Next Article in Journal
Design and Performance of CuNi-rGO and Ag-CuNi-rGO Composite Electrodes for Use in Fuel Cells
Previous Article in Journal
Biodiesel Synthesis from Coconut Oil Using the Ash of Citrus limetta Peels as a Renewable Heterogeneous Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photoinduced Mechanisms of C–S Borylation of Methyl(p-tolyl)Sulfane with Bis(Pinacolato)diboron: A Density Functional Theory Investigation

Chemical Synthesis and Pollution Control Key Laboratory of Sichuan Province, College of Chemistry and Chemical Engineering, China West Normal University, Nanchong 637001, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(8), 550; https://doi.org/10.3390/catal14080550
Submission received: 29 July 2024 / Revised: 19 August 2024 / Accepted: 20 August 2024 / Published: 22 August 2024
(This article belongs to the Section Photocatalysis)

Abstract

:
The reaction mechanisms of C–S borylation of aryl sulfides catalyzed with 1,4-benzoquinone (BQ) were investigated by employing the M06-2X-D3/ma-def2-SVP method and basis set. In this study, the SMD model was taken to simulate the solvent effect of 1,4-dioxane. Also, TD-DFT calculations of BQ and methyl(p-tolyl)sulfane were performed in an SMD solvent model. The computational results indicated that BQ and methyl(p-tolyl)sulfane, serving as a photo-catalyst, would be excited under a blue LED of 450 nm, aligning well with experimental observations. Additionally, the role of 3O2 was investigated, revealing that it could be activated into 1O2 from the released energy of 1[BQ + methyl(p-tolyl)sulfane]* or 3[BQ + methyl(p-tolyl)sulfane]*→BQ + methyl(p-tolyl)sulfane process. Then, 1O2, bis(pinacolato)diboron, and methyl(p-tolyl)sulfane would, through a series of reactions, yield the final product, P. The Gibbs free energy surface shows that path a2-2 is optimal, and this path has fewer steps and a lower energy barrier. Electron spin density isosurface graphs were employed to analyze the structures and elucidate the single electron distribution. These computational results offer valuable insights into the studied interactions and related processes and shed light on the mechanisms governing C–S borylation from aryl sulfides and b2pin2 catalyzed with BQ and methyl(p-tolyl)sulfane.

Graphical Abstract

1. Introduction

Over the past few decades, C–S bond cleavages and transformations of aryl sulfides have received considerable attention and have been rapidly developed [1,2]. Generally, the C–S bond activation of aryl sulfides could be achieved by guiding group-assisted or directing group-free transition metal-catalysis [3,4,5,6]. While the formed aromatic C–B bonds are a series of important complexes. Especially, arylboric acid has been widely used as a sensor of sugars in material science and medical research, an inhibitor of enzymes, and a carrier of nucleosides and sugars in biology [7,8,9,10,11]. Hence, great efforts have been devoted to the formation of aromatic C–B bonds in organic chemistry [12,13,14].
Traditionally, the Miyaura borylation reaction can be used to achieve the various C–heteroatom borylations using the corresponding aromatic electrophiles including aryl sulfides. For example, Miyaura used a palladium-catalyzed coupling reaction between halogenated aromatics and biborate pinacol ester to produce corresponding organic borates [15,16,17]; Yorimitsu [18] and Hosoya [19] finished the C–S borylation of aryl sulfides via Pd- and Rh-catalysis (Scheme 1a,b); and Yorimitsu [20] and Gao [21] realized the C–S borylation of aryl sulfonium salts via Pd-catalysis (Scheme 1c) and UV irradiation. Unfortunately, most synthetic methods for aryl boronate esters have various disadvantages, including the need for noble metals as catalysts, complex synthetic steps, and low yields [22,23,24].
In recent years, amidst the surge in popularity of green chemistry [25,26,27], considerable attention has been directed towards identifying synthesis methods that are eco-friendly, cost-effective, and highly efficient. Under the irradiation of blue LED, Xinqi Li et al. proposed a photocatalytic study that directs C–S bond activation of aryl sulfides via photoinduced aerobic borylation under transition-metal-free conditions in 2022 [28]. This synthetic methodology offers enhanced convenience, cleanliness, and efficacy, thereby bearing significant relevance in chemical research and practical applications. A thorough investigation of its reaction mechanism can undoubtedly facilitate experimenter comprehension and enable the design of similar reactions.

