You are on page 1of 6

Catalysis Today 303 (2018) 100–105

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Effect of acidity on Ni catalysts supported on P-modified Al2O3 for dry T


reforming of methane

Seonu Bang, Eunpyo Hong, Sung Woo Baek, Chae-Ho Shin
Department of Chemical Engineering, Chungbuk National University, Chungbuk 28644, Republic of Korea

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

Keywords: Phosphorus-modified Al2O3 supports containing different P content were prepared to control the acidity of Ni-
Dry reforming of methane impregnated catalysts for application in the dry reforming of methane (DRM). Results from the H2-temperature-
Nickel catalysts programmed reduction showed that an increase in P content promoted the formation of an AlPO4 phase, and
Phosphorus resulted in an increase in Ni particle size. Although the initial conversion of CH4 and CO2 decreased with
Carbon deposition
increasing P content, the addition of P influenced on the acidic properties of the support and the catalyst. Indeed,
Acid property
Support modification
a decrease in catalytic deactivation correlated with the decreasing acidity of the P-modified supports and the Ni/
P-Al2O3 catalysts, and the acidic properties were determined by the behaviors of the temperature-programmed
desorption of iso-propanol (IPA) and the dehydration of IPA. We found that the Ni/Al2O3 catalyst modified by
2 wt.% P exhibited the weakest acidity whilst demonstrating excellent coke resistance and stability over 100 h in
the DRM reaction.

1. Introduction by the reversed Boudouard reaction [3].


Alumina-supported metal catalysts are commonly employed for the
The dry reforming of methane (DRM) has been extensively studied DRM reaction, and various attempts have been conducted to suppress
as a method for producing synthesis gas from CO2 and CH4, as shown in carbon deposition. For example, alumina-supported noble metal cata-
Eq. (1). lysts effectively suppress carbon deposition [4,5]; however, they are
expensive. As an alternative, to minimize carbon deposition on Ni-
CO2 + CH4 → 2H2 + 2CO (△H°298 = 247.4 kJ/mol) (1)
based catalysts, the addition of alkali metals [6] and alkaline earth
As the products produced via the DRM have a lower H2/CO ratio metals [7], and the modification of supports such as CeO2-ZrO2,
than those of other processes such as steam reforming of methane and BaTiO3, and SiO2 [8,9] have been investigated. Amongst the various
partial oxidation, the DRM is suitable for use in Fischer-Tropsch characteristics of these modified catalysts, the acid-base functionalities
syntheses. Thus, as the DRM can reduce the quantity of undesirable on the catalyst surface influence on the deactivation phenomenon [10].
greenhouse gases while producing valuable chemical products, it could Indeed, the addition of basic oxides such as K2O and CaO can improve
be a promising method from both environmental and industrial aspects the stability of Ni-based catalysts by modifying the acidic functional-
[1]. ities [11].
Despite its advantages, catalyst deactivation by carbon deposition is We herein report the modification of alumina-supported Ni catalysts
a significant issue for the practical application of the DRM reaction. by the addition of phosphorus via a sequential impregnation method
Such carbon deposition is mainly caused by methane decomposition using P and Ni precursors on boehmite support. The effects of varying
and by the Boudouard reaction, as indicated in Eqs. (2) and (3), re- the P content on the DRM reaction will then be examined. In previously
spectively [2]. reported papers, the addition of P to alumina inhibits alumina phase
transition from γ to α phase, thereby improving thermal stability [12]
CH4 → C + 2H2 (△H°298 = 75.0 kJ/mol) (2) and preventing the reduction of specific surface area at high tempera-
2CO → C + CO2 (△H°298 = −172.0 kJ/mol) (3) tures [13]. In addition, it alters the acidity of the support by modifying
chemical properties of alumina surface [14,15]. Although the changes
As such, catalyst deactivation takes place unless the deposition of thermal stability and acid-base properties by the addition of P have
mechanism is inhibited or the deposited carbon is effectively removed been reported, only a few examples exist regarding application to the


Corresponding author.
E-mail address: chshin@chungbuk.ac.kr (C.-H. Shin).

