Next Article in Journal
Efficient Synthesis of 4,8-Dibromo Derivative of Strong Electron-Deficient Benzo[1,2-d:4,5-d’]bis([1,2,3]thiadiazole) and Its SNAr and Cross-Coupling Reactions
Previous Article in Journal
Progression of Antiviral Agents Targeting Viral Polymerases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

New Insight into CO2 Reduction to Formate by In Situ Hydrogen Produced from Hydrothermal Reactions with Iron

State Key Laboratory of Pollution Control and Resources Reuse, College of Environmental Science and Engineering, Tongji University, 1239 Siping Road, Shanghai 200092, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(21), 7371; https://doi.org/10.3390/molecules27217371
Submission received: 23 June 2022 / Revised: 13 July 2022 / Accepted: 20 July 2022 / Published: 29 October 2022

Abstract

:
To reveal the nature of CO2 reduction to formate with high efficiency by in situ hydrogen produced from hydrothermal reactions with iron, DFT calculations were used. A reaction pathway was proposed in which the formate was produced through the key intermediate species, namely iron hydride, produced in situ in the process of hydrogen gas production. In the in situ hydrogenation of CO2, the charge of H in the iron hydride was −0.135, and the Fe–H bond distance was approximately 1.537 Å. A C-H bond was formed as a transition state during the attack of Hδ− on Cδ+. Finally, a HCOO species was formed. The distance of the C-H bond was 1.107 Å. The calculated free energy barrier was 16.43 kcal/mol. This study may provide new insight into CO2 reduction to formate in hydrothermal reactions with metal.

1. Introduction

Global temperature has risen noticeably during the past few decades, mainly due to CO2 emissions [1]. As a result, many negative effects climate change affairs have been felt, which has caused a severe threat to human survival. Fortunately, many efforts have been made to reduce CO2 emissions. For example, the transformation of CO2 into chemicals or fuels has received much attention, which would not only reduce CO2 emissions but also alleviate the shortage of fossil fuels [2]. Therefore, establishing the most commercially feasible method for the hydrogenation of CO2 has become highly desirable [3]. However, the high kinetic and thermodynamic stability of CO2 hindered the reduction of CO2 emissions. Solar technologies are considered to be the ideal solutions to the greenhouse effect problem. However, the utilization of solar energy normally has the disadvantages of low efficiency and low production selectivity. On the other hand, the hydrogenation of CO2 with gaseous hydrogen is regarded as the most commercially feasible option. However, the gaseous hydrogen process is expensive due to the costs of production, storage, and transportation. Therefore, once the in situ produced hydrogen can be used for the hydrogenation of CO2, the problem of hydrogen storage and transportation can be solved easily. Thus, establishing an efficient CO2 conversion process is a necessary and fascinating challenge [4,5].
Hydrothermal reactions play an important role in the formation of fossil fuels, due to their special characteristics such as their low dielectric constant and high ion product [6]. Water under high temperature and high pressure can be used as a reaction medium and hydrogen source. The production of formate from CO2 reduction under hydrothermal conditions has been reported extensively [7,8,9]. Recently, a two-step CO2 reduction process based on water-splitting hydrogen production with the redox cycle of metals/metal oxides has been reported (for example, in relation to Fe/Fe3O4) [10]. In this process, the iron was used as a reductant, and the iron oxides were used to reproduce iron using solar energy, which is a kind of renewable energy. The hydrogen production efficiency from the dissociation of water under hydrothermal conditions was much higher than that with solar energy [11]. Thus, hydrogen production from hydrothermal reactions was assumed to be a sustainable method for the reduction of CO2. Consequently, the dissociation of water with metal under hydrothermal conditions would represent one of the most promising approaches to decrease CO2 emissions. Obviously, the conversion of CO2 to chemicals or fuels is the key step in this process. However, the hydrogenation of CO2 in hydrothermal reactions has not been studied in sufficient detail.
In recent years, we have studied extensively the metal hydrothermal reaction for CO2 reduction. Zinc hydride (Zn–H) is reported as a key intermediate species during the production of formate. The formate was produced through an SN2-like reaction [12]. In an experimental study, Zn–H species, produced from the reaction of Zn and H2O, were detected by using FT-IR and XPS analysis. In a theoretical study, the Zn–H species was detected as a key intermediate by using the quantum chemical calculations method. The reaction pathway detection can be used to explain the production of formate with Zn in hydrothermal reactions.
In our previous experimental study, we investigated the dissociation of H2O for reducing CO2 (or NaHCO3 as a CO2 source) to formic acid with metallic iron [13]. In this process, H2O is used as a source of hydrogen and Fe is used as a reductant, which can rapidly produce hydrogen under hydrothermal conditions. With the in situ produced hydrogen, CO2 was reduced to formic acid efficiently. After the reaction, Fe was oxidized to Fe3O4. It was also reported that NaHCO3 was reduced into formate with high efficiency by using Fe as a reductant [14]. The overall equation can be shown as:
3 Fe + NaHCO 3 + 3 H 2 O Fe 3 O 4 + 3 H 2 ( g ) + HCOONa
Δ G ( 573   K ) = 132.776   kJ mol 1
In this study, a theoretical study was conducted by using the DFT calculations. Under hydrothermal reaction conditions, the in situ technique for experimentally studying the reaction mechanism of formate production is very hard to use. However, the ab initio calculation is very useful for studying the mechanism including reliable structures as well as the energies of the reactants, products, intermediates, and transition states.

