You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/309759153

Effective hydrogen production from propane steam reforming over bimetallic


co-doped NiFe/Al2O3 catalyst

Article · November 2016


DOI: 10.1016/j.jiec.2016.10.046

CITATIONS READS

20 1,127

6 authors, including:

Byeong Kwak No-Kuk Park


Yeungnam University Yeungnam University
53 PUBLICATIONS   727 CITATIONS    147 PUBLICATIONS   1,048 CITATIONS   

SEE PROFILE SEE PROFILE

Tae Jin Lee Misook Kang


Chung-Ang University Yeungnam University
195 PUBLICATIONS   2,551 CITATIONS    327 PUBLICATIONS   5,030 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Investigations on rutile TiO2 Nanorod Thin Films for Dye Sensitised Solar cell Applications View project

photo-catalyst field View project

All content following this page was uploaded by Misook Kang on 16 January 2018.

The user has requested enhancement of the downloaded file.


Journal of Industrial and Engineering Chemistry 46 (2017) 324–336

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Effective hydrogen production from propane steam reforming over


bimetallic co-doped NiFe/Al2O3 catalyst
Kang Min Kima , Byeong Sub Kwaka , No-Kuk Parkb , Tae Jin Leeb , Sang Tae Leec,
Misook Kanga,*
a
Department of Chemistry, College of Science, Yeungnam University, Gyeongsan, Gyeongbuk 38541, Republic of Korea
b
School of Chemical Engineering, Yeungnam University, Gyeongsan, Gyeongbuk 38541, Republic of Korea
c
Wooshin Com., Gyeongsan, Gyeongbuk 38470, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Article history:
Received 3 September 2016 In this study, we investigated the role of Fe oxide as a promoter to improve the redox properties of Ni in
Received in revised form 18 October 2016 PSR (propane steam reforming), thereby extending its lifetime and enabling it to be more easily oxidized
Accepted 29 October 2016 by the CO generated to produce CO2. Bimetallic NiFe supported on g-Al2O3 (30NixFeyO/70Al2O3) samples
Available online 8 November 2016 are prepared as catalysts and characterized by XRD, TEM, H2-TPR, TPO, and XPS. Moreover, a mechanism
for propane steam reforming over the 30NixFeyO/70Al2O3 catalyst is proposed on the basis of the GC and
Keywords: mass spectroscopy results and CO-, C3H8-, and H2O-temperature programmed desorption analysis.
Propane steam reforming Consequently, the Fe components in the 30NixFeyO/70Al2O3 catalysts suppressed the agglomeration
Hydrogen production
between the Ni and Al particles. The catalytic performances on the 30NixFeyO/70Al2O3 catalysts are
Fe-promoter
improved by the reduced deposition of carbon compared to that on the 30NiO/70Al2O3 one: at 700  C, the
30NixFeyO/70Al2O3
Particle agglomeration hydrogen selectivity amounted to 86% in the 30Ni0.8Fe0.2O/70Al2O3 catalyst compared to 79% in the
30NiO/70Al2O3 one.
ã 2016 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Introduction compressible to a transportable liquid. Thus, propane is a potential


candidate as a hydrogen carrier, because of its common storage (LP
Hydrogen has attracted interest as an efficient and environmen- gases) and existing widely spread infrastructure. Additionally,
tally friendly energy carrier. The methods of hydrogen production propane steam reforming (PSR) is the most economical pathway
include water electrolysis [1], the partial oxidation or gasification of in terms of the hydrogen yield, since hydrogen is produced from
heavy oil or coal [2], and steam reforming of hydrocarbons [3]. It steam as well as propane, as shown in the following equation for
should be noted however that from an economic viewpoint, the propane overall steam reforming:
steam reforming of hydrocarbons has been highlighted as one of the
C3H8 + 6H2O ! 3CO2 + 10H2 DH 298 = 499 kJ/mol (1)
most reasonable methods of producing hydrogen. Currently, the
hydrocarbons used as the hydrogen sources in the steam reforming Since industrial operations always use excess steam to
reaction can be separated into two kinds: firstly, light paraffins such minimize catalyst deactivation, the maximum yield of hydrogen
as methane [4], ethane [5], propane [6] and butane [7], and secondly, per mole of propane fed can reach 10.
oxidized compounds such as methanol [8], dimethylether [9], In practice, PSR is performed at high temperatures over Ni-
ethanol [10] and acetic acid [11]. In particular, propane has many based catalysts. However, the high temperature required may favor
advantages as a hydrogen source candidate; it is a gas at standard several routes to the formation of carbon deposits and the
temperature and pressure and a by-product of natural gas processing decomposition of the hydrocarbons. In addition, Ni catalysts tend
and petroleum refining, is commonly used as a fuel for engines, oxy- to agglomerate and then lose their active surface area under PSR
gas torches, portable stoves and residential central heating, but is conditions, resulting in short catalyst lifetimes. Consequently,
novel preparation and promotion techniques that can resist rapid
catalyst deactivation by coking and sintering are essential. Much of
* Corresponding author. Fax: +82 53 815 5412. this research effort has focused on developing Ni catalysts with
E-mail address: mskang@ynu.ac.kr (M. Kang). improved resistance to coke formation by adding promoters. Alkali

http://dx.doi.org/10.1016/j.jiec.2016.10.046
1226-086X/ã 2016 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336 325

