Next Article in Journal
Design and Implementation of Real-Time Kitchen Monitoring and Automation System Based on Internet of Things
Previous Article in Journal
Analysis on the Effect of Obstruction on Sprinkler Skipping in Air Conditioner Room Fire
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon Dioxide Assisted Conversion of Hydrolysis Lignin Catalyzed by Nickel Compounds

by
Artem A. Medvedev
1,
Daria A. Beldova
1,
Konstantin B. Kalmykov
1,
Alexey V. Kravtsov
1,
Marina A. Tedeeva
1,
Leonid M. Kustov
1,2,3,*,
Sergey F. Dunaev
1 and
Alexander L. Kustov
1,2,3
1
Chemistry Department, Moscow State University, 119992 Moscow, Russia
2
N.D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, Leninsky Prospect 47, 119991 Moscow, Russia
3
Institute of Ecotechnologies and Engineering, National University of Science and Technology MISiS, Leninsky Prospect 4, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Energies 2022, 15(18), 6774; https://doi.org/10.3390/en15186774
Submission received: 18 August 2022 / Revised: 9 September 2022 / Accepted: 13 September 2022 / Published: 16 September 2022

Abstract

:
In this work, hydrolysis lignin with nickel compounds deposited on the surface was prepared. The resulting material was introduced into the process of carbon dioxide assisted conversion and the catalytic activity of the deposited nickel compounds in this reaction was evaluated. Use of the obtained catalytic system increases CO2 conversion by more than 30% in the temperature range 450–800 °C. After the conversion process, the material was subjected to a study using a variety of physico-chemical analysis methods (TEM, SEM-EDX, and X-ray phase analysis). Physico-chemical methods of analysis of a sample calcined at 300 °C to decompose nickel nitrate revealed NiO nanoparticles with an average particle size of 16.9 nm.

1. Introduction

Currently, an extremely important technological problem from the point of view of ecology is the development of ways to use waste from the woodworking industry, as well as low-grade coals, pitches, and spent carbon materials, including their efficient conversion into renewable energy sources and other demanded materials with high consumer properties [1,2]. With the current global trend of a constant growth in energy consumption, there is an increasing interest in the use of alternative sources of organic raw materials, including wastes from the woodworking industry [3,4,5], as well as coal [6,7,8,9] and low-quality carbon materials such as soot [10], spent carbon materials in the form of sorbents [11], and municipal waste [12,13].
One of these unclaimed carbon materials is hydrolytic lignin. Lignin is a mixture of aromatic polymers contained in the cells of woody plants [14]. Lignin is a large-tonnage product of the woodworking industry with ineffective usage due to its low reactivity and poor solubility [15]. The development of methods for the usage of lignin is promising based on the fact that almost 95% of lignin is utilized for low-value purposes, such as a low-grade fuel for thermal power engineering or as an additive to concrete (lignosulfonate) [16]. Unlike many other sources of biomass, such as corn, starch, etc., lignin has no nutritional value, and its use as a precursor for the chemical or fuel industry will have no effect on the food security of society.
Researchers are currently considering various methods of using lignin, such as gasification [16,17,18,19], pyrolysis [20,21,22], acid [23], and basic [24] hydrolysis, reductive [25], and oxidative [26] conversion. Using lignin as a basis for value-added functional materials is also being developed at the present time [27]. Lignin is an aromatic biopolymer, and one of the goals of researchers is to obtain industrially important industrial aromatic products, such as phenols, vanillin, aromatic acids, etc [28]. However, due to its heterogeneous structure, the direct preparation of well-defined compounds from lignin is a difficult task. To obtain well-defined products, it is necessary to pre-treat the material to reduce the complexity of intermediate products [16].
Thus, for all strategies other than gasification, the first stage of lignin conversion usually involves the selective breaking of the macromolecular polymer into small fragments and their subsequent separation. Furthermore, the structure and composition of lignin depends both on the type of biomass from which it was extracted and on the deposit of one or another type of biomass [29]. Thus, for each batch of lignin, it is necessary to select conditions for processing, which negatively affect the unification of the process and its introduction into industry.
The gasification process is devoid of the above disadvantages and makes it possible to obtain carbon monoxide, the main component of synthesis gas, which can be used to produce liquid fuels (the Fischer–Tropsch process) [30] and can be used as a gaseous fuel. An important goal at the moment is to develop methods for the reasonable use of this valuable renewable resource.
Gasification with carbon dioxide as a gasification agent allows one to obtain carbon monoxide [17,31,32,33] (reaction 1), which is an important industrial by-product of the chemical industry. For example, an industrial method for obtaining formic acid is the interaction of methanol with carbon monoxide catalyzed by cobalt compounds [26]. The modern trend is also the utilization of the greenhouse gas carbon dioxide [34,35,36,37,38,39,40].
CO 2 + C 2 CO Δ H = 176   kJ / mol
Due to the fact that carbon dioxide conversion is a highly endothermic process, a significant conversion of carbon dioxide to carbon monoxide occurs at high temperatures of about 1000 °C [41]. To maintain such high temperatures, additional energy is required, which negatively affects the economic performance of this process. Catalysts applied to the surface of the carbon material can reduce the process temperature by 150–200 °C [17,42]. Compounds of alkaline, alkaline earth, and transition metals exhibit catalytic activity in such processes [17,37,43,44,45]. Iron triad compounds are of particular interest to researchers due to their availability, high catalytic activity in various processes, and low environmental impact.
The researchers presented a number of works [46,47] in which the biomass is subjected to a two-stage process. The first being pyrolysis with the formation of gaseous products, and then the conversion of these gaseous products using various catalysts to obtain synthesis gas, hydrocarbons, etc. This approach has a drawback due to the fact that the catalyst in the second stage must be regenerated in the same way because as coking of the oligomeric pyrolysis products occurs, the catalytic activity decreases. In the case of gasification using a supported catalyst, there is no such problem due to the fact that solid residues remain after the process, the content of the metal catalyst compound (usually a metal oxide) can be dissolved in nitric acid, and a new portion of the carbon-containing material can be impregnated with the resulting solution. Thus, the catalytic activity of the system is always at the optimum level. Another disadvantage of the processes is that during two-stage pyrolysis processes followed by conversion, a mixture of gaseous products is always obtained, which is in contrast to gasification in carbon dioxide at an atmospheric pressure where the only by-product of the process is methane, with the maximum selectivity being less than 5% at temperatures of 300–400 °C, and where the water gas shift process constant passes through a maximum. At higher temperatures (700–800 °C), the methane selectivity is less than 0.5%.
It was previously shown that iron and cobalt compounds demonstrate catalytic activity in the process of carbon dioxide assisted conversion of hydrolysis lignin [17]. Nickel as a catalyst in CO2 reactions is widely studied in such processes such as low temperature dry reforming of methane (DRM) [48,49,50], CO2 methanation, [51] etc. The activity of supported catalysts in the gasification process has been widely discussed for various grades and types of coals [52]; however, such data are not presented for such an important unclaimed carbon material as hydrolytic lignin. The aim of this work is to show the catalytic activity of nickel compounds deposited on the surface of hydrolytic lignin in the process of carbon dioxide conversion.