2. Results and Discussion

According to reference [28], under the irradiation of blue LED, the reactants methyl(p-tolyl)sulfane (R1) and bis(pinacolato)diboron (R2) would go through a C–S borylation reaction to yield the final product, methyl boronic acid pinacol ester (P), in the solvent of 1,4-dioxane, in which 1,4-benzoquinone (BQ) is the photosensitizer, as depicted in Scheme 2.

2.1. Photocatalysis Process

As we all know, the circular conjugated large π-structure is the preferred site for photo-catalytic reactions. The structures of BQ and R1 have the conjugated π-bond, and BQ is usually used as the photosensitizer. Hence, there are two models designed to investigate the photo-catalytic process. When BQ was employed to explore the excitation in the first model, the computational results suggest that S0→S1 of BQ in Table 1 is required to absorb 265.6 kcal/mol energy from the wavelength of 450.2 nm, which accords with the experimental results (450.0 nm) within error, and the S0→S1 transition mainly depends on HOMO→LUMO (91.6%). Moreover, the Frontier Molecular Orbital Theory (FMOT) in Figure 1b,c shows that this S0→S1 excitation focuses on the intramolecular transfer from the σ-orbital of the benzene ring to the π-orbital of the benzene ring, which is consistent with ρele and ρhole in Figure 2a, while the S0→S2 and S0→S3 excitations depend on the wavelengths of 405.2 nm and 286.3 nm, respectively, which mainly come from the HOMO-2→LUMO (89.0%) and HOMO-1→LUMO (99.9%) in Table 1, and they cannot be achieved with blue LED (450.0 nm). In the second model, BQ and R1 form a complex BQ + R1 [29,30,31]. The TD-DFT calculations suggest that a 449.6 nm wavelength can finish the S0→S1 excitation of BQ + R1, which accords with the experimental results (450.0 nm) within error, and the S0→S1 transition mainly depends on HOMO→LUMO (57.5%) and HOMO-3→LUMO (37.5%). Frontier molecular orbital theory in Figure 1e–g displayed that HOMO→LUMO transition focuses on the intermolecular transfer from R1 to BQ. Meanwhile, the computational results in Table 1 indicate that S0→S2 with 436.4 nm could have the possibilities to accord with the experimental results (450.0 nm) within error. All of the above descriptions could suggest that the photo-catalytic reaction would successfully achieve the excitation process of BQ.

2.2. R1 + BQ→IM1(1O2)

As highlighted in reference [28], BQ, as an organic photocatalyst, is easy to become an excited state in the irradiation of blue LED (450 nm). The computed oscillator strengths (fos = 0) of BQ model in Section 2.1 indicates that it can be successfully changed into a singlet excited state 1BQ*, which is very active. Then, the energy transfer process between 1BQ* and 3O2 happened, and it is an exothermic process (purple line in Scheme 3 and Figure 3). Finally, the obtained intermediate 1O2 can continue to react with R1, while in the second model, the R1 + BQ model would absorb the energy of photons to finish the S0→S1 excitation. The calculations suggest that this photo-excited process requires 51.2 kcal/mol to become excitation state 1[R1 + BQ]*, which is very activate and has two possible paths, as depicted in Figure 3. In the first path, the singlet excited state 1[R1 + BQ]* would become the ground state R1 + BQ by releasing much energy, which results in the triplet oxygen 3O2 changing into singlet oxygen 1O2. Generally, the structure of triplet excited state 3[R1 + BQ]* is much more stable than that of singlet excited state 1[R1 + BQ]*. So, in the second path, Intersystem Crossing (ISC) would transfer 1[R1 + BQ]* into 3[R1 + BQ]*, and this process releases 7.8 kcal/mol of energy. Meanwhile, there is a possibility that 3[R1 + BQ]* and triplet oxygen 3O2 could go through energy transfer to become R1 + BQ and singlet oxygen 1O2 via releasing about 5.9 kcal/mol energy. The obtained 1O2, as IM1, is one excited state and very active, which continues to participate the following reaction.

2.3. IM1(1O2)→P

The formed IM1(1O2) is very active, which could participate the following reactions. Hence, 1O2 reacts with R1 and R2 in paths a1 and a2, respectively. Moreover, the reaction between 3O2 and R1 (or R2) has also been investigated in path a1, which is elaborated on in the following section.

