Next Article in Journal
C–O Hydrogenolysis of C3–C4 Polyols Selectively to Terminal Diols over Pt/W/SBA-15 Catalysts
Previous Article in Journal
Enhancing Acetophenone Tolerance of Anti-Prelog Short-Chain Dehydrogenase/Reductase EbSDR8 Using a Whole-Cell Catalyst by Directed Evolution
Previous Article in Special Issue
Achievements and Perspectives in Metal–Organic Framework-Based Materials for Photocatalytic Nitrogen Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Semi-Hydrogenation of Acetylene to Ethylene Catalyzed by Bimetallic CuNi/ZSM-12 Catalysts

1
State Key Laboratory of Green Chemical Engineering and Industrial Catalysis, Department of Process Development and Design, Shanghai Research Institute of Petrochemical Technology, Shanghai 201208, China
2
Key Laboratory of Jiangxi Province for Environment and Energy Catalysis, Institute of Applied Chemistry, College of Chemistry and Chemical Engineering, Nanchang University, Nanchang 330031, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(9), 1072; https://doi.org/10.3390/catal12091072
Submission received: 2 August 2022 / Revised: 10 September 2022 / Accepted: 15 September 2022 / Published: 19 September 2022

Abstract

:
The purpose of this work is to develop a low-cost and high-performance catalyst for the selective catalytic hydrogenation of acetylene to ethylene. Non-precious metals Cu and Ni were selected as active ingredients for this study. Using ZSM-12 as a carrier, Cu-Ni bimetallic catalysts of CuNix/ZSM-12 (x = 5, 7, 9, 11) with different Ni/Cu ratios were prepared by incipient wetness impregnation method. The total Cu and Ni loading were 2 wt%. Under the optimal reaction conditions, the acetylene conversion was 100%, and the ethylene selectivity was 82.48%. The CuNi7/ZSM-12 prepared in this work exhibits good performance in the semi-hydrogenation of acetylene to ethylene with low cost and has potential for industrial application.

Graphical Abstract

1. Introduction

Ethylene is a chemical in great demand for the preparation of downstream products [1,2]. Ethylene is usually obtained through the petrochemical route; that is, the catalytic cracking of petrochemical raw materials [3]. In recent years, with the continuous reduction of petroleum resources and the continuous increase of the demand for ethylene, people are exploring other methods to replace the petrochemical route to prepare ethylene. As we all know, China is a big country producing coal [4,5]. Thus, the synthesis of ethylene from the coal chemical route is a good alternative. The acetylene route based on calcium carbide is considered to be a very green and feasible production route [6].
At present, the most used active metal in hydrogenation reaction is Pd. Pd has extremely high activity for the conversion of acetylene, and 100% acetylene conversion can be achieved at relatively low temperatures [7]. “Recently, plenty of research focuses on the hydrogenation of acetylene to ethylene, such as the shale gas method [8], the electrochemical method [9], and Methanol to Olefins [10]. However, alloy catalysts are still the most widely used catalysts due to their excellent catalytic activity and selectivity [11].”
Besides active Pd being used in the selective hydrogenation of acetylene, it has been reported that other metals such as Pt [12,13], Ni [14,15,16], Au [17,18,19], Cu [20,21] can also be used as acetylene selection. The active metals of the hydrogenation catalysts and the order of their hydrogenation activity are: Pd > Pt > Ni > Au > Cu [22]. Pd and Pt are precious metals, and the resources are scarce, which is not conducive to extensive industrial application. Ni is a kind of non-precious metal discovered in recent years with good hydrogenation activity of acetylene, which has been widely studied by researchers [23]. Liu et al. [24] systematically studied the effect of different doping and different active metals on Ni-based catalysts, and adding IB (Group IB of the Periodic Table) metals to Ni-based catalysts can greatly improve the activity. The order of reactivity was: AuNi0.5/SiO2 > AgNi0.5/SiO2 > CuNi0.125/SiO2. Copper-based catalysts are less studied because of the low activity of Cu for acetylene hydrogenation. However, copper-based catalysts exhibit excellent selectivity towards ethylene at high temperatures. Monometallic Cu catalysts are easy to cause the polymerization of acetylene, and the polymer of acetylene would block the catalyst pores [25]. Thus, copper catalysts generally need to be modified by adding doping metals.
There are also many studies on nickel-based catalysts for acetylene hydrogenation, especially for the bimetallic catalysts. Many metal promoters have been added to Ni-based catalysts toward regulating the performance of acetylene hydrogenation. Wang et al. [26] prepared Ni-based catalysts and doped Ga with different Ni-Ga ratios. The results show that the selectivity of ethylene was better when Ni/Ga was 5. The enhanced performance was attributed to the formed Ni-Ga alloy and Ni3Ga intermetallic compound. Currently, there is charge transfer between Ni and Ga, which is beneficial to reduce the adsorption strength and adsorption capacity of ethylene on Ni atoms, then inhibit the peroxidation reaction and ethylene polymerization, and improve the selectivity of ethylene. However, the acidity and strength of the catalyst tended to increase with the increase of Ga content, which promotes polymerization and carbon deposition, which easily leads to catalyst deactivation. Chen et al. [27] prepared Ni/SiO2, Ni-In/SiO2 and Ni-In-K/SiO2 catalysts by the impregnation method and found that the addition of tin can significantly improve the conversion rate of acetylene, the selectivity of ethylene and also the catalytic stability. The main reason was that the geometric effect of In causes continuous Ni atoms to be isolated, and the electronic effect of In and Ni produces electron transfer to put Ni in an electron-rich state, while adding K would improve the selectivity of ethylene. In fact, due to the isolation effect and the transfer of electrons, the hydrogenolysis of C=C and the polymerization of acetylene were inhibited, and the surface acid strength and acid content of the catalyst were weakened. Hu et al. [14] prepared a Ni-Cu/Al2O3 catalyst with a Ni/Cu ratio of 1:1 by co-precipitation for the selective hydrogenation of acetylene to ethylene, resulting in nearly 100% acetylene conversion and 71% ethylene selectivity.
Although the metal oxide supported bimetallic NiCu catalyst has been investigated, the zeolites supported NiCu catalysts have been less investigated [28]. Komatsu et al. used MCM-41 molecular sieve with a mesoporous diameter as the support to prepare a uniformly dispersed Ni3Ge bimetallic catalyst. Compared with the traditional SiO2 carrier, MCM-41-supported Ni3Ge exhibited greater activity for the hydrogenation of cyclohexene. This is mainly due to the confinement effect between the Ni3Ge particles and the pore wall. This constraint effect is more obvious for MCM-41 with a smaller pore diameter [29]. Huang et al. modified β-zeolite with K+, and the modified β-zeolite had higher ethylene selectivity compared with γ-Al2O3 and Na+-β-zeolite supports, mainly due to the preference for acetylene over ethylene. Adsorbed on the modified β-zeolite [30].
In this study, ZSM-12 was chosen as the support due to its good thermal stability, unique pore structure and regulated acidity. Besides, a single molecular sieve was not active for acetylene hydrogenation. We choose the non-precious metal Ni as the active component, but the noble metal Ni often leads to excessive hydrogenation of acetylene to ethane due to its high activity. Therefore, in this work, we use NiCu bimetallic as the active component of the catalyst, which is supported on ZSM-12 support, and compared with single Ni/ZSM-12, the effect of Cu composition on catalyst performance was studied.