http://dx.doi.org/10.1016/j.cattod.2017.08.013
Received 24 May 2017; Received in revised form 2 August 2017; Accepted 7 August 2017
Available online 10 August 2017
0920-5861/ © 2017 Elsevier B.V. All rights reserved.
S. Bang et al. Catalysis Today 303 (2018) 100–105

DRM reaction. We therefore attempt to clarify the effect of the P con- Fi Ci


i Component selectivity(%)= × 100%
tent on the catalyst characteristics and on coke resistance in the DRM ΣFi Ci (5)
reaction.
where, FIPA and Fi are the molar flow rates of IPA and the products i,
which are propylene, di-isopropyl ether (DIPE), and acetone, and Ci is
2. Experimental the carbon number of each product, i.
To analyze the deposited carbon present following the DRM reac-
2.1. Catalyst preparation tion, thermogravimetric analysis (TGA), temperature-programmed
oxidation (TPO), and transmission electron microscopy (TEM) were
Boehmite (Catapal B, Sasol) was modified with H3PO4 (85%, Junsei) carried out. The quantity of deposited carbon was determined using a
using an incipient wetness method. Following impregnation, the re- Mettler-Toledo TGA/DSC1 instrument under 5% O2/Ar
sulting powder was dried at 60 °C overnight, then calcined at 850 °C for (100 cm3 min−1) with a heating rate of 10 °C min−1. The TPO experi-
10 h under a flow of air. The P-modified supports will be referred to as ments were conducted for the used catalysts using a QMS to determine
xPAl, where x indicates the weight percentage of P (i.e., 0.5, 1, 2, 4, and on the nature of the carbonaceous species. The fragment of %CO2 (m/
6 wt.%), while the P-free support will be referred to as a Al2O3. z = 44) was measured at 35–850 °C with a heating rate of 10 °C min−1
Catalyst preparation was conducted via an incipient wetness under 5% O2/Ar (100 cm3 min−1). TEM images of the used catalysts
method on the P-modified and pure Al2O3 supports using an Ni were obtained using a JEM-2100F instrument.
(NO3)2·6H2O (98%, Samchun) precursor. Following impregnation, the
Ni-impregnated catalysts were dried at 60 °C overnight, then calcined at
2.3. Dry reforming of methane (DRM)
800 °C for 3 h under a flow of air. The calcined catalysts will be referred
to as xPAl-10Ni and 10Ni/Al2O3, where the Ni loading was fixed at
The catalytic DRM reactions were conducted at atmospheric pres-
10 wt.% (Ni/(Ni + P-Al2O3)).
sure in a fixed bed quartz reactor. The catalyst (0.1 g) was loaded into
the reactor using quartz wool. Prior to testing, the catalyst was reduced
2.2. Characterization under H2 (50 cm3 min−1) at 750 °C for 1 h, then purged with N2
(50 cm3 min−1) at 750 °C for 20 min. The feed gas stream consisted of
The specific surface area, total pore volume, average pore size, and CH4 (40 vol.%) and CO2 (40 vol.%) diluted with N2, and the total flow
pore size distribution were analyzed by N2-physisorption (ASAP 2020, rate was maintained constant at 100 cm3 min−1. Long-term stability
Micromeritics). Prior to analysis, the adsorbed impurities on the cata- tests were performed at 750 °C for 20 and 100 h under identical con-
lyst were removed at 200 °C for 4 h. The specific surface area was then ditions. Reaction products were analyzed using a Varian CP-3800 GC
calculated at P/P0 = 0.05–0.20 using the Brunauer–Emmett–Teller equipped with a packed Carbosphere column (1/8" × 3 m). The con-