2.3.1. Path a1

According to Figure 4, the Gibbs free energy surface suggests that it is an SN2 reaction with the energy barrier of 54.7 kcal/mol, which is very high and cannot happen. Considering the large amount of 3O2, it can also have the possibility to react with R1 or R2 in the system. Figure 4 suggests that R1 + 3O2→TSR1-IM2a1-2→IM2a1-1 + IM3a1-1 (blue line) process has an energy barrier of 55.4 kcal/mol and cannot also happen in path a1-2. Finally, the reaction between 3O2 and R2 has a relative lower energy barrier than those in paths a1-1 and a1-2; 44.2 kcal/mol is still a higher barrier, and it cannot continue to occur. The formed IM2a1-1 is a free radical and the single electron is distributed on the S atom (Figure 5c) which is important in the following process.

2.3.2. Path a2

Based on the analysis of the Fukui function and dual descriptor, the B5(R2) atom with the positive value ( f B 5 ( R 2 ) ( 2 ) = 0.0746) in Table 2 is susceptible to nucleophilic attack. In Path a2 of Figure 6 and Figure 7, the B5 atom of R2 reacts with the O1(IM1) atom via transition state TSIM1-IM2a2, which has an energy barrier of 19.5 kcal/mol and the distance of O1(IM1)⋯B5(R1) in TSIM1-IM2a2 is 1.78 Å. Additionally, the energy of IM2a2 is 113.8 kcal/mol lower than that of IM1 + R2, indicating that this step is an exothermic process. Then, the following step would convert intermediate IM2a2 into IM3a2 via an intramolecular homolytic reaction by absorbing energy of 32.3 kcal/mol. The electron spin density isosurface graph of IM3a2 in Figure 5d shows that the single electron is distributed on the O1 atom, which is the reactive site participating in the following reaction. The condensed dual descriptor (CDD) values of the C4 atom of R1 and the B5 (or B6) atom of R2 are f C 4 ( R 1 ) ( 2 ) = 0.0025 and f B 5 ( R 2 ) ( 2 ) = 0.0746 in Table 2, respectively. Therefore, there are two paths (a2-1 and a2-2) that can yield the final product, P.
In path a2-1 of Scheme 4, IM3a2 would react with R2 at the sites of the O1(IM3a2) atom and the C4(R2) atom via transition state TSIM3a2-IM5a2-1. The computational results displayed that it is a SN2 reaction, and the distances of O1(IM3a2)⋯C4(R2) and C4(R2)⋯S3(R2) are 2.12 Å and 1.79 Å in TSIM3a2-IM5a2-1, respectively. Moreover, this IM3a2 + R1→IM4a2-1 + IM5a2-1 process is an exothermic procedure, and 24.0 kcal/mol would be released. The formed IM5a2-1 is a radical, and the single electron is distributed in S3 atom in Figure 5c. The next process is an intramolecular ring-closure reaction happened between the B5-B6(R2) bond and the S3 atom via a three-membered(B5, B6, and S3) transition state TSIM5a2-1-IM6a2-1, which needs 17.9 kcal/mol energy, and the distances of B5⋯S3, B6⋯S3, and B5⋯B6 are 2.61 Å, 2.60 Å, and 1.74 Å, respectively. The generated IM6a2-1 is one three-membered intermediate, and it would go through a ring-opening reaction to generate a byproduct of BP and IM7a2-1 radicals via transition state TSIM6a2-1-IM7a2-1. It is a fast step, and 4.5 kcal/mol is the energy barrier in Figure 6. Furthermore, Figure 5e,f,j reveal it is one single electron transfer process, and the single electron is distributed in the B5 atom of IM7a2-1 in Figure 5f. Finally, an SN2 reaction happened between the C4(R1) and B5(IM7a2-1) atoms via TSIM7a2-1-P, which needs 15.5 kcal/mol energy. The distance of C4(R1)⋯B5(IM7a2-1) decreased to 1.56 Å in P from 2.27 Å in TSIM7a2-1-P, indicating that the IM7a2-1 intermediate could successfully move to the C4(R1) atom of R1. The energy of the obtained IM5a2-1 + P was 31.6 kcal/mol lower than that of IM7a2-1 + R1, showing that this process is the exothermic reaction in Figure 6. Meanwhile, IM5a2-1 can continue to also attack the R2 reaction to realize recycling.
The B5 or B6 atom of R2 has an electron-deficient structure. So, in Path a2-2, IM3a2 with one single electron can also attack the B5 or B6 atom of R2 to complete the SN2 reaction. The calculations show that this IM3a2 + R2→IM5a2-2 process with the energy barrier of 11.5 kcal/mol via TSIM3a2-IM5a2-2 is an exothermic procedure (45.5 kcal/mol) in Figure 7. Meanwhile, the B5–B6 bond of R2 is broken, and an intermediate IM4a2-2 is formed. The structure of IM5a2-2 is as the same as IM7a2-1 in path a2-1, so the final reaction of the IM5a2-2 + R1→P + IM6a2-2 process is as the same as the IM7a2-1 + R1→P + IM5a2-1 procedure in path a2-1.

