Next Article in Journal
Recent Advancements and Future Prospects in Ultrathin 2D Semiconductor-Based Photocatalysts for Water Splitting
Next Article in Special Issue
Preparation and Characterization of Ni/ZrTiAlOx Catalyst via Sol-Gel and Impregnation Methods for Low Temperature Dry Reforming of Methane
Previous Article in Journal
Heterogeneous Photocatalysis Scalability for Environmental Remediation: Opportunities and Challenges
Previous Article in Special Issue
Mechanistic Insights for Dry Reforming of Methane on Cu/Ni Bimetallic Catalysts: DFT-Assisted Microkinetic Analysis for Coke Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CeZrOx Promoted Water-Gas Shift Reaction under Steam–Methane Reforming Conditions on Ni-HTASO5

1
Key Laboratory of Green Chemistry and Technology, Ministry of Education, College of Chemistry, Sichuan University, Chengdu 610064, China
2
College of Chemical Engineering, Sichuan University, Chengdu 610065, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(10), 1110; https://doi.org/10.3390/catal10101110
Submission received: 29 August 2020 / Revised: 20 September 2020 / Accepted: 23 September 2020 / Published: 25 September 2020
(This article belongs to the Special Issue Catalytic Steam Reforming)

Abstract

:
Ni-based catalysts (Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx) were prepared by impregnation method and characterized by BET, AAS, XRD, H2-TPR, CO-TPD, NH3-TPD, XPS, TG-DSC-MS and Raman spectroscopies. Using CeZrOx-modified Al2O3 (HTASO5) as support, the catalyst exhibited good catalytic performance (TOFCH4 = 8.0 × 10−2 s−1, TOFH2 = 10.5 × 10−2 s−1) and carbon resistance for steam-methane reforming (SMR) reaction. Moreover, CeZrOx was able to enhance water-gas shift (WGS) reaction for more hydrogen production. It was found that the addition of CeZrOx could increase the content of active nickel precursor on the surface of the catalyst, which was beneficial to the decomposition of water and methane on Ni-HTASO5. Furthermore, Ni-HTASO5 could decrease the strong acid sites of the catalyst, which would not only contribute to the formation of low graphited carbon, but also decrease the amount of carbon deposition.

1. Introduction

Hydrogen is considered an important part of future energy systems. With the development of the hydrogen fuel cell, the application of H2 in vehicles and energy fields has aroused the interest of many researchers. Using Ni-based catalysts, methane can react with H2O, O2 or CO2 to produce hydrogen and carbon monoxide [1,2,3]. Because of its high H2/CO ratio, steam–methane reforming reaction (Equation (1)) is the main approach of hydrogen production in industry. When water-gas shift (WGS) reaction (Equation (2)) occurs simultaneously, it will increase the yield of hydrogen. Removing the products (H2, CO2) or enhancing adsorption of CO and H2O on the catalyst are both beneficial to WGS reaction for hydrogen production.
CH4 + H2O ⇌ 3H2 + CO
CO + H2O ⇌ H2 + CO2
Although compared with precious metals, nickel is not the most active catalyst, it is the most attractive because of its low cost and promising catalytic performance [4]. It has been reported that coke formation and metal sintering are the main reasons that lead to the deactivation of Ni-based catalysts in SMR reaction [5]. With the stoichiometric steam-to-methane ratio (H2O/CH4 = 1), graphite carbon is formed on the nickel-based catalyst, leading to reactor blockage and further deactivation of the catalyst [6]. High water–methane ratio can decrease the formation of carbon. In actual industrial production, the molar ratio of steam to methane is in the range of 2–5 [7]. Meanwhile, the reaction conditions can also result in technical problems associated with industrial catalysts. In order to obtain a high methane conversion and avoid the formation of carbon deposits through methane cracking or CO disproportionation reaction, one approach is controlling the reaction conditions, e.g., temperature, water–methane ratio, gaseous hourly space velocity, etc. [8]
Another approach is controlling the preparation of the catalyst. The support plays an important role in controlling the properties of the catalyst. The most commonly used support for commercial steam–methane reforming catalyst is α-Al2O3, which has good mechanical properties and thermal stability. Nevertheless, alumina is an acidic support that can catalyze secondary reactions such as polymerization and cracking, resulting in carbon deposition on the catalyst surface and blockage of the active sites [9]. In recent years, Mg, Fe, K, Ag, Pt, Pd, Rh, Ru and other elements have been used as additives to inhibit catalyst deactivation in catalytic systems of methane steam reforming [1,10,11,12,13]. Some researchers have studied different supports such as CeO2, MgAl2O4, MgO, hydrotalcite, NiAl2O4, ZrO2 and perovskite [11,14,15]. CeO2, ZrO2 and CexZr1−xO2 are known for their properties of storing, releasing and exchanging oxygen in lattice structures, and then enhancing methane steam reforming activity in terms of catalytic activity, stability and carbon deposition resistance [16,17,18,19]. Furthermore, CeO2 could contribute to avoid the active metal sintering and phase inversion from γ-Al2O3 to α-Al2O3, which could promote the stability of the catalyst. As a promising support or promoter, ZrO2 is proved to have acid–basic sites [20,21,22,23]. It has been demonstrated that cerium doped with zirconia has higher oxygen storage properties, which is beneficial to the oxidation process [24,25]. A solid mixture of CeO2–ZrO2 deposited on Al2O3, used as support for a Ni-based catalyst during SMR reaction was studied by de Abreu et al. [26]. It was found that the high content of Ce would promote the reduction of nickel species and the adsorption of H2O, which inhibited carbon deposition. Roh et al. [27] found that stable NiOx, which was the precursor of active Ni site, would form on Ni/Ce–ZrO2/θ–Al2O3 catalyst. The high catalytic activity and stability were owed to Ce–ZrO2 promoted formation of NiOx species. Macroporous CeO2–ZrO2 oxygen support for SMR reaction was studied [28], and it was discovered that good oxygen mobility and higher reducibility were due to the porous structure, which could be favorable for the release of oxygen from bulk to surface. Meanwhile, it also promoted methane and steam to diffuse into the catalyst. Wang et al. [29] found that a CexZr1−xO2 layer precoated on SBA-15 could promote the formation of high activity Ni species rather than bulk NiO and was beneficial to the formation of mobile oxygen species. Therefore, Ni/CexZr1−xO2/SBA-15 catalyst exhibited high catalytic activity and stability. Iglesias et al. [14] studied nickel catalysts supported on Ce1−xZrxO2−δ, the Ce/Zr ratio was optimized for SMR at severe reaction conditions, i.e., low temperature and low H2O/CH4 ratio. They found that zirconium was effectively incorporated into ceria cubic lattice and the support’s reducibility and metallic dispersion could be enhanced by the addition of Zr. These characteristics were determined by the structure, and then the physical and chemical properties of the catalyst were adjusted in the preparation of the catalyst.
Obviously, reducible CeZrOx has good catalytic effects and anti-carbon properties when it was used as support or promoter on heterogeneous catalytic reaction [30,31,32]. In order to combine the characteristics of Al2O3: high specific surface, good mechanical property, thermal stability, and the oxygen transfer ability of cerium and zirconium oxide, the CeZrOx-modified Al2O3 support (HTASO5) is applied on SMR reaction so as to enhance the reaction performance.

