Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2433 2023-08-29 13:20:43 |
2 layout & references Meta information modification 2433 2023-08-30 03:56:15 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Mazurova, K.; Miyassarova, A.; Eliseev, O.; Stytsenko, V.; Glotov, A.; Stavitskaya, A. Fischer–Tropsch Synthesis Catalysts. Encyclopedia. Available online: https://encyclopedia.pub/entry/48595 (accessed on 14 May 2024).
Mazurova K, Miyassarova A, Eliseev O, Stytsenko V, Glotov A, Stavitskaya A. Fischer–Tropsch Synthesis Catalysts. Encyclopedia. Available at: https://encyclopedia.pub/entry/48595. Accessed May 14, 2024.
Mazurova, Kristina, Albina Miyassarova, Oleg Eliseev, Valentine Stytsenko, Aleksandr Glotov, Anna Stavitskaya. "Fischer–Tropsch Synthesis Catalysts" Encyclopedia, https://encyclopedia.pub/entry/48595 (accessed May 14, 2024).
Mazurova, K., Miyassarova, A., Eliseev, O., Stytsenko, V., Glotov, A., & Stavitskaya, A. (2023, August 29). Fischer–Tropsch Synthesis Catalysts. In Encyclopedia. https://encyclopedia.pub/entry/48595
Mazurova, Kristina, et al. "Fischer–Tropsch Synthesis Catalysts." Encyclopedia. Web. 29 August, 2023.
Fischer–Tropsch Synthesis Catalysts
Edit

The Fischer–Tropsch process is considered one of the most promising eco-friendly routes for obtaining synthetic motor fuels. Fischer–Tropsch synthesis is a heterogeneous catalytic process in which a synthesis gas (CO/H2) transforms into a mixture of aliphatic hydrocarbons, mainly linear alkanes. An important research direction has been to increase the selectivity of the process for the diesel fraction. Diesel fuel synthesized via the Fischer–Tropsch method has a number of advantages over conventional fuel, including the high cetane number, the low content of aromatic, and the practically absent sulfur and nitrogen impurities. One of the possible ways to obtain a high yield of diesel fuel via the Fischer–Tropsch process is the development of selective catalysts.

diesel fuel Fischer–Tropsch bifunctional catalysts process factors

1. Introduction

More than eighty percent of the world’s energy consumption is met by crude oil [1]. Due to the depletion of petroleum resources, the search for alternative fuel production technologies is of great demand. The Fischer–Tropsch process opens up the possibility for the production of fuels and chemicals from biomass, natural gas, or other resources and plays an increasingly important role in the energy sector.
Fischer–Tropsch synthesis is a catalytic process that could be tuned to meet various needs. The main products of this process include a wide range (C1–C70+) of hydrocarbons, primarily n-alkanes and linear olefins. FTS products also include iso-alkanes and cyclic hydrocarbons. Oxygenated species, such as aldehydes, ketones, acids, and alcohols, are formed during reactions under specific conditions together with CO2 [2]. Technologies such as gas to liquid (GTL) and the Fischer–Tropsch synthesis to olefins (FTO) are now the most industrially demanded due to perspectives about their use as ecologically friendly alternatives to traditional production methods [3][4].
Gas-to-liquid (GTL) processes produce high-quality environmentally friendly fuels, in particular, diesel, which in terms of its performance, fully complies with Euro-5 requirements [5]. It is known that synthetic products have better characteristics compared with refined petroleum products. Restrictions on the concentration of sulfur and aromatic hydrocarbons make diesel produced via the Fischer–Tropsch process even more interesting. In addition, synthetic diesel fuel is characterized by a cetane number of about 70, which is much higher compared with the same parameter of fuel derived from an oil refinery (about 50), a low boiling point of 90% of fuel, low density, and better biodegradability [6].
For the selective production of diesel fuels in industry, third-generation technology developed by Shell is used. A distinctive feature from the first two generations is associated with a two-stage method for obtaining motor components. The first stage is to obtain solid hydrocarbons with maximum selectivity [7]. At the second stage, after separation from liquid hydrocarbons, the paraffin fraction is sent to hydrocracking and hydroisomerization to obtain high-quality fuel. However, the multi-stage process, as a rule, leads to large energy, capital, and economic costs. In addition, the presence of hydrogen plants for the corresponding hydroprocesses on offshore facilities can create a number of safety problems [8].