3. Computational Details

All of the calculations were performed in Gaussian 09 program package [32]. Every structure was optimized at the M06-2X-D3/ma-def2-SVP [33,34,35] level in 101.325 KPa at the temperature of 298.15 K, and the frequency analysis confirmed that every transition state with only one imaginary frequency and the other structures with no imaginary frequency are correct. Intrinsic reaction coordinates (IRCs) have been calculated to make sure each transition state is correct [36] and the detailed results can be found in Supplementary Materials. The SMD model was employed to simulate the solvent effect of 1,4-dioxane, which is defined by the static dielectric constant (eps = 2.21) and the dynamic dielectric constant (epsinf = 1.90) [37,38]. The model has been widely used in the process of investigating the mechanisms of organic chemistry [30,39,40,41,42,43,44,45,46]. To obtain more accurate energy values, single point energies were calculated with the M06-2X-D3/ma-def2-TZVP level in the SMD model. Additionally, a hole-electron isosurface map, Spin density isosurface maps, a frontier molecular orbital (FMO), TD-DFT [47,48,49,50,51], and conceptual density functional theory (CDFT) data were obtained using the Multiwfn program (version 3.3.8) [52]. FMO was drawn using VMD program 1.9.3 [53]. All optimized three-dimensional (3D) structures were displayed in CYLview (version 1.0b) [54].

4. Conclusions

In summary, this paper systematically investigates the detailed formation mechanisms of methyl boronic acid pinacol ester (P) through a photoinduced aerobic borylation reaction between methyl(p-tolyl)sulfane (R1) and bis(pinacolato)diboron (R2). The computations were performed at the M06-2X-D3/ma-def2-TZVP level, with an SMD model to simulate the solvent effect of 1,4-dioxane. Additionally, the photo-catalytic mechanisms of BQ and BQ + R1 models were explored using the TDDFT method at the M06-2X-D3/ma-def2-TZVP level. The computational results revealed that both of these two photo-catalytic models could achieve excitation. Moreover, the activated excited states 1BQ* and 1[R1 + BQ]* can finish the transformation from 3O2 into 1O2. Meanwhile, 1[R1 + BQ]* has the possibility to change into 3[R1 + BQ]* via intersystem crossing (ISC). Hence, there are three existing possibilities that can change 3O2 into 1O2 via an energy transfer process. Then, 1O2 reacts with bis(pinacolato)diboron to complete the insert reaction, and the formed intermediate goes through a homolytic process to generate a radical intermediate. Subsequently, two paths (a2-1 and a2-2) were proposed to convert R2 into product P. In path a2-1, IM3a2, and R1 go through an SN2 reaction, ring-closure reaction, ring-opening reaction, and SN2 reaction to yield final product P. While in path a2-2, IM3a2 goes through two SN2 reactions to yield P. Compared these two paths, the highest point of path a2-2 is −70. 0 kcal/mol, lower than that of path a2-2 with −66.9 kcal/mol, implying that path a2-2 was identified as the optimal route with lower energy barriers, and the electron spin density isosurface graphs can reveal that it is a free radical transfer process. The clear mechanisms offer valuable insights for the design of similar reactions and showcase the potential of computational methods in advancing catalysis research.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14080550/s1.

Author Contributions

Investigation, data curation and writing—original draft preparation, Y.M.; writing—review and editing, T.F. and B.C.; supervision and funding acquisition, D.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Sichuan Province [grant number 2022NSFSC0629].

