Next Article in Journal
Bacterial Nanocellulose in Dentistry: Perspectives and Challenges
Next Article in Special Issue
Substituent Effects on the Solubility and Electronic Properties of the Cyanine Dye Cy5: Density Functional and Time-Dependent Density Functional Theory Calculations
Previous Article in Journal
In Silico and In Vitro Analysis of Major Cannabis-Derived Compounds as Fatty Acid Amide Hydrolase Inhibitors
Previous Article in Special Issue
Computational Insights into Allosteric Conformational Modulation of P-Glycoprotein by Substrate and Inhibitor Binding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogenation of CO2 Promoted by Silicon-Activated H2S: Origin and Implications

College of Chemistry and Chemical Engineering, Southwest University, Chongqing 400715, China
Molecules 2021, 26(1), 50; https://doi.org/10.3390/molecules26010050
Submission received: 3 November 2020 / Revised: 19 December 2020 / Accepted: 21 December 2020 / Published: 24 December 2020
(This article belongs to the Special Issue 25th Anniversary of Molecules—Recent Advances in Physical Chemistry)

Abstract

:
Unlike the usual method of COx (x = 1, 2) hydrogenation using H2 directly, H2S and HSiSH (silicon-activated H2S) were selected as alternative hydrogen sources in this study for the COx hydrogenation reactions. Our results suggest that it is kinetically infeasible for hydrogen in the form of H2S to transfer to COx at low temperatures. However, when HSiSH is employed instead, the title reaction can be achieved. For this approach, the activation of CO2 is initiated by its interaction with the HSiSH molecule, a reactive species with both a hydridic Hδ− and protonic Hδ+. These active hydrogens are responsible for the successive C-end and O-end activations of CO2 and hence the final product (HCOOH). This finding represents a good example of an indirect hydrogen source used in CO2 hydrogenation through reactivity tuned by silicon incorporation, and thus the underlying mechanism will be valuable for the design of similar reactions.

Graphical Abstract

1. Introduction

In the past two decades, global warming caused by excessive CO2 emission has received much attention [1,2]. An effective way to alleviate this environmental problem is to reduce carbon dioxide to value-added chemicals through hydrogen [3,4]. In addition to hydrogen, some indirect hydrogen sources such as silanes have also been used to achieve this goal [5]. Can hydrogen sulfide (H2S), as an indirect hydrogen source, work as well? If so, it will be very important from both environmental and public safety perspectives since H2S is known as one of the most toxic, corrosive, and malodorous gases in various industrial processes [6,7]. Research in this field is still in its infancy compared with the area of COx hydrogenation reactions that use H2 directly [8,9,10]. Nevertheless, I note that H2S has previously been modeled to act as a source of hydrogen atoms in the first excited state, and two pathways leading to HCOO + SH and HCOOH + S have been revealed [11]. However, UV radiation is required to trigger these reactions, which is hardly cost effective. Experimentally, although the reduction of CO2 with H2S in a simulated deep sea hydrothermal vent system has been observed [12], it is not H2S but H2O that acts as a hydrogen donor in HCO3- hydrogenation to formate. Different roles of H2S during CO2 hydrogenation on the MoS2 catalyst have also been probed [13]. It was observed that under high co-feeds conditions, H2S could react with CO2 to form COS. The same product has also been reported for the H2S + CO2 reaction under a metal free dual organocatalyst [14]. The emerging role of COS deserves attention. In the presence of H2, this intermediate competes well with CO2 or CO (in H2S + CO reaction) [15,16,17] in the hydrogenation step to produce CH3SH [18] and CH3SCH3 [19] molecules. Clearly, if active hydrogens (from H2S) are available for COx hydrogenation, the formation of COS intermediate should be minimized. So, how to ensure this when H2S is employed as a potential hydrogen source? In a recent matrix isolation study [20], it was found that Si atoms can react spontaneously with H2S to produce a HSiSH molecule that is stable until exposure to UV radiation. If the two hydrogens (from HSiSH) are active toward COx hydrogenation, SiS, a very stable molecule with Si=S multiple bonds [21,22,23], will be one of the final products, which suggests little chance for the production of COS, and hence contributes to the target reaction.
In this study, with the aid of quantum chemical calculations, the reactions of H2S and HSiSH with COx were investigated. I found that silicon incorporation increases the dehydrogenation of H2S while passivating the sulfur atom, which increases the possibility of CO2 hydrogenation. Furthermore, the reaction mechanisms, especially the structures and interconversion of some important reactive intermediates in the relevant reactions, were explored in detail, thus offering new ideas for the design of similar reactions involving S-H and C=O bond activation and transformation.