2. Results and Discussion

2.1. The Catalytic Activity of the Catalysts

The turnover frequency (TOF) of CH4 and H2 at 600 °C for 9 h are depicted in Figure 1. The initial TOF of CH4 conversion (for one hour) on Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx are 5.1 × 10−2 s−1, 8.0 × 10−2 s−1 and 5.5 × 10−2 s−1, respectively. The corresponding initial TOF of H2 (for one hour) are 7.0 × 10−2 s−1, 10.5 × 10−2 s−1 and 6.8 × 10−2 s−1, respectively. The higher TOF of H2 implies that WGS reaction may occur simultaneously. In the activity test, the Ni-HTASO5 exhibited the best catalytic performance while that of the other two catalysts were close. The repeatability of the catalyst was good, and the relative error between the two repeated experiments was 3.22% (Supplementary Materials Figure S2, Table S2). In terms of stability, the TOF of H2 on Ni-γ-Al2O3 and Ni-HTASO5 were stable at around 6.5 × 10−2 s−1 and 8.7 × 10−2 s−1, respectively after reaction for five hours, while that of Ni-CeZrOx gradually decreased. Meanwhile, the ratio of the content of CO2 to CO on Ni-CeZrOx was higher than those on the other two catalysts and that on Ni-HTASO5 was higher than on Ni-γ-Al2O3. It could be inferred that CeZrOx as additive or support can promote WGS reaction, which is beneficial to H2 production. Compared to other results from the literature, the catalytic effect of CeZrOx on WGS reaction was generally observed, and the modified Ni-HTASO5 catalyst was more conducive to the simultaneous production of hydrogen in the two-step reaction [14,26,28].

2.2. BET Results of the Catalysts

The specific surface area (SBET) and dispersion of pore volume (Vp) and pore size (Dp) of the catalysts are shown in Figure 2 and Table 1. The isotherm curves of the three catalysts are all type IV and the ratio of P to P0 for the hysteresis loop is over the range of 0.45 to 1.0, which indicates that they are all mesoporous materials [33]. The BET result (Table 1) showed that HTASO5 retained a relatively larger surface area of the catalyst than CeZrOx. The specific surface area of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx are 141.9 m2 g−1, 67.6 m2 g−1 and 49.0 m2 g−1, respectively. The addition of Ce, Zr in Al2O3 support decreased the surface area of the catalyst. The pore volume also decreased significantly compared with Ni-γ-Al2O3. However, the Ni-HTASO5 catalyst remained relatively larger surface area of than Ni-CeZrOx catalyst, which could promote the dispersion of nickel. The pore size distribution of Ni-HTASO5 was concentrated around 10–20 nm, while the other two catalysts were more widely distributed from 10 to 80 nm. As seen in Figure 2A, after being reduced in H2 atmosphere at 600 °C, the mesopores in Ni-γ-Al2O3 and Ni-CeZrOx could be classified into H3-type, which indicated an irregular pore structure and it is similar with that observed in nanorod and/or nanofiber [34]. However, the shape of hysteresis loop for Ni-HTASO5 was grouped within H1-type, which is typically found in spheroidal particles of uniform size and array [35,36].
The actual loading of nickel was characterized by atom adsorption spectrum, and the results are shown in Table 1. The content of Ni on Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx are 9.8%, 8.9% and 7.0%, respectively. Combining this result with that of surface area, the lower surface area was responsible for lower Ni-loading, which would influence the conversion of CH4.