Data Availability Statement

Data are contained within the article or Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lou, J.; Wang, Q.; Wu, P.; Wang, H.; Zhou, Y.G.; Yu, Z. Transition-metal mediated carbon–sulfur bond activation and transformations: An update. Chem. Soc. Rev. 2020, 49, 4307–4359. [Google Scholar] [CrossRef] [PubMed]
  2. Xia, X.L.; Zhu, Q.L.; Dong, Z.B.; Chen, J.Q.; Shi, Z. Synthesis of Aryl Dithiocarbamates from Tetramethylthiuram Monosulfide (TMTM) and Aryl Boronic Acids: Copper-Catalyzed Construction of C(sp2)–S Bonds. Synthesis 2021, 54, 475–482. [Google Scholar] [CrossRef]
  3. Delcaillau, T.; Boehm, P.; Morandi, B. Nickel-Catalyzed Reversible Functional Group Metathesis between Aryl Nitriles and Aryl Thioethers. J. Am. Chem. Soc. 2021, 143, 3723–3728. [Google Scholar] [CrossRef]
  4. Boehm, P.; Müller, P.; Finkelstein, P.; Rivero-Crespo, M.A.; Ebert, M.O.; Trapp, N.; Morandi, B. Mechanistic Investigation of the Nickel-Catalyzed Metathesis between Aryl Thioethers and Aryl Nitriles. J. Am. Chem. Soc. 2022, 144, 13096–13108. [Google Scholar] [CrossRef] [PubMed]
  5. Delcaillau, T.; Woenckhaus-Alvarez, A.; Morandi, B. Nickel-Catalyzed Cyanation of Aryl Thioethers. Org. Lett. 2021, 23, 7018–7022. [Google Scholar] [CrossRef]
  6. Delcaillau, T.; Morandi, B. Nickel-Catalyzed Thiolation of Aryl Nitriles. Chem. Eur. J. 2021, 27, 11823. [Google Scholar] [CrossRef]
  7. Yorimitsu, H. Catalytic Transformations of Sulfonium Salts via C–S Bond Activation. Chem. Rec. 2021, 21, 3356–3369. [Google Scholar] [CrossRef]
  8. Fan, R.; Tan, C.; Liu, Y.; Wei, Y.; Zhao, X.; Liu, X.; Tan, J.; Yoshida, H. A leap forward in sulfonium salt and sulfur ylide chemistry. Chin. Chem. Lett. 2020, 32, 299–312. [Google Scholar] [CrossRef]
  9. Ma, N.N.; Ren, J.A.; Liu, X.; Chu, X.Q.; Rao, W.; Shen, Z.L. Nickel-Catalyzed Direct Cross-Coupling of Aryl Sulfonium Salt with Aryl Bromide. Org. Lett. 2022, 24, 1953–1957. [Google Scholar] [CrossRef]
  10. Ilardi, E.A.; Vitaku, E.; Njardarson, J.T. Data-Mining for Sulfur and Fluorine: An Evaluation of Pharmaceuticals To Reveal Opportunities for Drug Design and Discovery. J. Med. Chem. 2013, 57, 2832–2842. [Google Scholar] [CrossRef]
  11. Feng, M.; Tang, B.; Liang, S.H.; Jiang, X. Sulfur Containing Scaffolds in Drugs: Synthesis and Application in Medicinal Chemistry. Curr. Top. Med. Chem. 2016, 16, 1200–1216. [Google Scholar] [CrossRef] [PubMed]
  12. Pattison, G. Fluorination of organoboron compounds. Org. Biomol. Chem. 2019, 17, 5651–5660. [Google Scholar] [CrossRef] [PubMed]
  13. Dhital, R.N.; Sakurai, H. Oxidative Coupling of Organoboron Compounds. Asian, J. Org. Chem. 2014, 3, 668–684. [Google Scholar] [CrossRef]
  14. Wang, J. When diazo compounds meet with organoboron compounds. Pure Appl. Chem. 2017, 90, 617–623. [Google Scholar] [CrossRef]
  15. Wang, M.; Shi, Z. Methodologies and Strategies for Selective Borylation of C–Het and C–C Bonds. Chem. Rev. 2020, 120, 7348–7398. [Google Scholar] [CrossRef]
  16. Rout, L.; Punniyamurthy, T. Recent advances in transition-metal-mediated Csp2-B and Csp2-P cross-coupling reactions. Co-ord. Chem. Rev. 2020, 431, 213675. [Google Scholar] [CrossRef]
  17. Gao, M.Y.; Gosmini, C. Cobalt-Catalyzed Reductive Cross-Coupling to Construct Csp3–Csp3 Bonds via Csp3–S and Csp3–X Bonds Activation. Org. Lett. 2023, 25, 7689–7693. [Google Scholar] [CrossRef]
  18. Bhanuchandra, M.; Baralle, A.; Otsuka, S.; Nogi, K.; Yorimitsu, H.; Osuka, A. Palladium-Catalyzed ipso-Borylation of Aryl Sulfides with Diborons. Org. Lett. 2016, 18, 2966–2969. [Google Scholar] [CrossRef] [PubMed]