2. Results and Discussions

2.1. Catalyst Characterizations

To understand the phase composition of the catalysts, XRD characterization of H-ZSM-12 support, CuNix/ZSM-12 and reduced catalyst is shown in Figure 1. The XRD pattern of H-ZSM-12 support shows that it presents typical diffraction peaks of H-ZSM-12, indicating that the support has been successfully synthesized. After loading different ratio CuNix/ZSM-12 catalysts, they also only presented the characteristic diffraction peaks of H-ZSM-12 support, without showing the diffraction peaks of NiO or CuO. It suggests that the metal with low loading was highly dispersed on the carrier and the impregnation method would not destroy the support’s structure. The reduction process at 500 °C did not affect the structure of the catalyst, only showing H-ZSM-12. The characteristic diffraction peak and the structure of the catalyst did not change after the stability test, indicating the excellent stability of the catalyst.
As shown in Figure 2, we carried out SEM and TEM analyses for the CuNi7/ZSM-12 catalyst. According to the SEM images, the CuNi7/ZSM-12 catalyst is a square nanoparticle with a rough surface. It can be seen from the TEM image that the metal compounds are uniformly distributed in the ZSM-12 on the surface of the carrier, which is consistent with the XRD and XPS results.
Furthermore, N2-adsorption and desorption experiments were carried out to characterize their specific surface area, pore structure and pore volume. As can be seen from Figure 3, N2-adsorption and desorption curves of both the H-ZSM-12 support and CuNix/ZSM-12 (x = 5, 7, 9, 11) catalysts are IV isotherms with H4-type hysteresis loops, indicating that they have both microporous and mesoporous structures. According to the pore size distribution curves of the samples (Figure 3b), their pores are in the range of 3–4.5 nm, 4.5–15 nm, and below 2 nm. The pore size distribution is consistent with the N2-adsorption and desorption curve and indicates that the support and catalysts exhibit both microporous and mesoporous. The specific surface area (SBET), pore volume (VP) and average pore size (DP) are listed in Table 1. It can be seen that the SBET, VP, and DP values of the CuNix/ZSM-12 catalysts are slightly lower than those of the H-ZSM-12 support. Therefore, the Cu and Ni metals have been successfully loaded into the zeolite support. The inductively coupled plasma emission spectroscopy (ICP-AES) was further tested as the actual metal loading, as shown in Table 1. The actual atomic ratio of each catalyst is close to the theoretical metal loading and atomic ratio.
The H2-TPR curves of Ni/ZSM-12, Cu/ZSM-12 and CuNix/ZSM-12 catalysts were studied, as shown in the Figure 4, and Ni/ZSM-12 showed a broad peak at 250.69 °C to 530.19 °C that attributed to the reduction of NiO species [3,31]. Cu/ZSM-12 appeared at two peaks at 198.42 °C and 246.96 °C, which were assigned to the reduction of Cu2+ and Cu+ species, respectively [21]. The CuNix/ZSM-12 series catalysts exhibited a large broad peak. The peak at low temperature belongs to the reduction of Cu species, and the peak at higher temperature is owing to the reduction of Ni species. Moreover, all the peaks shifted toward lower temperatures with the increase in Cu loading. It is shown that there is a strong interaction between Ni and Cu species, forming intermetallic compounds [15].
Furthermore, we used XPS to analyze the chemical states and surface properties of Ni/ZSM-12, Cu/ZSM-12, and CuNi7/ZSM-12 catalysts. As illustrated in Figure 5a, the Ni2p peaks of Ni/ZSM-12 and CuNi7/ZSM-12 catalysts can be deconvoluted into three components consisting of Ni satellite, Ni2+, and Ni0 [32]. The Ni/ZSM-12 exhibits a higher intensity than CuNi7/ZSM-12 and the three peaks have a slight red shift after CuNi7 formation, which means that the Cu doping has a strong interaction with Ni and produces an apparent electronic transfer. The Cu 2p peaks of Cu/ZSM-12 and CuNi7/ZSM-12 catalysts are presented in Figure 5b, the binding energy peaks around 935 and 955 eV belong to the Cu 2p3/2 and Cu 2p1/2 states, respectively [33]. The CuNi7/ZSM-12 contains fewer Cu species, and the ratio of Cu2+, Cu0/Cu+ is different from Cu/ZSM-12, which suggests the chemical states of Cu are different.
To investigate the acidity of these catalysts, the catalysts were characterized by NH3-TPD. As can be seen from the Figure 6 and Table 2, all the catalysts perform two similar peaks. The low-temperature peak is around 110 °C, while the high-temperature peak appears at 356 °C, which corresponds to weak and moderately strong acids, respectively [26]. Compared with Ni/ZSM-12, the medium and strong acid sites of Cu/ZSM-12 prefer lower temperatures. It indicates that the Ni-based catalyst contains less acidic sites and adding Cu component on Ni-based catalyst would reduce the total acid amount and catalyst acid strength.