2. Computational Details

DFT calculations, including a consideration of van der Waals (VDW) interactions, were adopted by using Gaussian 16 [15]. The B3LYP density functional method [16] with the D3 (GJ) dispersion correction was employed to carry out all of the computations. An SDD basis set was used for Fe atoms and a 6-31G+G(d,p) basis set was used for remaining atoms. Polarizable continuum model (PCM) calculations of water as the solvent were used in this work. Vibrational frequency analyses were conducted at the same level of theory to characterize stationary points as local minima. Intrinsic reaction coordinate (IRC) computations were carried out to confirm the transition states.

3. Results and Discussion

3.1. Analysis of Geometric Parameters

The geometric parameters of initial state and intermediate states are shown in Figure 1. Bond lengths are in Å and bond angles are in degrees. As shown in Figure 1, in the initial state, the bond distance of Fe–H was approximately 1.537 Å, and the Mulliken charge of H in Fe–H was −0.135. This resulting charge means that a hydride was formed, not a proton. The geometry and Mulliken charge distribution implied that the Fe–H species was similar to our previous study with zinc. Since metal hydride has high catalytic activity, it could be used as a reduction agent. The Fe–O bond distances between Fe and HCO3- was approximately 2.0~2.1 Å. When the intermediate state is formed, the Fe–H bond distance was about 1.482 Å. The Mulliken charge of H in Fe–H was 0.022. The C–H bond distance was about 1.905 Å. In the initial state, the Fe–Hδ− species was formed via the dissociation of water. Based on the calculation results, it was assumed that the CO2 was produced from bicarbonate under hydrothermal conditions as an intermediate. Then, the intermediate state was used for the calculation for the formate production.
The geometric parameters of the transition state and the final state are shown in Figure 2. The Fe–H bond distance in the transition state was approximately 1.505 Å, and the Mulliken charge of H in the iron hydride species was 0.029, as shown in Figure 2. The distance between C and H atoms was 1.569 Å. By comparing them with the distance in the intermediate state, one can see that the distance is closer. The distance between metal and H atoms is similar to our previous study on CO2 reduction by zinc, which was 1.490 Å. The Fe–O bond distance was approximately 1.85~1.86 Å. In the final state, the bond distance of C–H was 1.107 Å. The Mulliken charge of H in the C–H species was 0.099. Compared with the distance in the intermediate and the transition state, the distance of C–H bond was closer, at only 1.107 Å. It means that the C-H bond is stronger, and the final state is more stable. The Fe–O bond distances were approximately 1.8~1.9 Å, and the charge of Fe was 0.674. The bond angle between the H–C–O bond was 115.5°. Based on these bond distance results, we proposed the mechanism of HCOO production as follows. In the hydrogen production process, the Fe–Hδ− species was formed via the dissociation of water. Due to the high reducibility of Hδ−, the distance between H and C atoms became closer. When a transition state during the attack of Hδ− on Cδ+ formed, the HCOO species could be produced subsequently.