2.3. The Reducibility and Distribution of Ni Species on the Catalysts

The reducibility of the Ni-based catalysts with different supports is shown in Figure 3A. Two peaks around 286 °C and 401 °C are observed on Ni-CeZrOx catalyst. The reducibility curve of Ni-HTASO5 is similar to Ni-γ-Al2O3. The peaks are around 450 °C (A peak) and 750 °C (B peak). Compared to the other two catalysts, the Ni species on Ni-CeZrOx could be reduced at lower temperature, which is attributed to weak interaction between metal Ni and CeZrOx support. The two reduction peaks of Ni-HTASO5 both shifted to lower temperatures when compared with Ni-γ-Al2O3. Furthermore, the ratio of A peak to B peak on Ni-γ-Al2O3 was lower than that of Ni-HTASO5. They both indicated that the addition of CeZrOx to Al2O3 can weaken the interaction between Ni species and the support.
The XRD result of Ni-based catalysts with different supports is shown in Figure 3B. It can be seen that the diffraction peak of 2θ = 44.5°, 51.9°, 76.4° (PDF# 04-0850) of metal Ni0 appears on all the catalysts after reduction and reaction. The Ce0.6Zr0.4O2 crystalline phase appeared on Ni-CeZrOx catalyst, while Ce0.16Zr0.84O2 crystalline phase was shown on Ni-HTASO5. Therefore, different CeZrOx were formed on Ni-CeZrOx and Ni-HTASO5. There was no significant change of the crystal size of Ni0 on the catalysts before and after the reaction (see in Table 2). The crystal size of Ni0 on Ni-γ-Al2O3 was smaller than the other two catalysts, which is owed to the high dispersion of Ni on the support. Furthermore, the carbon diffraction peak (PDF# 41-1487) appeared on Ni-γ-Al2O3 after reaction while it cannot be seen on the other catalysts, which means graphite carbon did form on Ni-γ-Al2O3. The grain size of Ni did not change significantly before and after the reaction of Ni-HTASO5 and Ni-CeZrOx and the grain size of them was similar.
The information of different surface species on the catalyst was determined by X-ray photoelectron spectrum (XPS). As seen in Figure 4, for Ni 2p, all the catalysts have three peaks. The binding energy at 854.5 eV, 855.8 eV were regarded as NiOx species and Ni2+ species, respectively [14,25,37]. Furthermore, NiOx species was regarded as the precursor of active nickel species [27,38]. From quantitative analysis results in Table 3, among the three catalysts, Ni-HTASO5 catalyst had the highest content of precursor of the active species nickel (NiOx), which play very important roles in the catalytic activity and stability [27,38]. Comparing the binding energy of Zr, Ce and O on Ni-HTASO5 and Ni-CeZrOx, it was found that the peak location of Ce had little change while that of Zr on Ni-HTASO5 shifted to higher binding energy. This means that Zr on Ni-HTASO5 had a stronger ability to donate electrons, which may contribute to the reduction of Ni species. The result of O 1s-binding energy showed that the ratio of lattice oxygen (O2−, 529.6 eV) to surface oxygen (OH, 531.6 eV) on Ni-CeZrOx was much higher than that on Ni-HTASO5 [39,40]. Therefore, CeZrOx had the property of storing oxygen while the HTASO5 support contained Al2O3, which made the surface of Ni-HTASO5 contain more OH. It has been reported that the lattice oxygen could promote activation of water [21,23], which may promote WGS reaction. The high surface content of active nickel could promote methane decomposition and hydrogen production, which made high catalytic activity of SMR reaction. Although there was more active lattice oxygen to promote the WGS reaction on Ni-CeZrOx, the low content of active nickel limited the SMR reaction for H2 production. Therefore, the TOF of hydrogen on Ni-HTASO5 was higher than that on Ni-CeZrOx.

2.4. The Acidity and Carbon Deposition of the Catalysts

The acidity of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts was characterized by temperature programmed NH3 desorption experiment. Three NH3 desorption peaks, which are ascribed to weak, medium and strong acid sites, over the range of 50–500 °C on all the catalysts are shown in Figure 5. The desorption peak at the range of 50–150 °C (peak 1) was attributed to weak Lewis acid site, and the peak with maximum at 150–250 °C (peak 2) was regarded as NH4+ bound to medium acid site [41]. The peak at above 250 °C (peak 3) was attributed to NH3 bound to strong Lewis acid site and NH4+ bound to strong Brønsted acid site [41,42,43], which may be due to the presence of surface OH on the support. Moreover, the amount of desorbed ammonia was calculated from NH3 desorption peak area and the result is shown in Table 4. The tendency of acid strength was as follows: Ni-γ-Al2O3 > Ni-HTASO5 > Ni-CeZrOx. The medium and strong acid site (peak 2 and peak 3) gradually increased with the increase of Al2O3 content. That was due to the acidity of Al2O3 support. The weak acid site (peak 1) increased by the addition of CeZrOx. The acidity of Ni-HTASO5 was lower than Ni-γ-Al2O3, which was due to the addition of CeZrOx to the support.
The amount and type of carbon deposition were investigated by thermogravimetric analysis combined with differential scanning calorimeter and mass spectrometer (TG-DSC-MS). The result is shown in Figure 6. In terms of the amount of deposited carbon, there was the highest amount of coke on Ni-γ-Al2O3 (24.01%) catalyst and the least on Ni-CeZrOx (9.43%) catalyst. It had previously been indicated that carbon was more likely to deposit on Ni-γ-Al2O3. This may be attributed to the acidic site of Al2O3, which could catalyze polymerization and cracking, resulting in more carbon deposition on the catalyst surface [44]. Meanwhile, the addition of Zr or Ce could increase the ability of carbon resistance [44,45,46]. Comparing with Ni-γ-Al2O3 catalyst, Ni-HTASO5 catalyst had less carbon deposition. Meanwhile, the CO2 MS signal on Ni-HTASO5 shifted to lower temperature. It means that the deposited carbon on Ni-HTASO5 was more easily removed. There was the least amount of coke on Ni-CeZrOx, furthermore the coke was of easy removal character. Combined with the result of NH3-TPD, it was inferred that the increase of weak acid sites would decrease the amount of carbon deposition and inhibit the growth of coke, which could then easily to be removed. That is to say, the Ni-HTASO5 modified by CeZrOx could not only form easily removable carbon, but also reduce the amount of carbon formation.
In order to identify the graphitization degree of coke, the Raman spectroscopy analysis was performed (Figure 6D). It could be seen that there were four peaks on all the catalysts. The peak at 1343 cm−1 (D bond) was considered to C–C stretch vibration of disordered carbon while the peak at 1579 cm−1 (G bond) was regarded as C–C stretch vibration of well-ordered carbon. The ratio of IG/ID of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx was 0.58, 0.47 and 0.20 (see in Table 2), respectively. As is shown, the graphitization degree increased with the increase content of Al2O3 in the support. Adding CeO2 and ZrO2 to the composite support could change the property of carbon deposition, making it decrease the graphitization, which was beneficial to carbon removal.
Combined with the characterization of the catalysts and the results of activity, it could be seen that the highest turnover frequency of hydrogen on Ni-HTASO5 is owed to its high surface content of active nickel and promising carbon resistance.