2. Materials and Methods

The hydrolysis lignin taken for the study (Table 1) has an internal surface area according to BET (S(lignin) = 163.8 m2/g). Elemental analysis showed a high carbon content in the material (52.8%) [23]. Four samples of hydrolysis lignin weighing 3 g each were impregnated with an aqueous solution of nickel (II) nitrate so that the mass fraction of Ni in the resulting material after complete drying was 1, 3, 5, and 7% by weight, respectively. All obtained samples were dried at room temperature for 24 h.
After deposition and preparation, the materials were subjected to SEM-EPMA analysis with a Leo Supra 50VP scanning electron microscope at a reduced pressure in a nitrogen atmosphere with detectors of secondary electrons (VPSE), reflected electrons (QBSD), and an INCA Energy+ X-ray spectrometer (Oxford Instruments, X-Max- 80, Abingdon, UK). To assess the distribution of elements, the samples were mapped. The surface compositions of the obtained materials were determined by the XMA method at various points.
The powder XRD diffraction pattern was obtained using a Rigaku IV Ultra device using CuKα radiation. The samples were examined in the region 2θ = 5–60 o at a sweep rate of 1 degree per minute.
The study by transmission electron microscopy was performed with a transmission electron microscope JEM-2100 (JEOL, Tokyo, Kanto, Japan).
The obtained and initial materials were introduced into the interaction with carbon dioxide at atmospheric pressure. Thus, the evaluation of the catalytic activity of the obtained materials in the process of carbon dioxide assisted conversion relative to the source material was carried out.
The catalytic tests were carried out in a quartz flow reactor with an internal diameter of 8 mm. The loading weight was 1 g. The CO2 flow rate was 30 mL/min at a total pressure of 1 atm. A Bronkhorst EL-FLOW SELECT F-111B gas flow controller was used to determine the gas flow rate. The temperature was increased at a constant rate: 600 °C/h in the range from 100 to 850 °C. During the reaction, the reaction gas products were analyzed using a Chromatek Crystal 5000 gas chromatograph with thermal conductivity detectors, M ss316 3 m * 2 mm columns, Hayesep Q 80/100 mesh, and CaA molecular sieves. The ratios of the amounts of substances were determined from the data of gas chromatography by the method of absolute calibration. The efficiency of the catalyst was evaluated by the conversion of carbon dioxide in the temperature range from 100 to 850 °C. The following reactions took place in the reaction zone for lignin (denoted as C):
CO 2 + C 2 CO
CO + H 2 O CO 2 + H 2
2 H 2 + C CH 4
The conversion of carbon dioxide during the tests with lignin was calculated by the formula:
Conv . CO 2 = 0.5 n ( CO ) + n ( CH 4 ) n ( CO 2 ) + 0.5 n ( CO ) 2 n ( CH 4 )
The main parameters of the gasification process for the samples of lignin and activated carbon are given in the table below:
Table 1. The main parameters of the gasification process.
Table 1. The main parameters of the gasification process.
ParameterDescription
CO2 pressure1 bar
Test protocolTemperature ramp up from 100 °C up to 850 °C with the rate 600°/h
CO2 WHSV at the inlet500.94 h−1
WHSV for 1% wt. of metal353.6 h−1
WHSV for 3% wt. of metal117.9 h−1
WHSV for 5% wt. of metal70.7 h−1
WHSV for 7% wt. of metal50.5 h−1
Analytical methodsGas chromatography for gaseous products. SEM-EDX, TEM and XRD analysis for solid residue.