metals, such as K2O [12] and MgO [13], have been shown to nickel-filtered CuKa radiation (30.0 kV, 30.0 mA) over the 2u range
improve the coking resistance by enhancing the carbon gasifica- of 10–100 . The core and shell shapes of the catalyst particles were
tion, but at the cost of reduced catalytic activity. The effects of determined by high-resolution TEM (H-7600, Hitachi, Japan)
novel metals, such as Pt, Rh and Ru, have also been investigated operated at 120 kV. The BET (Brunauer, Emmett and Teller) surface
[14–16], and their introduction into Ni-based catalysts to increase areas of the catalysts were measured using a Belsorp II instrument.
their hydrogen selectivity and decrease their coke deposition The BET surface measurement was performed by nitrogen gas
showed dramatic results. More recently, bi or tri metallic catalytic adsorption using a continuous flow method with a mixture of
species have been investigated. Laosiripojana et al. [17] reported nitrogen and helium as the carrier gas. A tube was filled with 0.2 g
that adding CeO2 to Ni/Al2O3 catalysts enhanced the nickel of the sample under an N2 atmosphere and then out-gassed for 1 h
dispersion and reactivity of the carbon deposits, leading to at 200  C before the measurements. After pre-treatment, the
improved catalytic activity and stability in the steam reforming samples were cooled down to room temperature and then exposed
of propane. Althenayan et al. [18] researched the use of bimetallic to liquid nitrogen for 2 h. In the a-plot method, the adsorption
Co–Ni/Al2O3 catalysts for propane dry reforming, in order to volume, Vads (p/p0), which was normalized to Vads for the reference
determine the intrinsic reaction rate simultaneously with the material, is used as a new x-axis to plot the adsorption isotherms
carbon-induced deactivation coefficient from the transient rate for the samples of interest. The discrete adsorption data for the
data over an extended period of time (up to 72 h), for propane dry reference material was interpolated numerically to generate a
reforming over a Co–Ni catalyst at 823–973 K. Malaibari et al. [19] continuous x-axis. The XPS (AXIS-NOVA Kratos Inc.) spectroscopy
also investigated the effect of Mo as a promoter in alumina- measurements of Ni2p, Fe2p Al2p, O1s, and C1s using a non-
supported Ni catalysts for PSR at 450  C. Mo promotion showed a monochromatic AlKa (1486.6 eV) X-ray source were performed at
beneficial effect by both decreasing the rate of carbon deposition the Yeungnam University instrumental center, Korea. The powders
and increasing the catalytic activity. were pelletized at 1.2  104 kPa for 1 min and the 1.0-mm pellets
In order to develop more durable and cheaper catalysts for use were stored overnight in a vacuum (1.0  107 Pa) to remove the
in PSR, this study broke new ground. Fe oxide was introduced as a water molecules from the surface before the measurements. The
promoter to improve the redox properties of Ni, thereby extending base pressure in the system was less than 1 109 Pa. The
its lifetime and enabling it to be more easily oxidized by the CO experiments were performed using a 200-W source power and
generated to produce CO2. Iron oxide can be thermally reduced by an angular acceptance of 5 . The analyzer axis formed a 90 angle
CO during PSR as follows: Fe2O3 + 3CO ! 2Fe + 3CO2. Additionally, with the specimen surface. The Shirley function was used to
partial reduction with hydrogen at about 400  C gives magnetite, a subtract the background for the XPS data analysis. The signals were
black magnetic material that contains both Fe(III) and Fe(II) fitted using the mixed Lorentzian–Gaussian curves.
through the reaction: 3Fe2O3 + H2 $ 2Fe3O4 + H2O and this reaction
can occur reversely also. Thus, in this study, we attempted to apply Gas adsorption abilities of the 30NixFeyO/70Al2O3 catalysts
bimetallic NiFe catalysts supported on g-Al2O3 to PSR. The
prepared catalysts were characterized by XRD, TEM, H2O-, C3H8- The H2-TPR experiments for the as-synthesized 30NixFeyO/
and CO-TPD, H2-TPR, TPO, and XPS. In addition, the nature of the 70Al2O3 samples were conducted using the same equipment as
promoter action afforded by Fe oxide was examined. The results that used for the thermo-gravimetric analysis (TGA) experiment
were applied to the design of practical catalysts for PSR. (Shinco com., Korea). Approximately 0.05 g of the catalyst was
pre-treated under flowing argon gas (30 mL min1) at 300  C for
Experimental 1 h and then cooled to 50  C. The analysis was carried out by
increasing the catalyst temperature from room temperature to
Preparations of the 30NixFeyO/70Al2O3 catalysts 900  C at a rate of 5  C min1 under H2 (5 vol.%)/Ar with a flow rate
of 50 mL min1.
The bimetallic NiFe catalysts supported on g-Al2O3 (30NixFeyO/ The adsorption abilities of the catalysts for the CO and C3H8
70Al2O3) were prepared using impregnation methods [20] with a gases were measured from the CO- and C3H8-TPD experiments
Ni-based catalyst content of 30 wt.% and Fe/Ni atomic ratios of 0, performed using a BELCAT (Bel Japan Inc., Japan). Each catalyst
0.1, 0.25, and 0.43. The 70 wt.% g-Al2O3 (Junsei Co., Japan) was (0.05 g) was placed in the quartz reactor of the TPD apparatus. The
mixed in ethanol solvent at room temperature and the suspension catalysts were pretreated at 300  C for 1 h under a He flow
was stirred for 1 h. Then, NiCl2 (99.99%, Junsei Co., Japan) and FeCl2 (30 mL min1) to remove the physically absorbed water and
(99.99%, Junsei Co., Japan) as the Ni and Fe sources, respectively, impurities. CO (5 vol.% CO/He) and C3H8 (25 vol.% C3H8/He) gases
were added to the suspension containing 70 wt.% g-Al2O3 powder, were injected into the reactor for 1 h at a rate of 50 mL min1 at
and the mixture was stirred at 40  C. At this time, the catalysts 50  C. The physically absorbed CO and C3H8 gases were removed by
were labeled NixFey, where x and y indicate the atomic %. The evacuating the catalyst samples at 50  C for 30 min. The furnace
NixFey main catalytic species were loaded onto the external temperature was increased from 50 to 900  C at a rate of
surfaces of alumina and the total loaded weights in all of the 10  C min1 under He flow. The desorbed CO and C3H8 gases were
samples were the same, viz. 30 wt.%. After aging for 3 h, the final detected using a TCD detector.
solutions were evaporated at 70  C for 6 h and dried at 50  C for The activation energy for water desorption in the catalysts was
24 h in an oven. The final samples were treated thermally at 800  C determined by a TGA apparatus equipped with a micro thermo-
for 3 h in air, and reduced by H2/argon (1:10 ratio) at 700  C for 2 h differential and gravimetric analyzer (Shinco com., Korea). The
to generate the NiFe oxide composites. The following four samples were analyzed after coming into contact with saturated
experimental bimetallic NiFe catalysts supported on Al2O3 were (NH4)2SO4 for 24 h to maintain identical water vapor conditions.
prepared for comparison: 30NiO/70Al2O3, 30Ni0.9Fe0.1O/70Al2O3, Blaine and Kissinger [21] presented a useful equation to calculate
30Ni0.8Fe0.2O/70Al2O3, and 30Ni0.7Fe0.3O/70Al2O3. the activation energies of various thermal reactions based on the
shifts of the maximum deflection temperature of the DTA
Physical properties of the 30NixFeyO/70Al2O3 catalysts thermograms upon the changing of the heat rates, as follows.
The Kissinger equation was selected and is given as ln(C/Tm2) =
The 30NiO/70Al2O3 and 30NixFeyO/70Al2O3 catalysts were Ea/RTm + a constant, where C is the DTA heating rate ( C/min),
examined by powder XRD (model MPD from PANalytical) using Tm is the crystallization peak temperature, Ea is the activation
326 K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336