2. Results

The global minimum (Figure 1) obtained on the H2S + CO potential energy surface (PES) was H2 + COS, which was computed to be exothermic by 0.6 kcal·mol−1 relative to the entry channel. However, similar calculations confirmed an H2S···CO2 complex as a global minimum, indicating the inertness of CO2 in the hydrogen transfer reaction. Interestingly, when HSiSH was used instead, C-H bonds were formed for the most stable species (Figure 1, right panel), exhibiting prominent hydrogen transfer characteristics. Are these structures kinetically accessible? I first explored the energy landscape to access H2 + COS, which suggested a preference for breaking the S-H and C-H bonds for H2 production (Figure 2, upper panel). In this pathway, three sequential steps are involved: H2S addition to CO to give trans-HCOSH; isomerization of this molecule to cis-HCOSH; and finally, cleavage of the C-H and S-H bonds to afford H2 + COS. The high activation energy predicted for this reaction implies that it is kinetically infeasible for hydrogens in the form of H2S to transfer to CO at low temperatures, and hence explains the difficulty in generating the H2 + COS species. For completeness, the pathways leading to stable but much less favorable products like CS + H2O were also explored, and are summarized in Figure S3 in the Supporting Information (SI).
Although the H2S···CO2 complex is the most stable species on the CO2 + H2S PES, the reaction channels leading to the COS + H2O product were explored and are presented in Figure 2 (lower panel) and Figure S6. It is clear that the initial step of hydrogen transfer that occurs during CO2 reduction is energy demanding (48.7 kcal·mol−1), which is comparable to the value of 49.6 kcal·mol−1 reported in a previous study [11]. Additionally, the step for COS···H2O production is difficult at low temperatures. With an efficient catalyst [14], these barriers can be reduced significantly. However, the conversion of COx to COS makes it less likely that the COx hydrogenation reaction will occur. Furthermore, the active hydrogens are lost in the form of H2 or H2O. Interestingly, when H2S is replaced with HSiSH, disparate behaviors are predicted toward COx activation. In the case of CO, a global minimum with a four-membered ring can form through concerted double hydrogen transfer from HSiSH to CO (Figure 3). Since the hydrogen transfer involves simultaneous HSiSH dehydrogenation and CO hydrogenation, “coupled reaction pathway” is employed here to denote it. If the hydrogen is released (from HSiSH) instead, “decoupled reaction pathway” is more appropriate. In the coupled reaction pathway, an alternative step to produce the kinetically more stable SiS + HCHO is clear. However, competitive channels for hydrogen release (in the decoupled reaction pathway) are, from an energetic point of view, more effective (Figure 3, lower panel), and CO even acts as a catalyst in this reaction. As can be inferred from the figure, the Gibbs free energy barrier for two-step dehydrogenation is about 18.8 kcal·mol−1 on average while that for one-step dehydrogenation amounts to 34.8 kcal·mol−1. In other words, the thermodynamically favored hydrogen transfer reaction to reach the global minimum is decoupled or at least partially decoupled for H2 production. As for the global minimum for CO2, the two pathways toward its formation are presented in the SI (Figure S13), and the rate-limiting step is appreciably higher: at least 57.7 kcal·mol−1 at the M06-2X/aug-cc-pVTZ level, revealing the difficulties in its generation.
Remarkably, HSiSH exhibits a significant reactivity toward CO2 in the optimal path (Figure 4). It can easily form a precursor complex with CO2 by ƞ1-O coordination, and a short Si-O bond distance (2.855 Å) is obtained as expected. The subsequent attack by Si-H on the carbon of CO2 appears to be key to initiating CO2 reduction. In this step, a transition state (TS) containing a near-planar ring as well as a bent CO2 is predicted. The contraction of Si-O and the elongation of both Si-H and C=O in this TS structure are consistent with the activation of CO2, which then produces an insertion product with a HCO2 moiety. It is worth noting that the Gibbs free energy barrier (15.6 kcal·mol−1) to reach this TS is the highest among the saddle points, indicating that the hydrosilylation of CO2 is the rate-limiting step.
To fulfill the reductive elimination of HCOOH, two successive rotations—one of the OCH moiety around the C-O bond and the other of the thiol group around the Si-S bond—are required. Such structural rearrangements give a reactive intermediate with a short O···H distance (2.120 Å) and small O-C-O-H dihedral angle (21.0°). Clearly, these geometric features are favorable for subsequent O-end activation of the formate ligand, and it is therefore not surprising that the Gibbs free energy barrier to reach the last TS is only 9.2 kcal·mol−1. The result is then a η2-O,H-dihapto complex, as shown in Figure 4, where the oxygen and hydrogen of the formate ligand are bound to the silicon atom and sulfur atom, respectively. Free HCOOH can be liberated from this complex with an energy demand of 9.6 kcal·mol−1. Although this optimal path is slightly exothermic (0.2 kcal·mol−1), it should be mentioned that this energy does not include the silicon incorporation step. In fact, the insertion of an Si atom into the S-H bond of H2S is spontaneous and highly exothermic (66.2 kcal·mol−1 at the CCSD(T)/6-311++g(3df,3pd)//B3LYP/6-311++g(3df,3pd) level) as proved in a previous study [20].
The pathways explored in this reaction can simply be classified into three categories according to the initial activation site(s): C-end activation pathways, O-end activation pathways, and double-end (C, O) activation pathways. The optimal pathway associated with C-end activation, as presented above, involves an HCOO intermediate and gives HCOOH as the final product. Figure S12 in the SI illustrates some less favorable products in this category; details for their formation are summarized in Figure S13. As shown in the figure, SSiO + HCHO, SiO2 + H2CS, or SiO + HCOSH lies even above the TS in the rate-limiting step of the optimal path, suggesting the unavailability of these species. Although SiS + H2 + CO2 is more stable than SiS + HCOOH by 3.5 kcal·mol−1, the predicted rise in the barrier height shows again the preference for HCOOH species. In addition to the C-end activation pathway, HCOOH can, alternatively, form through the double-end activation pathway as shown above. However, the pathway in the category of O-end activation produces predominantly SiS + CO + H2O via a COOH intermediate (SI, Figures S14 and S15). There are, indeed, several molecules that one can conceive as products in this category. However, as far as the thermodynamics of the reaction are concerned, SSiO + HCHO, SSiO + CO + H2, and SiO + CO + H2S all lie above SiS + CO + H2O by 8.3, 7.1, and 4.7 kcal·mol−1, respectively. Besides, the channels leading to these species are kinetically less favored (SI, Figure S15). A similar situation is encountered for products HCOOH + SiS and SiS + CO2 + H2 though they are more stable than SiS + CO + H2O by 7.3 and 10.8 kcal·mol−1, respectively (SI, Figure S15). Since the activation energies of O-end activation and double-end (C, O) activation pathways are approximately 10.0 kcal·mol−1 higher than that of C-end activation pathways, and the interconversion between the HCOO and COOH intermediates encounters a large Gibbs free energy barrier (M06-2X/aug-cc-pVTZ: 103.7 kcal·mol−1), the efficient production of formic acid can be expected.