3. Results and Discussions

3.1. Study of Catalytic Activity

Figure 1 shows the dependence of the carbon dioxide conversion on temperature for the initial sample of hydrolysis lignin and for lignin samples with deposited nickel compounds with a metal mass fraction of 1, 3, 5, and 7%. The use of supported catalysts makes it possible to transfer the curve to a lower temperature region, and the use of 7% wt. nickel makes it possible to increase the conversion of carbon dioxide by about 30% in the range 450–800 °C.
The material balance during this process converges by more than 85%. Such significant losses are explained by the fact that in the course of reaching the temperature regime, thermal destruction of hydrolysis lignin occurs, during which fragmented monomers are formed. These monomers condense in the trap and in the cold parts of the pipelines of the installation, and it is not possible to collect them and to determine the mass.
At temperatures of 350 to 500 °C, a noticeable conversion of carbon dioxide was observed not into carbon monoxide, but into methane. This is explained by the fact that at these temperatures the water gas shift reaction (WGSR) occurs (reaction 2). In the course of this reaction, hydrogen is formed, which interacts with carbon from the carbon material to form methane (reaction 3). However, as the temperature rises above 650–700 °C, the proportion of methane in the products decreases significantly and tends to zero due to the fact that the water gas shift reaction constant decreases with increasing temperature.

3.2. Physico-Chemical Analysis of Obtained Materials

The obtained materials were examined by a number of physico-chemical analysis methods: XRD and SEM-EDX. Nickel (Ka1) mapping was also carried out for a sample of lignin with a mass fraction of nickel of 7% (Figure 2). Judging by the mapping, nickel compounds were evenly distributed over the surface of the hydrolysis lignin. According to the EDX data from the 7% Ni/lignin sample, the metal content by weight was 6.9%. The proximity of the average value of the nickel fraction on the surface and a small standard deviation (1.92) indicates the uniformity of the distribution of nickel over the surface of the sample.
Figure 3 shows the XRD data for the initial lignin sample (a), a sample of hydrolysis lignin impregnated with nickel (II) nitrate (7 wt. % Ni) after calcination in a CO2 atmosphere for 1 h at a temperature of 300 °C (b), and residues of a sample of hydrolysis lignin with supported nickel compounds (7 wt. % Ni) after the carbon dioxide conversion process (c). The diffraction patterns obtained on all samples contain reflections of silicon dioxide on the diffraction patterns (a and b) SiO2 (100) at 20.90 and (011) at 26.69° (JCPDS card No. 01-085-1054); and on the diffraction pattern (c) SiO2 (100) at 20.83°, (011) at 26.62°, (200) at 42.4°, and (112) at 50.11° (JCPDS card No. 01-078-1252). Diffraction patterns (a–c) contain a reflection corresponding to graphite: C (002) at 26.43° (JCPDS card No. 00-008-0415). The diffraction pattern (b) contains a reflection corresponding to the compound (H3O)2NiO2: (003) at 13.34° (JCPDS card No. 01-078-1252). The diffraction pattern (c) also contains reflections corresponding to nickel (II) oxide: NiO (003) at 37.25° and (012) at 43.30° (JCPDS card No. 00-022-1189) and pure nickel: Ni (111) at 44.6° and (200) at 51.91° (JCPDS card No. 00-001-1260). It is widely known that in a similar process of gasification of activated carbon and graphite catalyzed by metal nanoparticles, waste is dissolved in metal particles with the formation of carbides [52]. Thus, diffusion of carbon atoms occurs through the metal particle to the surface, where the interaction of carbides occurs; however, XRD of the sample after conversion does not show the presence of reflections related to nickel carbides. Apparently, nickel carbides decompose and transform into the oxide form when the system is cooled to room temperature. The same phenomenon was observed during the gasification of activated carbon earlier [41].
A sample of the initial hydrolysis lignin after gasification, as well as one with deposited nickel compounds (7% wt.) after calcination in a CO2 atmosphere at 300 °C after gasification were examined by transmission electron microscopy. According to the TEM data of the Ni/lignin sample after calcination, the average nickel particle size was determined to be 16.9 ± 4.4 nm (Figure 4 and Figure 5). It is also possible to note the uniform distribution of nickel nanoparticles on the surface of the material (Figure 4). After the conversion process, nickel oxide forms spherical nanoparticles with an average particle size of 127 ± 52.3 nm shown on the TEM microphotographs of the lignin sample with Ni compounds (Figure 6); however, there are no such spherical nanoparticles on the TEM images of pure lignin after gasification (Figure 7). Elemental analysis carried out during the study using transmission electron microscopy showed the presence of nickel in materials with supported nickel compounds after calcination in a CO2 atmosphere and after the gasification process (Figure 8). Elemental analysis also showed the presence of calcium compounds, which also have catalytic activity in the process of carbon dioxide conversion of carbon materials [53].
Previously presented works devoted to the catalytic carbon dioxide conversion of hydrolytic lignin demonstrate a rather high activity of iron triad metals in this process. The papers consider supported catalysts containing iron and cobalt [17,31,32], with heating occurring both in a furnace and under the action of microwave radiation [31]. Scanning electron microscopy studies of residues after carbon dioxide conversion showed similar spherical structures, as in the case obtained in this work. This additionally confirms the mechanism of this reaction, which occurs through the dissolution of carbon in a metal nanoparticle with subsequent interaction with CO2 on the surface of the particle [52].
Thus, systems containing nickel are promising for the utilization of carbon dioxide and lignin with the production of carbon monoxide as a precursor for further synthesis of compounds with a higher added value.

