannides, T., 2006. In situ Catalysis A: Chemical 191, 123–134.
combustion synthesis of structured Cu–Ce–O and Cu–Mn–O catalysts Yan, X., Wang, S., Li, X., Hou, M., Yuan, Z., Li, D., Pan, L., Zhang, C.,
for the production and purification of hydrogen. Applied Catalysis B: Liu, J., Ming, P., Yi, B., 2006. A 75-kW methanol reforming fuel cell
Environmental 66, 168–174. system. Journal of Power Sources 244, 1–9.

Please cite this article as: Patel, S., Pant, K.K., Selective production of hydrogen via oxidative steam reforming of methanol using Cu–Zn–Ce–Al oxide
catalysts. Chemical Engineering Science (2007), doi: 10.1016/j.ces.2007.01.066

You might also like