(BET) equation, the total pore volume was measured at P/P0 = 0.995, versions and the H2/CO ratios were calculated using Eqs. (6)–(8):
and the pore size distribution was calculated using the Barrett-Joyner-
Halenda method from the desorption branch of the N2-isotherm. X-ray (CH 4)in − (CH 4)out
CH 4 conversion(%) = × 100%
diffraction (XRD) patterns were obtained using a Bruker X-ray dif- (CH 4)in (6)
fractometer with Cu-Kα radiation in the 2θ range of 10–80° (scan
(CO2)in − (CO2)out
rate = 4° min−1). Prior to analysis of the reduced catalysts, it was CO2 conversion(%) = × 100%
passivated at room temperature under 1% O2/Ar (50 cm3 min−1) for (CO2)in (7)
0.5 h to prevent reoxidation in the atmosphere.
(H2)out (CO)out
The H2-temperature-programmed reduction (H2-TPR) profiles of the H2 : CO ratio = /
2(CH 4)in (CH 4)in+(CO2)in (8)
calcined catalysts were collected using a quadrupole mass spectrometer
(QMS, Thermo-Star GSD 301T, Pfeiffer). Prior to all experiments, each
sample (0.15 g) was pretreated at 400 °C for 1 h under Ar 3. Results and discussion
(50 cm3 min−1). Following pretreatment, the sample was cooled to
35 °C, and then heated to 1100 °C at a rate of 10 °C min−1 under a flow 3.1. Characterization
of 5% H2/Ar. Chemisorption measurements were performed by using a
dynamic pulse method (ASAP 2020C, Micromeritics). After reduction at Table 1 shows the BET surface areas of the supports and catalysts
750 °C under a flow of 10% H2/Ar for 1 h, the catalyst (0.25 g) was calculated from N2 adsorption-desorption isotherms. All N2-sorption
cooled to 40 °C under Ar, and H2-chemisorption was performed at this isotherms, pore size distributions, and detailed texture properties are
temperature. shown in Fig. S1 and Table S1. As indicated in Fig. S1, all supports and
To investigate the acidic properties of the catalysts (0.1 g), the
temperature-programmed desorption of iso-propanol (IPA-TPD) was Table 1
carried out on a QMS. After pretreatment at 750 °C for 1 h under H2, the Physicochemical properties of the Al2O3 and xPAl supports and the 10Ni/Al2O3 and xPAl-
sample was cooled to 35 °C, and IPA was introduced into the reactor 10Ni catalysts with different P content.
with a constant partial pressure (3 kPa IPA/Ar balanced) over 0.5 h.
Sample Support Catalyst (Ni/xP-Al2O3)
The sample was then purged by Ar (30 cm3 min−1) for 1 h. Finally, the
temperature was elevated to 750 °C at 10 °C min−1, and the products SBET SBET Reduced
characterized by acidic sites were calculated from measurement of the (m2 g−1) (m2 g−1)
%C3H5 fragment (m/z = 41). In addition, all samples (0.1 g) were pre- Metallic Ni Amount of H2
particle size chemisorbed
treated at 400 °C for 1 h under N2 (50 cm3 min−1) prior to conducting (nm)a (cm3 g−1)
the IPA dehydration reaction at 220 °C for 5 h. During the reaction, the
total flow rate and IPA partial pressure were maintained at 10Ni/Al2O3 144 108 7.6 0.185
50 cm3 min−1 and 3 kPa, respectively. Following the reduction process, 0.5PAl-10Ni 150 125 8.3 0.179
1PAl-10Ni 160 133 8.4 0.070
the gases passing through the reactor were analyzed on-line using a
2PAl-10Ni 162 151 9.0 0.032
Varian CP-3800 gas chromatograph (GC). 4PAl-10Ni 154 138 10.0 0.021
6PAl-10Ni 139 122 12.7 0.020
FIPA reacted
Conversion(%) = × 100%
FIPA fed (4) a
Determined by the Scherrer equation using the Ni(111) diffraction peak.