Author Contributions

Conceptualization, L.M.K. and A.L.K.; methodology, A.L.K.; validation, A.A.M., D.A.B. and M.A.T.; formal analysis, A.A.M.; investigation, A.A.M. and D.A.B.; resources, A.L.K., K.B.K. and A.V.K.; writing—original draft preparation, A.A.M. and D.A.B.; writing—review and editing, A.L.K. and L.M.K.; supervision, L.M.K. and S.F.D.; project administration, L.M.K.; funding acquisition, L.M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Foundation for Basic Research (grant no. 20-33-90323) in the part related to the catalyst’s preparation and catalytic tests. This research was funded by the Ministry of Science and Higher Education of the Russian Federation (project no. 075-15-2021-591) in the part of study of the catalysts by physico-chemical methods.

Institutional Review Board Statement

Not acceptable.

Informed Consent Statement

Not acceptable.

Data Availability Statement

The data are available from the authors upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Titirici, M.-M.; White, R.J.; Brun, N.; Budarin, V.L.; Su, D.S.; del Monte, F.; Clark, J.H.; MacLachlan, M.J. Sustainable Carbon Materials. Chem Soc. Rev. 2015, 44, 250–290. [Google Scholar] [CrossRef]
  2. Zhuo, C.; Levendis, Y.A. Upcycling Waste Plastics into Carbon Nanomaterials: A Review. J. Appl. Polym. Sci. 2014, 131, 39931. [Google Scholar] [CrossRef]
  3. Arapova, O.V.; Chistyakov, A.V.; Palankoev, T.A.; Bondarenko, G.N.; Tsodikov, M.V. Microwave-Assisted Lignin Conversion to Liquid Products in the Presence of Iron and Nickel. Pet. Chem. 2020, 60, 1019–1025. [Google Scholar] [CrossRef]
  4. Kustov, L.M.; Tarasov, A.L.; Beletskaya, I.P. Microwave-Assisted Conversion of Lignin into Aromatic Compounds. Russ. J. Org. Chem. 2015, 51, 1677–1680. [Google Scholar] [CrossRef]
  5. Singh, O.; Sharma, T.; Ghosh, I.; Dasgupta, D.; Vempatapu, B.P.; Hazra, S.; Kustov, A.L.; Sarkar, B.; Ghosh, D. Converting Lignocellulosic Pentosan-Derived Yeast Single Cell Oil into Aromatics: Biomass to Bio-BTX. ACS Sustain. Chem. Eng. 2019, 7, 13437–13445. [Google Scholar] [CrossRef]
  6. Zhang, D.-K.; Poeze, A. Variation of Sodium Forms and Char Reactivity during Gasification of a South Australian Low-Rank Coal. Proc. Combust. Inst. 2000, 28, 2337–2344. [Google Scholar] [CrossRef]
  7. Zhang, F.; Xu, D.; Wang, Y.; Argyle, M.D.; Fan, M. CO2 Gasification of Powder River Basin Coal Catalyzed by a Cost-Effective and Environmentally Friendly Iron Catalyst. Appl. Energy 2015, 145, 295–305. [Google Scholar] [CrossRef]
  8. De, S.; Agarwal, A.K.; Moholkar, V.S.; Thallada, B. Coal and Biomass Gasification. In Recent Advances and Future Challenges; Springer: Singapore, 2018; ISBN 978-981-10-7334-2. [Google Scholar]
  9. Huang, X.; Zhang, F.; Wang, Y. Chapter 7—Catalytic Coal Gasification. In Sustainable Catalytic Processes; Saha, B., Fan, M., Wang, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2015; pp. 179–199. ISBN 978-0-444-59567-6. [Google Scholar] [CrossRef]
  10. Lee, W.-J.; Kim, D.-Y.; Choi, J.-H.; Lee, J.-W.; Kim, J.-S.; Son, K.; Ha, M.-J.; Kang, J. Utilization of Petroleum Coke Soot as Energy Storage Material. Energies 2019, 12, 3195. [Google Scholar] [CrossRef]
  11. Frías, M.; Sanchez de Rojas, M.I.; García, R.; Juan Valdés, A.; Medina, C. Effect of Activated Coal Mining Wastes on the Properties of Blended Cement. Cem. Concr. Compos. 2012, 34, 678–683. [Google Scholar] [CrossRef]