2.2. Catalytic Performance of 2 wt% Cu/ZSM-12 and 2 wt% Ni/ZSM-12

The catalytic performance of Ni and Cu supported by ZSM-12 is shown in Figure 7. The activity of pure 2 wt% Cu/ZSM-12 catalyst for acetylene hydrogenation was very low, and there was almost no activity before 150 °C. Then it increased with the increase of temperature above 150 °C, but the acetylene conversion was still only 59.59% at 250 °C over 2 wt% Cu/ZSM-12. The selectivity toward ethylene was also increased with increasing the reaction temperature, but it could reach up to 78.54% at 250 °C over 2 wt% Cu/ZSM-12. For the 2.0 wt% Ni/ZSM-12 catalyst, the acetylene conversion was 100% at above 90 °C. However, the selectivity of ethylene was not high, and it increased with the increase in temperature, which was only 70.53% at 250 °C. Thus, it was necessary to combine 2.0 wt% Cu/ZSM-12 and 2.0 wt% Ni/ZSM-12 catalysts.

2.3. Catalytic Performance of CuNix/ZSM-12 Catalysts

Figure 8 shows the catalytic performance of ZSM-12 with different CuNi ratios, CuNix/ZSM-12 (x = 5, 7, 9, 11) catalysts have high catalytic activity for acetylene hydrogenation, with the acetylene conversion reaching more than 95% at 90 °C and 100% after 110 °C. For ethylene selectivity, the CuNi5/ZSM-12 and CuNi7/ZSM-12 catalysts presented higher selectivity to ethylene than the other samples. At 250 °C, CuNi7/ZSM-12 catalyst exhibited the highest selectivity to ethylene (82.48%) and acetylene conversion reached up to 100%. Compared with pure 2.0 wt% Cu/ZSM-12 and 2.0 wt% Ni/ZSM-12, the catalytic performance—especially the ethylene yield of CuNi7/ZSM-12 for acetylene hydrogenation—was much better. Thus, the Cu promoter could increase the ethylene yield. At the same time, we also compared with the catalysts reported in other literatures (see Table S1), the CuNi bimetallic catalysts studied in this paper show better catalytic performance than other Ni-based catalysts, which are close to the catalytic performance of noble metal catalysts performance, providing an interesting proposal for the study of non-precious metal catalysts.
In addition, the influence of the ratio of hydrogen to alkyne on catalytic performance of CuNi7/ZSM-12 catalyst was investigated to find the optimal reaction conditions. Figure 9 shows that CuNi7/ZSM-12 catalyzes the semi-hydrogenation of acetylene under a different ratio of hydrogen and alkyne. When the ratio of hydrogen to alkyne was one, acetylene conversion rate was lower than 20% and the selectivity of ethylene was also lower than 10%. This may be caused by a lack of hydrogen, and a large amount of acetylene was unable to be hydrogenated. Instead, some polymer would be generated and blocked catalysts. After 170 °C, the catalytic performance became deactivated. When the ratio of H2/C2H2 increased to greater than or equal to two, the acetylene conversion could be largely increased to 100%. However, the selectivity of ethylene decreased with the increase of H2/C2H2 ratio, which may be caused by the increase of hydrogen concentration and the excessive hydrogenation of acetylene to ethane. In general, the CuNi7/ZSM-12 catalyst could selectively hydrogenate acetylene to ethylene with the ratio of H2/C2H2 at two.

2.4. Catalytic Stability Test of CuNi7/ZSM-12 Catalysts

The CuNi7/ZSM-12 catalyst was tested at the highest yield the stability test, as shown in the Figure 10a, shows that the conversion of CuNi7/ZSM-12 to acetylene has been maintained at 100%, and the selectivity of ethylene has been maintained at more than 80% before three hours, and after seven hours, the ethylene Selectivity was kept at 55%. TG analysis was performed on the catalysts before and after the reaction, as shown in the Figure 10b; for the fresh catalyst, the weight loss before 100 °C was attributed to the evaporation of water and remained constant thereafter, indicating the stability of the catalyst structure. For the reacted catalyst, the weight loss at 300–600 °C was attributed to the combustion of carbon deposits, which was the main reason for catalyst deactivation. The samples after stability test were analyzed by XRD and SEM. As shown in Figure 1 and Figure 2, it can be seen that the crystal structure of the catalyst has not changed, but the catalyst agglomerates together, which is the main factor for the decrease of catalyst selectivity.

3. Materials and Methods

3.1. Materials

CH3CH2OH (AR. China Sinopharm Group Chemical Reagent Co., Ltd., Ningbo, China), silica sol (40% SiO2, China Shandong Haoyao New Material Co., Ltd., Shandong, China), tetraethylammonium hydroxide aqueous solution (25%, China Yancheng Fanan Chemical Co., Ltd., Yancheng, China), NaAlO2 (AR. China Sinopharm Group Chemical Reagent Co., Ltd.), Ni(NO3)2·6H2O (AR. China Tianjin Damao Chemical Reagent Factory, Tianjin, China), NH4NO3 (AR. China Taishan Yueqiao Reagent Plastic Co., Ltd., Taishan, China), Cu(NO3)2·3H2O (AR. China Tianjin Damao Chemical Reagent Factory).