3.2. Energy Diagram for HCOO Production

An energy diagram of the C–H bond formation is shown in Figure 3. The optimized geometry of the initial, intermediate, and final states is depicted in Figure 1 and Figure 2. As shown in Figure 3, the calculated activation energy is 16.43 kcal/mol. Compared with our previous study, when HCO3 was used as the initial state, the calculated activation energy was 24.1 kcal/mol, which means that the CO2 was easier to reduce than HCO3 under reaction conditions. An important implication of the calculated activation energy is that the in situ produced Fe–H species has high activity regarding reduction properties.

3.3. IRC and HOMO/LUMO Calculation of Transition State

The IRC calculations and the HOMO and LUMO orbital shapes of the transition state (TS) are shown in Figure 4 and Figure 5, respectively. The IRC curve is smooth, as shown in Figure 4. The formate was produced in the vertex of the curve. It can easily be seen that the the reaction energy barrier was not very high, which means that formate could easily produced once the Fe–H species was formed. As shown in Figure 5, the occupied HOMO of the transition state exhibits a bonding interaction between the C atom and the H atom of the iron hydride species. On the other hand, the unoccupied LUMO of the transition state shows an anti-bonding character. These results clearly show the formation of a C–H bond of the formate.

4. Conclusions

In this study, CO2 reduction to formate by in situ hydrogen produced from hydrothermal reactions with iron was investigated through a theoretical study to provide new insights. A possible reaction pathway was proposed in which iron hydride, produced in situ under hydrothermal conditions, was a key intermediate species in formate production. The structures and energies were obtained, and the IRC and HOMO/LUMO orbit shapes were built using the 6-31G+G(d,p) basis set. In the hydrogen production process, the Fe–Hδ− species was formed via the dissociation of water. In the in situ hydrogenation process of CO2, the Mulliken charge of H in the iron hydride was −0.135, and the Fe–H bond distance was approximately 1.537 Å. A C–H bond was formed as a transition state during the attack of Hδ− on Cδ+. Finally, a HCOO species was formed. The distance of the C-H bond was 1.107 Å. The calculated free energy barrier was 16.43 kcal/mol. These results showed that the reaction could take place with a lower formation barrier. This study may provide new insight into CO2 reduction to formate in hydrothermal reactions with metal, which will promote new ideas in the field of CO2 utilization.

Author Contributions

Conceptualization, X.Z.; methodology, X.Z.; software, X.Z. and G.-D.Y.; validation, X.Z.; investigation, X.Z.; writing—original draft preparation, X.Z. and Y.-Y.Z.; writing—review and editing, X.Z. and J.-F.Z.; supervision, J.-F.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by State Key Laboratory of Pollution Control and Resource Reuse Foundation of China (No. PCRRE20001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