  12. Tursunov, O. A Comparison of Catalysts Zeolite and Calcined Dolomite for Gas Production from Pyrolysis of Municipal Solid Waste (MSW). Ecol. Eng. 2014, 69, 237–243. [Google Scholar] [CrossRef]
  13. Zheng, X.; Ying, Z.; Wang, B.; Chen, C. Hydrogen and Syngas Production from Municipal Solid Waste (MSW) Gasification via Reusing CO2. App. Eng. 2018, 144, 242–247. [Google Scholar] [CrossRef]
  14. Lebo, S.E., Jr.; Gargulak, J.D.; McNally, T.J. Lignin. In Kirk-Othmer Encyclopedia of Chemical Technology; American Cancer Society: Atlanta, GA, USA, 2001; ISBN 9780471238966. [Google Scholar]
  15. Chio, C.; Sain, M.; Qin, W. Lignin Utilization: A Review of Lignin Depolymerization from Various Aspects. Renew. Sustain. Energy Rev. 2019, 107, 232–249. [Google Scholar] [CrossRef]
  16. Li, C.; Zhao, X.; Wang, A.; Huber, G.W.; Zhang, T. Catalytic Transformation of Lignin for the Production of Chemicals and Fuels. Chem. Rev. 2015, 115, 11559–11624. [Google Scholar] [CrossRef] [PubMed]
  17. Medvedev, A.A.; Kustov, A.L.; Beldova, D.A.; Kravtsov, A.V.; Kalmykov, K.B.; Sarkar, B.; Kostyukhin, E.M.; Kustov, L.M. Gasification of Hydrolysis Lignin with CO2 in the Presence of Fe and Co Compounds. Mendeleev Commun. 2022, 32, 402–404. [Google Scholar] [CrossRef]
  18. Tian, H.; Wei, Y.; Cheng, S.; Huang, Z.; Qing, M.; Chen, Y.; Yang, H.; Yang, Y. Optimizing the Gasification Reactivity of Biochar: The Composition, Structure and Kinetics of Biochar Derived from Biomass Lignocellulosic Components and Their Interactions during Gasification Process. Fuel 2022, 324, 124709. [Google Scholar] [CrossRef]
  19. Zhang, S.; Yu, S.; Li, Q.; Mohamed, B.A.; Zhang, Y.; Zhou, H. Insight into the Relationship between CO2 Gasification Characteristics and Char Structure of Biomass. Biomass Bioenergy 2022, 163, 106537. [Google Scholar] [CrossRef]
  20. Carrier, M.; Windt, M.; Ziegler, B.; Appelt, J.; Saake, B.; Meier, D.; Bridgwater, A. Quantitative Insights into the Fast Pyrolysis of Extracted Cellulose, Hemicelluloses, and Lignin. ChemSusChem 2017, 10, 3212–3224. [Google Scholar] [CrossRef]
  21. Collard, F.-X.; Blin, J. A Review on Pyrolysis of Biomass Constituents: Mechanisms and Composition of the Products Obtained from the Conversion of Cellulose, Hemicelluloses and Lignin. Renew. Sustain. Energy Rev. 2014, 38, 594–608. [Google Scholar] [CrossRef]
  22. Guo, H.; Qi, Z.; Liu, Y.; Xia, H.; Li, L.; Huang, Q.; Wang, A.; Li, C. Tungsten-Based Catalysts for Lignin Depolymerization: The Role of Tungsten Species in C–O Bond Cleavage. Catal. Sci. Technol. 2019, 9, 2144–2151. [Google Scholar] [CrossRef]
  23. Deepa, A.K.; Dhepe, P.L. Lignin Depolymerization into Aromatic Monomers over Solid Acid Catalysts. ACS Catal. 2015, 5, 365–379. [Google Scholar] [CrossRef]
  24. Dabral, S.; Engel, J.; Mottweiler, J.; Spoehrle, S.S.M.; Lahive, C.W.; Bolm, C. Mechanistic Studies of Base-Catalysed Lignin Depolymerisation in Dimethyl Carbonate. Green Chem. 2018, 20, 170–182. [Google Scholar] [CrossRef]
  25. Huang, Y.-B.; Zhang, J.-L.; Zhang, X.; Luan, X.; Chen, H.-Z.; Hu, B.; Zhao, L.; Wu, Y.-L.; Lu, Q. Catalytic Depolymerization of Lignin via Transfer Hydrogenation Strategy over Skeletal CuZnAl Catalyst. Fuel Process. Technol. 2022, 237, 107448. [Google Scholar] [CrossRef]