101
S. Bang et al. Catalysis Today 303 (2018) 100–105

Fig. 1. XRD patterns of the (a) calcined, (b) reduced, and (c) used 10Ni/Al2O3 and xPAl-10Ni catalysts with different P content. ▲: Ni2Al18O29, ●: NiO, ■: metallic Ni, ▼: Carbon.

catalysts exhibited type IV isotherms with a type H2 hysteresis loop at


relative pressures of 0.6–0.9, which corresponded to a typical meso-
porous structure. Upon increasing the P content, the hysteresis loop
became broader, indicating that different pore sizes developed. In ad-
dition, the BET surface areas of the Al2O3 and xPAl supports (Table 1)
produce a volcano-shaped curve as a function of P content, with a
maximum value of 162 m2 g−1 being recorded for the 2PAl support.
However, this value decreased at higher P concentrations due to partial
pore blockage on the alumina support. In addition, the variation in BET
surface area with the P content was closely related to the thermal sta-
bility of the alumina support. Stanislaus et al. [16] reported improved
thermal stabilities upon the addition of P to γ-Al2O3. They reported that
the BET surface area of 5 wt.% P-modified alumina was rather lower
than that of the unmodified alumina after calcination at 500 °C, while it
was higher at calcination temperatures > 700 °C. This is because P Fig. 2. H2-TPR profiles of the catalysts calcined at 800 °C for 3 h: (a) 10Ni/Al2O3, (b)
0.5PAl-10Ni, (c) 1PAl-10Ni, (d) 2PAl-10Ni, (e) 4PAl-10Ni, and (f) 6PAl-10Ni.
prevents sintering of the alumina particles during calcination [15]. It
was therefore considered that 2 wt.% P-modified alumina was optimal
for improving the thermal stability of the xPAl-10Ni catalysts. aluminate (Ni2Al18O29) [19]. As the P content was increased, this peak
Fig. 1a shows the XRD patterns of the xPAl-10Ni catalysts after intensity gradually decreased, and a third peak appeared at
calcination at 800 °C for 3 h. XRD peaks associated with Ni2Al18O29 920–1100 °C corresponding to the formation of AlPO4 [20]. Although
(JCPDS No. 22-0451) were observed for all catalysts at 2θ = 37.6, 45.7, AlPO4 was not detected by XRD analysis, its presence was confirmed by
and 66.4°, indicating that the impregnated Ni interacted strongly with H2-TPR measurements. This increase in AlPO4 formation with in-
Al2O3 at high calcination temperatures. In addition, peaks corre- creasing P content interfered with the interactions between the active
sponding to NiO (JCPDS No. 44-1159) were also observed at 2θ = 37.3 component and the support, and thus improved the NiO reducibility
and 43.2°. Upon increasing the P content, these NiO peaks became [21]. However, a decrease in interactions between Ni and the support
sharper, while those corresponding to Ni2Al18O29 became broader, led to reduction of the sintering resistance, thereby decreasing the
thereby suggesting that P suppressed the diffusion of Ni into the Al2O3 metal dispersion following reduction [20]. Therefore, as a function of
lattice, thus increasing the NiO content on the Al2O3 surface. Although the P content, the quantity of chemisorbed H2 on reduced Ni sites de-
the crystalline AlPO4 phase was not observed in the XRD patterns, its creased due to the crystal growth of the metallic Ni (Table 1).
formation was possible by the incorporation of P into the alumina Fig. 3 shows the IPA-TPD results of the Al2O3 and P-modified Al2O3
support. The AlPO4 phase is also produced by the sequential impreg- supports. These experiments are often used to determine the surface
nation of Ni and P precursors. It was therefore predicted that any AlPO4 acidic properties of catalysts [22], as IPA molecules can react at dif-
present in the catalysts was highly dispersed and likely in its amorphous ferent types of active sites to produce different products. To determine
form. In this respect, compounds containing Ni and P (e.g., Ni3(PO4)2) the acidic functionalities present, the dehydrated product propylene
can also be produced, but the formation of AlPO4 is dominant in the fragment, i.e., %C3H5, m/z = 41, was detected. In addition, the acidic
sequential impregnation process [17]. Furthermore, metallic Ni crys-
tallites (JCPDS No. 04-0850) were detected after reduction at 750 °C for
1 h in Fig. 1b. The crystal size of the metallic Ni was calculated from the
diffraction peak at 2θ = 44.5° using the Scherrer equation (Table 1).
Upon increasing the P content, the crystal size gradually increased from
7.6 to 12.7 nm due to sintering of the Ni species.
The H2-TPR profiles of the calcined xPAl-10Ni catalysts are shown
in Fig. 2. The reducibility of the Ni-based catalysts was greatly influ-
enced by the addition of P, and three reduction peaks were observed at
400–650, 700–1000, and 920–1100 °C. The first reduction peak at
400–650 °C corresponded to the reduction of NiO weakly interacting
with the support [18]. However, the reduction peaks at higher tem-
peratures were more intense, as large quantities of Ni were in-
corporated into the alumina support due to the high calcination tem-
perature (800 °C). In the 10Ni/Al2O3 catalyst, only one peak was
detected above 700 °C, which corresponded to the reduction of nickel Fig. 3. IPA-TPD profiles of the Al2O3 and xPAl supports calcined at 850 °C for 10 h.