3. Materials and Methods

3.1. Synthesis of Catalysts

The γ-Al2O3 support was provided by Adamas-beta, Shanghai, China (99.99%). The HTASO5 support with the composition of CeO2:ZrO2:Al2O3 = 1:2:7 was provided by Zhongzi Environmental Protection Technology Co., LTD, Chengdu, China.
The CeZrOx support was prepared by co-precipitation method. Cerium nitrate (99.5%, Kelong, Chengdu, China) and zirconium nitrate (99.5%, Kelong, Chengdu, China) with the mole-ratio of Ce:Zr = 3:2 were dissolved in deionized water. The resulting solution was transferred to a flask and the pH adjusted to 9 using 2.5 M NH3·H2O with constantly stirring. The mixture was aged for 2 h at 70 °C, then cooled to room temperature and stored for 12 h. The resulting gel was rinsed thoroughly with deionized water and then it was dried in air overnight. The support was finally calcined at 450 °C for 4 h.
Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts were prepared by impregnation method. First, Ni(NO3)4·6H2O (2.1846 g) was added to 20 mL deionized water. The above support (4 g) was added after 20 min ultrasound treatment of the solution. The mixture was impregnated by stirring constantly at room temperature for 24 h, then dried at 80 °C using oil bath and at 110 °C in oven for 4 h. The three catalysts precursors were obtained after being calcined at 600 °C for 4 h.

3.2. Catalytic Activity Test

The catalyst activity was tested in a stainless steel fixed bed reactor at atmospheric pressure. Ni-γ-Al2O3 and Ni-HTASO5 were reduced at 800 °C for 1 h under the atmosphere of H2 and Ar mixture (H2/Ar = 1) and passivated in the mixture of oxygen and argon (5% O2 in Ar) at room temperature, while the reduction temperature of Ni-CeZrOx was 600 °C and passivated under the same conditions. Before the activity test, all the catalysts were reduced again at 600 °C for 1 h. The reaction gases methane (F = 30 mL/min) and H2O (nH2O/nCH4 = 3) were injected after ten min of argon purge. Meanwhile, the gas products were analyzed by gas chromatography (TDX-01 packed column).
The CH4 (or H2) turnover frequencies were calculated by the molar number of CH4 (or H2) converted or produced per second per mole exposed Ni atom. The number of exposed Ni atoms was determined by CO temperature programmed experiments (seen in Table S1). The methane conversion rate (XCH4), H2 production rate (YH2) and the CO2, CO (CCO2, CCO) content were calculated as follows:
X CH 4 = F CH 4 , in F CH 4 , out F CH 4 , in × 100 %
Y H 2 = F H 2 , out 2 × ( F CH 4 , in F CH 4 , out ) × 100 %
C CO 2 = F CO 2 , out F CO 2 , out + F CO , out + F H 2 , out × 100 %
C CO = F CO , out F CO 2 , out + F CO , out + F H 2 , out × 100 %
TOF(CH4) = m(CH4)/n(Ni)
TOF(H2) = m(H2)/n(Ni)
where F referred to the flow of gas, mL/min and m(CH4) (or m(H2)) was the molar number of CH4 (or H2) converted per second and n(Ni) was the molar number of exposed Ni atoms per gram of catalyst.

3.3. Characterization Methods

The BET specific area and the distribution of pore volume and pore size of samples were tested in a Micromeritics Tristar II 3020 instrument (Micromeritics, Norcross, GA, USA) by adsorption–desorption of N2 at −196 °C. Before the test, the sample was activated at 120 °C and 300 °C to eliminate any adsorbed substance.
The actual loading was tested by atom adsorption spectrum in a SpectrAA 220FS instrument (Varian, Palo Alto, CA, USA). The sample was dissolved in aqua regia (HCl:HNO3 = 3:1) and a small amount of HF. Deionized water was added while heating to nearly complete dissolution. Then water was added repeatedly to remove the acid completely. The grating was holographic diffraction grating with 1200 lines/mm (240 nm).
The temperature-programmed reduction (TPR) was performed to attain the reduction property of the catalysts by the Micromeritics Autochem II 2920 instrument (Micromeritics, Norcross, GA, USA). The test was from 50 °C to 900 °C with a heating rate of 5 °C/min under an atmosphere of 10% H2/Ar.
The temperature programmed NH3 desorption (NH3-TPD) was used to study the acid sites of the sample. First, NH3 was adsorbed on the reduced catalysts at 50 °C for 1 h under a mixture of 10% NH3 in He. Then, He was used to clean the excessive unadsorbed NH3 for 1 h. After this, with a heating rate of 10 °C/min under He flow, the sample was heated to 600 °C. The NH3 desorbed curve at different temperature was presented.
The CO temperature programmed desorption (CO-TPD) was carried out to study the amount of active centers on the sample. First, the catalyst was reduced at 600 °C under an atmosphere of 10% H2/Ar. After reduction, the samples were blown with He for 2 h. Then, CO was adsorbed on the reduced catalysts at 50 °C for 1 h under a mixture of 3% CO in He. He was used to clean the excessive unadsorbed CO for 1 h. After this, with a heating rate of 10 °C/min under a He flow, the sample was heated to 800 °C. The CO desorbed curve was presented.
X-ray diffraction (XRD) was conducted to probe the type and the size of formed crystal by an XRD-6100 (SHIMADZU, Japan) instrument. Cu Kα radiation of 40 kV and 25 mA was used. The diffraction angle ranged from 5° to 80°.
Thermogravimetric, differential scanning calorimeter and mass combination (TG-DSC-MS) analysis were performed to characterize carbon deposition on used catalysts by TG209F1 (NETZSCH, Selb, Germany) instrument. With the heating rate of 10 °C/min, the test started from 30 °C to 800 °C under an atmosphere of air/N2 (20/60 mL/min).
X-ray photoelectron spectroscopy (XPS) was carried out to study the surface state of the element on AXIS Ultra DLD (KRATOS, Manchester, UK) instrument equipped with a neutralizer. With monochromatic Al Kα as the light source, the acceleration power of 25 W, the binding energies were calibrated using C1s 284.6 eV, and the peak separation was performed with a Lorenz–Gaussian ratio (L/G) of 20%.
Raman spectroscopy analysis was conducted to characterize the deposited carbon on used catalysts. He–Ne laser source (532 nm) was used on LabRAM HR instrument (HORIBA, Kyoto, Japan) for the test. The filter was D1 and the aperture was 200 mm.