  26. Zhang, Z.; Yin, G.; Andrioletti, B. Advances in Value-Added Aromatics by Oxidation of Lignin with Transition Metal Complexes. Transit. Metal. Chem. 2022, 47, 189–211. [Google Scholar] [CrossRef]
  27. Liu, S.; Li, Q.; Zuo, S.; Xia, H. Facile Synthesis of Lignosulfonate-Derived Sulfur-Doped Carbon Materials for Photocatalytic Degradation of Tetracycline under Visible-Light Irradiation. Microporous Mesoporous Mater. 2022, 336, 111876. [Google Scholar] [CrossRef]
  28. Upton, B.M.; Kasko, A.M. Strategies for the Conversion of Lignin to High-Value Polymeric Materials: Review and Perspective. Chem. Rev. 2016, 116, 2275–2306. [Google Scholar] [CrossRef]
  29. Ha, J.M.; Hwang, K.R.; Kim, Y.M.; Jae, J.; Kim, K.H.; Lee, H.W.; Kim, J.Y.; Park, Y.K. Recent Progress in the Thermal and Catalytic Conversion of Lignin. Renew. Sustain. Energy Rev. 2019, 111, 422–441. [Google Scholar] [CrossRef]
  30. van der Laan, G.P.; Beenackers, A.A.C.M. Kinetics and Selectivity of the Fischer–Tropsch Synthesis: A Literature Review. Catal. Rev. 1999, 41, 255–318. [Google Scholar] [CrossRef]
  31. Tsodikov, M.V.; Ellert, O.G.; Nikolaev, S.A.; Arapova, O.V.; Bukhtenko, O.V.; Maksimov, Y.V.; Kirdyankin, D.I.; Vasil’kov, A.Y. Fe-Containing Nanoparticles Used as Effective Catalysts of Lignin Reforming to Syngas and Hydrogen Assisted by Microwave Irradiation. J. Nanoparticle Res. 2018, 20, 86. [Google Scholar] [CrossRef]
  32. Kustov, L.M.; Tarasov, A.L.; Nissenbaum, V.D.; Kustov, A.L. Dry Reforming of Lignin: The Effect of Impregnation with Iron. Mendeleev Commun. 2021, 31, 376–378. [Google Scholar] [CrossRef]
  33. Tsodikov, M.D.; Ellert, O.G.; Arapova, O.V.; Nikolaev, S.A.; Chistyakov, A.V.; Maksimov, Y.V. Benefit of Fe-Containing Catalytic Systems for Dry Reforming of Lignin to Syngas under Microwave Radiation. Chem. Eng. Trans. 2018, 65, 367–372. [Google Scholar] [CrossRef]
  34. Hietala, J.; Vuori, A.; Johnsson, P.; Pollari, I.; Reutemann, W.; Kieczka, H. Formic Acid. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2016; pp. 1–22. ISBN 9783527316021. [Google Scholar]
  35. Omae, I. Aspects of Carbon Dioxide Utilization. Catal Today 2006, 115, 33–52. [Google Scholar] [CrossRef]
  36. Hunt, A.J.; Sin, E.H.K.; Marriott, R.; Clark, J.H. Generation, Capture, and Utilization of Industrial Carbon Dioxide. ChemSusChem 2010, 3, 306–322. [Google Scholar] [CrossRef] [PubMed]
  37. Tursunov, O.; Zubek, K.; Czerski, G.; Dobrowolski, J. Studies of CO2 Gasification of the Miscanthus Giganteus Biomass over Ni/Al2O3-SiO2 and Ni/Al2O3-SiO2 with K2O Promoter as Catalysts. J. Anal. Calorim 2020, 139, 3481–3492. [Google Scholar] [CrossRef]
  38. Bogdan, V.I.; Pokusaeva, Y.A.; Koklin, A.E.; Savilov, S.V.; Chernyak, S.A.; Lunin, V.V.; Kustov, L.M. Carbon Dioxide Reduction with Hydrogen on Carbon-Nanotube-Supported Catalysts under Supercritical Conditions. Energy Technol. 2019, 7, 1900174. [Google Scholar] [CrossRef]
  39. Pokusaeva, Y.A.; Koklin, A.E.; Lunin, V.V.; Bogdan, V.I. CO2 Hydrogenation on Fe-Based Catalysts Doped with Potassium in Gas Phase and under Supercritical Conditions. Mendeleev Commun. 2019, 29, 382–384. [Google Scholar] [CrossRef]
  40. Bogdan, V.I.; Koklin, A.E.; Kustov, A.L.; Pokusaeva, Y.A.; Bogdan, T.V.; Kustov, L.M. Carbon Dioxide Reduction with Hydrogen on Fe, Co Supported Alumina and Carbon Catalysts under Supercritical Conditions. Molecules 2021, 26, 2883. [Google Scholar] [CrossRef]