  19. Uetake, Y.; Niwa, T.; Hosoya, T. Rhodium-Catalyzed ipso-Borylation of Alkylthioarenes via C–S Bond Cleavage. Org. Lett. 2016, 18, 2758–2761. [Google Scholar] [CrossRef]
  20. Minami, H.; Otsuka, S.; Nogi, K.; Yorimitsu, H. Palladium-Catalyzed Borylation of Aryl Sulfoniums with Diborons. ACS Catal. 2017, 8, 579–583. [Google Scholar] [CrossRef]
  21. Huang, C.; Feng, J.; Ma, R.; Fang, S.; Lu, T.; Tang, W.; Du, D.; Gao, J. Redox-Neutral Borylation of Aryl Sulfonium Salts via C–S Activation Enabled by Light. Org. Lett. 2019, 21, 9688–9692. [Google Scholar] [CrossRef] [PubMed]
  22. Xu, Y.; Fang, H. Research Progress towards Synthesis of Aryl Boronic Acid Compounds. Chin. J. Org. Chem. 2018, 38, 738–751. [Google Scholar] [CrossRef]
  23. Fang, H.P.; Fu, C.C.; Tai, C.K.; Chang, K.H.; Yang, R.H.; Wu, M.J.; Chen, H.C.; Li, C.J.; Huang, S.Q.; Lien, W.H.; et al. Synthesis and stability study of isocyano aryl boronate esters and their synthetic applications. RSC Adv. 2016, 6, 30362–30371. [Google Scholar] [CrossRef]
  24. Zhou, J.; Berthel, J.H.J.; Kuntze Fechner, M.W.; Friedrich, A.; Marder, T.B.; Radius, U. NHC Nickel-Catalyzed Suzuki–Miyaura Cross-Coupling Reactions of Aryl Boronate Esters with Perfluorobenzenes. J. Org. Chem. 2016, 81, 5789–5794. [Google Scholar] [CrossRef] [PubMed]
  25. Ludwig, J.K. Green Chemistry: An Introductory Text. Green Chem. Lett. Rev. 2017, 10, 30–31. [Google Scholar] [CrossRef]
  26. Al-Shatti, B.J.; Alsairafi, Z.; Al-Tannak, N.F. Green chemistry and its implementation in pharmaceutical analysis. Rev. Anal. Chem. 2023, 42, 20230069. [Google Scholar] [CrossRef]
  27. Kar, S.; Sanderson, H.; Roy, K.; Benfenati, E.; Leszczynski, J. Green Chemistry in the Synthesis of Pharmaceuticals. Chem. Rev. 2021, 122, 3637–3710. [Google Scholar] [CrossRef] [PubMed]
  28. Li, X.; Wan, Z.; Hu, X.; Zhang, H. Photoinduced aerobic C–S borylation of aryl sulfides. Org. Chem. Front. 2022, 9, 3034–3038. [Google Scholar] [CrossRef]
  29. Yang, B.; Zhou, D.G. Dft Investigation for the Mechanisms of γ-Butyrolactones from Styrene and Acrylic Acid, Catalyzed by Nma*Bf 4 and Phssph. Comput. Theor. Chem. 2024, 1237, 114654. [Google Scholar] [CrossRef]
  30. Zheng, X.F.; Zhou, D.G.; Yang, L.J. DFT investigation of the DDQ-catalytic mechanism for constructing C–O bonds. Org. Biomol. Chem. 2024, 22, 3693–3707. [Google Scholar] [CrossRef]
  31. Belguidoum, K.; Boulmokh, Y.; Hamamdia, F.Z.; Madi, F.; Nouar, L.; Amira Guebailia, H. Tetradentate square-planar acety-lumbelliferone–nickel (II) complex formation: A DFT and TD-DFT study. Theor. Chem. Acc. 2022, 141, 48. [Google Scholar] [CrossRef]
  32. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B. Gaussian 09, R.E. 01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  33. 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] [CrossRef]
  34. Saheb, V.; Bahadori, A. Theoretical studies on the kinetics of the hydrogen-abstraction reactions from 1,3,5-trioxane and 1,4-dioxane by OH radicals. Prog. React. Kinet. Mech. 2020, 45, 1–13. [Google Scholar] [CrossRef]
  35. Cheng, X. Computational insights into the coupling mechanism of benzoic acid, phenoxy acetylene and dihydroisoquinoline catalyzed by silver ion as polarizer and stabilizer. Appl. Organomet. Chem. 2020, 34, e5903. [Google Scholar] [CrossRef]
  36. Kenichi, F. The Path of Chemical Reactions—The Irc Approach. Acc. Chem. Res. 1981, 14, 363–368. [Google Scholar] [CrossRef]