All authors thank the financial support of the State Key Laboratory of Pollution Control and Resource Reuse.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meng, M.; Niu, D. Modeling CO2 emissions from fossil fuel combustion using the logistic equation. Energy 2011, 36, 3355–3359. [Google Scholar] [CrossRef]
  2. Wang, W.; Wang, S.; Ma, X.; Gong, J.L. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Huff, C.A.; Sanford, M.S. Cascade catalysis for the homogeneous hydrogenation of CO2 to methanol. J. Am. Chem. Soc. 2011, 133, 18122–18125. [Google Scholar] [CrossRef] [PubMed]
  4. Roy, S.C.; Varghese, O.K.; Paulose, M.; Grimes, C.A. Toward solar fuels: Photocatalytic conversion of carbon dioxide to hydrocarbons. ACS Nano 2010, 4, 1259–1278. [Google Scholar] [CrossRef] [PubMed]
  5. Takeda, H.; Koike, K.; Inoue, H.; Ishitani, O. Development of an efficient photocatalytic system for CO2 reduction using rhenium(I) complexes based on mechanistic studies. J. Am. Chem. Soc. 2008, 130, 2023–2031. [Google Scholar] [CrossRef] [PubMed]
  6. Jin, F.M.; Enomoto, H. Rapid and highly selective conversion of biomass into value-added products in hydrothermal conditions: Chemistry of acid/base-catalysed and oxidation reactions. Energy Environ. Sci. 2011, 4, 382–397. [Google Scholar] [CrossRef]
  7. Takahashi, H.; Kori, T.; Onoki, T.; Tohji, K.; Yamasaki, N. Hydrothermal processing of metal based compounds and carbon dioxide for the synthesis of organic compounds. J. Mater. Sci. 2008, 43, 2487–2491. [Google Scholar] [CrossRef]
  8. Tian, G.; He, C.; Chen, Y.; Yuan, H.M.; Liu, Z.W.; Shi, Z.; Feng, S.H. Hydrothermal Reactions from Carbon Dioxide to Phenol. ChemSusChem 2010, 3, 323–324. [Google Scholar] [CrossRef] [PubMed]
  9. Martin, G.; Gabor, L. Formic Acid as a Hydrogen Source-Recent Developments and Future Trends. Energy Environ. Sci. 2012, 5, 8171–8181. [Google Scholar]
  10. Galvez, M.E.; Loutzenhiser, P.G.; Hischier, I.; Steinfeld, A. CO2 splitting via two-step solar thermochemical cycles with Zn/ZnO and FeO/Fe3O4 redox reactions: Thermodynamic analysis. Energy Fuels 2008, 22, 3544–3550. [Google Scholar] [CrossRef]
  11. Hu, Y.H. A highly efficient photocatalyst hydrogenated black TiO2 for the photocatalytic splitting of water. Angew. Chem. Int. Ed. 2012, 51, 12410–12412. [Google Scholar] [CrossRef] [PubMed]
  12. Zeng, X.; Hatakeyama, M.; Ogata, K.; Liu, J.K.; Wang, Y.Q.; Gao, Q.; Fujii, K.; Fujihira, M.; Jin, F.M.; Nakamura, S. New insights into highly efficient reduction of CO2 to formic acid by using zinc under mild hydrothermal conditions: A joint experimental and theoretical study. Phys. Chem. Chem. Phys. 2014, 16, 19836–19840. [Google Scholar] [CrossRef] [PubMed]
  13. Zhong, H.; Gao, Y.; Yao, G.D.; Zeng, X.; Li, Q.J.; Huo, Z.B.; Jin, F.M. Highly efficient water splitting and carbon dioxide reduction into formic acid with iron and copper powder. Chem. Eng. J. 2015, 280, 215–221. [Google Scholar] [CrossRef]
  14. Duo, J.; Jin, F.M.; Wang, Y.Q.; Zhong, H.; Lyu, L.Y.; Yao, G.D.; Huo, Z.B. NaHCO3-enhanced hydrogen production from water with Fe and in situ highly efficient and autocatalytic NaHCO3 reduction into formic acid. Chem. Commun. 2016, 52, 3316–3319. [Google Scholar] [CrossRef] [PubMed]
  15. Frisch, M.E.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.P.G.A.; Petersson, G.A.; Nakatsuji, H.J.R.A.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, UK, 2016. [Google Scholar]
  16. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
Figure 1. Optimized geometric parameters of the initial state and the intermediate state.
Figure 1. Optimized geometric parameters of the initial state and the intermediate state.
Molecules 27 07371 g001
Figure 2. Optimized geometric parameters of the transition state and the final state.
Figure 2. Optimized geometric parameters of the transition state and the final state.
Molecules 27 07371 g002
Figure 3. Potential energy diagram for HCOO production.
Figure 3. Potential energy diagram for HCOO production.
Molecules 27 07371 g003
Figure 4. Intrinsic reaction coordinate calculation results.
Figure 4. Intrinsic reaction coordinate calculation results.
Molecules 27 07371 g004
Figure 5. HOMO and LUMO orbital shapes of the transition state.
Figure 5. HOMO and LUMO orbital shapes of the transition state.
Molecules 27 07371 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zeng, X.; Yin, G.-D.; Zhou, Y.-Y.; Zhao, J.-F. New Insight into CO2 Reduction to Formate by In Situ Hydrogen Produced from Hydrothermal Reactions with Iron. Molecules 2022, 27, 7371. https://doi.org/10.3390/molecules27217371

AMA Style

Zeng X, Yin G-D, Zhou Y-Y, Zhao J-F. New Insight into CO2 Reduction to Formate by In Situ Hydrogen Produced from Hydrothermal Reactions with Iron. Molecules. 2022; 27(21):7371. https://doi.org/10.3390/molecules27217371

Chicago/Turabian Style

Zeng, Xu, Guo-Dong Yin, Yang-Yuan Zhou, and Jian-Fu Zhao. 2022. "New Insight into CO2 Reduction to Formate by In Situ Hydrogen Produced from Hydrothermal Reactions with Iron" Molecules 27, no. 21: 7371. https://doi.org/10.3390/molecules27217371

Article Metrics

Back to TopTop