  41. Kurbatova, N.A.; El’man, A.R.; Bukharkina, T.V. Application of Catalysts to Coal Gasification with Carbon Dioxide. Kinet. Catal. 2011, 52, 739–748. [Google Scholar] [CrossRef]
  42. Huang, Y.; Yin, X.; Wu, C.; Wang, C.; Xie, J.; Zhou, Z.; Ma, L.; Li, H. Effects of Metal Catalysts on CO2 Gasification Reactivity of Biomass Char. Biotechnol Adv. 2009, 27, 568–572. [Google Scholar] [CrossRef]
  43. Kustov, L.M.; Tarasov, A.L.; Kustov, A.L. Steam Reforming of Lignin Modified with Iron. Mendeleev Commun. 2020, 30, 76–77. [Google Scholar] [CrossRef]
  44. Tarasov, A.L.; Kostyukhin, E.M.; Kustov, L.M. Gasification of Metal-Containing Coals and Carbons via Their Reaction with Carbon Dioxide. Mendeleev Commun. 2018, 28, 530–532. [Google Scholar] [CrossRef]
  45. Domazetis, G.; Liesegang, J.; James, B.D. Studies of Inorganics Added to Low-Rank Coals for Catalytic Gasification. Fuel Process. Technol. 2005, 86, 463–486. [Google Scholar] [CrossRef]
  46. Zsinka, V.; Miskolczi, N.; Juzsakova, T.; Jakab, M. Pyrolysis-Gasification of Biomass Using Nickel Modified Catalysts: The Effect of the Catalyst Regeneration on the Product Properties. J. Energy Inst. 2022, 105, 16–24. [Google Scholar] [CrossRef]
  47. Zhao, Y.; Deng, L.; Liao, B.; Fu, Y.; Guo, Q.-X. Aromatics Production via Catalytic Pyrolysis of Pyrolytic Lignins from Bio-Oil. Energy Fuels 2010, 24, 5735–5740. [Google Scholar] [CrossRef]
  48. Wang, Y.; Yao, L.; Wang, S.; Mao, D.; Hu, C. Low-Temperature Catalytic CO2 Dry Reforming of Methane on Ni-Based Catalysts: A Review. Fuel Process. Technol. 2018, 169, 199–206. [Google Scholar] [CrossRef]
  49. Yao, L.; Shi, J.; Xu, H.; Shen, W.; Hu, C. Low-Temperature CO2 Reforming of Methane on Zr-Promoted Ni/SiO2 Catalyst. Fuel Process. Technol. 2016, 144, 1–7. [Google Scholar] [CrossRef]
  50. Kustov, L.M.; Tarasov, A.L.; Tkachenko, O.P.; Kapustin, G.I. Nickel–Alumina Catalysts in the Reaction of Carbon Dioxide Re-Forming of Methane under Thermal and Microwave Heating. Ind. Eng. Chem. Res. 2017, 56, 13034–13039. [Google Scholar] [CrossRef]
  51. Lv, C.; Xu, L.; Chen, M.; Cui, Y.; Wen, X.; Li, Y.; Wu, C.; Yang, B.; Miao, Z.; Hu, X.; et al. Recent Progresses in Constructing the Highly Efficient Ni Based Catalysts With Advanced Low-Temperature Activity Toward CO2 Methanation. Front. Chem. 2020, 8, 269. [Google Scholar] [CrossRef]
  52. Lobo, L.S.; Carabineiro, S.A.C. Kinetics and Mechanism of Catalytic Carbon Gasification. Fuel 2016, 183, 457–469. [Google Scholar] [CrossRef]
  53. Kapteijn, F.; Porre, H.; Moulijn, J.A. CO2 Gasification of Activated Carbon Catalyzed by Earth Alkaline Elements. AIChE J. 1986, 32, 691–695. [Google Scholar] [CrossRef]
Figure 1. Dependence of the carbon dioxide conversion on temperature in the process of carbon dioxide assisted conversion for samples of the initial hydrolysis lignin and samples of hydrolysis lignin with deposited nickel compounds: 1%, 3%, 5%, and 7% by weight.
Figure 1. Dependence of the carbon dioxide conversion on temperature in the process of carbon dioxide assisted conversion for samples of the initial hydrolysis lignin and samples of hydrolysis lignin with deposited nickel compounds: 1%, 3%, 5%, and 7% by weight.
Energies 15 06774 g001
Figure 2. Micrographs (a,c,e,g) and nickel mapping (b,d,f,h) of the samples of hydrolysis lignin with a mass fraction of nickel of 1%, 3%, 5%, and 7%, respectively.