3.2. Carrier Preparation

ZSM-12 was prepared by hydrothermal synthesis. TEAOH, H2O, NaAlO2 and silica sol were added to the Teflon hydrothermal reaction kettle in turn and stirred for 0.5 h, then crystallized in an oven at 160 °C for six days. After crystallization, the products were washed and filtered with deionized water. The obtained solid products were dried at 60 °C overnight. The dried samples were roasted in a muffle furnace at 550 °C for 2 h and ground to obtain Na/ZSM-12, which was recorded as Na/ZSM-12. Finally, H-ZSM-12 was obtained by an ammonium ion exchange reaction of Na/ZSM-12.

3.3. Catalyst Preparation

The catalysts in this paper were prepared by the incipient wetness impregnation method. Preparation of single metal Cu or Ni/ZSM-12 catalysts with 2 wt% metal loading was as follows: 0.5 g of above ZSM-12 support was put in a crucible, then 0.5 mL solution of copper nitrate or nickel nitrate were slowly added to the carrier and dried naturally. The dried samples were calcined in a muffle furnace at 500 °C for 2 h. Finally, the samples were named 2 wt% Cu/ZSM-12 and 2 wt% Ni/ZSM-12, respectively. For preparation of the bimetallic catalyst, take 0.5 g carrier in the crucible, mix the required volume of copper nitrate solution and nickel nitrate solution and take 0.5 mL of the mixed solution slowly added to the carrier; the solution should be a completely soaked carrier. Then, the samples are placed in a cool place for natural drying. After natural drying, the samples are roasted in a muffle furnace at 500 °C for 2 h. After roasting, the samples are ground to obtain a fresh catalyst. The CuNix/ZSM-12 catalysts (x is atomic ratio of Ni/Cu, x = 5, 7, 9, 11) were also prepared by the above method. The difference was that the single metal nitrate was changed to the corresponding copper nitrate and nickel nitrate. The total loading of Cu and Ni in the bimetallic catalysts were 2 wt%.

3.4. Catalyst Evaluation

The catalyst activity evaluation device used in this paper is a model 6010 small fixed-bed reactor. The flow speed of the reaction gas is controlled by the mass flowmeter developed by Beijing Seven-star Hua Chuang Electronics (Beijing, China). After the reaction of the gas raw material, the tail gas is separated and analyzed online by FULI 9790 gas chromatograph with packed column of the Hao han 790 GC System (Tengzhou, China). For evaluation, a 0.1 g fresh catalyst was installed in a straight reaction quartz tube, and both ends were fixed with an appropriate amount of quartz cotton. The reactor temperature was raised to 500 °C at a rate of 5 °/min under hydrogen flow of 20 mL/min, and the catalyst was reduced at 500 °C for 2 h. After the reduction treatment, the reaction gas acetylene and hydrogen (H2: 10 mL/min, C2H2: 5 mL/min) were introduced after the temperature was cooled to 30 °C, and the space speed was 9000 mL·gcat−1·h−1. For the detection of outlet steam composition at different reaction temperatures, acetylene conversion (XC2H2) was calculated as follows:
X C 2 H 2 = n C 2 H 2 ( I N ) n C 2 H 2 ( O U T ) n C 2 H 2 ( I N ) × 100 %
In the above formula, X C 2 H 2 is the acetylene conversion rate, n C 2 H 2 ( I N ) represents the acetylene injection amount, and n C 2 H 2 ( O U T ) represents the acetylene outflow amount detected by gas chromatography.
Ethylene selectivity ( S C 2 H 4 ) was calculated as follows:
S C 2 H 4 = n C 2 H 4 ( O U T ) n C 2 H 2 ( I N ) n C 2 H 2 ( O U T ) × 100 %
In the above formula, S C 2 H 4 is ethylene selectivity, n C 2 H 4 ( O U T ) is ethylene outflow detected by gas chromatography, n C 2 H 2 ( I N ) is acetylene injection volume, n C 2 H 2 ( O U T ) is acetylene outflow detected by gas chromatography.
Ethylene yield is calculated as follows:
  Y C 2 H 4 = X C 2 H 2 × S C 2 H 4 × 100 %
In the above formula,   Y C 2 H 4 is the yield of ethylene, X C 2 H 2 is the conversion rate of acetylene and S C 2 H 4 is the selectivity of ethylene.