3. Discussion

Without the incorporation of Si into H2S, the H2S···CO2 complex is calculated as the global minimum on the PES, and triggering the reaction at the hydrogen transfer step is predicted to be difficult. However, when H2S is replaced by HSiSH, the successive C-end and O-end activation of CO2 can proceed smoothly to produce HCOOH with appreciably lower Gibbs free energy barriers. What is the difference between H2S and HSiSH that accounts for these distinct chemical behaviors? I first note that the incorporation of Si into H2S forms Si-H bonds, and hydridic Hδ− may be generated by the electronegativity difference between the Si and H atoms. Interestingly, previous studies have shown that the presence of a negatively charged hydride is crucial for the activation of CO2 [24], even if it is from the lattice [25]. I thus computed the atomic net charge distribution for the HSiSH molecule (Table 1), which confirmed the hydridic character of the hydrogen bonded directly to silicon. In addition, the condensed dual descriptors are presented in the same table, which allow for rationalization of the reactivity in terms of the simultaneous knowledge of the nucleophilic and electrophilic sites within a molecule. It was found that one of the hydrogen atoms experiences a substantial change from electrophilic (in H2S) to nucleophilic (in HSiSH) with the incorporation of Si into H2S, and one new electrophilic site (Si) is created as well (Table 1 and Figure 5). The combination of these two sites is exactly the Si-H chemical bond, which initiates the C-end activation of CO2 that would otherwise be O-end activation without silicon incorporation.
It is worth mentioning here that much attention has been paid to the Si-H bond in the field of CO2 activation recently [26,27,28,29,30]. Another benefit from incorporating silicon into H2S is the formation of Si=S multiple bonds when the protonic Hδ+ attacks HCOO. Not only can this avoid COS formation, but it also gives HCOOH as an important product [31] without requiring much extra energy. Considering that the activation energy of the rate-limiting step may be affected by the δ+M − Hδ− bond polarity [32], lead, as a heavier element in the same group as silicon, was used instead to test this effect. Interestingly, our results demonstrated a 3.4 kcal·mol−1 decrease in activation energy. Although this effect may not be highly pronounced, it could contribute to the design and optimization of this kind of reaction.
Of course, the abovementioned species is not the only one to activate H2S in CO2 hydrogenation and may be replaced by appropriate clusters like “superatom” species [33,34,35,36] that mimic elements of the periodic table with similar or even greater functionalities. Studies in this field will be intriguing, considering that there is plenty of room at the bottom; that is, novel species with tailored reactivity can be discovered by finely tuning their size, shape, composition, and charge. Active surfaces mimicking the chemistry of silicon might also exist. A reassessment of the key factors of the CO2 hydrogenation reaction suggests that these surfaces should have a very weak interaction with CO2 but exhibit high reactivity toward H2S dissociation, so as to form species resembling HSiSH first. The fact that the bond dissociation energy of S-H in H2S is ~381 kJ·mol−1 [37] is much smaller than that of C=O in CO2 (~750 kJ·mol−1) [38] indicates that this criterion is not difficult to meet. If the surface atom (or certain species deposited onto the surface) is capable of polarizing hydrogen to produce hydridic Hδ− while at the same time holding sulfur throughout the reaction, the C-end activation of CO2 and subsequent O-end activation to produce HCOOH can be expected. Clearly, surface defects, doping, or alloying would contribute further to the reactivity.