Since the beginning of the 21st century, the one-stage Fischer–Tropsch process using a bifunctional catalyst has become the most attractive method for the direct conversion of synthesis gas into middle distillates with high selectivity [9]. Many scientific groups from different countries are actively working on the implementation of the fourth generation of catalysts that combine the functions of chain growth and selective hydrogenolysis.

2. Bifunctional Catalysts

The yield of liquid hydrocarbons and feedstock conversion can be adjusted by changing the chemical state and the crystalline phase of the active metal; adjusting the particle and pore size of the support; as well as choosing the appropriate promoters and, of course, the Fischer–Tropsch synthesis conditions. However, new strategies are required to obtain higher selectivity for middle distillate hydrocarbons such as diesel fuel.
Bifunctional catalysts combine a traditional FT catalyst for the hydrogenation of CO to heavier hydrocarbons and an acid site on which hydrocarbons undergo hydrocracking. And, recently, such catalysts have been considered for the process of direct conversion of synthesis gas with the production of middle distillates with high selectivity [10][11][12][13][14][15][16].
Usually, bifunctional catalysts are prepared by impregnating Brønsted acid supports with transition (Co and Ni) or noble metals (Pt and Pd) [17]. The active metal in the catalyst performs the dehydrogenating and hydrogenating functions, and the catalyst carrier has an acidic function. The acidic properties of the catalyst determine its cracking and isomerizing ability. According to the most common mechanism, the resulting n-paraffin is first dehydrogenated on the surface of the metal site to n-olefins. The molecules diffuse into the Brønsted acid sites, where they are protonated to form carbocations. The carbocation undergoes isomerization or cracking reactions to form an olefin and a carbocation with fewer carbon atoms in the chain. After deprotonation, the olefins are displaced via diffusion to the metal centers, where they are hydrogenated [9][18][19].
The development of effective bifunctional catalysts involves several important aspects. Firstly, the conditions of Fischer–Tropsch synthesis and hydrocracking are significantly different. It is optimal to carry out FTS using cobalt-based catalysts at temperatures lower than hydrocracking. It is necessary to choose the optimal temperature for the two processes simultaneously. Secondly, continuous running of two types of reaction requires close contact between active phase of FT and acidic sites, which is difficult because of catalyst deactivation. For example, a hybrid catalyst consisting of an alkali-promoted molten iron catalyst and HZSM-5 rapidly deactivates due to the migration of the alkali metal from the iron catalyst into the zeolite [20]. Thirdly, the deactivation of the zeolite may be faster due to the deposition of carbon on the acid sites. Thus, improving catalyst stability is a key issue.
To solve the problems of rapid deactivation, a team of authors [21][22] developed a kind of bifunctional catalyst with a core–shell structure. Theoretically, the heavier hydrocarbons formed on the core, which is a conventional FT catalyst, diffuse through the shell, which is a zeolite. Such catalysts have an appropriate acidity to promote the cracking of long chain alkanes and show better stability in the Fischer–Tropsch synthesis. Catalysts with a core–shell structure were also obtained in [23] by depositing a ZSM-5 film of controlled thickness on the Co-Al2O3 surface. As a result of catalytic tests, an increase in selectivity for motor fuels was observed (the yield of gasoline fraction reached 75.5%), as was a feedstock conversion up to 78.7% due to effective mass and heat transfer, as well as the outer layer of ZSM-5, which is used as a cracking agent and isomerization of hydrocarbons. Core–shell catalysts based on halloysite aluminosilicate nanotubes were also obtained in [24]. The formation of bimetallic RuCo nanoparticles inside the support was achieved using microwave radiation. The result was systems with uniform metal deposition and a narrower particle size distribution compared with conventional synthesis methods. In addition, the formation of particles inside the pores of the mineral carrier made it possible to increase the selectivity for liquid fuels (up to 90.0%) and a chain growth index (ASF) of α = 0.853, which is typical for the formation of diesel fraction components.