  37. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  38. Feng, T.T.; Lin, Y.; Chen, B.; Zhou, D.G.; Li, R. Alkali metal hydroxide-catalyzed mechanisms of Csp–H silylation of alkynes: A DFT investigation. Org. Biomol. Chem. 2024, 22, 6352–6361. [Google Scholar] [CrossRef] [PubMed]
  39. Zhou, D.G.; Wang, P. Mechanisms of the reaction between benzonitrile and 4-octyne catalyzed by Ni(PMe3)2: A theoretical investigation. J. Phys. Org. Chemistry. 2019, 32, 3932. [Google Scholar] [CrossRef]
  40. Zhou, D.G.; Li, Y.Q. Mechanistic Study of 1,4-Benzodiazepine-2,5-diones from Diphenylamine and Diethyl 2-Phenylmalonate by Density Functional Theory. J. Phys. Chem. A 2019, 124, 395–408. [Google Scholar] [CrossRef]
  41. Zhou, D.G.; Zhou, P.P.; Jing, H.W. Mechanisms of Csp3–H functionalization of ethyl 2-(methyl(p-tolyl)amino)acetate: A theoretical investigation. Comput. Theor. Chem. 2017, 1118, 144–152. [Google Scholar] [CrossRef]
  42. Zhou, D.G. DFT investigation on the mechanism of catalytic reaction between 3-diazoindolin-2-imines and N-ethylaniline catalyzed by Rh2(Oct)4. Chem. Phys. 2019, 531, 110661. [Google Scholar] [CrossRef]
  43. Zheng, X.F.; Zhou, D.G. Mechanisms of asymmetric sulfa-Michael additions between phenylacetylene and thiolacetic acid: A DFT investigation. Comput. Theor. Chem. 2021, 1207, 113523. [Google Scholar] [CrossRef]
  44. Zhang, M.; Wang, Y.; Li, S.J.; Wang, X.H.; Shi, Q.Q.; Li, X.; Qu, L.B.; Wei, D.H.; Lan, Y. Multiple Functional Organocata-lyst-Promoted Inert C-C Activation: Mechanism and Origin of Selectivities. ACS Catal. 2021, 11, 3443–3454. [Google Scholar] [CrossRef]
  45. Zhou, D.G. Mechanisms of Csp3-H functionalization of acetonitrile or acetone with coumarins: A DFT investigation. Mol. Catal. 2020, 498, 111246. [Google Scholar] [CrossRef]
  46. Chen, B.; Zhou, D.G.; Yang, L.J. Reaction mechanism of acetonitrile, olefins, and amines catalyzed by Ag2CO3: A DFT inves-tigation. J. Phys. Org. Chem. 2023, 37, e4594. [Google Scholar] [CrossRef]
  47. Budyka, M.F. Density functional theory study of the styrylbenzoquinoline dyad and the related dibenzoquinolylcyclobutane formed in the [2 + 2] photocycloaddition reaction. Int. J. Quantum Chem. 2023, 124, e27264. [Google Scholar] [CrossRef]
  48. Liu, Z.; Lu, T.; Chen, Q. An sp-hybridized all-carboatomic ring, cyclo[18]carbon: Electronic structure, electronic spectrum, and optical nonlinearity. Carbon 2020, 165, 461–467. [Google Scholar] [CrossRef]
  49. Cui, J.; Li, W.; Fang, C.; Su, S.; Luan, J.; Gao, T.; Hu, L.; Lu, Y.; Chen, G. AdaBoost Ensemble Correction Models for TDDFT Calculated Absorption Energies. IEEE Access 2019, 7, 38397–38406. [Google Scholar] [CrossRef]
  50. Lyon, K.; Preciado-Rivas, M.R.; Zamora-Ledezma, C.; Despoja, V.; Mowbray, D.J. LCAO-TDDFT-k-ω: Spectroscopy in the optical limit. J. Phys. Condens. Matter. 2020, 32, 415901. [Google Scholar] [CrossRef]
  51. Tiwary, A.S. Prediction of Charge Transfer Transition Energies of the Molecular Complexes of Pmda with a Series of Methylbenzenes by Tddft. J. Indian Chem. Soc. 2019, 96, 659–665. [Google Scholar]
  52. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  53. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef] [PubMed]
  54. Legault, C.Y. CYLview, version 1.0b; Université de Sherbrooke: Quebec, QC, Canada, 2009.
Scheme 1. C–S borylation of aryl sulfides.
Scheme 1. C–S borylation of aryl sulfides.
Catalysts 14 00550 sch001
Scheme 2. The total reaction from R1 + R2→P.