102
S. Bang et al. Catalysis Today 303 (2018) 100–105

Table 2 and a volcano-shaped curve was obtained for the deactivation rate as a
Correlation between IPA-TPD and catalytic activities in the dehydration of IPA over Al2O3 function of P content, which corresponds with the IPA-TPD (Figs. 3 and
and xPAl supports.
S2) and IPA-dehydration (Table 2) results.
Support IPA-TPD Temp. (°C) Conversion (%) Selectivity (%) As catalyst deactivation in the DRM reaction is mainly caused by
carbon deposition from methane decomposition (Eq. (2)) and the
Propylene DIPE Boudouard reaction (Eq. (3)), such deactivation occurs rapidly when
the rate of carbon deposition is higher than the coke gasification rate
Al2O3 190.7 47.8 89.9 10.1
0.5PAl 194.7 39.1 89.0 11.0 [24]. Corthals et al. [25] reported that when the basic strength of the
1PAl 200.7 36.4 88.6 11.4 support increased, acidic CO2 molecules are more easily adsorbed on
2PAl 204.3 27.6 89.5 10.5 the catalyst surface. In this respect, catalyst acidity can be tuned by the
4PAl 195.8 35.8 92.0 8.0 addition of P. Therefore, CO2 adsorption is more favorable at lower
6PAl 188.0 37.3 92.3 6.8
acidities, in which the reversed Boudouard reaction can effectively take
place and remove the deposited carbon [26]. From the DRM reaction
strength is closely related to the peak temperature [22,23]. As shown in equation (Eq. (1)), a 1:1 ratio of H2/CO should be produced; however, a
Fig. 3, the peak temperature varied upon increasing the P content, with somewhat smaller ratio was actually obtained, likely due to the re-
the 2PAl support exhibiting the highest peak temperature, thereby versed Boudouard reaction (Eq. (9)) and the reverse water-gas shift
suggesting the weakest acidic strength. As shown in Fig. S2, the acidity reaction (Eq. (10)) [26]:
of the xPAl-10Ni catalysts also varied in a similar manner with the P C + CO2 → 2CO (ΔH°298 = 172.0 kJ/mol) (9)
content.
To further analyze the acidic properties of the supports, the IPA- CO2 + H2 → CO + H2O (ΔH°298 = 41.0 kJ/mol) (10)
dehydration reaction was performed (see Table 2 and Fig. S3). As in-
The long-term stabilities of the 10Ni/Al2O3 and 2PAl-10Ni catalysts
dicated, the main products were propylene and DIPE, while acetone
were then tested over 100 h (Table 3 and Fig. 5), giving deactivation
formation was negligible. The production of DIPE via an intermolecular
rates of 31.6 (CH4) and 30.1% (CO2) for the 10Ni/Al2O3 catalyst, and
dehydration reaction also characterizes the presence of acidic func-
6.6 (CH4) and 4.5% (CO2) for the 2PAl-10Ni catalyst. Therefore, 2PAl-
tionalities on the catalyst surface. Therefore, most IPA molecules re-
10Ni catalyst showed relatively stable performance than 10Ni/Al2O3
acted on acid sites give propylene and DIPE under the given reaction
catalyst.
condition. The obtained conversions for the IPA-dehydration reaction
as a function of P content thereby increased in the following order:
2PAl < 4PAl < 1PAl < 6PAl < 0.5PAl < Al2O3, which corre- 3.3. Carbon deposition on the used catalysts
sponded with the variation in IPA-TPD peak temperatures. Overall, the
results demonstrated that the 2PAl support exhibited the lowest acidity To investigate the degree of carbon deposition on the catalysts, TGA
and maintained this tendency even after Ni impregnation. was performed using the catalysts employed in the 20 h DRM reaction
(Fig. 6a). The first weight loss at 100–280 °C was associated with the
loss of water, while the weight gain between 280 and 400 °C was caused
3.2. Catalytic activity by oxidation of the metallic Ni to its corresponding oxides [27]. The
second weight loss at 400–720 °C was attributed to the combustion of
The conversions, deactivation rates, and the degrees of coke for- carbon deposited during the DRM reaction. This signal was then em-
mation for the DRM reaction over the 10Ni/Al2O3 and xPAl-10Ni cat- ployed to quantify the degree of carbon deposition (Table 3), which
alysts are summarized in Table 3. The results of the DRM reaction with increased in the following order: 2PAl-10Ni < 1PAl-10Ni < 4PAl-
time on stream for 20 h are shown in Fig. 4. As indicated, the initial 10Ni < 0.5PAl-10Ni < 6PAl-10Ni < 10Ni/Al2O3. This tendency
conversions for CH4 and CO2 over the 10Ni/Al2O3 catalyst were 70.2 corresponded to that of the deactivation rate indicated in Table 3,
and 68.4%, respectively. These initial conversions gradually decreased thereby confirming that catalyst deactivation during the DRM reaction
upon increasing the P content, indicating a decrease in the number of was caused by carbon deposition on the catalyst surface. These results
active sites (i.e., the metallic Ni surface area) in the initial reaction were also consistent with the intensities of the carbon (JCPDS No. 41-
stage. This is likely due to Ni particle agglomeration caused by AlPO4 1487) diffraction peak in the XRD patterns (Fig. 1c).
formation, thereby decreasing the interactions between Ni and the TPO experiments were also conducted to further investigate the
support (see Fig. 2 and Table 1). Although the initial CH4 conversion deposited carbon species present. As shown in Fig. 6b, three distinct
was the highest for 10Ni/Al2O3, rapid deactivation was observed as the oxidation peaks (α, β, and γ) were observed, indicating the presence of
reaction proceeded from 70.2 at initial step to 54.6% after 20 h. In different carbon species. As the α peak is the active species or reaction
contrast, 2PAl-10Ni exhibited the most stable catalytic performance, intermediate, it was omitted from further investigations. In contrast,