4. Computational Methods

The quantum chemical calculations consisted mainly of two parts: low-lying isomer predictions and PES construction for the gas-phase reaction of H2S and HSiSH with COx. First, to obtain the global minimum and some low-lying isomers, a global (or local) search based on a particle swarm optimization algorithm within an evolutionary scheme was carried out using the CALYPSO package [39]. Owing to the interfacing of this package with Gaussian [40], DFT calculations at the B3LYP/6-311+G(d) level of theory [41,42,43] were performed, which guaranteed the reliability and precision of our searches in the initial stage. To enhance the structural diversity and search efficiency, identical structures were then excluded by a structure fingerprinting technique with a bond characterization matrix [44]. For the obtained structures, further optimization at the M062X/aug-cc-pVTZ level of theory [45,46,47,48] was performed, which has been shown to provide good results for main group thermochemistry and kinetics with accuracy comparable to more sophisticated and expensive correlated molecular orbital methods [49,50]. Correlation of these optimized structures provides direct structural insight into the reaction mechanism, and the PES can then be constructed by exploring the lacking intermediates and TSs along the reaction coordinate. The presence of zero or a single imaginary frequency was used to identify all the stationary points as minima or TSs. To ascertain that the identified TSs connected reactants and products smoothly, intrinsic reaction coordinate (IRC) calculations [51] were performed. Finally, the energetics of the optimal path were further refined using the coupled cluster single and double substitution method with a perturbative treatment of triple excitations (CCSD(T)) [52]. Specifically, the M06-2X/aug-cc-pVTZ calculated vibrational frequencies were used to estimate the zero-point corrections of all the converged structures, and the zero-point-corrected CCSD(T)/aug-cc-pVTZ electronic energies (Gibbs free energies at zero temperature) at the M06-2X/aug-cc-pVTZ-optimized geometries were used for the presented analyses.
To understand the unique reactivity of HSiSH as compared to H2S in CO2 hydrogenation, the Fukui function [53], dual descriptor [54], and their condensed forms [55] of all the reactants were computed at the M062X/aug-cc-pVTZ level of theory. As one of the widely used local descriptors in conceptual DFT [56], the Fukui function is used here to assign a certain type of reactivity such as nucleophilicity (f+) or electrophilicity (f) to a specific region within a molecule, while the dual descriptor can be interpreted as the difference between the Fukui functions for nucleophilic and electrophilic attacks (∆f). This latter descriptor thus gives a combination of both Fukui functions: it is negative for locations where an electrophilic attack is more probable than a nucleophilic attack and positive where nucleophilic attack is more probable. In this way, the new index is dual and can be used to simultaneously detect the electrophile or nucleophile behavior of a given atomic region in the molecule. In the present work, the values of the dual descriptor were obtained at the 0.01 a.u. isodensity surface and visualized with the GaussView software [57].

5. Conclusions

The current study reveals that it is kinetically infeasible for hydrogens in the form of H2S to transfer to COx at low temperatures. However, when HSiSH is employed instead, successive C-end and O-end activations of CO2 can be achieved, with HCOOH as the final product. This observation can be well explained by the unique structure of the HSiSH molecule, in which Si-Hδ− with both electrophilic and nucleophilic sites can initiate the hydrosilylation of CO2, followed by S-Hδ+ cleavage to facilitate the hydrogenation of HCOO. The strong tendency of Si to form a strong bond with sulfur contributes further to the release of HCOOH. Although this finding confirms only the role of the silicon atom in tuning the reactivity of H2S, future works will focus more on the underlying chemistry between atoms, clusters, and surface systems for investigations and improved understanding.

Supplementary Materials

The following are available online, Figure S1, Cartesian coordinates of stationary points and transition states in the reactions of H2S with CO; Figure S2, Intrinsic reaction coordinate calculation for the reactions of H2S with CO; Figure S3, Reaction profiles for the H2S + CO; Figure S4, Cartesian coordinates of stationary points and transition states in the reactions of H2S with CO2; Figure S5, Intrinsic reaction coordinate calculation for the reactions of H2S with CO2; Figure S6, Reaction profiles for the H2S + CO2; Figure S7, Cartesian coordinates of stationary points and transition states in the reactions of HSiSH with CO; Figure S8, Intrinsic reaction coordinate calculation for the reactions of HSiSH with CO; Figure S9, Reaction profiles for the HSiSH + CO; Figure S10, Cartesian coordinates of stationary points and transition states in the reactions of HSiSH with CO2; Figure S11, Intrinsic reaction coordinate calculation for the reactions of HSiSH with CO2; Figure S12, A brief summary (TS omitted) of C-end activation pathways for the reaction of HSiSH + CO2; Figure S13, C-end activation pathways for the reaction of HSiSH + CO2; Figure S14, A brief summary (TS omitted) of O-end activation pathways for the reaction of HSiSH + CO2; Figure S15, O-end activation pathways for the reaction of HSiSH + CO2; Figure S16, Double-end activation pathway for the reaction of HSiSH + CO2.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 21803044) and Frontier and Applied Basic Research of Chongqing (Grant No. cstc2018jcyjAX0191).