4. Conclusions

Nickel-based catalysts were prepared using γ-Al2O3, HTASO5 and CeZrOx as supports and used for steam–methane reforming reaction. Ni-HTASO5 showed good catalytic performance at 600 °C for 9 h while being more responsive to WGS reaction and having a promoting effect on hydrogen production. The high catalytic activity of Ni-HTASO5 was due to the presence of a high amount of active Ni precursor species on its surface as compared with Ni-γ-Al2O3 and Ni-CeZrOx. This could contribute to the decomposition of methane and water. The presence of CeZrOx promoted the WGS reaction under steam–methane reforming conditions. Furthermore, it was seen that the weak acid sites would decrease on the Ni-based catalyst-doping with CeZrOx in Al2O3 as support, which was beneficial to decrease the amount of carbon deposition and make it easier to be removed with a low graphited structure.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/10/1110/s1, Figure S1: CO-TPD of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts. Figure S2: The turnover frequency of CH4. (A)The first activity result and (B) the second activity result. Table S1: The amount of nickel active site determined by CO temperature programmed desorption. Table S2: Two activity experiment error of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts.

Author Contributions

Conceptualization, methodology, Q.Z. and C.H.; software, validation, formal analysis, investigation, resources, data curation, writing—original draft preparation, Q.Z.; writing—review and editing, visualization, supervision, Y.W., G.L. and C.H.; project administration, funding acquisition, C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key R and D Program of China (2018YFB1501404), the 111 program (B17030) and Fundamental Research Funds for the Central Universities.