Table 3
Effect of P content in the DRM reaction over 10Ni/Al2O3 and xPAl-10Ni catalysts.

Catalyst Conversion (%) H2/CO Deactivation ratea (%) Amount of cokeb (wt.%)

CH4 CO2 CH4 CO2

Initial 20 h 100 h Initial 20 h 100 h 20 h 100 h 20 h 100 h

10Ni/Al2O3 70.2 54.6 48.0 68.4 57.1 47.8 0.887 22.2 31.6 16.5 30.1 25.4
0.5PAl-10Ni 68.0 62.3 – 65.0 59.8 – 0.888 8.4 – 8.0 – 11.5
1PAl-10Ni 66.4 63.2 – 63.0 60.8 – 0.883 4.9 – 3.4 – 8.0
2PAl-10Ni 62.5 60.9 58.4 60.1 58.6 57.4 0.871 2.5 6.6 2.5 4.5 6.2
4PAl-10Ni 52.6 50.7 – 51.7 49.8 – 0.836 3.7 – 3.7 – 10.6
6PAl-10Ni 37.8 32.4 – 34.9 27.6 – 0.845 14.4 – 21.0 – 23.6

a
Defined as (initial conversion – 20 h conversion)/initial conversion × 100%.
b
Coke deposition was quantified by TGA of the used catalysts after the DRM reaction at 750 °C for 20 h.

103
S. Bang et al. Catalysis Today 303 (2018) 100–105

Fig. 4. The catalytic performance of 10Ni/Al2O3 and xPAl-10Ni catalysts in the DRM reaction at 750 °C as a function of time on stream: (a) CH4 conversion, (b) CO2 conversion and (c)
H2/CO ratio.