The Co/ZSM-5 catalyst obtained via impregnation based on moisture capacity was encapsulated with a microporous shell of silicalite-1 using hydrothermal synthesis [25]. The synthesized microcapsule catalyst showed a significantly high degree of CO conversion (68.9%) and selectivity for gasoline range hydrocarbons (74.7%) at a low level of methane content due to a homogeneous microporous structure with an additional residence zone for both reactants and product inside its channels. The cobalt-embedded zeolite catalyst described by Liu et al. [26] also showed a high yield on gasoline fuels compared with a conventional Co/SiO2 catalyst and a zeolite supported catalyst due to the limited reaction environment, the high diffusion efficiency, and suitable acidic properties. High selectivity for the diesel fraction (66.2%) was obtained on a Fischer–Tropsch catalyst with cobalt nanocrystals of uniform size embedded in mesoporous SiO2 [27]. The spatial limitation of metal nanoparticles contributes to the inhibition of the aggregation of cobalt nanocrystals during FTS reactions and also leads to an increase in the feedstock conversion. In addition, the contact time between entrapped reaction intermediates and active sites can be increased inside the enclosed space, which further enhances the growth of long chain hydrocarbons. It should be noted that the core–shell catalyst is synthesized mainly using the hydrothermal method, which has some limitations [22]. First, this method is not easy for some zeolites such as HY and Hbeta. Second, the highly alkaline conditions used to synthesize the zeolite and the vulnerability of the FT catalyst core (especially the SiO2 carrier, which can degrade in an alkaline environment) further complicate this method. Thirdly, this method requires large energy and economic costs compared with traditional methods for the synthesis of a Fischer–Tropsch catalyst.
In recent years, many bifunctional zeolite–metal catalysts have also been developed and synthesized for the production of gasoline and diesel fuel. Many studies have found that adjusting the acidity and pore structure of the catalyst results in the selective production of a particular liquid fuel.
Thus, the authors of [10] showed that the higher selectivity for the gasoline fraction in a cobalt catalyst on the carrier H-ZSM-5 is explained by its higher acidity compared with HY and H-modernite. The pore structure of the carrier also plays an important role in the molecular weight distribution of the synthesis products. For example, among cobalt catalysts supported on various zeolites, Co/HZSM-12 showed the greatest efficiency in the synthesis of hydrocarbons in the gasoline range and the Co/HZSM-34 catalyst showed the greatest efficiency in the synthesis of n-paraffins, although the first is characterized by lower acidity. This was explained by the difference in the zeolite pore structure: H-ZSM-12 had larger pore channels (0.57–0.61 nm) than H-ZSM-34 (0.50 nm). It was shown that the availability of acid sites located inside the zeolite pores is also significant for the secondary reactions of products.
In [28], cobalt-containing catalytic systems were studied in the synthesis of Fischer–Tropsch fuel based on a SAPO-11 microporous molecular sieve with different characteristics of the secondary porous structure. When SAPO-11 is used as a base, CO conversion is at least 76.8% and selectivity for C5+ hydrocarbons is more than 59%, so SAPO-11 can be successfully used as an acid component in a bifunctional catalyst. At the same time, the activity of cobalt catalysts based on SAPO-11, as well as the selectivity and stability of the catalytic system, are significantly affected by the characteristics of the secondary porous structure of SAPO-11. Catalyst systems based on SAPO-11 with a more developed secondary pore structure improve the listed process characteristics due to efficient diffusion of the initial reagents and rapid desorption of reaction products. According to the authors, the creation of catalytic systems based on SAPO-11 molecular sieves, which have an even more developed secondary mesoporous structure, opens up the possibility of further increasing the selectivity for the C11–C18 diesel fraction with good low-temperature properties.
It should be emphasized that diffusion limitations arise in zeolites with micropores, since access to active centers is limited, which leads to restrictions on the activity, selectivity, and service life of the catalyst [29]. To solve this problem, to date, a number of works have been published on the preparation of mesoporous catalysts based on cobalt/zeolite systems for the Fischer–Tropsch process. Comparing microporous zeolites and hierarchical zeolites, it was found that the use of hierarchical zeolites makes it possible to increase CO conversion and selectivity for diesel fractions due to more efficient hydrocracking and isomerization reactions inside mesopores.
It was found that high selectivity for the diesel fraction can be achieved using a cobalt catalyst based on mesoporous zeolite Y in sodium form (Na-meso-Y). The selectivity reaches 60%, which is significantly higher than that with the classical ASF distribution (39%). The researchers also revealed that the key factors influencing the FTS are the average size of Co particles and mesopores [13].
Kang et al. [15] compared two methods of depositing cobalt on the carrier: impregnation and melt infiltration. The second method produces cobalt particles with a narrower size distribution. In melt infiltration, cobalt can be selectively deposited in mesopores, which increases the reducibility and dispersion of the particles. The authors suggest that the probability of repeated hydrogenolysis of reaction products can be reduced due to larger pores, which increases selectivity for high-molecular-weight products. The average Na-meso-Y mesopore size also affected the product selectivity; with an increase in the average mesopore size, the selectivity for C10–C20 increased, while the selectivity for methane decreased. The highest selectivity for the diesel fraction is typical for the Co/Na-meso-Y catalyst with a pore size of 15 nm prepared via melt infiltration. The authors suggest that the narrower size distribution of Co promotes selective hydrogenolysis and hence C10–C20 selectivity in the Fischer–Tropsch synthesis.
In one of the studies by Wang et al. [30], the influence of the porous structure of macro- and mesoporous ZSM-5 on the selectivity for gasoline and diesel fractions was studied by using various organic templates in the preparation of the catalyst. The authors emphasized the importance of the number of acid sites, since some acid sites may not perform well during hydrocracking, and an excess of acid sites can lead to excessive hydrocracking in the FTS.
Chen et al. [31] studied a series of Co/MZ hybrid catalysts that differ in the number of acid sites. The systems were obtained by mixing various amounts of nano-H-ZSM-5 zeolite with a Co/MCF catalyst in which nano-H-ZSM-5 zeolite performed the acidic function. As the mass fraction of nano-H-ZSM-5 zeolite increased to 80 wt.%, the number of acid sites increased to 1.96·10−4 mol·g·cat−1, which led to a decrease in the formation of C21+ products. As the acidity of the catalyst increased, the selectivity for C10–C20 hydrocarbons decreased due to excessive hydrocracking reactions; the Co/M catalyst without acid sites obeyed the traditional ASF distribution.
In [32], a series of ZSM-5/SBA-15 catalysts with a Co load of 15 wt.% was prepared. The supports were prepared by physically mixing ZSM-5 and SBA-15 in various proportions. The catalytic performance of composite-supported catalysts has been shown to be significantly better than that of the corresponding single-material-supported catalysts. An increase in the content of ZSM-5 from 0 to 20% led to an increase in the selectivity for C5–C22 hydrocarbons from 60.5 to 70.0% due to the large pore size of the catalyst, which ensures optimal accessibility to acid sites. A further increase in the proportion of ZSM-5 led to a decrease in liquid products.
In [33], a mesoporous cobalt catalyst based on ZSM-5 with a bimodal structure was synthesized via the double-matrix method. As a result, it was shown that the obtained catalyst with moderate acidity has a higher diesel fraction selectivity and a lower methane yield than the traditional Co/SBA-15 catalyst.

References

  1. AlNouss, A.; McKay, G.; Al-Ansari, T. A techno-economic-environmental study evaluating the potential of oxygen-steam biomass gasification for the generation of value-added products. Energy Convers. Manag. 2019, 196, 664–676.
  2. Shafer, W.D.; Gnanamani, M.K.; Graham, U.M.; Yang, J.; Masuku, C.M.; Jacobs, G.; Davis, B.H. Fischer–Tropsch: Product selectivity–the fingerprint of synthetic fuels. Catalysts 2019, 9, 259.
  3. Wang, C.; Fang, W.; Liu, Z.; Wang, L.; Liao, Z.; Yang, Y.; Li, H.; Liu, L.; Zhou, H.; Qin, X.; et al. Fischer–Tropsch synthesis to olefins boosted by MFI zeolite nanosheets. Nat. Nanotechnol. 2022, 17, 714–720.
  4. Kim, K.J.; Kim, K.Y.; Rhim, G.B.; Youn, M.H.; Lee, Y.L.; Chun, D.H.; Roh, H.S. Nano-catalysts for gas to liquids: A concise review. Chem. Eng. J. 2023, 468, 143632.
  5. Gill, S.S.; Tsolakis, A.; Dearn, K.D.; Rodríguez-Fernández, J. Combustion characteristics and emissions of Fischer–Tropsch diesel fuels in IC engines. Prog. Energy Combust. Sci. 2011, 37, 503–523.
  6. Aut Zhang, S.; Yang, X.; Zhang, H.; Chu, C.; Zheng, K.; Ju, M.; Liu, L. Liquefaction of biomass and upgrading of bio-oil: A review. Molecules 2019, 24, 2250.
  7. Eliseev, O.L.; Savost’yanov, A.P.; Sulima, S.I.; Lapidus, A.L. Recent development in heavy paraffins synthesis from CO and H2. Mendeleev Commun. 2018, 28, 345–351.
  8. Kibby, C.; Jothimurugesan, K.; Das, T.; Lacheen, H.S.; Rea, T.; Saxton, R.J. Chevron’s gas conversion catalysis-hybrid catalysts for wax-free Fischer–Tropsch synthesis. Catal. Today 2013, 215, 131–141.
  9. Tan, Y.; Hu, W.; Du, Y.; Li, J. Species and impacts of metal sites over bifunctional catalyst on long chain n-alkane hydroisomerization: A review. Appl. Catal. A Gen. 2021, 611, 117916.
  10. Zhang, Q.; Cheng, K.; Kang, J.; Deng, W.; Wang, Y. Fischer–Tropsch catalysts for the production of hydrocarbon fuels with high selectivity. ChemSusChem 2013, 7, 1251–1264.
  11. Li, J.; He, Y.; Tan, L.; Zhang, P.; Peng, X.; Oruganti, A.; Yang, G.; Abe, H.; Wang, Y.; Tsubaki, N. Integrated tuneable synthesis of liquid fuels via Fischer–Tropsch technology. Nat. Catal. 2018, 1, 787–793.
  12. Ryu, J.H.; Kang, S.H.; Kim, J.H.; Lee, Y.J.; Jun, K.W. Fischer-Tropsch synthesis on Co-Al2O3-(promoter)/ZSM5 hybrid catalysts for the production of gasoline range hydrocarbons. Korean J. Chem. Eng. 2015, 32, 1993–1998.
  13. Peng, X.; Cheng, K.; Kang, J.; Gu, B.; Yu, X.; Zhang, Q.; Wang, Y. Impact of Hydrogenolysis on the Selectivity of the Fischer–Tropsch Synthesis: Diesel Fuel Production over Mesoporous Zeolite-Y-Supported Cobalt Nanoparticles. Angew. Chem. Int. Ed. Engl. 2015, 54, 4553–4556.
  14. Kang, J.; Wang, X.; Peng, X.; Yang, Y.; Cheng, K.; Zhang, Q.; Wang, Y. Mesoporous Zeolite Y-Supported Co Nanoparticles as Efficient Fischer–Tropsch Catalysts for Selective Synthesis of Diesel Fuel. Ind. Eng. Chem. Res. 2016, 55, 13008–13019.
  15. Sun, B.; Qiao, M.; Fan, K.; Ulrich, J.; Tao, F. Fischer–Tropsch synthesis over molecular sieve supported catalysts. ChemCatChem 2011, 3, 542–550.
  16. Abelló, S.; Montané, D. Exploring iron-based multifunctional catalysts for Fischer–Tropsch synthesis: A review. ChemSusChem 2011, 4, 1538–1556.
  17. Tomasek, S.; Lonyi, F.; Valyon, J.; Wollmann, A.; Hancsók, J. Hydrocracking of Fischer–Tropsch paraffin mixtures over strong acid bifunctional catalysts to engine fuels. ACS Omega 2020, 5, 26413–26420.
  18. Weitkamp, J. Catalytic Hydrocracking—Mechanisms and Versatility of the Process. ChemCatChem 2012, 4, 292–306.
  19. Bouchy, C.; Hastoy, G.; Guillon, E.; Martens, J.A. Fischer-Tropsch Waxes Upgrading via Hydrocracking and Selective Hydroisomerization. Oil Gas Sci. Technol. 2009, 64, 91–112.
  20. Botes, F.G.; Böhringer, W. The addition of HZSM-5 to the Fischer–Tropsch process for improved gasoline production. Appl. Catal. A Gen. 2004, 267, 217–225.
  21. Yang, G.; Xing, C.; Hirohama, W.; Jin, Y.; Zeng, C.; Suehiro, Y.; Wang, T.; Yoneyama, Y.; Tsubaki, N. Tandem catalytic synthesis of light isoparaffin from syngas via Fischer–Tropsch synthesis by newly developed core–shell-like zeolite capsule catalysts. Catal. Today 2013, 215, 29–35.
  22. Sartipi, S.; van Dijk, J.E.; Gascon, J.; Kapteijn, F. Toward bifunctional catalysts for the direct conversion of syngas to gasoline range hydrocarbons: H-ZSM-5 coated Co versus H-ZSM-5 supported Co. Appl. Catal. A Gen. 2013, 456, 11–12.
  23. Zhu, C.; Bollas, G.M. Gasoline selective Fischer-Tropsch synthesis in structured bifunctional catalysts. Appl. Catal. B Environ. 2018, 235, 92–102.
  24. Mazurova, K.; Glotov, A.; Kotelev, M.; Eliseev, O.; Gushchin, P.; Rubtsova, M.; Kazantsev, R.; Vinokurov, V.; Stavitskaya, A. Natural aluminosilicate nanotubes loaded with RuCo as nanoreactors for Fischer-Tropsch synthesis. Sci. Technol. Adv. Mater. 2021, 23, 17–30.
  25. Javed, M.; Cheng, S.; Zhang, G.; Dai, P.; Cao, Y.; Lu, C.; Yang, R.; Xing, C.; Shan, S. Complete encapsulation of zeolite supported Co based core with silicalite-1 shell to achieve high gasoline selectivity in Fischer–Tropsch synthesis. Fuel 2018, 215, 226–231.
  26. Liu, J.; Wang, D.; Chen, J.F.; Zhang, Y. Cobalt nanoparticles imbedded into zeolite crystals: A tailor-made catalyst for one-step synthesis of gasoline from syngas. Int. J. Hydrogen Energy 2016, 41, 21965–21978.
  27. Cheng, Q.; Tian, Y.; Lyu, S.; Zhao, N.; Ma, K.; Ding, T.; Jiang, Z.; Wang, L.; Zhang, J.; Zheng, L.; et al. Confined small-sized cobalt catalysts stimulate carbon-chain growth reversely by modifying ASF law of Fischer–Tropsch synthesis. Nat. Commun. 2018, 9, 3250.
  28. Yakovenko, R.E.; Agliullin, M.R.; Zubkov, I.N.; Papeta, O.P.; Khliyan, G.T.; Savostyanov, A.P. Bifunctional cobalt-containing catalytic systems based on SAPO-11 molecular sieves in Fischer–Tropsch synthesis of fuels. Pet. Chem. 2021, 61, 378–387.
  29. Abello, S.; Bonilla, A.; Perez-Ramirez, J. Mesoporous ZSM-5 zeolite catalysts prepared by desilication with organic hydroxides and comparison with NaOH leaching. Appl. Catal. A Gen. 2009, 364, 191–198.
  30. Wang, Y.; Yu, J.; Qiao, J.; Sun, Y.; Jin, W.; Zhang, H.; Ma, J. Effect of mesoporous ZSM-5 morphology on the catalytic performance of cobalt catalyst for Fischer–Tropsch synthesis. J. Energy Inst. 2020, 93, 1187–1194.
  31. Chen, Y.; Zhang, J.; Jiang, X.; Wei, L.; Li, Z.; Liu, C. Nano-ZSM-5 decorated cobalt based catalysts for Fischer–Tropsch synthesis to enhance the gasoline range products selectivity. J. Taiwan Inst. Chem. Eng. 2020, 116, 153–159.
  32. Wu, L.; Li, Z.; Han, D.; Wu, J.; Zhang, D. A preliminary evaluation of ZSM-5/SBA-15 composite supported Co catalysts for Fischer–Tropsch synthesis. Fuel Process. Technol. 2015, 134, 449–455.
  33. Wang, Y.; Jiang, Y.; Huang, J.; Liang, J.; Wang, H.; Li, Z.; Wu, J.; Li, M.; Zhao, Y.; Niu, J. Effect of hierarchical crystal structures on the properties of cobalt catalysts for Fischer–Tropsch synthesis. Fuel 2016, 174, 17–24.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , ,
View Times: 330
Revisions: 2 times (View History)
Update Date: 30 Aug 2023
1000/1000