Acknowledgments

We would like to thank the Analytical and Testing center of Sichuan University for characterization and we would be grateful to Yunfei Tian for his help of XPS experiments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gonçalves, J.F.; Souza, M.M.V.M. Effect of doping niobia over Ni/Al2O3 catalysts for methane steam reforming. Catal. Lett. 2018, 148, 1478–1489. [Google Scholar] [CrossRef]
  2. Wang, Y.; Li, L.; Wang, Y.; Da Costa, P.; Hu, C. Highly carbon-resistant Y doped NiO–ZrOm catalysts for dry reforming of methane. Catalysts 2019, 9, 1055. [Google Scholar] [CrossRef] [Green Version]
  3. Yabe, T.; Yamada, K.; Oguri, T.; Higo, T.; Ogo, S.; Sekine, Y. Ni-Mg supported catalysts on low-temperature electrocatalytic tri-reforming of methane with suppressed oxidation. ACS Catal. 2018, 8, 11470–11477. [Google Scholar] [CrossRef]
  4. Wei, J.M.; Iglesia, E. Structural requirements and reaction pathways in methane activation and chemical conversion catalyzed by rhodium. J. Catal. 2004, 225, 116–127. [Google Scholar] [CrossRef]
  5. Khzouz, M.; Gkanas, E.I. Experimental and numerical study of low temperature methane steam reforming for hydrogen production. Catalysts 2017, 8, 5. [Google Scholar]
  6. Wu, H.; La Parola, V.; Pantaleo, G.; Puleo, F.; Venezia, A.; Liotta, L. Ni-based catalysts for low temperature methane steam reforming: Recent results on Ni-Au and comparison with other bi-metallic systems. Catalysts 2013, 3, 563–583. [Google Scholar] [CrossRef] [Green Version]
  7. Berman, A.; Karn, R.K.; Epstein, A. Kinetics of steam reforming of methane on Ru/Al2O3 catalyst promoted with Mn oxides. Appl. Catal. A Gen. 2005, 282, 73–83. [Google Scholar] [CrossRef]
  8. Park, H.-G.; Han, S.-Y.; Jun, K.-W.; Woo, Y.; Park, M.-J.; Kim, S.K. Bench-scale steam reforming of methane for hydrogen production. Catalysts 2019, 9, 615. [Google Scholar] [CrossRef] [Green Version]
  9. Yung, M.M.; Jablonski, W.S.; Magrini-Bair, K.A. Review of catalytic conditioning of biomass-derived syngas. Energy Fuels 2009, 23, 1874–1887. [Google Scholar] [CrossRef]
  10. Koo, K.Y.; Roh, H.-S.; Seo, Y.T.; Seo, D.J.; Yoon, W.L.; Park, S.B. Coke study on MgO-promoted Ni/Al2O3 catalyst in combined H2O and CO2 reforming of methane for gas to liquid (GTL) process. Appl. Catal. A Gen. 2008, 340, 183–190. [Google Scholar] [CrossRef]
  11. Katheria, S.; Deo, G.; Kunzru, D. Rh-Ni/MgAl2O4 catalyst for steam reforming of methane: Effect of Rh doping, calcination temperature and its application on metal monoliths. Appl. Catal. A Gen. 2019, 570, 308–318. [Google Scholar] [CrossRef]
  12. De Lima, S.P.; Vicentini, V.; Fierro, J.L.G.; Rangel, M.C. Effect of aluminum on the properties of lanthana-supported nickel catalysts. Catal. Today 2008, 133–135, 925–930. [Google Scholar] [CrossRef]
  13. Roh, H.S.; Koo, K.Y.; Jeong, J.H.; Seo, Y.T.; Seo, D.J.; Seo, Y.S.; Yoon, W.L.; Park, S.B. Combined reforming of methane over supported Ni catalysts. Catal. Lett. 2007, 117, 85–90. [Google Scholar] [CrossRef]
  14. Iglesias, I.; Baronetti, G.; Alemany, L.; Mariño, F. Insight into Ni/Ce1−xZrxO2−δ support interplay for enhanced methane steam reforming. Int. J. Hydrogen Energy 2019, 44, 3668–3680. [Google Scholar] [CrossRef]
  15. Halabi, M.H.; de Croon, M.H.J.M.; van der Schaaf, J.; Cobden, P.D.; Schouten, J.C. Low temperature catalytic methane steam reforming over ceria–zirconia supported rhodium. Appl. Catal. A Gen. 2010, 389, 68–79. [Google Scholar] [CrossRef]
  16. Lim, Y.S.; Lee, M.-J.; Lee, K.-J.; Lee, S.; Hwang, H. Fabrication of Ce-promoted Ni/Al2O3 methane steam reforming catalysts by impregnation. J. Nanosci. Nanotechnol. 2020, 20, 4327–4330. [Google Scholar] [CrossRef]
  17. Dewoolkar, K.D.; Vaidya, P.D. Tailored Ce- and Zr-doped Ni/hydrotalcite materials for superior sorption-enhanced steam methane reforming. Int. J. Hydrogen Energy 2017, 42, 21762–21774. [Google Scholar] [CrossRef]
  18. Ghungrud, S.A.; Dewoolkar, K.D.; Vaidya, P.D. Cerium-promoted bi-functional hybrid materials made of Ni, Co and hydrotalcite for sorption-enhanced steam methane reforming (SESMR). Int. J. Hydrogen Energy 2019, 44, 694–706. [Google Scholar] [CrossRef]
  19. Koo, K.Y.; Lee, S.-h.; Jung, U.H.; Roh, H.-S.; Yoon, W.L. Syngas production via combined steam and carbon dioxide reforming of methane over Ni–Ce/MgAl2O4 catalysts with enhanced coke resistance. Fuel Process. Technol. 2014, 119, 151–157. [Google Scholar] [CrossRef]
  20. Matsumura, Y.; Nakamori, T. Steam reforming of methane over nickel catalysts at low reaction temperature. Appl. Catal. A Gen. 2004, 258, 107–114. [Google Scholar] [CrossRef]
  21. Kumar, P.; Sun, Y.; Idem, R.O. Nickel-based ceria, zirconia, and ceria-zirconia catalytic systems for low-temperature carbon dioxide reforming of methane. Energy Fuels 2007, 21, 3113–3123. [Google Scholar] [CrossRef]
  22. Ren, P.; Zhao, Z. Unexpected coke-resistant stability in steam-CO2 dual reforming of methane over the robust Mo2C-Ni/ZrO2 catalyst. Catal. Commun. 2019, 119, 71–75. [Google Scholar] [CrossRef]
  23. Zhao, Q.; Wang, Y.; Wang, Y.; Li, L.; Zeng, W.; Li, G.; Hu, C. Steam reforming of CH4 at low temperature on Ni/ZrO2 catalyst: Effect of H2O/CH4 ratio on carbon deposition. Int. J. Hydrogen Energy 2020, 45, 14281–14292. [Google Scholar]