3.5. Catalyst Characterization

The X-ray diffractometer (XRD) (Beijing, China) was tested on an XD-3 X-ray diffractometer produced by Beijing General Analysis Instrument. The test conditions were as follows: the radiation source was Cu Kα (λ = 0.154 nm) with the tube current of 30 mA and the radiation tube voltage of 40 kV. The data were collected in the 2θ range of 10–50° at a scan speed of 4°/min. Determination of Ni and Cu content of samples was tested by the inductively coupled plasma emission spectroscopy (ICP-AES). The ICP instrument used in this paper is an Agilent 5100 ICP OES type inductively coupled plasma atomic emission spectrometer produced by Agilent Technologies (Santa Clara, CA, USA). The physical adsorption tester used in this paper was done on an Micromeritics ASAP2020 physical adsorption tester. The physical adsorption tester used for the BET test was the Micromeritics ASAP2020 physical adsorption tester (Norcross, GA, USA). Hydrogen temperature-programmed reduction (H2-TPR) was performed on a Chemisorb 2750 chemisorption instrument (Micromeritics Instrument Company, Norcross, GA, USA) equipped with a thermal conductivity detector (TCD). For the measurement of H2-TPR, 50 mg of catalyst was charged in a U-shaped quartz tube reactor and reduced from 45 °C to 800 °C at a heating rate of 10 °C/min under 10 V% H2-Ar. The SEM electron microscope tested in this paper is a JSM-6701F Field Emission Scanning Electron Microscope (FE-SEM) instrument produced by Hitachi, Tokyo, Japan. The secondary electron image resolution of the electron microscope is as follows: 1 nm (15 kV) and 2.2 nm (1 kV), magnifications are × 25–100,000, acceleration voltages are 0.5–30 kV. In this paper, the samples were characterized using Talos F200X field emission transmission electron microscope (TEM), the point resolution was less than 0.24 nm, the line resolution was less than 0.14 nm, and the voltage was 200 kV. Thermogravimetric (TG) analysis was performed on a TGA 550 instrument (TA Instruments, New Castle, DE, USA) under an air flow with a heating rate of 10 °C/min. NH3 temperature programmed desorption (NH3-TPD) was tested by Micromeritics AutoChem II 2920, the reduction temperature was 500 °C, and the heating rate was 10 °C/min. The measurement of the degree of dispersion was carried out using Micromeritics AutoChem II 2920 in the United States, and it was determined by hydrogen-oxygen titration. X-ray photoelectron spectroscopy (XPS) was performed on an Thermo Scientific K-Alpha instrument (Waltham, MA, USA) using Al Kα radiation (1486.6 eV).

4. Conclusions

The pure Cu/ZSM-12 and Ni/ZSM-12 catalysts, as well as bimetallic CuNix/ZSM-12 catalysts, were prepared by impregnation method. The different nickel-copper ratio and hydrogen-acetylene ratio on the hydrogenation of acetylene to ethylene were investigated in detail. For the CuNix/ZSM-12 (x = 5, 7, 9, 11) catalysts, the CuNi7/ZSM-12 catalyst with nickel-copper ratio of 7 displayed the best performance at 250 °C, with acetylene conversion reaching up to 100% and ethylene selectivity reaching up to 82.48%, which is much better than the performance of single-metal nickel and single-metal copper catalysts. In addition, it was found that the optimal ratio of H2/C2H2 was two. The higher ratio of H2/C2H2 would decrease the selectivity of ethylene, while the lower ratio of H2/C2H2 would decrease the acetylene conversion. In summary, the copper promoter and the ratio of reactants would significantly affect the semi-hydrogenation of acetylene.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12091072/s1, Table S1: Comparison of the activity and TOF of different catalysts that previous reports. References [34,35,36,37,38,39,40,41,42,43] are cited in Supplementary Materials.

Author Contributions

Writing—original draft preparation, S.H. and C.Z.; writing—review and editing, M.W., R.Y., D.S., M.L., P.Z., G.F. and R.Z.; project administration, S.H. and R.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is grateful for the support of the National Natural Science Foundation of China (Grants No. 21875096 and U19B60021005505), the Foundation of State Key Laboratory of Coal Conversion (Grant No. J22-23-903) and the Natural Science Foundation of Jiangxi Province, China (Grants No. 20181BCD40004).

Data Availability Statement

The data presented in this study are available on request from corresponding authors.