the β and γ peaks were attributed to carbon encapsulating the Ni par-


ticles and filamentous carbon, respectively [28,29]. As indicated, the
peak areas corresponded to reversed volcano-shaped curve, which
varied with the P content. More specifically, the smallest peak area was
observed for the 2PAl-10Ni catalyst. Furthermore, the β peak was
dominant when the P content was ≤1 wt.%, while the γ peak was
mainly observed above this point. In addition, encapsulating carbon, β,
led to faster deactivation than filamentous carbon because it strongly
interfered with contact between the active Ni metal and the reactant
[30]. Furthermore, although 0.5PAl-10Ni and 4PAl-10Ni contained si-
milar quantities of deposited carbon (Table 3), deactivation of the β-
rich 0.5PAl-10Ni catalyst was more pronounced. These results clearly
demonstrate that P modification affected not only the degree of carbon
deposition but also the types of deposited carbon present through
changes in the support acidity.
Fig. 5. Long-term stabilities of 10Ni/Al2O3 and 2PAl-10Ni in the DRM reaction over The above trend was also confirmed by TEM. More specifically, as
100 h.
shown in the TEM image of 10Ni/Al2O3 (Fig. 7a), where deactivation
was the most pronounced, many encapsulating and filamentous carbon
species were observed, while mainly filamentous carbon existed on the
6PAl-10Ni catalyst surface (Fig. 7d). In contrast, significantly lower
quantities of carbon were observed for the 2PAl-10Ni catalyst (Fig. 7c),
again demonstrating that carbon deposition was inhibited by mod-
ification with optimal P content.

4. Conclusions

We herein reported our investigation into the DRM over P-modified


Al2O3 supports containing different P contents to control the acidity of
Ni-impregnated catalysts (i.e., xPAl-10Ni, x = 0.5, 1, 2, 4, and 6 wt.%).
From the H2-TPR results, it was confirmed that AlPO4 was formed upon
increasing the P content, which weakened interactions between the Ni
species and the support. The Ni particle size also increased as a function
of the P content, which corresponded with a decrease in the initial
conversion of the DRM reaction. However, the deactivation rate also
varied with the P content, and was closely related to the changes in
acidity of the P-modified support. The sample acidities, confirmed by
IPA-TPD and the IPA dehydration reaction, exhibited volcano-shaped
curves as a function of the P content, with the lowest acidity being
observed for the 2PAl-10Ni catalyst. As a result, the deactivation rates
during the 100 h DRM reaction of the 2PAl-10Ni catalyst (i.e., 6.6% for
CH4 and 4.5% for CO2) were significantly lower than those of the un-
modified 10Ni/Al2O3 catalyst (i.e., 31.6% for CH4 and 30.1% for CO2).
This is likely due to acidic CO2 molecules becoming more easily ad-
sorbed on the catalyst surface, thereby allowing the effective removal of
the deposited carbon by the reversed Boudouard reaction.

Acknowledgements
Fig. 6. (a) TGA and (b) TPO results for the used 10Ni/Al2O3 and xPAl-10Ni catalysts after
the DRM reaction at 750 °C for 20 h. This research was supported by Basic Science Research Program
through the National Research Foundation of Korea (NRF) funded by
the Ministry of Science and ICT (2017R1A2B3011316).

104
S. Bang et al. Catalysis Today 303 (2018) 100–105

Fig. 7. TEM images of the used catalysts after the


DRM reaction at 750 °C for 20 h: (a) 10Ni/Al2O3, (b)
1PAl-10Ni, (c) 2PAl-10Ni, and (d) 6PAl-10Ni.

Appendix A. Supplementary data (2000) 439–446.