energy required for crystallization and R is the gas constant Results and discussion
(1.987 cal/mol).
Characteristics of the 30NixFeyO/70Al2O3 catalysts
Propane steam reforming (PSR) reaction on the 30NixFeyO/70Al2O3
catalysts Based on the mechanism of the hydrocarbon reforming
reaction, it was confirmed that its rate increases in the presence
The reactor for PSR in this study was designed as a fixed bed- of metallic gradients [24]. Therefore, many studies have examined
type [22], and the devices that were employed are shown in Fig. 1: the reforming reaction after reducing the catalysts. The catalytic
it consists of a feed gas (C3H8) supply, a water supply, a quartz reduction temperature is also very important, and the activity of
reactor, and analysis equipment (GC). The catalytic activities were the catalyst is dependent on the reduction temperature. Therefore,
measured at 700  C for a reaction time of 10 h at a steam-to- to determine the proper pre-reduction temperature for the
propane ratio of 1:6 (mol.%) with a GHSV (gas hourly space samples before PSR, the changes corresponding to the reduction
velocity) of 6000 h1. The catalyst (0.4 g) was pelletized to a 20– of the Ni and Fe components were observed in the H2-TPR profiles
24 mesh and then packed with a small amount of quartz wool to of all of the fresh samples, as shown in Fig. 2. Generally, in H2-TPR,
prevent the catalyst from moving in the fixed-bed quartz reactor, the peak area corresponds to the hydrogen uptake and the peak
which was then mounted vertically inside the furnace. All of the location depends on how easily the catalyst species are reduced
catalysts were reduced in situ under hydrogen (10 mL min1) for [25]. The peak locations for the three 30NixFeyO/70Al2O3 catalysts
2 h at 700  C before each run. In this study, the amount of steam were almost the same, but the amounts of Ni reduced in them
was adjusted by regulating the temperature according to the decreased in proportion to the amount of Fe added. The NiO
partial pressure law [23]. The flow rate was kept constant at isolated from the 30NiO/70Al2O3 and 30NixFeyO/70Al2O3 samples
10.0 mL/min for propane gas (25.0 vol.%). Argon gas was used to was considered to be reduced to metallic Ni [26] at approximately
carry the vaporized mixture into the reactor. The samples were 380–550  C, and the peaks were separated into two curves,
pre-reduced by H2 (5 vol.%)/Ar gases at 700  C for 2 h before the corresponding to slightly different oxidation states. Another
reaction. The reaction products during PSR were measured by an sharper and larger Ni-reduction curve was exhibited at higher
on-line gas chromatograph (Donam DS6200, Donam Company, temperatures in the range of 700–900  C in both the 30Ni/70Al2O3
Korea) equipped with a thermal conductivity detector (TCD) and and 30NixFeyO/70Al2O3 samples, which corresponds to the
flame ionizing detector (FID). H2, CO, CO2, CH3CHO, CH3COCH3, and reduction of NiO to Ni in the spinel structured NiAl2O4, and the
CH3COOH were detected using the TCD, whereas the FID was used peak location was shifted to a slightly higher temperature in
to detect CH4, C2H4, C2H6, C3H8, and the other products. 30NixFeyO/70Al2O3 compared to that in 30NiO/70Al2O3. This
Additionally, mass spectroscopy (BelMass, Bel Japan Inc., Japan) appears to be due to the influence of the added iron. Furthermore,
was used to confirm the intermediates evolved during PSR. The there are three types of reduction curves at around 350, 500, and
C3H8 conversion and CH4, CO2, CO, and H2 selectivities were 750  C, respectively, on 30FeO/70Al2O3, which correspond to the
defined as: reductions of the Fe oxide in its various oxidation states and, in
particular, the isolated Fe components were reduced at higher
C3H8 conversion (%) = ([C3H8]in  [C3H8]out)/[C3H8]in  100, (2)
temperature compared to the Ni components. The temperature to
reduce the exposed NiO and FeO ingredients were kept at 700  C
before PSR. It is expected that some of the oxidized Ni or Fe
H2 (CO, CO2, or CH4) selectivity (%) = [H2 (CO, CO2, or CH4)]out/
components might remain after PSR, which would affect their
([CO2 + CO + CH4 + H2]out  100. (3)
catalytic activity.

Fig. 1. Equipment configuration of batch bed type reactor for propane steam reforming.
K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336 327

Fig. 2. H2-TPR profiles of the three fresh samples, 30NiO/70Al2O3, 30FeO/7070Al2O3 and 30Ni0.8Fe0.2O/70Al2O3.

The crystallinities of the synthesized samples before (A) and On the other hand, Fig. 3B shows the XRD patterns after the
after (B) hydrogen pre-treatment were examined by XRD, as shown hydrogen pre-treatment. The main XRD peaks for metallic cubic
in Fig. 3. In the hydrogen pre-treated samples (Fig. 3A), the main crystalline Ni [JSPDS no. 01-1258, Fm-3m] after reduction in all of
XRD peaks for the 30NiO/70Al2O3 samples were observed at the samples were observed at 2u = 44.37 (111), 51.60 (200), 76.08
2u = 37.25 (111), 43.28 (200), 63.88 (220), 75.42 (311), 79.41 (220), and 92.09 (311). Small peaks, which were assigned to spinel
(222), and 95.58 (400), which correspond to cubic crystalline NiO structured NiAl2O4, were also observed in all of the reduced
[27]. Special peaks for spinal structured NiAl2O4 [JSPDS no. 01-078- samples, which means that the Ni and Fe components are well
1601, Fd-3m, Cubic] and FeAl2O4 [JSPDS no. 00-003-0894, Fd-3m, combined with the Al2O3 support ingredients. The estimated
Cubic] were observed in all of the 30NixFeyO/70Al2O3 samples, crystallite sizes based on the (111) plane of the formed Ni were
however, there were no peaks for NiO or FeO, which means that the compared, and the values were 20.12, 23.76, 24.92, and 25.82 nm,
Ni and Fe components are very well dispersed over the surface of respectively, in the 30NiO/70Al2O3, 30Ni0.9Fe0.1O/70Al2O3,
Al2O3, causing them to have a perfect spinel structure. Generally, 30Ni0.8Fe0.2O/70 Al2O3, and 30Ni0.7Fe0.3O/70Al2O3 samples. In
the crystallite size decreases with increasing line-broadening of particular, the sizes were increased 2.23 fold compared to those in
the peaks. The crystallite size was estimated using Scherrer’s the 30Ni/70Al2O3 sample before H2-reduction.
equation [28], t = 0.9l/bcos u, where l is the wavelength of the Fig. 4 shows the TEM image and TEM-elemental mapping on the
incident X-rays, b is the full width at half maximum height in reduced 30Ni0.8Fe0.2O/70Al2O3 as a representative sample. The
radians, and u is the diffraction angle in radians. The estimated scale bars of the images are 100 nm. Rectangular-shaped Ni, NiFe
crystallite sizes based on the (111) and (311) planes of the NiO and alloy, and NiAl2O4 particles, approximately 30, 50, and 100 nm in
NiAl2O4 formed in the 30NiO/70Al2O3, 30Ni0.9Fe0.1O/70Al2O3, size, respectively, were observed in the reduced 30Ni0.8Fe0.2O/
30Ni0.8Fe0.2O/70 Al2O3, and 30Ni0.7Fe0.3O/70Al2O3 samples were 70Al2O3 sample, and there was no significant aggregation between
9.04, 7.29, 9.21, and 17.4 nm, respectively. Consequently, the the metal oxides. The particle sizes decreased with increasing
crystallites sizes decreased with increasing amount of Fe added. amount of Fe added, and there was little aggregation between the

Fig. 3. XRD patterns of 30NiO/70Al2O3 and 30Ni0xFeyO/70Al2O3 before and after pre-reduction.
328 K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336

Fig. 4. TEM and elements-mapping images of 30Ni0.8Fe0.2O/70Al2O3 after pre-reduction.

metal oxides. This suggests that the Fe components existed 30Ni0.8Fe0.2O/70Al2O3, and 30Ni0.7Fe0.3O/70Al2O3. The difference
between the Ni particles, preventing any strong agglomeration between the 2p3/2 and 2p1/2 peaks was 17.49 eV in each case.
between them. The elemental mapping showed that all of the The 2p3/2 spin-orbital photoelectron of Ni before PSR was located
components, Ni, Fe, Al, and O, were evenly distributed and their at a binding energy of 853.8–860.0 eV, which was assigned to
concentrations were reliable as expected. NiO in NiAl2O4 for all of the samples [29]. On the other hand,
Fig. 5 displays the survey spectra derived from the quantitative the curves were slightly changed to a lower binding energy in the
XPS analysis of the Ni2p, Fe2p, Al2p, and O1s peaks for all of the three samples, 30Ni0.9Fe0.1O/70Al2O3, 30Ni0.8Fe0.2O/70Al2O3, and
reduced samples, viz. 30NiO/70Al2O3, 30Ni0.9Fe0.1O/70Al2O3, 30Ni0.7Fe0.3O/70Al2O3, corresponding to lower oxidation states of

Fig. 5. Survey spectra from quantitative XPS of the Ni2p, Fe2p, Al2p, and O1s peaks for the 30NiO/70Al2O3 and 30Ni0xFeyO/70Al2O3 catalysts after pre-reduction.
K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336 329

the Ni components, even though the area of the peaks was smaller. Fig. 6 shows the adsorption–desorption isotherm curves of N2
From this result, it was deduced that the reduction potentials of Ni2 at 77 K for the reduced samples, 30NiO/70Al2O3, 30Ni0.9Fe0.1O/
+
! Ni0 and Fe2+ ! Fe0 were 0.25 and 0.44 eV, respectively, and, 70Al2O3, 30Ni0.8Fe0.2O/70Al2O3, and 30Ni0.7Fe0.3O/70Al2O3. Using
thus, that the reduction of Ni is performed more rapidly than that Kelvin’s equation [31], the radius of the pores, in which capillary
of Fe in the 30NixFeyO/70Al2O3 samples, so the added Fe condensation occurs actively, can be determined as a function of
ingredients should facilitate the reduction of Ni oxide. With the relative pressure (P/P0). The mean pore diameter, Dp, was
increasing concentration of Fe added, the peak intensities calculated from Dp = 4VT/S, where VT is the total volume of the
according to Fe oxide increased. During PSR, Fe oxides can strongly pores and S is the BET surface area. According to the IUPAC
attract propane molecules, and Ni oxides can act as an oxygen classification, all of the isotherms belonged to type IV [32]. The
donor to the thermally cracked CH2CH4 or CH4 molecules from hysteresis slopes were observed at higher relative pressures in all
propane in the formation of CH3CHO and CO as intermediates. In of the samples, indicating the presence of bulk mesopores formed
contrast, the spin-orbital spectra of Fe2p3/2 and Fe2p1/2 for FeO in between the particles. The BET specific surface areas were
FeAl2O4 over the 30NixFeyO/70Al2O3 samples revealed broad peaks significantly higher in the 30NixFeyO/70Al2O3 samples (94.5–
at 712.0 and 725.6 eV, respectively [30], and the difference 100.9 m2 g1) than in the 30NiO/70Al2O3 one (56.67 m2 g1). In
between the two orbitals, HOMO and LUMO, was 13.6 eV. The general, the specific surface areas in regular particles are strongly
peak locations were not changed, but the intensities were more related to the particle sizes: the lower the particle size, the higher
sharply dependent on the amount of Fe added. The Al2p and O1s the surface area. These results suggest that the particle sizes were
spin-orbital photoelectrons were located at binding energies of smaller in the 30NixFeyO/70Al2O3 samples due to the addition of
76.4 and 533.0 eV, respectively, which were assigned to Al and O in Fe. The total pore volumes increased with increasing amount of Fe
Al2O3 or NiAl2O4 for all of the samples. The curves were shifted to a added, showing the same trend in all of the samples. The pore size
very slightly lower binding energy in the 30NixFeyO/70Al2O3 distribution (PDS) is an important characteristic for porous
samples compared to 30NiO/70Al2O3. Additionally, the atomic materials. Among these methods, the BJH (Barrett–Joyner–
compositions for the O, Al, Ni, and Fe components obtained by XPS Halenda) plot provides a suitable method for the measurement
analysis (%) after reduction are also compared in the right-hand of micro- or meso-pores [33]. Through the BJH plot, the pore size
table. The atomic ratios for Ni/Fe were 3.26, 2.63, and 1.59 in distributions and average pore diameters in the 30NixFeyO/
30Ni0.9Fe0.1O/70Al2O3, 30Ni0.8Fe0.2O/70Al2O3, and 30Ni0.7Fe0.3O/ 70Al2O3 samples were found to be slightly sharper and smaller,
70Al2O3, respectively, which differ from the initial amounts used in depending on the amount of Fe added, than those in 30NiO/
the synthesis step. The amounts of Ni and Fe present on the surface 70Al2O3.
were not affected greatly by the amount inserted during synthesis,
which also indicates their high dispersion. In addition, the accurate Gas adsorption abilities on the 30NixFeyO/70Al2O3 catalysts
quantitative analysis of a surface by XPS analysis alone is difficult.
However, the atomic amount of Fe increased proportionally in the The reactivity of the catalyst is strongly associated with its
order of 30Ni0.9Fe0.1O/70Al2O3, 30Ni0.8Fe0.2O/70Al2O3, and adsorption capacity for the reactant or intermediate gases. In
30Ni0.7Fe0.3O/70Al2O3, thus these values seem to be somewhat general, a catalyst having a strong adsorption capacity for these
consistent. gases exhibits superior catalytic activity. Chemisorption is a kind of

Fig. 6. N2-adsorption/desorption isotherm curves (A) and pore size distributions determined from the BJH (Barrett–Joyner–Halenda) plots (B) of the 30NiO/70Al2O3 and
30Ni0xFeyO/70Al2O3 catalysts after pre-reduction.
330 K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336

adsorption which involves a chemical reaction between the surface


and the adsorbent: new chemical bonds are generated at the
adsorbent surface. An important aspect of chemisorption is in
heterogeneous catalysis, which involves molecules reacting with
each other via the formation of chemisorbed intermediates [34].
After the chemisorbed species combine, the product desorbs from
the surface. This study attempted the thermal programmed
chemisorption for propane gas and steam as the reactant feeds
and CO which is generated in the final stage. In the first step, tests
of C3H8-chemisorption were performed and the desorption curves
for chemisorbed C3H8 are exhibited in Fig. 7A. Additionally, the
evolution of the mass spectrometer signals (m/z) recorded during
propane desorption with O2 valence carriers over 30NixFeyO/
70Al2O3 is shown in Fig. 7B. Propane chemisorption experiments
were conducted in order to investigate the interaction of the
molecular species with the surface oxygen in the catalytic surfaces
derived from the 30NixFey/70Al2O3 samples. The C3H8 desorption
profiles were characterized over one broad temperature range, Fig. 8. CO-TPD profiles of the pre-reduced 30NiO/70Al2O3 and 30Ni0xFeyO/70Al2O3
400–600  C, corresponding to C3H8 desorption on the metal catalysts.
species. The temperature locations were almost the same for all
of the samples, but the areas of the curves were significantly larger similar to the intermediates and products produced during the
in the case of 30NixFeyO/70Al2O3. In particular, the desorbed area reforming reaction for propane.
was the largest in the case of 30Ni0.8Fe0.2O/70Al2O3, where it was CO-TPD experiments were performed for all of the samples, as
2.8 times larger than that for 30Ni/70Al2O3, which means that shown in Fig. 8. The majority of the adsorbed CO desorbs as CO2 at
considerably more C3H8 molecules were adsorbed on the surface of high temperatures, but some molecular desorption of CO occurs at
the bimetallic NiFeO surfaces than on the Ni surface. Otherwise, temperatures as low as 50–250  C and, indeed, CO desorption
the desorption curves were separated into three curves, and on the begins immediately upon the initiation of the temperature ramp
basis of mass spectroscopy, assigned to H2, CH4, CH3CHO, CO, and [36]. In this study, the CO desorption profiles were characterized
CO2 desorptions [35]. Fig. 7B presents the mass spectra for the over two broad temperature ranges in 30NiO/70Al2O3 centered at
desorbed molecules in O2 valence for the three desorption 350 and 600  C, corresponding to CO and CO2 desorption on the
temperatures, 500, 750, and 850  C. Signals at m/z = 2 (H2), 16 metal species, respectively. The temperatures decreased and the
(CH4), 28 (CO), 43 (CH3CHO), and 44 (CO2) were exhibited over the curve intensities increased significantly in the 30NixFeyO/70Al2O3
30Ni0.8Fe0.2O/70Al2O3 sample. The mass spectroscopy signals did samples, which mean that considerably more CO molecules were
not vary with the desorption temperature in any of the segmental adsorbed on the surface of the metallic NiFe than on the mono Ni
compounds, except for CO2. These segmental compounds are component. Additionally the adsorbed area was the largest in the

Fig. 7. C3H8-TPD profiles (A) and mass spectra (B) on the pre-reduced 30NiO/70Al2O3 and 30Ni0xFeyO/70Al2O3 catalysts.
K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336 331

Fig. 9. Correlation graphs of ln(C/Tm2) versus Tm deduced by Kissinger equation for the pre-reduced 30NiO/70Al2O3 and 30Ni0.8Fe0.2O/70Al2O3 catalysts.

30Ni0.8Fe0.2O/70Al2O3 sample. In general, a rapid catalytic reaction reaction temperature, both the propane conversion and hydrogen
occurs when many reactants are well-adsorbed over the catalyst. selectivity increased in the two samples, and the propane
On the other hand, water (H2O) molecules constitute the main conversion was 100% above 650  C in both samples. However,
gas in hydrocarbon steam reforming. Water molecules are involved the rate of increase of the hydrogen selectivity was greater in the
in the water gas shift reaction to convert CO to CO2. The H2O- 30Ni0.8Fe0.2O/70Al2O3 catalyst than that in 30NiO/70Al2O3 and the
desorption experiments for two representative samples, 30NiO/ increase was particularly marked at 700  C in both samples. At
70Al2O3 and 30Ni0.8Fe0.2/70Al2O3, were performed using a TG 700  C, the hydrogen selectivity amounted to 86% in the
instrument, and their activation energies were calculated from the 30Ni0.8Fe0.2O/70Al2O3 catalyst, whereas it was only 79% in
slopes, which were obtained from the plots [37] of ln(C/Tm2) 30NiO/70Al2O3 at the same temperature. Thus, on the basis of
versus 1/Tm, as shown in Fig. 9. The activation energy was larger in this result, the optimum temperature was fixed at 700  C in this
30Ni0.8Fe0.2O/70Al2O3 than in 30NiO/70Al2O3: the difference in the study.
activation energies reached 2.78 fold. Additionally, the water Fig. 11 exhibits the product distributions and the absolute
desorption rate was higher in the 30Ni0.8Fe0.2O/70Al2O3 sample. values for all of the samples evolved during propane steam
This means that the water molecules were absorbed better on the reforming for 1 h at 700  C for a GHSV of 6000 h1. Here, the
surface of the catalyst having the larger desorption activation propane conversions were close to 100% in all of the samples,
energy. It is expected that the steam added to the propane feed gas however the absolute amounts of hydrogen generated differed
will adsorb on the surface of the Fe component rather than that of depending on the catalyst (Fig. 11A); the absolute value exceeded
Ni during the propane steam reforming reaction. 80 mL in the 30Ni0.8Fe0.2O/70Al2O3 catalyst, which is more than the
value of 20 mL emitted in the 30NiO/70Al2O3 one. Regarding the
PSR performances over the 30NixFeyO/70Al2O3 catalysts product distributions during PSR, there are only three molecules in
the 30NiO/70Al2O3 catalyst, viz. H2, CH4, and CO, as the main
Fig. 10 shows the propane conversion and hydrogen selectivity products, whereas more CO2 gas was emitted in the case of the
during propane steam reforming on the two representative 30NixFeyO/70Al2O3 catalysts. This is evidence that the water gas
samples of 30NiO/70Al2O3 and 30Ni0.8Fe0.2O/70Al2O3, according shift reaction proceeds readily during propane steam reforming
to the reaction temperature in the range from 500 to 800  C with an over the 30NixFeyO/70Al2O3 catalysts. However, no intermediates,
interval of 50  C for 1 h at a GHSV 6000 h1. Depending on the except for acetaldehyde and ethylene molecules, were observed in

Fig. 10. The propane conversion and hydrogen selectivity on 30NiO/70Al2O3 and 30Ni0.8Fe0.2O/70Al2O3 catalysts according to the reaction temperatures.
332 K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336

Fig. 11. Product distributions (A) for all of the catalysts, and gas chromatography (B) and mass spectroscopy (C) results for the 30Ni0.8Fe0.2O/70Al2O3 catalyst during PSR.

the GC signal during PSR at 700  C, possibly because the reaction CO + *Ox ! CO2 + *Ox1, (9)
temperature was too high. Thus, the reaction temperature was
reduced to 450  C in an attempt to identify any intermediates. As a
H2O + *Ox1 ! H2 + *Ox (10)
result, a small amount of ethylene (C2H4) molecules was seen in
the GC signal during PSR at 450  C, as shown in Fig. 11B. In order to Fig. 12 exhibits the C3H8 conversion and hydrogen selectivity
observe all of the intermediates evolved during PSR, mass during propane steam reforming for all of the catalysts, 30NiO/
spectroscopy was used and the resultant signals are shown in 70Al2O3, 30Ni0.9Fe0.1O/70Al2O3, 30Ni0.8Fe0.2O/70Al2O3, and
Fig. 11C: the signals at m/z = 14, 26, 28, and 29 were assigned to 30Ni0.7Fe0.3O/70Al2O3, at a time on stream of 10 h at 700  C for a
CH2, C2H2, C2H4, and CH3CHO, respectively. The mechanism of CH4 GHSV of 6000 h1. The propane was perfectly converted to the
steam reforming has been extensively studied [38], however, less products with 100% conversion and these perfect conversions were
research has focused on the propane and higher hydrocarbon maintained for 10 h in all of the samples. The hydrogen production
steam reforming mechanisms [39]. In the present work, the solid– was the highest over the 30Ni0.8Fe0.2O/70Al2O3 sample, reaching
gas mechanism involves the reactions between hydrocarbons 89% after 10 h, whereas it was 67% in the case of 30NiO/70Al2O3.
(C3H8, CH4, and C2H4) and/or an intermediate surface hydrocarbon Generally, the presence of CO degrades the active catalyst, due to
species with lattice oxygen at the NiFeO surface. Based on this catalyst poisoning by CO molecules [40]. The CO evolution was the
result, the mechanism in PSR was expected to be as follows: The smallest in the 30Ni0.8Fe0.2O/70Al2O3 sample. On the other hand,
main reaction is accompanied by cracking to produce methane and the hydrogen selectivity decreased significantly in the
ethylene on the surface of metallic Ni in the catalysts (*) with the 30Ni0.7Fe0.3O/70Al2O3 sample. This is possibly because the
possible hydrogenation of the latter. The cracking reaction is increase in the amount of Fe led to the acceleration of the
C3H8 + * ! C2H4 + CH4. Here, C2H6 produced by the hydrogenation Fisher–Tropsch reaction (CO + H2 ! CH4), leading to the reverse
of C2H4 was not detected in this study. The production of CH4 as an methane reforming reaction. Therefore, there is an optimal
intermediate was favored under our conditions and various amount of Fe of 0.2 mol. Based on the product distribution, the
reactions proceeded on the surface of the catalysts: the introduction of an appropriate amount of Fe into the 30Ni/70Al2O3
C2H4formed was converted to CH3CHO on the surface of the catalyst during PSR has a favorable effect on the catalytic
oxidized metal in the catalyst and then converted to CH4 and CO. performance, due to the enhancement of the water gas shift
After that, the CH4 was converted to H2 and CO. Finally, the steam reaction, resulting in less catalytic deactivation. Therefore, these
(H2O) enables the reduced metal species in the catalysts to recover results highlight the synergic effect of Fe and Ni on the catalytic
by delivering oxygen with the production of H2 (water gas shift performance in PSR.
reaction).
Characteristics of the catalysts after PSR
C2H4 + *Ox ! CH3CHO + *Ox1, (4)
The XRD patterns of the used 30NixFeyO/70Al2O3 catalysts were
CH3CHO + * ! CH4 + CO*, (5) analyzed, in order to determine the structural changes in the metal
species of the catalysts after PSR, as shown in Fig. 13. The
CH4 + *Ox ! 2H2 + C*Ox, (6) diffraction lines for all of the used catalysts were similar to those of
the fresh ones (the pre-reduced catalysts). Metallic Ni species after
C*Ox ! CO + *Ox1, (7) PSR were seen in all of the catalysts, and the intensities were the
strongest in the 30NiO/70Al2O3 sample, however the peaks were
CO* ! CO + *, (8) smaller and the (111) planes were shifted to lower angles in the
K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336 333

Fig. 12. Propane conversions and hydrogen productions versus reaction times for the 30NiO/70Al2O3 and 30Ni0xFeyO/70Al2O3 catalysts.