Acknowledgments

The authors are also thankful for Shiyanjia Lab (www.shiyanjia.com) on the TEM, XPS, and TG analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, T.-L.; Wang, H.; Li, B.; Krishna, R.; Wu, H.; Zhou, W.; Zhao, Y.; Han, Y.; Wang, X.; Zhu, W. Microporous metal–organic framework with dual functionalities for highly efficient removal of acetylene from ethylene/acetylene mixtures. Nat. Commun. 2015, 6, 1–9. [Google Scholar] [CrossRef] [PubMed]
  2. Gao, Y.; Neal, L.; Ding, D.; Wu, W.; Baroi, C.; Gaffney, A.M.; Li, F. Recent advances in intensified ethylene production—a review. ACS Catal. 2019, 9, 8592–8621. [Google Scholar] [CrossRef]
  3. Li, Q.; Wang, Y.; Skoptsov, G.; Hu, J. Selective hydrogenation of acetylene to ethylene over bimetallic catalysts. Ind. Eng. Chem. Res. 2019, 58, 20620–20629. [Google Scholar] [CrossRef]
  4. Gui, X.; Liu, J.; Cao, Y.; Miao, Z.; Li, S.; Xing, Y.; Wang, D. Coal preparation technology: Status and development in China. Energy Environ. 2015, 26, 997–1013. [Google Scholar] [CrossRef]
  5. Zhang, X.; Song, X.; Wang, J.; Su, W.; Zhou, B.; Bai, Y.; Yu, G. Physico-chemical structure evolution characteristics of coal char during gasification in the presence of iron-based waste catalyst. Int. J. Coal Geol. 2020, 7, 456–463. [Google Scholar] [CrossRef]
  6. Schobert, H. Production of acetylene and acetylene-based chemicals from coal. Chem. Rev. 2014, 114, 1743–1760. [Google Scholar] [CrossRef]
  7. Ball, M.R.; Rivera-Dones, K.R.; Gilcher, E.B.; Ausman, S.F.; Hullfish, C.W.; Lebrón, E.A.; Dumesic, J.A. AgPd and CuPd catalysts for selective hydrogenation of acetylene. ACS Catal. 2020, 10, 8567–8581. [Google Scholar] [CrossRef]
  8. Thiruvenkataswamy, P.; Eljack, F.T.; Roy, N.; Mannan, M.S.; El-Halwagi, M.M. Safety and techno-economic analysis of ethylene technologies. J. Loss Prev. Process Ind. 2016, 39, 74–84. [Google Scholar] [CrossRef]
  9. Wang, S.; Uwakwe, K.; Yu, L.; Ye, J.; Zhu, Y.; Hu, J.; Chen, R.; Zhang, Z.; Zhou, Z.; Li, J. Highly efficient ethylene production via electrocatalytic hydrogenation of acetylene under mild conditions. Nat. Commun. 2021, 12, 1–9. [Google Scholar] [CrossRef]
  10. Tian, P.; Wei, Y.; Ye, M.; Liu, Z. Methanol to olefins (MTO): From fundamentals to commercialization. ACS Catal. 2015, 5, 1922–1938. [Google Scholar] [CrossRef]
  11. Pei, G.X.; Liu, X.Y.; Wang, A.; Lee, A.F.; Isaacs, M.A.; Li, L.; Pan, X.; Yang, X.; Wang, X.; Tai, Z. Ag alloyed Pd single-atom catalysts for efficient selective hydrogenation of acetylene to ethylene in excess ethylene. ACS Catal. 2015, 5, 3717–3725. [Google Scholar] [CrossRef]
  12. Ma, H.-Y.; Wang, G.-C. Selective Hydrogenation of Acetylene on Pt n/TiO2 (n = 1, 2, 4, 8) Surfaces: Structure Sensitivity Analysis. ACS Catal. 2020, 10, 4922–4928. [Google Scholar] [CrossRef]
  13. Chesnokov, V.; Svintsitskii, D.; Chichkan, A.; Parmon, V. Effect of the structure of carbon support on the selectivity of Pt/C catalysts for the hydrogenation of acetylene to ethylene. Nanotechnol. Russ. 2018, 13, 246–255. [Google Scholar] [CrossRef]
  14. Hu, N.; Yang, C.; He, L.; Guan, Q.; Miao, R. Ni–Cu/Al2O3 catalysts for the selective hydrogenation of acetylene: A study on catalytic performance and reaction mechanism. New J. Chem. 2019, 43, 18120–18125. [Google Scholar] [CrossRef]
  15. Zhou, S.; Kang, L.; Zhou, X.; Xu, Z.; Zhu, M. Pure acetylene semihydrogenation over Ni–Cu bimetallic catalysts: Effect of the Cu/Ni ratio on catalytic performance. Nanomaterials 2020, 10, 509. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Y.; Zhao, J.; Feng, J.; He, Y.; Du, Y.; Li, D. Layered double hydroxide-derived Ni-Cu nanoalloy catalysts for semi-hydrogenation of alkynes: Improvement of selectivity and anti-coking ability via alloying of Ni and Cu. J. Catal. 2018, 359, 251–260. [Google Scholar] [CrossRef]
  17. Yang, B.; Burch, R.; Hardacre, C.; Headdock, G.; Hu, P. Origin of the Increase of Activity and Selectivity of Nickel Doped by Au, Ag, and Cu for Acetylene Hydrogenation. ACS Catal. 2012, 2, 1027–1032. [Google Scholar] [CrossRef]
  18. Liu, X.; Li, Y.; Lee, J.W.; Hong, C.-Y.; Mou, C.-Y.; Jang, B.W. Selective hydrogenation of acetylene in excess ethylene over SiO2 supported Au–Ag bimetallic catalyst. Appl. Catal. A Gen. 2012, 439, 8–14. [Google Scholar] [CrossRef]
  19. Gluhoi, A.C.; Bakker, J.W.; Nieuwenhuys, B.E. Gold, still a surprising catalyst: Selective hydrogenation of acetylene to ethylene over Au nanoparticles. Catal. Today 2010, 154, 13–20. [Google Scholar] [CrossRef]
  20. Chen, Y.; Liu, L.; Liu, B.-J. Selective hydrogenation of acetylene on CuNi/Al2O3 catalyst. Chem. Eng. J. 2012, 32, 50–52. [Google Scholar]