Scheme 2. The total reaction from R1 + R2→P.
Catalysts 14 00550 sch002
Figure 1. Frontier molecular orbital of BQ and R1 + BQ models.
Figure 1. Frontier molecular orbital of BQ and R1 + BQ models.
Catalysts 14 00550 g001
Figure 2. The ρhole (blue) and ρele (green) of S0→S1, S0→S2, and S0→S3 of R1 + BQ.
Figure 2. The ρhole (blue) and ρele (green) of S0→S1, S0→S2, and S0→S3 of R1 + BQ.
Catalysts 14 00550 g002
Scheme 3. The detailed reaction photocatalysis process.
Scheme 3. The detailed reaction photocatalysis process.
Catalysts 14 00550 sch003
Figure 3. The Gibbs free energy surfaces in photocatalysis process.
Figure 3. The Gibbs free energy surfaces in photocatalysis process.
Catalysts 14 00550 g003
Figure 4. The Gibbs free energy surfaces of Paths a1-1, a1-2, and a1-3.
Figure 4. The Gibbs free energy surfaces of Paths a1-1, a1-2, and a1-3.
Catalysts 14 00550 g004
Figure 5. The electron spin density isosurface graphs of some structures in Path a1 and a2.
Figure 5. The electron spin density isosurface graphs of some structures in Path a1 and a2.
Catalysts 14 00550 g005
Scheme 4. The detailed reaction process from IM1 to P via two probable Paths, a2-1 and a2-2.
Scheme 4. The detailed reaction process from IM1 to P via two probable Paths, a2-1 and a2-2.
Catalysts 14 00550 sch004
Figure 6. The Gibbs free energy surfaces from IM1 to P via probable Path a2-1.
Figure 6. The Gibbs free energy surfaces from IM1 to P via probable Path a2-1.
Catalysts 14 00550 g006
Figure 7. The Gibbs free energy surfaces from IM1 to P via probable Path a2-2.
Figure 7. The Gibbs free energy surfaces from IM1 to P via probable Path a2-2.
Catalysts 14 00550 g007
Table 1. The excitation analysis of several models with at M06-2X-D3/ma-def2-SVP level.
Table 1. The excitation analysis of several models with at M06-2X-D3/ma-def2-SVP level.
StateE (kcal/mol)Λ (nm)fOSOrbital (Coefficient)
BQS1265.6450.20.0000H > L (91.6%)
S2295.1405.20.0000H-2 > L (89.0%)
S3417.7286.30.0000H-1 > L (99.9%)
BQ + R1S1266.0449.60.0033H-3 > L (37.5%)
H > L (57.5%)
S2274.0436.40.0029H-3 > L (51.5%)
H > L (42.0%)
S3299.9398.70.0000H-5 > L (79.4%)
H-4 > L (9.0%)
fOS: oscillator strength.
Table 2. Fukui functions and dual descriptors of R1, R2, and IM3a2 at the M06-2X-D3/ma-def2-SVP level.
Table 2. Fukui functions and dual descriptors of R1, R2, and IM3a2 at the M06-2X-D3/ma-def2-SVP level.
Atomq(N)q(N+1)q(N−1)ff+CDD
R1S3−0.0355−0.10540.28340.31890.0699−0.2491
C4−0.0227−0.07260.02460.04730.04990.0025
R2B50.16500.04700.20850.04350.11810.0746
B60.16500.04700.20850.04350.11810.0746
IM3a2O1−0.1258−0.1258−0.07590.04990.0176−0.0323
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ming, Y.; Feng, T.; Chen, B.; Zhou, D. Photoinduced Mechanisms of C–S Borylation of Methyl(p-tolyl)Sulfane with Bis(Pinacolato)diboron: A Density Functional Theory Investigation. Catalysts 2024, 14, 550. https://doi.org/10.3390/catal14080550

AMA Style

Ming Y, Feng T, Chen B, Zhou D. Photoinduced Mechanisms of C–S Borylation of Methyl(p-tolyl)Sulfane with Bis(Pinacolato)diboron: A Density Functional Theory Investigation. Catalysts. 2024; 14(8):550. https://doi.org/10.3390/catal14080550

Chicago/Turabian Style

Ming, Yuxiao, Tiantian Feng, Bin Chen, and Dagang Zhou. 2024. "Photoinduced Mechanisms of C–S Borylation of Methyl(p-tolyl)Sulfane with Bis(Pinacolato)diboron: A Density Functional Theory Investigation" Catalysts 14, no. 8: 550. https://doi.org/10.3390/catal14080550

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