30NixFeyO/70Al2O3 samples. Using Scherrer’s equation, the crys- did not change compared to those in the fresh catalysts, but the
tallite sizes estimated based on the (111) plane of the Ni formed in intensities were greatly decreased in the 30NiO/70Al2O3 sample.
the 30NiO/70Al2O3, 30Ni0.9Fe0.1O/70Al2O3, 30Ni0.8Fe0.2O/70Al2O3, This trend was also observed for the Fe2p3/2 spin-orbitals, and the
and 30Ni0.7Fe0.3O/70Al2O3 samples were 33.70, 27.86, 26.14, and peak decreased significantly when there were no Fe ions in the
26.21 nm, respectively, which provide proof that the Ni particles catalysts. In particular, the peak intensity remains the strongest in
after the reaction were strongly aggregated in the 30NiO/70Al2O3 the 30Ni0.8Fe0.2O/70Al2O3 sample. Eventually, the reduced Ni was
sample. On the other hand, the peaks which were assigned to re-oxidized by the Fe oxide during PSR, resulting in the formation
NiAl2O4 became somewhat bigger in the case of 30NixFeyO/ of CH3CHO and CO as intermediates. The observation of peak
70Al2O3, which means that the NiO components remained in the separation in the Al2p spectra in the 30NiO/70Al2O3 sample is quite
30NixFeyO/70Al2O3 samples. unusual: the amount of aluminum oxide with reduced states at
Fig. 14 shows the survey spectra derived from the quantitative around 75.0 eV was lower, but a new peak was observed at 78 eV
XPS spectra of the Ni2p, Fe2p, Al2p, and O1s peaks for the used with a more highly oxidized state. This is considered to be due to
catalyst after PSR. The peak locations for the 2p3/2 spin-orbital the conversion of the isolated aluminum from the Ni components
photoelectron, which was assigned to Ni2+ in all of the catalysts, to Al(OH)3 during PSR [41]. This trend is the same in the O1s
spectra. It likely corresponds to more highly oxidized oxygen;
probably the O of the OH in Al(OH)3.
To determine the amounts and shapes of the carbon deposited
on each catalyst, the amounts of carbon deposited were measured
from the C1s-XPS analysis, as shown in Fig. 15. There is a peak at
286 eV which is assigned to carbon deposited on the surface of the
catalysts after PSR. It can be assumed to correspond to three main
kinds of carbon species, viz. C–O, C–C or C–H, as well as a small
amount of CNTs [42,43]. The peak locations were almost the same,
except for the 30Ni0.8Fe0.2O/70Al2O3 sample, where they were
shifted to a higher binding energy, corresponding to a carbon
component in a more highly oxidized state. It is estimated that the
carbon deposited on the catalysts was in the form of lumps of
carbon, and a very small amount of CNTs was produced. The peak
area, which corresponds to the amount of carbon deposited, was
the largest in the used 30NiO/70Al2O3 sample and the smallest in
the 30Ni0.8Fe0.2O/70Al2O3 sample, indicating that the level of
catalytic degradation caused by coke formation was reduced over
the latter. This shows that the carbon did not grow over the Fe sites,
Fig. 13. XRD patterns of the used 30NiO/70Al2O3 and 30Ni0xFeyO/70Al2O3 catalysts suggesting that water could adsorb easily over them, indicating
after PSR.
334 K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336

Fig. 14. Survey spectra from quantitative XPS of the Ni2p, Fe2p, Al2p, and O1s peaks for the 30NiO/70Al2O3 and 30Ni0xFeyO/70Al2O3 catalysts after PSR.

that Fe plays a role in changing CO to CO2, promoting the CO–water gas shift reaction over Ni and FeO, respectively. Their
production of hydrogen with less catalytic deterioration. gas adsorption abilities were compared through H2O-TPD experi-
On the basis of the PSR performances and the physicochemical ments (Fig. 9). On the basis of the CO- and H2O-TPD experiments,
analysis on the catalysts, in particular the mass spectroscopy it can be inferred that water molecules are well adsorbed on the
results, Scheme 1 presents the proposed PSR mechanism over Fe components and, thus, the CO-water gas shift reaction takes
the30NixFeyO/70Al2O3 catalysts. During propane reforming, the place more predominantly over 30NixFeyO/70Al2O3 than over
metallic Ni component plays an important role in the dehydro- 30NiO/70Al2O3. Finally, the CO molecules obtained by CH4-SR
genation and thermal cracking resulting from the propane gases. should be converted to CO2 and H2 via a secondary CO-WGS
These gases are transformed into C2H4 and CH4 as intermediates reaction. On the other hand, the amount of CO2 molecules evolved
and acetaldehyde (CH3CHO) is generated by ethylene oxidation over the 30NiO/70Al2O3 sample was too small and, consequently,
over NiFeO. Then, acetaldehyde is thermally cracked catalytically CO molecules are deposited as carbon lumps on the 30NiO/
to produce CH4 and CO. The CH4 and CO obtained participate in 70Al2O3 catalyst surface, resulting in catalytic deactivation.
other reactions, such as the CH4-stem reforming reaction and the Contrary to our expectations, however, no deterioration of the

Fig. 15. Survey spectra from quantitative XPS of the C1s peaks for the 30NiO/70Al2O3 and 30Ni0xFeyO/70Al2O3 catalysts after PSR.
K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336 335

Scheme 1. The expected model for propane steam reforming over the 30Ni0xFeyO/70Al2O3 catalyst.

catalyst occurred. In this experiment, the addition of Fe oxide Planning (No. 2015R1A1A3A04001268), for which the authors are
contributed significantly to preventing sintering between the very grateful.
inter-Ni particles. Therefore, Fe oxide as a promoter helps produce
a high hydrogen yield in the PSR reaction while inducing the References
emission of CO2.
[1] S. Mohsen Mousavi Ehteshami, S. Vignesh, R.K.A. Rasheed, S.H. Chan, Appl.
Energy 170 (2016) 388.
Conclusions [2] Y. Wang, S. Wang, G. Zhao, Y. Guo, Y. Guo, Int. J. Hydrogen Energy 41 (2016)
2238.
The goal of this study was to improve the catalytic stability of [3] J.Y.Z. Chiou, C.L. Lai, S.-W. Yu, H.-H. Huang, C.-L. Chuang, C.-B. Wang, Int. J.
Hydrogen Energy 39 (2014) 20689.
the 30NiO/70Al2O3 catalyst through the introduction of Fe as a sub- [4] D.H. Choi, S.M. Chun, S.H. Ma, Y.C. Hong, J. Ind. Eng. Chem. 34 (2016) 286.
catalytic species. Three types of 30NixFeyO/70Al2O3 catalysts were [5] X. Huang, R. Reimert, Fuel 106 (2013) 380.
prepared and there existed an optimum Ni/Fe ratio, which was [6] Y. Li, X. Wang, C. Song, Catal. Today 263 (2016) 22.
[7] A. Mosayebi, R. Abedini, J. Ind. Eng. Chem. 20 (2014) 1542.
Ni0.8:Fe0.2. The hydrogen yield in PSR was the highest on the [8] A. Hossein, H. Mohammad, Energy Convers. Manag. 118 (2016) 231.
30Ni0.8Fe0.2O/70Al2O3 catalyst, reaching 83% with 100% propane [9] F. Mahdi, A. Mohsen, R. Ali, R.R. Mohammad, J. Nat. Gas Sci. Eng. 14 (2013) 158.
conversion. In addition, more C3H8 and CO gases adsorbed over the [10] B.S. Kwak, K.M. Kim, S.W. Jo, J.Y. Do, S. Kang, M. Park, N.-K. Park, T.J. Lee, S.T. Lee,
M. Kang, J. Ind. Eng. Chem. 37 (2016) 57.
30Ni0.8Fe0.2O/70Al2O3 catalyst, resulting in the formation of a
[11] G. Saioa, E. Heike, L.A. Pedro, K. Norbert, J. Power Source 279 (2015) 312.
greater amount of CO2 gas. During PSR, C2H4, CH4, and CH3CHO [12] T. Ahmad, B. Mohammad, K. Ali, Biomass Bioenergy 80 (2015) 63.
molecules were identified as intermediates. A trace amount of [13] S. Kang, B.S. Kwak, M. Kang, Ceram. Int. 40 (2014) 14197.
carbon lumps was deposited over the 30Ni0.8Fe0.2/70Al2O3 catalyst [14] M. Surendar, T.V. Sagar, G. Raveendra, M. Ashwani Kumar, N. Lingaiah, K.S.
Rama Rao, P.S. Sai Prasad, Int. J. Hydrogen Energy 41 (2016) 2285.
after the PSR reaction. Overall, this study proved that the [15] O.-V. Paula, H.C. Cristian, M.N. Rufino, L.G.F. Jose, R. Patricio, Appl. Catal. A: Gen.
introduction of Fe along with Ni in PSR has a favorable effect on 505 (2015) 159.
the stable production of hydrogen gas, with significantly less [16] H. Iida, N. Onuki, T. Numa, A. Igarashi, Fuel Process. Technol. 142 (2016) 397.
[17] N. Laosiripojana, W. Sangtongkitcharoen, S. Assabumrungrat, Fuel 85 (2006)
catalytic deactivation due to catalytic poisoning by CO molecules, 323.
since these are transformed to CO2 through the water gas shift [18] F.M. Althenayan, S.Y. Foo, E.M. Kennedy, B.Z. Dlugogorski, A.A. Adesina, Chem.
reaction. Eng. Sci. 65 (2010) 66.
[19] Z.O. Malaibari, E. Croiset, A. Amin, W. Epling, Appl. Catal. A: Gen. 490 (2015) 80.
[20] G. Lee, D. Kim, B.S. Kwak, M. Kang, Catal. Today 232 (2014) 139.
Acknowledgments [21] R.L. Blaine, H.E. Kissinger, Thermochim. Acta 540 (2012) 1.
[22] B.S. Kwak, G. Lee, S.-M. Park, M. Kang, Appl. Catal. A: Gen. 503 (2015) 165.
[23] S.W. Jo, B.S. Kwak, K.M. Kim, J.Y. Do, N.-K. Park, T.J. Lee, S.-T. Lee, M. Kang, Chem.
This work was supported by the Human Resource Training
Eng. J. 288 (2016) 858.
Program for Regional Innovation and Creativity through the [24] H. Harjua, J. Lehtonena, L. Lefferts, Appl. Catal. B: Environ. 182 (2016) 33.
Ministry of Education and National Research Foundation of Korea [25] H. Zhang, J. Wang, Y. Zhang, Y. Jiao, C. Ren, M. Gong, Y. Chen, Appl. Surf. Sci. 377
(2016) 48.
(NRF-2015H1C1A1035639), and by the Basic Research Science and
[26] J. Ashok, Y. Kathiraser, M.L. Ang, S. Kawi, Appl. Catal. B: Environ. 172–173 (2015)
Technology Projects through the National Research Foundation of 116.
Korea Grant funded by the Ministry of Science, ICT & Future
336 K.M. Kim et al. / Journal of Industrial and Engineering Chemistry 46 (2017) 324–336

[27] A. Qurashi, Z. Zhang, M. Asif, T. Yamazaki, Int. J. Hydrogen Energy 40 (2015) [36] Y. Zeng, H. Ma, H. Zhang, W. Ying, D. Fang, Fuel 137 (2014) 155.
15801. [37] J. Orava, A.L. Greer, Thermochim. Acta 603 (2015) 63.
[28] M. Ramzan Parra, F.Z. Haque, J. Mater. Res. Technol. 3 (2014) 363. [38] V. Kyriakou, I. Garagounis, A. Vourros, E. Vasileiou, A. Manerbino, W.G. Coors,
[29] N.F.M. Salleh, A.A. Jalila, S. Triwahyono, J. Efendi, R.R. Mukti, B.H. Hameed, Appl. M. Stoukides, Appl. Catal. B: Environ. 186 (2016) 1.
Surf. Sci. 349 (2015) 485. [39] K. Lee, E. Lee, C. Song, J.J. Michael, J. Catal. 309 (2014) 248.
[30] H. Zhou, Y. Su, W. Liao, W. Deng, F. Zhong, Appl. Catal. A: Gen. 505 (2015) 402. [40] M.V. Gil, J. Fermoso, C. Pevida, D. Chen, F. Rubiera, Appl. Catal. B: Environ. 184
[31] D.K. Panesar, J. Francis, Const. Build. Mater. 52 (2014) 52. (2016) 64.
[32] H.M. Jennings, A. Kumar, G. Sant, Cem. Concr. Res. 76 (2015) 27. [41] G. Garbarino, C. Wang, I. Valsamakis, S. Chitsazan, P. Riani, E. Finocchio, M.
[33] M. Khajenoori, M. Rezaei, F. Meshkani, J. Ind. Eng. Chem. 21 (2015) 717. Flytzani-Stephanopoulos, G. Busca, Appl. Catal. B: Environ. 174–175 (2015) 21.
[34] M. Pori, B. Likozar, M. Marinšek, Z. Crnjak Orel, Fuel Process. Technol. 146 [42] C. Montero, A. Ochoa, P. Castaño, J. Bilbao, A.G. Gayubo, J. Catal. 331 (2015) 181.
(2016) 39. [43] W. Wohlleben, M.W. Meier, S. Vogel, R. Landsiedel, G. Cox, S. Hirth, Ž. Tomovic’,
[35] M.M. Zyryanova, P.V. Snytnikov, A.B. Shigarov, V.D. Belyaev, V.A. Kirillov, V.A. Nonoscale 5 (2013) 369.
Sobyanin, Fuel 135 (2014) 76.

View publication stats

You might also like