Figure 2. Micrographs (a,c,e,g) and nickel mapping (b,d,f,h) of the samples of hydrolysis lignin with a mass fraction of nickel of 1%, 3%, 5%, and 7%, respectively.
Energies 15 06774 g002
Figure 3. X-ray diffraction patterns of the original lignin (a), lignin coated with nickel (II) nitrate (7% wt. Ni) after calcination (b), and residues after carbon dioxide conversion (7% wt. Ni) (c).
Figure 3. X-ray diffraction patterns of the original lignin (a), lignin coated with nickel (II) nitrate (7% wt. Ni) after calcination (b), and residues after carbon dioxide conversion (7% wt. Ni) (c).
Energies 15 06774 g003
Figure 4. TEM micrographs of a sample of hydrolysis lignin coated with nickel compounds (7% wt.) after calcination in a CO2 atmosphere.
Figure 4. TEM micrographs of a sample of hydrolysis lignin coated with nickel compounds (7% wt.) after calcination in a CO2 atmosphere.
Energies 15 06774 g004
Figure 5. Distribution of nickel particle sizes on the surface of hydrolysis lignin after calcination in a CO2 atmosphere according to the TEM data.
Figure 5. Distribution of nickel particle sizes on the surface of hydrolysis lignin after calcination in a CO2 atmosphere according to the TEM data.
Energies 15 06774 g005
Figure 6. Transmission electron micrographs of a sample of the initial hydrolysis lignin after gasification.
Figure 6. Transmission electron micrographs of a sample of the initial hydrolysis lignin after gasification.
Energies 15 06774 g006
Figure 7. Transmission electron micrographs of a sample of hydrolysis lignin coated with nickel compounds (7% wt.) after gasification.
Figure 7. Transmission electron micrographs of a sample of hydrolysis lignin coated with nickel compounds (7% wt.) after gasification.
Energies 15 06774 g007
Figure 8. Elemental analysis of images obtained during the study using a transmission electron microscope of the original lignin after the process (a), post-process nickel-supported lignin (7 wt. % Ni) (b), and lignin coated with nickel compounds after calcination (7 wt. % Ni) (c).
Figure 8. Elemental analysis of images obtained during the study using a transmission electron microscope of the original lignin after the process (a), post-process nickel-supported lignin (7 wt. % Ni) (b), and lignin coated with nickel compounds after calcination (7 wt. % Ni) (c).
Energies 15 06774 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Medvedev, A.A.; Beldova, D.A.; Kalmykov, K.B.; Kravtsov, A.V.; Tedeeva, M.A.; Kustov, L.M.; Dunaev, S.F.; Kustov, A.L. Carbon Dioxide Assisted Conversion of Hydrolysis Lignin Catalyzed by Nickel Compounds. Energies 2022, 15, 6774. https://doi.org/10.3390/en15186774

AMA Style

Medvedev AA, Beldova DA, Kalmykov KB, Kravtsov AV, Tedeeva MA, Kustov LM, Dunaev SF, Kustov AL. Carbon Dioxide Assisted Conversion of Hydrolysis Lignin Catalyzed by Nickel Compounds. Energies. 2022; 15(18):6774. https://doi.org/10.3390/en15186774

Chicago/Turabian Style

Medvedev, Artem A., Daria A. Beldova, Konstantin B. Kalmykov, Alexey V. Kravtsov, Marina A. Tedeeva, Leonid M. Kustov, Sergey F. Dunaev, and Alexander L. Kustov. 2022. "Carbon Dioxide Assisted Conversion of Hydrolysis Lignin Catalyzed by Nickel Compounds" Energies 15, no. 18: 6774. https://doi.org/10.3390/en15186774

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