  21. McCue, A.J.; McRitchie, C.J.; Shepherd, A.M.; Anderson, J.A. Cu/Al2O3 catalysts modified with Pd for selective acetylene hydrogenation. J. Catal. 2014, 319, 127–135. [Google Scholar] [CrossRef]
  22. Ravanchi, M.T.; Sahebdelfar, S.; Komeili, S. Acetylene selective hydrogenation: A technical review on catalytic aspects. Rev. Chem. Eng. 2018, 34, 215–237. [Google Scholar] [CrossRef]
  23. Ye, R.-P.; Liao, L.; Reina, T.R.; Liu, J.; Chevella, D.; Jin, Y.; Fan, M.; Liu, J. Engineering Ni/SiO2 catalysts for enhanced CO2 methanation. Fuel 2021, 285, 119151. [Google Scholar] [CrossRef]
  24. Liu, H.; Chai, M.; Pei, G.; Liu, X.; Li, L.; Kang, L.; Wang, A.; Zhang, T. Effect of IB-metal on Ni/SiO2 catalyst for selective hydrogenation of acetylene. Chin. J. Catal. 2020, 41, 1099–1108. [Google Scholar] [CrossRef]
  25. Zhang, R.; Zhang, J.; Zhao, B.; He, L.; Wang, A.; Wang, B. Insight into the effects of Cu component and the promoter on the selectivity and activity for efficient removal of acetylene from ethylene on Cu-based catalyst. J. Phys. Chem. C 2017, 121, 27936–27949. [Google Scholar] [CrossRef]
  26. Wang, L.; Li, F.; Chen, Y.; Chen, J. Selective hydrogenation of acetylene on SiO2-supported Ni-Ga alloy and intermetallic compound. J. Energy Chem. 2019, 29, 40–49. [Google Scholar] [CrossRef]
  27. Chen, Y.; Chen, J. Selective hydrogenation of acetylene on SiO2 supported Ni-In bimetallic catalysts: Promotional effect of In. Appl. Surf. Sci. 2016, 387, 16–27. [Google Scholar] [CrossRef]
  28. Zhao, J.; Hua, Z.; Liu, Z.; Li, Y.; Guo, L.; Bu, W.; Cui, X.; Ruan, M.; Chen, H.; Shi, J. Direct fabrication of mesoporous zeolite with a hollow capsular structure. ChemComm 2009, 48, 7578–7580. [Google Scholar] [CrossRef]
  29. Komatsu, T.; Kishi, T.; Gorai, T. Preparation and catalytic properties of uniform particles of Ni3Ge intermetallic compound formed inside the mesopores of MCM-41. J. Catal. 2008, 259, 174–182. [Google Scholar] [CrossRef]
  30. Huang, W.; Pyrz, W.; Lobo, R.F.; Chen, J.G. Selective hydrogenation of acetylene in the presence of ethylene on K+-β-zeolite supported Pd and PdAg catalysts. Appl. Catal. A Gen. 2007, 333, 254–263. [Google Scholar] [CrossRef]
  31. Zhou, S.; Kang, L.; Xu, Z.; Zhu, M. Catalytic performance and deactivation of Ni/MCM-41 catalyst in the hydrogenation of pure acetylene to ethylene. RSC Adv. 2020, 10, 1937–1945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Hu, F.; Chen, X.; Tu, Z.; Lu, Z.-H.; Feng, G.; Zhang, R. Graphene aerogel supported Ni for CO2 hydrogenation to methane. Ind. Eng. Chem. Res. 2021, 60, 12235–12243. [Google Scholar] [CrossRef]
  33. Li, A.; Yao, D.; Yang, Y.; Yang, W.; Li, Z.; Lv, J.; Huang, S.; Wang, Y.; Ma, X. Active Cu0–Cuσ+ Sites for the Hydrogenation of Carbon–Oxygen Bonds over Cu/CeO2 Catalysts. ACS Catal. 2022, 12, 1315–1325. [Google Scholar] [CrossRef]
  34. Zhou, H.; Yang, X.; Li, L.; Liu, X.; Huang, Y.; Pan, X.; Wang, A.; Li, J.; Zhang, T. PdZn intermetallic nanostructure with Pd-Zn-Pd ensembles for highly active and chemoselective semi-hydrogenation of acetylene. ACS Catal. 2016, 6, 1054–1061. [Google Scholar] [CrossRef]
  35. Kang, J.H.; Shin, E.W.; Kim, W.J.; Park, J.D.; Moon, S.H. Selective Hydrogenation of Acetylene on TiO2-Added Pd Catalysts. J. Catal. 2002, 208, 310–320. [Google Scholar] [CrossRef]
  36. Kang, L.; Cheng, B.; Zhu, M. Pd/MCM-41 catalyst for acetylene hydrogenation to ethylene. R. Soc. Open Sci. 2019, 6, 191155. [Google Scholar] [CrossRef]
  37. Zhou, S.; Shang, L.; Zhao, Y.; Shi, R.; Waterhouse, G.I.N.; Huang, Y.C.; Zheng, L.; Zhang, T. Pd Single-Atom Catalysts on Nitrogen-Doped Graphene for the Highly Selective Photothermal Hydrogenation of Acetylene to Ethylene. Adv. Mater. 2019, 31, e1900509. [Google Scholar] [CrossRef]
  38. Jia, J.F.; Haraki, K.; Kondo, J.N.; Domen, K.; Tamaru, K. Selective hydrogenation of acetylene over Au/Al2O3 catalyst. J. Phys. Chem. B 2000, 104, 11153–11156. [Google Scholar] [CrossRef]
  39. Sun, X.; Li, F.; Shi, J.; Zheng, Y.; Su, H.; Sun, L.; Peng, S.; Qi, C. Gold nanoparticles supported on MgOx-Al2O3 composite oxide: An efficient catalyst for selective hydrogenation of acetylene. Appl. Surf. Sci. 2019, 487, 625–633. [Google Scholar] [CrossRef]
  40. Pei, G.X.; Liu, X.Y.; Wang, A.; Su, Y.; Li, L.; Zhang, T. Selective hydrogenation of acetylene in an ethylene-rich stream over silica supported Ag-Ni bimetallic catalysts. Appl. Catal. A-Gen. 2017, 545, 90–96. [Google Scholar] [CrossRef]
  41. Simanullang, W.F.; Ma, J.; Shimizu, K.-I.; Furukawa, S. Silica-decorated Ni–Zn alloy as a highly active and selective catalyst for acetylene semihydrogenation. Catal. Sci. Technol. 2021, 11, 4016–4020. [Google Scholar] [CrossRef]