[15] J. Wang, Y. Wang, J. Wen, M. Shen, W. Wang, Microporous Mesoporous Mater. 121
(2009) 208–218.
Supplementary data associated with this article can be found, in the [16] A. Stanislaus, M. Absi-halabi, K. Al-dolama, Appl. Catal. 39 (1987) 239–253.
online version, at http://dx.doi.org/10.1016/j.cattod.2017.08.013. [17] H. Kraus, R. Prins, J. Catal. 170 (1997) 20–28.
[18] Y. Kathiraser, W. Thitsartarn, K. Sutthiumporn, S. Kawi, J. Phys. Chem. C 117
(2013) 8120–8130.
References [19] Z. Zhang, M. Tang, J. Chen, Appl. Surf. Sci. 360 (2016) 353–364.
[20] Y. Bang, S.J. Han, J. Yoo, S. Park, J.H. Choi, Y.J. Lee, J.H. Song, I.K. Song, Int. J.
Hydrogen Energy 39 (2014) 4909–4916.
[1] S.T. Oyama, P. Hacarlioglu, Y. Gu, D. Lee, J. Int, Hydrogen Energy 37 (2012)
[21] M.K. Gnanamani, G. Jacobs, V.R.R. Pendyala, U.M. Graham, S.D. Hopps,
10444–10450.
G.A. Thomas, W.D. Shafer, D.E. Sparks, Q. Xiao, Y. Hu, B.H. Davis, Appl. Catal.
[2] J.M. Ginsburg, J. Pina, T.E. Solh, H.I. De Lasa, Ind. Eng. Chem. Res. 44 (2005)
A—Gen. 523 (2016) 146–158.
4846–4854.
[22] E. Hong, H.I. Sim, C.-H. Shin, Chem. Eng. J. 292 (2016) 156–162.
[3] K. Nagaoka, K. Seshan, K. Aika, J.A. Lercher, J. Catal. 197 (2001) 34–42.
[23] A.G.J.M. Campelo, J.F. Herencia, D. Luna, J.M. Marinas, A.A. Romero, J. Catal. 151
[4] M. García-Diéguez, E. Finocchio, M.Á. Larrubia, L.J. Alemany, G. Busca, J. Catal.
(1995) 307–314.
274 (2010) 11–20.
[24] S. Damyanova, B. Pawelec, K. Arishtirova, M.V.M. Huerta, J.L.G. Fierro, Appl.
[5] Z. Hou, P. Chen, H. Fang, X. Zheng, T. Yashima, Int. J. Hydrogen Energy 31 (2006)
Catal. B—Environ. 89 (2009) 149–159.
555–561.
[25] S. Corthals, J. Van Nederkassel, J. Geboers, H. De Winne, J. Van Noyen, B. Moens,
[6] J.H. Lee, K.Y. Koo, U.H. Jung, J.E. Park, W.L. Yoon, Korean J. Chem. Eng. 31 (2016)
B. Sels, P. Jacobs, Catal. Today 138 (2008) 28–32.
3115–3120.
[26] H. Eltejaei, H. Reza Bozorgzadeh, J. Towfighi, M. Reza Omidkhah, M. Rezaei,
[7] M.H.A. Shiraz, M. Rezaei, F. Meshkani, Korean J. Chem. Eng. 33 (2016) 3359–3366.
R. Zanganeh, A. Zamaniyan, A. Zarrin Ghalam, Int. J. Hydrogen Energy 37 (2012)
[8] I.G. Osojnik Črnivec, P. Djinović, B. Erjavec, A. Pintar, Chem. Eng. J. 207–208
4107–4118.
(2012) 299–307.
[27] M.M. Nair, S. Kaliaguine, F. Kleitz, ACS Catal. 4 (2014) 3837–3846.
[9] L. Xiancai, W. Min, L. Zhihua, H. Fei, Appl. Catal. A—Gen. 290 (2005) 81–86.
[28] H.M. Swaan, V.C.H. Kroll, G.A. Martin, C. Mirodatos, Catal. Today 21 (1994)
[10] A.W. Budiman, S.H. Song, T.S. Chang, C.-H. Shin, M.J. Choi, Catal. Surv. Asia 16
571–578.
(2012) 183–197.
[29] Z. Zhang, X.E. Verykios, S.M. MacDonald, S. Affrossman, J. Phys. Chem. 100 (1996)
[11] S. Liu, L. Guan, J. Li, N. Zhao, W. Wei, Y. Sun, Fuel 87 (2008) 2477–2481.
744–754.
[12] Y.-J. Lee, J.M. Kim, J.W. Bae, C.-H. Shin, K.-W. Jun, Fuel 88 (2009) 1915–1921.
[30] I. Luisetto, S. Tuti, C. Battocchio, S. Lo Mastro, A. Sodo, Appl. Catal. A—Gen. 500
[13] L.J. Lakshmi, K. Narsimha, P.K. Rao, Appl. Catal. A—Gen. 94 (1993) 61–70.
(2015) 12–22.
[14] G.A.H. Mekhemer, A.K.H. Nohman, N.E. Fouad, H.A. Khalaf, Colloid Surf. A 161

105

You might also like