  24. Ramirez-Cabrera, E.; Atkinson, A.; Chadwick, D. Reactivity of ceria, Gd- and Nb-doped ceria to methane. Appl. Catal. B Environ. 2002, 36, 193–206. [Google Scholar] [CrossRef]
  25. Wang, Y.; Zhao, Q.; Wang, Y.; Hu, C.; Da Costa, P. One-step synthesis of highly active and stable Ni–ZrOx for dry reforming of methane. Ind. Eng. Chem. Res. 2020, 59, 11441–11452. [Google Scholar] [CrossRef]
  26. De Abreu, A.J.; Lucrédio, A.F.; Assaf, E.M. Ni catalyst on mixed support of CeO2-ZrO2 and Al2O3: Effect of composition of CeO2-ZrO2 solid solution on the methane steam reforming reaction. Fuel Process. Technol. 2012, 102, 140–145. [Google Scholar] [CrossRef]
  27. Roh, H. Methane-reforming reactions over Ni/Ce-ZrO2/θ-Al2O3 catalysts. Appl. Catal. A Gen. 2003, 251, 275–283. [Google Scholar] [CrossRef]
  28. Zheng, Y.E.; Zhu, X.; Wang, H.; Li, K.; Wang, Y.; Wei, Y. Characteristic of macroporous CeO2-ZrO2 oxygen carrier for chemical-looping steam methane reforming. J. Rare Earths 2014, 32, 842–848. [Google Scholar] [CrossRef]
  29. Wang, X.L.K.; Ji, S.; Shi, X.; Tang, J.J. Effect of CexZr1−xO2 promoter on Ni-Based SBA-15 catalyst for steam reforming of methane. Energy Fuels 2009, 23, 25–31. [Google Scholar]
  30. Iglesias, I.; Forti, M.; Baronetti, G.; Mariño, F. Zr-enhanced stability of ceria-based supports for methane steam reforming at severe reaction conditions. Int. J. Hydrogen Energy 2019, 44, 8121–8132. [Google Scholar] [CrossRef]
  31. Lertwittayanon, K.; Youravong, W.; Lau, W.J. Enhanced catalytic performance of Ni/α-Al2O3 catalyst modified with CaZrO3 nanoparticles in steam-methane reforming. Int. J. Hydrogen Energy 2017, 42, 28254–28265. [Google Scholar] [CrossRef]
  32. Zhao, C.; Zhou, Z.; Cheng, Z.; Fang, X. Sol-gel-derived, CaZrO3-stabilized Ni/CaO-CaZrO3 bifunctional catalyst for sorption-enhanced steam methane reforming. Appl. Catal. B Environ. 2016, 196, 16–26. [Google Scholar] [CrossRef]
  33. Pan, Y.; Liu, Y.Q.; Liu, C.G. Nanostructured nickel phosphide supported on carbon nanospheres: Synthesis and application as an efficient electrocatalyst for hydrogen evolution. J. Power Sources 2015, 285, 169–177. [Google Scholar] [CrossRef]
  34. Wang, T.; Guan, X.; Lu, H.; Liu, Z.; Ji, M. Nanoflake-assembled Al2O3-supported CeO2-ZrO2 as an efficient catalyst for oxidative dehydrogenation of ethylbenzene with CO2. Appl. Surf. Sci. 2017, 398, 1–8. [Google Scholar] [CrossRef]
  35. Li, S.; Deng, J.; Dan, Y.; Xiong, L.; Wang, J.; Chen, Y. Designed synthesis of highly active CeO2-ZrO2-Al2O3 support materials with optimized surface property for Pd-only three-way catalysts. Appl. Surf. Sci. 2020, 506, 144866. [Google Scholar] [CrossRef]
  36. Raju, V.; Jaenicke, S.; Chuah, G.-K. Effect of hydrothermal treatment and silica on thermal stability and oxygen storage capacity of ceria–zirconia. Appl. Catal. B Environ. 2009, 91, 92–100. [Google Scholar] [CrossRef]
  37. Yao, L.; Wang, Y.; Shi, J.; Xu, H.; Shen, W.; Hu, C. The influence of reduction temperature on the performance of ZrOx/Ni-MnOx/SiO2 catalyst for low-temperature CO2 reforming of methane. Catal. Today 2017, 281, 259–267. [Google Scholar] [CrossRef]
  38. Lu, Y.; Xue, J.Z.; Yu, C.C.; Liu, Y.; Shen, S.K. Mechanistic investigations on the partial oxidation of methane to synthesis gas over a nickel-on-alumina catalyst. Appl. Catal. A Gen. 1998, 174, 121–128. [Google Scholar] [CrossRef]
  39. Maneerung, T.; Hidajat, K.; Kawi, S. Co-production of hydrogen and carbon nanofibers from catalytic decomposition of methane over LaNi(1−x)MxO3−α perovskite (where M = Co, Fe and X = 0, 0.2, 0.5, 0.8, 1). Int. J. Hydrogen Energy 2015, 40, 13399–13411. [Google Scholar] [CrossRef]
  40. Oemar, U.; Hidajat, K.; Kawi, S. Pd–Ni catalyst over spherical nanostructured Y2O3 support for oxy-CO2 reforming of methane: Role of surface oxygen mobility. Int. J. Hydrogen Energy 2015, 40, 12227–12238. [Google Scholar] [CrossRef]
  41. Xu, H.; Sun, M.; Liu, S.; Li, Y.; Wang, J.; Chen, Y. Effect of the calcination temperature of cerium-zirconium mixed oxides on the structure and catalytic performance of WO3/CeZrO2 monolithic catalyst for selective catalytic reduction of NOx with NH3. RSC Adv. 2017, 7, 24177–24187. [Google Scholar] [CrossRef] [Green Version]
  42. Smoláková, L.; Kout, M.; Koudelková, E.; Čapek, L. Effect of calcination temperature on the structure and catalytic performance of the Ni/Al2O3 and Ni-Ce/Al2O3 catalysts in oxidative dehydrogenation of ethane. Ind. Eng. Chem. Res. 2015, 54, 12730–12740. [Google Scholar] [CrossRef]
  43. Wang, J.; Dong, X.; Wang, Y.; Li, Y. Effect of the calcination temperature on the performance of a CeMoOx catalyst in the selective catalytic reduction of NOx with ammonia. Catal. Today 2015, 245, 10–15. [Google Scholar] [CrossRef]
  44. Therdthianwong, S.; Siangchin, C.; Therdthianwong, A. Improvement of coke resistance of Ni/Al2O3 catalyst in CH4/CO2 reforming by ZrO2 addition. Fuel Process. Technol. 2008, 89, 160–168. [Google Scholar] [CrossRef]
  45. Halliche, D.; Bouarab, R.; Cherifi, O.; Bettahar, M.M. Carbon dioxide reforming of methane on modified Ni/α-A12O3 catalysts. Catal. Today 1996, 29, 373–377. [Google Scholar] [CrossRef]
  46. Zhao, Z.; Ren, P.; Li, W.; Miao, B. Effect of mineralizers for preparing ZrO2 support on the supported Ni catalyst for steam-CO2 bi-reforming of methane. Int. J. Hydrogen Energy 2017, 42, 6598–6609. [Google Scholar] [CrossRef]