Data Availability Statement

The data presented in this study are available in supplementary material.

Conflicts of Interest

The author declares no conflict of interest.

Sample Availability

Samples of the compounds are not available from the author.

References

  1. Meinshausen, M.; Meinshausen, N.; Hare, W.; Raper, S.C.B.; Frieler, K.; Knutti, R.; Frame, D.J.; Allen, M.R. Greenhouse-gas emission targets for limiting global warming to 2 °C. Nature 2009, 458, 1158–1196. [Google Scholar] [CrossRef] [PubMed]
  2. Seneviratne, S.I.; Rogelj, J.; Seferian, R.; Wartenburger, R.; Allen, M.R.; Cain, M.; Millar, R.J.; Ebi, K.L.; Ellis, N.; Hoegh-Guldberg, O.; et al. The many possible climates from the Paris Agreement’s aim of 1.5 °C warming. Nature 2018, 558, 41–49. [Google Scholar] [CrossRef] [PubMed]
  3. Zhou, W.; Cheng, K.; Kang, J.C.; Zhou, C.; Subramanian, V.; Zhang, Q.H.; Wang, Y. New horizon in C1 chemistry: Breaking the selectivity limitation in transformation of syngas and hydrogenation of CO2 into hydrocarbon chemicals and fuels. Chem. Soc. Rev. 2019, 48, 3193–3228. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, H.Y.; Zhang, C.; Gao, P.; Wang, H.; Li, X.P.; Zhong, L.S.; Wei, W.; Sun, Y.H. A review of the catalytic hydrogenation of carbon dioxide into value-added hydrocarbons. Catal. Sci. Technol. 2017, 7, 4580–4598. [Google Scholar] [CrossRef]
  5. Zhang, Y.; Zhang, T.; Das, S. Catalytic transformation of CO2 into C1 chemicals using hydrosilanes as a reducing agent. Green Chem. 2020, 22, 1800–1820. [Google Scholar] [CrossRef]
  6. Yant, W.P. Hydrogen sulphide in industry—Occurrence, effects, and treatment. Am. J. Public Health Nations Health 1930, 20, 598–608. [Google Scholar] [CrossRef] [Green Version]
  7. Shen, L.J.; Cao, Y.N.; Du, Z.J.; Zhao, W.T.; Lin, K.; Jiang, L.L. Illuminate the active sites of γ-FeOOH for low-temperature desulfurization. Appl. Surf. Sci. 2017, 425, 212–219. [Google Scholar] [CrossRef]
  8. Zhong, J.W.; Yang, X.F.; Wu, Z.L.; Liang, B.L.; Huang, Y.Q.; Zhang, T. State of the art and perspectives in heterogeneous catalysis of CO2 hydrogenation to methanol. Chem. Soc. Rev. 2020, 49, 1385–1413. [Google Scholar] [CrossRef]
  9. Liu, G.X.; Poths, P.; Zhang, X.X.; Zhu, Z.G.; Marshall, M.; Blankenhorn, M.; Alexandrova, A.N.; Bowen, K.H. CO2 hydrogenation to formate and formic acid by bimetallic palladium-copper hydride clusters. J. Am. Chem. Soc. 2020, 142, 7930–7936. [Google Scholar] [CrossRef]
  10. Jiang, X.; Nie, X.W.; Guo, X.W.; Song, C.S.; Chen, J.G. Recent advances in carbon dioxide hydrogenation to methanol via heterogeneous catalysis. Chem. Rev. 2020, 120, 7984–8034. [Google Scholar] [CrossRef]
  11. Baltrusaitis, J.; Patterson, E.V.; Hatch, C. Computational studies of CO2 activation via photochemical reactions with reduced sulfur compounds. J. Phys. Chem. A 2012, 116, 9331–9339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. He, R.T.; Hu, B.Y.; Zhong, H.; Jin, F.M.; Fan, J.J.; Hu, Y.H.; Jing, Z.Z. Reduction of CO2 with H2S in a simulated deep-sea hydrothermal vent system. Chem. Commun. 2019, 55, 1056–1059. [Google Scholar] [CrossRef] [PubMed]
  13. Upadhyay, L.R.; Rangarajan, S.; Baltrusaitis, J. Inhibitor, co-catalyst, or co-reactant? Probing the different roles of H2S during CO2 hydrogenation on the MoS2 catalyst. ACS Catal. 2019, 9, 10044–10059. [Google Scholar]
  14. Kumar, M.; Francisco, J.S. Hydrogen sulfide induced carbon dioxide activation by metal-free dual catalysis. Chem. Eur. J. 2016, 22, 4359–4363. [Google Scholar] [CrossRef] [PubMed]
  15. Mul, G.; Wachs, I.E.; Hirschon, A.S. Catalytic synthesis of methanethiol from hydrogen sulfide and carbon monoxide over vanadium-based catalysts. Catal. Today 2003, 78, 327–337. [Google Scholar] [CrossRef]
  16. Gutierrez, O.Y.; Kaufmann, C.; Hrabar, A.; Zhu, Y.Z.; Lercher, J.A. Synthesis of methyl mercaptan from carbonyl sulfide over sulfide K2MoO4/SiO2. J. Catal. 2011, 280, 264–273. [Google Scholar] [CrossRef]
  17. Gutierrez, O.Y.; Kaufmann, C.; Lercher, J.A. Synthesis of methanethiol from carbonyl sulfide and carbon disulfide on (Co)K-promoted sulfide Mo/SiO2 catalysts. ACS Catal. 2011, 1, 1595–1603. [Google Scholar] [CrossRef]
  18. Chen, A.P.; Wang, Q.; Li, Q.L.; Hao, Y.J.; Fang, W.P.; Yang, Y.Q. Direct synthesis of methanethiol from H2S-rich syngas over sulfided Mo-based catalysts. J. Mol. Catal. A Chem. 2008, 283, 69–76. [Google Scholar] [CrossRef]
  19. Zhang, B.J.; Taylor, S.H.; Hutchings, G.J. Catalytic synthesis of methanethiol from CO/H2/H2S mixtures using α-Al2O3. New J. Chem. 2004, 28, 471–476. [Google Scholar] [CrossRef]
  20. Liu, X.; Liu, X.R.; Wang, X.F. Splitting of hydrogen sulfide by group 14 elements (Si, Ge, Sn, Pb) in excess argon at cryogenic temperatures. J. Phys. Chem. A 2018, 122, 7023–7032. [Google Scholar] [CrossRef]
  21. Cockett, M.C.R.; Dyke, J.M.; Morris, A.; Niavaran, M.H.Z. High-temperature photoelectron spectroscopy. A study of SiS(X 1Σ+). J. Chem. Soc. Faraday Trans. 2 Mol. Chem. Phys. 1989, 85, 75–83. [Google Scholar]
  22. Muller, H.S.P.; McCarthy, M.C.; Bizzocchi, L.; Gupta, H.; Esser, S.; Lichau, H.; Caris, M.; Lewen, F.; Hahn, J.; Esposti, C.D.; et al. Rotational spectroscopy of the isotopic species of silicon monosulfide, SiS. Phys. Chem. Chem. Phys. 2007, 9, 1579–1586. [Google Scholar] [CrossRef] [PubMed]
  23. Schnockel, H.; Koppe, R. Matrix IR spectrum and ab initio SCF calculations of molecular SiS2. J. Am. Chem. Soc. 1989, 111, 4583–4586. [Google Scholar] [CrossRef]
  24. Gottle, A.J.; Koper, M.T.M. Determinant role of electrogenerated reactive nucleophilic species on selectivity during reduction of CO2 catalyzed by metalloporphyrins. J. Am. Chem. Soc. 2018, 140, 4826–4834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Tang, Q.; Lee, Y.J.; Li, D.Y.; Choi, W.; Liu, C.W.; Lee, D.; Jiang, D.E. Lattice-hydride mechanism in electrocatalytic CO2 reduction by structurally precise copper-hydride nanoclusters. J. Am. Chem. Soc. 2017, 139, 9728–9736. [Google Scholar] [CrossRef]
  26. Khandelwal, M.; Wehmschulte, R.J. Deoxygenative reduction of carbon dioxide to methane, toluene, and diphenylmethane with [Et2Al]+ as catalyst. Angew. Int. Ed. Chem. 2012, 51, 7323–7326. [Google Scholar] [CrossRef]
  27. Sun, W.; Qian, C.X.; He, L.; Ghuman, K.K.; Wong, A.P.Y.; Jia, J.; Jelle, A.A.; O’Brien, P.G.; Reyes, L.M.; Wood, T.E.; et al. Heterogeneous reduction of carbon dioxide by hydride-terminated silicon nanocrystals. Nat. Commun. 2016, 7, 1–9. [Google Scholar] [CrossRef] [Green Version]
  28. Dasog, M.; Kraus, S.; Sinelnikov, R.; Veinot, J.G.C.; Rieger, B. CO2 to methanol conversion using hydride terminated porous silicon nanoparticles. Chem. Commun. 2017, 53, 3114–3117. [Google Scholar] [CrossRef]
  29. Wong, A.P.Y.; Sun, W.; Qian, C.X.; Jelle, A.A.; Jia, J.; Zheng, Z.Q.; Dong, Y.C.; Ozin, G.A. Tailoring CO2 reduction with doped silicon nanocrystals. Adv. Sustain. Syst. 2017, 1, 1700118. [Google Scholar] [CrossRef]
  30. Qian, C.X.; Sun, W.; Hung, D.L.H.; Qiu, C.Y.; Makaremi, M.; Kumar, S.G.H.; Wan, L.L.; Ghoussoub, M.; Wood, T.E.; Xia, M.K.; et al. Catalytic CO2 reduction by palladium-decorated silicon-hydride nanosheets. Nat. Catal. 2019, 2, 46–54. [Google Scholar] [CrossRef]
  31. Grasemann, M.; Laurenczy, G. Formic acid as a hydrogen source—recent developments and future trends. Energy Environ. Sci. 2012, 5, 8171–8181. [Google Scholar] [CrossRef]
  32. Villegas-Escobar, N.; Ortega, D.E.; Cortés-Arriagada, D.; Durán, R.; Yepes, D.; Gutiérrez-Oliva, S.; Toro-Labbé, A. Why low valent lead (II) hydride complex would be a better catalyst for CO2 activation than its 14 group analogues? J. Phys. Chem. C 2017, 121, 12127–12135. [Google Scholar] [CrossRef]
  33. Castleman, A.W. From elements to clusters: The periodic table revisited. J. Phys. Chem. Lett. 2011, 2, 1062–1069. [Google Scholar] [CrossRef]
  34. Jena, P. Beyond the periodic table of elements: The role of superatoms. J. Phys. Chem. Lett. 2013, 4, 1432–1442. [Google Scholar] [CrossRef]
  35. Luo, Z.X.; Castleman, A.W. Special and general superatoms. Acc. Chem. Res. 2014, 47, 2931–2940. [Google Scholar] [CrossRef]
  36. Jena, P.; Sun, Q. Super atomic clusters: Design rules and potential for building blocks of materials. Chem. Rev. 2018, 118, 5755–5870. [Google Scholar] [CrossRef]
  37. Dean, J.A. Lange’s Handbook of Chemistry, 15th ed.; McGraw Hill: New York, NY, USA, 1999. [Google Scholar]
  38. Chang, X.X.; Wang, T.; Gong, J.L. CO2 photo-reduction: Insights into CO2 activation and reaction on surfaces of photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196. [Google Scholar] [CrossRef]
  39. Wang, Y.C.; Lv, J.; Zhu, L.; Ma, Y.M. CALYPSO: A method for crystal structure prediction. Comput. Phys. Commun. 2012, 183, 2063–2070. [Google Scholar] [CrossRef] [Green Version]
  40. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09 Revision d 01; Gaussian Inc.: Wallingford Center, CO, USA, 2009. [Google Scholar]
  41. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  42. Lee, C.; Yang, Y.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B Condens. Matter Mater. Phys. 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  43. Frisch, M.J.; Pople, J.A.; Binkley, J.S. Self-consistent molecular orbital methods 25. Supplementary functions for Gaussian basis sets. J. Chem. Phys. 1984, 80, 3265–3269. [Google Scholar] [CrossRef]
  44. Lv, J.; Wang, Y.C.; Zhu, L.; Ma, Y.M. Particle-swarm structure prediction on clusters. J. Chem. Phys. 2012, 137, 084104. [Google Scholar] [CrossRef] [PubMed]
  45. Kendall, R.A.; Dunning, T.H.; Harrison, R.J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef] [Green Version]
  46. Woon, D.E.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98, 1358–1871. [Google Scholar] [CrossRef] [Green Version]
  47. Peterson, K.A. Systematically convergent basis sets with relativistic pseudopotentials. I. Correlation consistent basis sets for the post-d group 13–15 elements. J. Chem. Phys. 2003, 119, 11099–11112. [Google Scholar] [CrossRef] [Green Version]
  48. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  49. Luo, S.J.; Zhao, Y.; Truhlar, D.G. Validation of electronic structure methods for isomerization reactions of large organic molecules. Phys. Chem. Chem. Phys. 2011, 13, 13683–13689. [Google Scholar] [CrossRef]
  50. Li, X.Y.; Xu, X.F.; You, X.Q.; Truhlar, D.G. Benchmark calculations for bond dissociation enthalpies of unsaturated methyl esters and the bond dissociation enthalpies of methyl linolenate. J. Phys. Chem. A 2016, 120, 4025–4036. [Google Scholar] [CrossRef]
  51. Hratchian, H.P.; Schlegel, H.B. Using Hessian updating to increase the efficiency of a Hessian based predictor-corrector reaction path following method. J. Chem. Theory Comput. 2005, 1, 61–69. [Google Scholar] [CrossRef]
  52. Bartlett, R.J.; Musial, M. Coupled-cluster theory in quantum chemistry. Rev. Mod. Phys. 2007, 79, 291–352. [Google Scholar] [CrossRef] [Green Version]
  53. Parr, R.G.; Yang, W.T. Density functional approach to the frontier-electron theory of chemical reactivity. J. Am. Chem. Soc. 1984, 106, 4049–4050. [Google Scholar] [CrossRef]
  54. Morell, C.; Grand, A.; Toro-Labbé, A. New dual descriptor for chemical reactivity. J. Phys. Chem. A 2005, 109, 205–212. [Google Scholar] [CrossRef] [PubMed]
  55. Gázquez, J.L. Chemical reactivity concepts in density functional theory. In Chemical Reactivity Theory—A Density Functional View; Chattaraj, P.K., Ed.; CRC Press: Boca Ratón, FL, USA, 2009. [Google Scholar]
  56. Geerlings, P.; De Proft, F.; Langenaeker, W. Conceptual density functional theory. Chem. Rev. 2003, 103, 1793–1873. [Google Scholar] [CrossRef] [PubMed]
  57. Dennington, R.; Keith, T.; Millam, J. Gauss View; Semichem Inc.: Shawnee, KS, USA, 2009. [Google Scholar]
Figure 1. Global minima obtained for the reactions of H2S and HSiSH with COx (x = 1, 2).
Figure 1. Global minima obtained for the reactions of H2S and HSiSH with COx (x = 1, 2).
Molecules 26 00050 g001
Figure 2. Reaction profiles for H2S + COx (x = 1, 2) (all energetics were refined by the CCSD(T) method).
Figure 2. Reaction profiles for H2S + COx (x = 1, 2) (all energetics were refined by the CCSD(T) method).
Molecules 26 00050 g002
Figure 3. Reaction profiles for HSiSH + CO, with coupled and decoupled reaction pathways illustrated in the lower panel (all energetics were refined by the CCSD(T) method).
Figure 3. Reaction profiles for HSiSH + CO, with coupled and decoupled reaction pathways illustrated in the lower panel (all energetics were refined by the CCSD(T) method).
Molecules 26 00050 g003
Figure 4. Optimal path obtained for the HSiSH + CO2 reaction; the activation energies for different pathways are shown in the upper panel (all energetics were refined by the CCSD(T) method).
Figure 4. Optimal path obtained for the HSiSH + CO2 reaction; the activation energies for different pathways are shown in the upper panel (all energetics were refined by the CCSD(T) method).
Molecules 26 00050 g004
Figure 5. Dual descriptors computed for the H2S, CO2, and cis-/trans-HSiSH molecules; nucleophilic and electrophilic zones are represented in red and blue, respectively (isovalue for surfaces: ρ = 0.01 a.u.).
Figure 5. Dual descriptors computed for the H2S, CO2, and cis-/trans-HSiSH molecules; nucleophilic and electrophilic zones are represented in red and blue, respectively (isovalue for surfaces: ρ = 0.01 a.u.).
Molecules 26 00050 g005
Table 1. Hirshfeld charges, condensed Fukui functions (f+/f), and condensed dual descriptors (∆f) for the CO2, H2S, and cis-/trans-HSiSH molecules.
Table 1. Hirshfeld charges, condensed Fukui functions (f+/f), and condensed dual descriptors (∆f) for the CO2, H2S, and cis-/trans-HSiSH molecules.
MoleculeAtomq (N) [a]q (N + 1)q (N − 1)f+ff
CO2C0.3618−0.29600.57770.65780.21590.4419
O−0.1808−0.31790.21120.13710.3920−0.2549
H2SH0.0616−0.08060.17200.14210.11040.0317
S−0.1231−0.80210.65600.67890.7792−0.1002
cis-HSiSHH [b]−0.1150−0.18740.02620.07240.1412−0.0688
Si0.2444−0.39870.79660.64300.55230.0908
S−0.1938-0.42960.04820.23580.2419−0.0061
H0.06440.01870.12900.04570.0646−0.0189
trans-HSiSHH [b]−0.1119−0.18410.02050.07220.1324−0.0602
Si0.2394−0.40240.77840.64180.53900.1029
S−0.1880−0.42520.08270.23730.2707−0.0334
H0.06050.01490.11840.04560.0580−0.0124
[a] q (N), q (N + 1) and q (N − 1) correspond to the charge values of each atom for neutral, anion, and cation, respectively. [b] The hydrogen atom that bonded with silicon.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, X. Hydrogenation of CO2 Promoted by Silicon-Activated H2S: Origin and Implications. Molecules 2021, 26, 50. https://doi.org/10.3390/molecules26010050

AMA Style

Liu X. Hydrogenation of CO2 Promoted by Silicon-Activated H2S: Origin and Implications. Molecules. 2021; 26(1):50. https://doi.org/10.3390/molecules26010050

Chicago/Turabian Style

Liu, Xing. 2021. "Hydrogenation of CO2 Promoted by Silicon-Activated H2S: Origin and Implications" Molecules 26, no. 1: 50. https://doi.org/10.3390/molecules26010050

Article Metrics

Back to TopTop