  42. Esmaeili, E.; Mortazavi, Y.; Khodadadi, A.A.; Rashidi, A.M.; Rashidzadeh, M. The role of tin-promoted Pd/MWNTs via the management of carbonaceous species in selective hydrogenation of high concentration acetylene. Appl. Surf. Sci. 2012, 263, 513–522. [Google Scholar] [CrossRef]
  43. Zhao, L.; Wei, Z.; Zhu, M.; Dai, B. Catalytic performance of a Ti added Pd/SiO2 catalyst for acetylene hydrogenation. J. Ind. Eng. Chem. 2012, 18, 45–48. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of H-ZSM-12 and CuNix/ZSM-12 catalysts, (b) XRD patterns of the catalyst after reduction.
Figure 1. (a) XRD patterns of H-ZSM-12 and CuNix/ZSM-12 catalysts, (b) XRD patterns of the catalyst after reduction.
Catalysts 12 01072 g001
Figure 2. (a,b) The SEM results of Ni7Cu/ZSM-12; (c,d) SEM results of Ni7Cu/ZSM-12 after stability test; (e,f) the TEM results of Ni7Cu/ZSM-12; (go) EDS line profile results of Ni7Cu/ZSM-12.
Figure 2. (a,b) The SEM results of Ni7Cu/ZSM-12; (c,d) SEM results of Ni7Cu/ZSM-12 after stability test; (e,f) the TEM results of Ni7Cu/ZSM-12; (go) EDS line profile results of Ni7Cu/ZSM-12.
Catalysts 12 01072 g002
Figure 3. N2-adsorption/desorption curves (a) and pore size distributions (b) of H-ZSM-12 and CuNix/ZSM-12 catalysts.
Figure 3. N2-adsorption/desorption curves (a) and pore size distributions (b) of H-ZSM-12 and CuNix/ZSM-12 catalysts.
Catalysts 12 01072 g003
Figure 4. H2-TPR results of CuNix/ZSM-12, Cu/ZSM-12 and Ni/ZSM-12 catalysts.
Figure 4. H2-TPR results of CuNix/ZSM-12, Cu/ZSM-12 and Ni/ZSM-12 catalysts.
Catalysts 12 01072 g004
Figure 5. XPS spectra of Ni2p (a) of Ni/ZSM-12 and CuNi7/ZSM-12 catalysts; Cu2p (b) of Cu/ZSM-12 and Ni/ZSM-12, Cu/ZSM-12 and CuNi7/ZSM-12 catalysts.
Figure 5. XPS spectra of Ni2p (a) of Ni/ZSM-12 and CuNi7/ZSM-12 catalysts; Cu2p (b) of Cu/ZSM-12 and Ni/ZSM-12, Cu/ZSM-12 and CuNi7/ZSM-12 catalysts.
Catalysts 12 01072 g005
Figure 6. NH3-TPD results of Ni/ZSM-12, CuNi7/ZSM-12 and Cu/ZSM-12.
Figure 6. NH3-TPD results of Ni/ZSM-12, CuNi7/ZSM-12 and Cu/ZSM-12.
Catalysts 12 01072 g006
Figure 7. Catalytic performance of 2 wt% Cu/ZSM-12 and 2 wt% Ni/ZSM-12.
Figure 7. Catalytic performance of 2 wt% Cu/ZSM-12 and 2 wt% Ni/ZSM-12.
Catalysts 12 01072 g007
Figure 8. Activity test results of catalyst CuNix/ZSM-12 (x = 5, 7, 9, 11).
Figure 8. Activity test results of catalyst CuNix/ZSM-12 (x = 5, 7, 9, 11).
Catalysts 12 01072 g008
Figure 9. Influence of the hydrogen alkyne ratio on catalytic performance of CuNi7/ZSM-12.
Figure 9. Influence of the hydrogen alkyne ratio on catalytic performance of CuNi7/ZSM-12.
Catalysts 12 01072 g009
Figure 10. The stability test of CuNi7/ZSM-12 catalyst for acetylene semi-hydrogenation to ethylene (a). Test conditions: P = 0.1 MPa, T = 250 °C, GHSV = 9000 mL·g1·h1, and TG results of CuNi7/ZSM-12 before and after reaction (b).
Figure 10. The stability test of CuNi7/ZSM-12 catalyst for acetylene semi-hydrogenation to ethylene (a). Test conditions: P = 0.1 MPa, T = 250 °C, GHSV = 9000 mL·g1·h1, and TG results of CuNi7/ZSM-12 before and after reaction (b).
Catalysts 12 01072 g010
Table 1. Structural information and metal loadings of prepared samples.
Table 1. Structural information and metal loadings of prepared samples.
SamplesSBET m2/gVp cm3/gDp nmNi (wt%)Cu (wt%)Cu/Ni
CuNi5/ZSM-12361.70.1426.9871.560.345.04
CuNi7/ZSM-12361.90.1397.2361.730.276.91
CuNi9/ZSM-12365.00.1406.8581.800.219.46
CuNi11/ZSM-12374.20.1446.9521.820.1711.65
H-ZSM-12385.80.1497.093---
Table 2. Relative acid amounts of different catalysts.
Table 2. Relative acid amounts of different catalysts.
SamplesRelative Acidity Amount
(a)(a)Total Acid
Ni/ZSM-1210.411.41
CuNi7/ZSM-120.940.381.32
Cu/ZSM-120.910.271.18
(a): stands for weak acid in the first peak while stands for medium strong acid in the second peak.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hu, S.; Zhang, C.; Wu, M.; Ye, R.; Shi, D.; Li, M.; Zhao, P.; Zhang, R.; Feng, G. Semi-Hydrogenation of Acetylene to Ethylene Catalyzed by Bimetallic CuNi/ZSM-12 Catalysts. Catalysts 2022, 12, 1072. https://doi.org/10.3390/catal12091072

AMA Style

Hu S, Zhang C, Wu M, Ye R, Shi D, Li M, Zhao P, Zhang R, Feng G. Semi-Hydrogenation of Acetylene to Ethylene Catalyzed by Bimetallic CuNi/ZSM-12 Catalysts. Catalysts. 2022; 12(9):1072. https://doi.org/10.3390/catal12091072

Chicago/Turabian Style

Hu, Song, Chong Zhang, Mingyu Wu, Runping Ye, Depan Shi, Mujin Li, Peng Zhao, Rongbin Zhang, and Gang Feng. 2022. "Semi-Hydrogenation of Acetylene to Ethylene Catalyzed by Bimetallic CuNi/ZSM-12 Catalysts" Catalysts 12, no. 9: 1072. https://doi.org/10.3390/catal12091072

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