Figure 1. (A) Turnover frequency of CH4, (B) turnover frequency of H2 and (C) the content ratio of CO2 to CO on Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts at 600 °C with the mixed flow of H2O:CH4 = 3:1.
Figure 1. (A) Turnover frequency of CH4, (B) turnover frequency of H2 and (C) the content ratio of CO2 to CO on Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts at 600 °C with the mixed flow of H2O:CH4 = 3:1.
Catalysts 10 01110 g001
Figure 2. (A) N2 adsorption–desorption isotherm curve and (B) pore-volume and size-distribution for different Ni-based catalysts after reduction.
Figure 2. (A) N2 adsorption–desorption isotherm curve and (B) pore-volume and size-distribution for different Ni-based catalysts after reduction.
Catalysts 10 01110 g002
Figure 3. (A) H2-TPR profile and (B) XRD result of Ni-based catalysts with different supports: Ni-γ-Al2O3 (a) after reduction (b) after reaction, Ni-HTASO5 (c) after reduction (d) after reaction and Ni-CeZrOx (e) after reduction (f) after reaction.
Figure 3. (A) H2-TPR profile and (B) XRD result of Ni-based catalysts with different supports: Ni-γ-Al2O3 (a) after reduction (b) after reaction, Ni-HTASO5 (c) after reduction (d) after reaction and Ni-CeZrOx (e) after reduction (f) after reaction.
Catalysts 10 01110 g003
Figure 4. Ni 2p, Ce 3d, Zr 3d and O 1s binding energy of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts.
Figure 4. Ni 2p, Ce 3d, Zr 3d and O 1s binding energy of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts.
Catalysts 10 01110 g004
Figure 5. Temperature-programmed NH3 desorption (NH3-TPD) profile of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts.
Figure 5. Temperature-programmed NH3 desorption (NH3-TPD) profile of Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts.
Catalysts 10 01110 g005
Figure 6. TG-DSC-MS profile. (A) Ni-γ-Al2O3; (B) Ni-HTASO5; (C) Ni-CeZrOx; (D) Raman spectra of different Ni-based catalysts after reaction at 600 °C for nine hours.
Figure 6. TG-DSC-MS profile. (A) Ni-γ-Al2O3; (B) Ni-HTASO5; (C) Ni-CeZrOx; (D) Raman spectra of different Ni-based catalysts after reaction at 600 °C for nine hours.
Catalysts 10 01110 g006
Table 1. Physical properties of different Ni-based catalysts and different supports, the surface area was determined by BET method, the pore size (Dp) and the pore volume (Vp) were both determined by BJH method. The loading content of Ni was detected by atomic absorption spectrum (AAS).
Table 1. Physical properties of different Ni-based catalysts and different supports, the surface area was determined by BET method, the pore size (Dp) and the pore volume (Vp) were both determined by BJH method. The loading content of Ni was detected by atomic absorption spectrum (AAS).
SamplesSBET (m2 g−1)Vp (cm3 g−1)Dp (nm)Ni (%)
Ni-γ-Al2O3141.90.65169.8
Ni-HTASO567.60.25128.9
Ni-CeZrOx49.00.22167.0
γ-Al2O3151.20.6816/
HTASO571.50.2211/
CeZrOx51.60.1715/
Table 2. Crystal size of nickel on Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx after reduction and reaction as determined by XRD. The intensity ratio of G bond to D bond of deposited carbon after reaction at 600 °C for 9 h, tested by Raman spectroscopy.
Table 2. Crystal size of nickel on Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx after reduction and reaction as determined by XRD. The intensity ratio of G bond to D bond of deposited carbon after reaction at 600 °C for 9 h, tested by Raman spectroscopy.
CatalystNi0 (nm)IG/ID
ReductionReaction
Ni-γ-Al2O312120.58
Ni-HTASO518190.47
Ni-CeZrOx19180.20
Table 3. Surface content of different nickel species on reduced catalysts determined by XPS.
Table 3. Surface content of different nickel species on reduced catalysts determined by XPS.
CatalystContent (%)
NiOxNi2+
Ni-γ-Al2O32872
Ni-HTASO53466
Ni-CeZrOx2575
Table 4. Acidity measured by temperature programmed NH3 desorption experiment on Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts and weight loss of catalysts determined by thermogravimetric analysis.
Table 4. Acidity measured by temperature programmed NH3 desorption experiment on Ni-γ-Al2O3, Ni-HTASO5 and Ni-CeZrOx catalysts and weight loss of catalysts determined by thermogravimetric analysis.
CatalystsPeak 1Peak 2Peak 3Total
(mmol/g)
Weight Loss (%)
T1Amount (mmol/g)T2Amount (mmol/g)T3Amount (mmol/g)
Ni-γ-Al2O3127108175311317906132524.01
Ni-HTASO513316118316128637369512.77
Ni-CeZrOx130169191702702214609.43
T1, T2, T3—center temperature of the peak.

Share and Cite

MDPI and ACS Style

Zhao, Q.; Wang, Y.; Li, G.; Hu, C. CeZrOx Promoted Water-Gas Shift Reaction under Steam–Methane Reforming Conditions on Ni-HTASO5. Catalysts 2020, 10, 1110. https://doi.org/10.3390/catal10101110

AMA Style

Zhao Q, Wang Y, Li G, Hu C. CeZrOx Promoted Water-Gas Shift Reaction under Steam–Methane Reforming Conditions on Ni-HTASO5. Catalysts. 2020; 10(10):1110. https://doi.org/10.3390/catal10101110

Chicago/Turabian Style

Zhao, Qing, Ye Wang, Guiying Li, and Changwei Hu. 2020. "CeZrOx Promoted Water-Gas Shift Reaction under Steam–Methane Reforming Conditions on Ni-HTASO5" Catalysts 10, no. 10: 1110. https://doi.org/10.3390/catal10101110

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop