Next Article in Journal
Argon Adsorption on Cationic Gold Clusters Aun+ (n ≤ 20)
Next Article in Special Issue
Recent Advances in Asymmetric Synthesis of Pyrrolidine-Based Organocatalysts and Their Application: A 15-Year Update
Previous Article in Journal
The Fruits of Paris polyphylla Inhibit Colorectal Cancer Cell Migration Induced by Fusobacterium nucleatum-Derived Extracellular Vesicles
Previous Article in Special Issue
New Mesoporous Silica-Supported Organocatalysts Based on (2S)-(1,2,4-Triazol-3-yl)-Proline: Efficient, Reusable, and Heterogeneous Catalysts for the Asymmetric Aldol Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

The Activating Effect of Strong Acid for Pd-Catalyzed Directed C–H Activation by Concerted Metalation-Deprotonation Mechanism

1
Lab of Computational Chemistry and Drug Design, State Key Laboratory of Chemical Oncogenomics, Peking University Shenzhen Graduate School, Shenzhen 518055, China
2
Shenzhen Bay Laboratory, Shenzhen 518132, China
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(13), 4083; https://doi.org/10.3390/molecules26134083
Submission received: 8 June 2021 / Revised: 29 June 2021 / Accepted: 1 July 2021 / Published: 4 July 2021
(This article belongs to the Collection Recent Advances in Organocatalysis)

Abstract

:
A computational study on the origin of the activating effect for Pd-catalyzed directed C–H activation by the concerted metalation-deprotonation (CMD) mechanism is conducted. DFT calculations indicate that strong acids can make Pd catalysts coordinate with directing groups (DGs) of the substrates more strongly and lower the C–H activation energy barrier. For the CMD mechanism, the electrophilicity of the Pd center and the basicity of the corresponding acid ligand for deprotonating the C–H bond are vital to the overall C–H activation energy barrier. Furthermore, this rule might disclose the role of some additives for C–H activation.

1. Introduction

Transition metal-catalyzed C–H activation to synthesize diverse organic molecules from simple hydrocarbon derivatives has emerged as a powerful tool for C–C and C-heteroatom bond formation and has received significant attention in recent years [1,2,3,4,5,6,7]. However, regioselectivity and reactivity have remained the most significant challenges in this active research field. Therefore, different strategies, especially directed C–H activation strategies, have been developed to improve regioselectivity and reactivity for C–H activation [8,9,10,11,12,13,14,15,16,17,18,19,20,21]. The directing groups (DGs) on substrates can chelate the metal catalyst and guide it to a specific position. Strong coordinating DGs usually contain strongly coordinating atoms, such as nitrogen, phosphorus, or sulfur atoms. They could widely promote C–H activation functionalization [22,23,24]. However, strong coordinating DGs are challenging to remove from the final products, limiting the utility of this strategy. Weak coordinating DGs (e.g., ketones, carboxylic acids, and ethers) are commonly occurring functional groups on the substrates and usually have much lower reactivity for C–H activation reaction [25,26,27,28,29,30]. Yu and co-workers have established several strategies to overcome the low reactivity of C–H activation by weak coordination, such as using counter cation effect, auxiliaries and monoprotected amino acid ligands [31,32,33,34,35]. Although impressive progress has been made, the scope of application of these strategies is still limited [36,37]. Using strong acid is also a widely used strategy to promote C–H activation; for example, palladium(II)-catalyzed ortho-selective C–H chlorination/bromination has demonstrated that proper strong acids (TFA, TfOH) could promote the reactivity of ortho-selective C–H bond cleavage (Scheme 1 and Scheme 2) [25,26,27,28,29,30,38,39,40]. These experimental results showed that higher catalytic activity occurred in reactions with lower pKa value acids.
The activating effect of a strong acid on Pd(II)-Catalyzed directed C–H activation has been known for a long time. Many kinds of C–H functionalization can be promoted by TFA or TfOH, such as carboxylation, [41] olefination, [42] arylation, [43,44] fluorination, [45] carbonylation, [46] trifluoromethylation, [47] amidation, [48] and oxygenation [49,50]. However, current understanding of the nature of strong acid-assisted C–H activation is still limited [51,52]. Fujiwara proposed that strong acid as a solvent facilitates the generation of highly cationic species([PdX]+) through ligand exchange (Scheme 3), which are very electrophilic. Cyclopalladium intermediates can be formed through the electrophilic aromatic substitution (SEAr) of the C–H bond [53]. Other researchers have had a similar opinion to Fujiwara regarding the activating effect of strong acid [41,42,43,44,45,46,47,48,49,50].
Besides SEAr, other frequently proposed mechanisms for C–H activation, including oxidative addition, σ-bond metathesis, and CMD mechanism, do not generate cationic species ([PdX]+) [54,55,56,57,58,59]. The CMD mechanism has been widely accepted as the best pathway for Pd-catalyzed C–H bond cleavage (Scheme 4) [60,61,62,63,64,65]. Therefore, we carried out density functional theory (DFT) [66] studies to explore the origin of the activating effect of strong acid on Pd-catalyzed directed C–H activation.

2. Results and Discussion

Since C–H activation is usually involved in the rate-determining step (RDS), [8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50] we hypothesized the relative free energy of transition states of C–H cleavage (ΔGTS) can determine the reactivity. The CMD mechanism was chosen for C–H activation, which usually has the lowest barrier among the frequently proposed mechanisms [57,58]. The transition state for the SEAr mechanism could not be located (see Figure S1, Supplementary Materials). Five substrates with different DGs were chosen; these substrates have previously been studied by experiments [38,39,40,49]. Trimeric [Pd(OAc)2]3 was chosen as the reference point of DFT calculations [67,68,69]. X-ray crystallography has provided evidence that when a strong acid such as TFA or TfOH is used, the OAc in the palladium acetate can be exchanged with TFA or OTf to form Pd(TFA)2 or Pd(OTf)2 [48,70,71,72]. As shown in Figure 1, for all of the five substrates, the ΔGTS using three different Pd catalysts is in the same order: ΔGTS[Pd(OTf)2] < ΔGTS[Pd(TFA)2] < ΔGTS[Pd(OAc)2]. The order of reactivity is consistent with the experimental results, [38,39,40,73,74] indicating our DFT calculation is reliable.
Next, energy decomposition strategy was used to explore the origin of the activating effect for directed C–H activation by strong acid [63,75,76,77,78]. As shown in Scheme 5, ΔGTS can be decomposed into two parts: ΔG1 and ΔG2. ΔG1 is the reaction energy caused by the coordination between the DG and the Pd catalyst. ΔG2 represents the energy needed to proceed with the C–H activation from int1. This strategy can reflect how the three different Pd catalysts influence ΔG1 and ΔG2, respectively.
For all of the five substrates, the order of ΔG1 using the three different Pd catalysts can be summarized as ΔG1[Pd(OTf)2] < ΔG1[Pd(TFA)2] < ΔG1[Pd(OAc)2] (see Figure 2a), which is in the reverse order of electrophilicity of the Pd catalysts: Pd(OTf)2 > Pd(TFA)2 > Pd(OAc)2 [48]. For ΔG1, the conclusion can be drawn that the more electrophilic Pd catalyst results in better coordination with DGs. For ΔG2, all of the five substrates have the consistent order: ΔG2[Pd(TFA)2] < ΔG2[Pd(OTf)2] < ΔG2[Pd(OAc)2] (see Figure 2b). The order of ΔG2 is different from that of ΔG1, and the C–H activation energy barrier of Pd(TFA)2 is the lowest. It was generally believed that a more electrophilic Pd catalyst would result in a lower barrier for the C–H activation step in the past [41,42,43,44,45,46,47,48,49,50]. However, our results do not support this belief: the electrophilicity of Pd(OTf)2 is strongest, but its C–H activation energy barrier is not the lowest. Therefore, the reason why the C–H activation energy barrier of Pd(TFA)2 is the lowest needs further study.
Inspection of the TS by the CMD mechanism demonstrated that ΔG2 is related to the metal center’s electrophilicity and the basicity of the acid ligand (Scheme 6a). To investigate the influence of electrophilicity and basicity on ΔG2 separately, an intermolecular model was built to decompose the effect of the two factors (Scheme 6b).
As shown in Figure 3, sub5 was chosen as an example for the intermolecular model study. In each row, the Pd catalyst in the three TSs is the same, but with three different external acid ligands, i.e., OAc, TFA and OTf. From each row, we can see how the basicity of acid ligands influences ΔG2. In each column, the external acid ligand in the three TSs is the same, but with three different Pd catalysts, i.e., Pd(OAc)2, Pd(TFA)2, Pd(OTf)2. From each column, we can see how electrophilicity of the Pd catalysts influences ΔG2. The three diagonal transition states, which have the same external acid ligand and the ligand of Pd catalysts, are most similar to our intramolecular mechanism.
For each row, the C–H activation energy barrier of the intermolecular model decreases with increasing basicity of the external acid ligand. The Pd(X)2_OAc transition states have the lowest energy barrier. The energy differences between the OAc and TFA ligands are about 4.6 and 6.3 kcal/mol, and the similar gaps between the TFA and OTf ligands are about 5.6 and 6.8 kcal/mol. For each column, the C–H activation energy barrier of the intermolecular model decreases with the increasing electrophilicity of the Pd catalyst, and the Pd(OTf)2_X transition states have the lowest energy barrier. The energy difference between the Pd(OAc)2 and Pd(TFA)2 catalysts is about 13.9~15.8 kcal/mol, much larger than the gap of 2.1~3.1 kcal/mol between the Pd(TFA)2 and Pd(OTf)2 catalysts. Therefore, the Pd(OTf)2_OAc transition state with the strongest basicity of the OAc ligand and the strongest electrophilicity of the Pd(OTf)2 catalyst has the lowest energy barrier among the nine intermolecular models.
However, for the three diagonal transition states with the same external acid ligand and ligand of Pd catalysts, electrophilicity and basicity show an opposite trend. For example, although the electrophilicity of the metal center in Pd(OTf)2 is the strongest, the basicity of the acid ligand (OTf) is the weakest. The Pd(TFA)2_TFA transition state has the lowest energy barrier considering the influence of the external acid ligand’s basicity and the Pd catalyst’s electrophilicity, consistent with the lowest C–H activation energy barrier of Pd(TFA)2 in the intramolecular CMD process. According to the above discussion, it can be concluded that for the CMD mechanism, the electrophilicity of Pd catalysts and the basicity of acid ligands are critical to C–H activation.
Inspired by the lowest activation energy of Pd(OTf)2_OAc, we hypothesized that C–H activation via an intermolecular CMD mechanism with a strong electrophilic Pd catalyst and strong external base may be favored, and some experiments support our hypothesis. In Yu’s work, the combination of Pd(OTf)2 and N-Methyl-2-pyrrolidone (NMP, a stronger base than TfO) is crucial for C–H fluorination [45]. Our calculations indicate that the NMP-assistant intermolecular C–H activation process is about 9 kcal/mol lower in energy than the intramolecular C–H deprotonation by TfO (see Figure 4a). Buchwald and coworkers found that the combination of Pd(OAc)2/TFA and DMSO can improve the yield of C–H arylation [44]. They proposed that palladium black formation could be slowed by the addition of DMSO (10 mol%). Our calculations indicate that DMSO, a stronger base than TFA, can also promote intermolecular deprotonation (see Figure 4b). As shown in Figure 4 and Figure S3 (Supplementary Materials), our calculations demonstrated that the intermolecular mechanism is more favorable than the intramolecular mechanism for the above two studies. Our findings might disclose the role of some additives for C–H functionalization.

3. Conclusions

In summary, the activating effect of strong acid for Pd(II)-catalyzed directed C–H functionalization was investigated with DFT calculations. Our results were consistent with previous experimental results and disclosed that the origin of the activating effect by strong acid comes from two parts: ΔG1 (coordination energy between the DG and the Pd catalyst) and ΔG2 (C–H activation energy). For the CMD mechanism, the electrophilicity of the Pd center and the basicity of the related acid ligand for deprotonation of the C–H bond is vital to the overall C–H activation energy barrier. This rule can be used to explain the role of some additives for C–H activation. It is hoped that our study could be used to improve the reactivity of some C–H functionalization reactions.

Supplementary Materials

Figure S1: The scan of C–H bond of cationic species ([PdOAc]+), Figure S2: The relative free energy of different intermediates of sub2, Figure S3: The relative free energy barrier of the intermolecular model.

Author Contributions

Computation, data curation, writing—original draft preparation, H.J.; writing—review and editing, supervision, project administration, T.-Y.S.; funding acquisition, H.J. and T.-Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Shenzhen Post-doctoral Special Funding, grant number S219201013 and China Postdoctoral Science Foundation, grant number 8206300344.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

More details about DFT calculation are in Supporting Information. This material is available free of charge via the Internet at http://pubs.acs.org.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Shilov, A.E.; Shul’pin, G.B. Activation of C–H Bonds by Metal Complexes. Chem. Rev. 1997, 97, 2879–2932. [Google Scholar] [CrossRef] [PubMed]
  2. Daugulis, O.; Do, H.-Q.; Shabashov, D. Palladium- and Copper-Catalyzed Arylation of Carbon−Hydrogen Bonds. Acc. Chem. Res. 2009, 42, 1074–1086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Lyons, T.W.; Sanford, M.S. Palladium-Catalyzed Ligand-Directed C–H Functionalization Reactions. Chem. Rev. 2010, 110, 1147–1169. [Google Scholar] [CrossRef] [Green Version]
  4. Neufeldt, S.R.; Sanford, M.S. Controlling Site Selectivity in Palladium-Catalyzed C–H Bond Functionalization. Acc. Chem. Res. 2012, 45, 936–946. [Google Scholar] [CrossRef] [Green Version]
  5. Ackermann, L. Carboxylate-Assisted Ruthenium-Catalyzed Alkyne Annulations by C–H/Het–H Bond Functionalizations. Acc. Chem. Res. 2014, 47, 281–295. [Google Scholar] [CrossRef]
  6. Gao, K.; Yoshikai, N. Low-Valent Cobalt Catalysis: New Opportunities for C–H Functionalization. Acc. Chem. Res. 2014, 47, 1208–1219. [Google Scholar] [CrossRef]
  7. Gandeepan, P.; Müller, T.; Zell, D.; Cera, G.; Warratz, S.; Ackermann, L. 3d Transition Metals for C–H Activation. Chem. Rev. 2019, 119, 2192–2452. [Google Scholar] [CrossRef] [PubMed]
  8. Phipps, R.J.; Gaunt, M.J. A Meta-Selective Copper-Catalyzed C–H Bond Arylation. Science 2009, 323, 1593–1597. [Google Scholar] [CrossRef]
  9. Wang, D.-H.; Engle, K.M.; Shi, B.-F.; Yu, J.-Q. Ligand-Enabled Reactivity and Selectivity in a Synthetically Versatile Aryl C–H Olefination. Science 2010, 327, 315–319. [Google Scholar] [CrossRef] [Green Version]
  10. Brückl, T.; Baxter, R.D.; Ishihara, Y.; Baran, P.S. Innate and Guided C–H Functionalization Logic. Acc. Chem. Res. 2011, 45, 826–839. [Google Scholar] [CrossRef] [Green Version]
  11. Rousseau, G.; Breit, B. Removable Directing Groups in Organic Synthesis and Catalysis. Angew. Chem. Int. Ed. 2011, 50, 2450–2494. [Google Scholar] [CrossRef]
  12. Luo, J.; Preciado, S.; Larrosa, I. Overriding Ortho–Para Selectivity via a Traceless Directing Group Relay Strategy: The Meta-Selective Arylation of Phenols. J. Am. Chem. Soc. 2014, 136, 4109–4112. [Google Scholar] [CrossRef] [PubMed]
  13. Chu, J.C.K.; Rovis, T. Complementary Strategies for Directed C(sp3)−H Functionalization: A Comparison of Transition-Metal-Catalyzed Activation, Hydrogen Atom Transfer, and Carbene/Nitrene Transfer. Angew. Chem. Int. Ed. 2018, 57, 62–101. [Google Scholar] [CrossRef] [PubMed]
  14. Dey, A.; Sinha, S.K.; Achar, T.K.; Maiti, D. Accessing Remote meta-and para-C(sp2)–HBonds withCovalently Attached Directing Groups. Angew. Chem. Int. Ed. 2019, 58, 10820–10843. [Google Scholar] [CrossRef] [PubMed]
  15. Tang, R.-Y.; Li, G.; Yu, J.-Q. Conformation-induced remote meta-C–H activation of amines. Nature 2014, 507, 215–220. [Google Scholar] [CrossRef] [Green Version]
  16. Zhao, C.; Crimmin, M.R.; Toste, F.D.; Bergman, R.G. Ligand-Based Carbon–Nitrogen Bond Forming Reactions of Metal Dinitrosyl Complexes with Alkenes and Their Application to C–H Bond Functionalization. Acc. Chem. Res. 2014, 47, 517–529. [Google Scholar] [CrossRef]
  17. Rouquet, G.; Chatani, N. Catalytic Functionalization of C(sp2)-H and C(sp3)-H Bonds by Using Bidentate Directing Groups. Angew. Chem. Int. Ed. 2013, 52, 11726–11743. [Google Scholar] [CrossRef] [PubMed]
  18. Leow, D.; Li, G.; Mei, T.-S.; Yu, J.-Q. Activation of remote meta-C–H bonds assisted by an end-on template. Nature 2012, 486, 518–522. [Google Scholar] [CrossRef]
  19. Lee, S.; Lee, H.; Tan, K.L. Meta-Selective C–H Functionalization Using a Nitrile-Based Directing Group and Cleavable Si-Tether. J. Am. Chem. Soc. 2013, 135, 18778–18781. [Google Scholar] [CrossRef] [Green Version]
  20. Negretti, S.; Narayan, A.R.H.; Chiou, K.C.; Kells, P.M.; Stachowski, J.L.; Hansen, D.A.; Podust, L.M.; Montgomery, J.; Sherman, D.H. Directing Group-Controlled Regioselectivity in an Enzymatic C–H Bond Oxygenation. J. Am. Chem. Soc. 2014, 136, 4901–4904. [Google Scholar] [CrossRef]
  21. Sambiagio, C.; Schönbauer, D.; Blieck, R.; Dao-Huy, T.; Pototschnig, G.; Schaaf, P.; Wiesinger, T.; Zia, M.F.; Wencel-Delord, J.; Besset, T.; et al. A comprehensive overview of directing groups applied in metal-catalysed C–H functionalisation chemistry. Chem. Soc. Rev. 2018, 47, 6603–6743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Chen, Z.; Wang, B.; Zhang, J.; Yu, W.; Liu, Z.; Zhang, Y. Transition metal-catalyzed C–H bond functionalizations by the use of diverse directing groups. Org. Chem. Front. 2015, 2, 1107–1295. [Google Scholar] [CrossRef]
  23. Xu, Y.; Dong, G. sp3 C–H activation via exo-type directing groups. Chem. Sci. 2018, 9, 1424–1432. [Google Scholar] [CrossRef] [Green Version]
  24. Rej, S.; Ano, Y.; Chatani, N. Bidentate Directing Groups: An Efficient Tool in C–H Bond Functionalization Chemistry for the Expedient Construction of C–C Bonds. Chem. Rev. 2020, 120, 1788–1887. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, Y.-H.; Shi, B.-F.; Yu, J.-Q. Palladium(II)-Catalyzed ortho Alkylation of Benzoic Acids with Alkyl Halides. Angew. Chem. Int. Ed. 2009, 48, 6097–6100. [Google Scholar] [CrossRef] [Green Version]
  26. Dai, H.-X.; Stepan, A.F.; Plummer, M.S.; Zhang, Y.-H.; Yu, J.-Q. Divergent C–H Functionalizations Directed by Sulfonamide Pharmacophores: Late-Stage Diversification as a Tool for Drug Discovery. J. Am. Chem. Soc. 2011, 133, 7222–7228. [Google Scholar] [CrossRef] [PubMed]
  27. Engle, K.M.; Mei, T.-S.; Wasa, M.; Yu, J.-Q. Weak Coordination as a Powerful Means for Developing Broadly Useful C–H Functionalization Reactions. Acc. Chem. Res. 2012, 45, 788–802. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, X.-C.; Hu, Y.; Bonacorsi, S.; Hong, Y.; Burrell, R.; Yu, J.-Q. Pd(II)-Catalyzed C–H Iodination Using Molecular I2 as the Sole Oxidant. J. Am. Chem. Soc. 2013, 135, 10326–10329. [Google Scholar] [CrossRef] [Green Version]
  29. Ma, S.; Villa, G.; Thuy-Boun, P.S.; Homs, A.; Yu, J.-Q. Palladium-Catalyzed ortho-Selective C–H Deuteration of Arenes: Evidence for Superior Reactivity of Weakly Coordinated Palladacycles. Angew. Chem. Int. Ed. 2014, 53, 734–737. [Google Scholar] [CrossRef] [PubMed]
  30. Gandeepan, P.; Ackermann, L. Transient Directing Groups for Transformative C–H Activation by Synergistic Metal Catalysis. Chem 2018, 4, 199–222. [Google Scholar] [CrossRef] [Green Version]
  31. Chan, K.S.L.; Wasa, M.; Wang, X.; Yu, J.-Q. Palladium(II)-Catalyzed Selective Monofluorination of Benzoic Acids Using a Practical Auxiliary: A Weak-Coordination Approach. Angew. Chem. Int. Ed. 2011, 50, 9081–9084. [Google Scholar] [CrossRef] [PubMed]
  32. Tran, A.T.; Yu, J.-Q. Practical Alkoxythiocarbonyl Auxiliaries for Iridium(I)-Catalyzed C–H Alkylation of Azacycles. Angew. Chem. Int. Ed. 2017, 56, 10530–10534. [Google Scholar] [CrossRef] [PubMed]
  33. Lu, Y.; Wang, D.-H.; Engle, K.M.; Yu, J.-Q. Pd(II)-Catalyzed Hydroxyl-Directed C–H Olefination Enabled by Monoprotected Amino Acid Ligands. J. Am. Chem. Soc. 2010, 132, 5916–5921. [Google Scholar] [CrossRef] [Green Version]
  34. Plata, R.E.; Hill, D.E.; Haines, B.E.; Musaev, D.G.; Chu, L.; Hickey, D.P.; Sigman, M.S.; Yu, J.-Q.; Blackmond, D.G. A Role for Pd(IV) in Catalytic Enantioselective C–H Functionalization with Monoprotected Amino Acid Ligands under Mild Conditions. J. Am. Chem. Soc. 2017, 139, 9238–9245. [Google Scholar] [CrossRef]
  35. Mei, T.-S.; Giri, R.; Maugel, N.; Yu, J.-Q. PdII-Catalyzed Monoselective ortho Halogenation of C–H Bonds Assisted by Counter Cations: A Complementary Method to Directed ortho-Lithiation. Angew. Chem. Int. Ed. 2008, 47, 5215–5219. [Google Scholar] [CrossRef] [PubMed]
  36. Yu, Y.; Lv, H.; Li, S. The C–H functionalization of organic cations: An interesting and fresh journey. Org. Biomol. Chem. 2020, 18, 8810–8826. [Google Scholar] [CrossRef]
  37. Pototschnig, G.; Maulide, N.; Schnürch, M. Direct Functionalization of C–H Bonds by Iron, Nickel, and Cobalt Catalysis. Chem. Eur. J. 2017, 23, 9206–9232. [Google Scholar] [CrossRef]
  38. Sun, X.; Shan, G.; Sun, Y.; Rao, Y. Regio- and Chemoselective C-H Chlorination/Bromination of Electron-Deficient Arenes by Weak Coordination and Study of Relative Directing-Group Abilities. Angew. Chem. Int. Ed. 2013, 52, 4440–4444. [Google Scholar] [CrossRef]
  39. Tischler, O.; Kovács, S.; Érsek, G.; Králl, P.; Daru, J.; Stirling, A.; Novák, Z. Study of Lewis acid accelerated palladium catalyzed Csingle bondH activation. J. Mol. Catal. A Chem. 2017, 426, 444–450. [Google Scholar] [CrossRef]
  40. Vana, J.; Bartacek, J.; Hanusek, J.; Roithová, J.; Sedlak, M. C–H Functionalizations by Palladium Carboxylates: The Acid Effect. J. Org. Chem. 2019, 84, 12746–12754. [Google Scholar] [CrossRef]
  41. Fujiwara, Y.; Jia, C. New developments in transition metal-catalyzed synthetic reactions via C–H bond activation. Pure Appl. Chem. 2001, 73, 319–324. [Google Scholar] [CrossRef]
  42. Boele, M.D.K.; van Strijdonck, G.P.F.; de Vries, A.H.M.; Kamer, P.C.J.; de Vries, J.G.; van Leeuwen, P.W.N.M. Selective Pd-Catalyzed Oxidative Coupling of Anilides with Olefins through C–H Bond Activation at Room Temperature. J. Am. Chem. Soc. 2002, 124, 1586–1587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Daugulis, O.; Zaitsev, V.G. Anilide ortho-Arylation by Using C–H Activation Methodology. Angew. Chem. Int. Ed. 2005, 44, 4046–4048. [Google Scholar] [CrossRef] [PubMed]
  44. Brasche, G.; García-Fortanet, J.; Buchwald, S.L. Twofold C–H Functionalization: Palladium-Catalyzed Ortho Arylation of Anilides. Org. Lett. 2008, 10, 2207–2210. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, X.; Mei, T.-S.; Yu, J.-Q. Versatile Pd(OTf)2·2H2O-Catalyzed ortho-Fluorination Using NMP as a Promoter. J. Am. Chem. Soc. 2009, 131, 7520–7521. [Google Scholar] [CrossRef] [PubMed]
  46. Houlden, C.E.; Hutchby, M.; Bailey, C.D.; Ford, J.G.; Tyler, S.N.G.; Gagné, M.R.; Lloyd-Jones, G.C.; Booker-Milburn, K.I. Room-Temperature Palladium-Catalyzed C–H Activation: Ortho-Carbonylation of Aniline Derivatives. Angew. Chem. Int. Ed. 2009, 48, 1830–1833. [Google Scholar] [CrossRef] [Green Version]
  47. Wang, X.; Truesdale, L.; Yu, J.-Q. Pd(II)-Catalyzed ortho-Trifluoromethylation of Arenes Using TFA as a Promoter. J. Am. Chem. Soc. 2010, 132, 3648–3649. [Google Scholar] [CrossRef]
  48. Xiao, B.; Gong, T.-J.; Xu, J.; Liu, Z.-J.; Liu, L. Palladium-Catalyzed Intermolecular Directed C–H Amidation of Aromatic Ketones. J. Am. Chem. Soc. 2011, 133, 1466–1474. [Google Scholar] [CrossRef]
  49. Shan, G.; Yang, X.; Ma, L.; Rao, Y. Pd-Catalyzed C-H Oxygenation with TFA/TFAA: Expedient Access to Oxygen-Containing Heterocycles and Late-Stage Drug Modification. Angew. Chem. Int. Ed. 2012, 51, 13070–13074. [Google Scholar] [CrossRef]
  50. Nishikata, T.; Abela, A.R.; Huang, S.; Lipshutz, B.H. Cationic Palladium(II) Catalysis: C–H Activation/Suzuki-Miyaura Couplings at Room Temperature. J. Am. Chem. Soc. 2010, 132, 4978–4979. [Google Scholar] [CrossRef] [Green Version]
  51. Yang, Y.; Yang, X.; Zhang, Y.; Xue, Y. Computational Mechanism Study on Allylic Oxidation of cis-Internal Alkenes: Insight into the Lewis Acid-Assisted Brønsted Acid (LBA) Catalysis in Heteroene Reactions. J. Org. Chem. 2018, 83, 13344–13355. [Google Scholar] [CrossRef]
  52. Ghaderi, A.; Iwasaki, T.; Kambe, N. Pivalic Acid-Assisted Rh(III)-Catalyzed C–H Functionalization of 2-Arylpyridine Derivatives Using Arylsilanes. Asian J. Org. Chem. 2019, 8, 1344–1347. [Google Scholar] [CrossRef]
  53. Jia, C.; Kitamura, T.; Fujiwara, Y. Catalytic Functionalization of Arenes and Alkanes via C–H Bond Activation. Acc. Chem. Res. 2001, 34, 633–639. [Google Scholar] [CrossRef] [PubMed]
  54. Engle, K.M.; Wang, D.-H.; Yu, J.-Q. Ligand-Accelerated C–H Activation Reactions: Evidence for a Switch of Mechanism. J. Am. Chem. Soc. 2010, 132, 14137–14151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Labinger, J.A.; Bercaw, J.E. Understanding and exploiting C–H bond activation. Nature 2002, 417, 507–514. [Google Scholar] [CrossRef]
  56. Ackermann, L. Carboxylate-Assisted Transition-Metal-Catalyzed C–H Bond Functionalizations: Mechanism and Scope. Chem. Rev. 2011, 111, 1315–1345. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, S.; Shi, L.; Ding, Y. Theoretical Analysis of the Mechanism of Palladium(II) Acetate-Catalyzed Oxidative Heck Coupling of Electron-Deficient Arenes with Alkenes: Effects of the Pyridine-Type Ancillary Ligand and Origins of the meta-Regioselectivity. J. Am. Chem. Soc. 2011, 133, 20218–20229. [Google Scholar] [CrossRef]
  58. Yang, Y.-F.; Cheng, G.-J.; Liu, P.; Leow, D.; Sun, T.-Y.; Chen, P.; Zhang, X.; Yu, J.-Q.; Wu, Y.-D.; Houk, K.N. Palladium-Catalyzed Meta-Selective C–H Bond Activation with a Nitrile-Containing Template: Computational Study on Mechanism and Origins of Selectivity. J. Am. Chem. Soc. 2014, 136, 344–355. [Google Scholar] [CrossRef]
  59. Cheng, G.-J.; Yang, Y.-F.; Liu, P.; Chen, P.; Sun, T.-Y.; Li, G.; Zhang, X.; Houk, K.N.; Yu, J.-Q.; Wu, Y.-D. Role of N-Acyl Amino Acid Ligands in Pd(II)-Catalyzed Remote C–H Activation of Tethered Arenes. J. Am. Chem. Soc. 2014, 136, 894–897. [Google Scholar] [CrossRef] [PubMed]
  60. Gorelsky, S.I.; Lapointe, D.; Fagnou, K. Analysis of the Concerted Metalation-Deprotonation Mechanism in Palladium-Catalyzed Direct Arylation Across a Broad Range of Aromatic Substrates. J. Am. Chem. Soc. 2008, 130, 10848–10849. [Google Scholar] [CrossRef] [PubMed]
  61. Potavathri, S.; Pereira, K.C.; Gorelsky, S.I.; Pike, A.; LeBris, A.P.; DeBoef, B. Regioselective Oxidative Arylation of Indoles Bearing N-Alkyl Protecting Groups: Dual C–H Functionalization via a Concerted Metalation-Deprotonation Mechanism. J. Am. Chem. Soc. 2010, 132, 14676–14681. [Google Scholar] [CrossRef] [Green Version]
  62. Anand, M.; Sunoj, R.B. Palladium(II)-Catalyzed Direct Alkoxylation of Arenes: Evidence for Solvent-Assisted Concerted Metalation Deprotonation. Org. Lett. 2011, 13, 4802–4805. [Google Scholar] [CrossRef]
  63. Gorelsky, S.I.; Lapointe, D.; Fagnou, K. Analysis of the Palladium-Catalyzed (Aromatic)C–H Bond Metalation–Deprotonation Mechanism Spanning the Entire Spectrum of Arenes. J. Org. Chem. 2012, 77, 658–668. [Google Scholar] [CrossRef] [PubMed]
  64. Davies, D.; Macgregor, S.A.; McMullin, C.L. Computational Studies of Carboxylate-Assisted C–H Activation and Functionalization at Group 8–10 Transition Metal Centers. Chem. Rev. 2017, 117, 8649–8709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Shan, C.; Zhu, L.; Qu, L.-B.; Bai, R.; Lan, Y. Mechanistic view of Ru-catalyzed C–H bond activation and functionalization: Computational advances. Chem. Soc. Rev. 2018, 47, 7552–7576. [Google Scholar] [CrossRef]
  66. Geometries were optimized with M06/(LANL2DZ+f: Pd; 6-31G(d): Other atoms). Single-point energies were calculated with M06/(SDD: Pd; 6-311++G(d, p): Other atoms) with the SMD solvation model (dichloroethane was chosen as the solvent). All calculations were performed with Gaussian 09. (Frisch, M.J.; et al.) See SI for more computational details.
  67. Skapski, A.C.; Smart, M.L. The crystal structure of trimeric palladium(II) acetate. J. Chem. Soc. D 1970, 11, b658–b659. [Google Scholar] [CrossRef]
  68. Kurzeev, S.A.; Kazankov, G.M.; Ryabov, A.D. Second- and inverse order pathways in the mechanism of orthopalladation of primary amines. Inorg. Chim. Acta 2002, 340, 192–196. [Google Scholar] [CrossRef]
  69. Bakhmutov, V.I.; Berry, J.F.; Cotton, F.A.; Ibragimov, S.; Murillo, C.A. Non-trivial behavior of palladium(ii) acetate. Dalton Trans. 2005, 1989–1992. [Google Scholar] [CrossRef]
  70. Gandeepan, P.; Parthasarathy, K.; Cheng, C.-H. Synthesis of Phenanthrone Derivatives from sec-Alkyl Aryl Ketones and Aryl Halides via a Palladium-Catalyzed Dual C–H Bond Activation and Enolate Cyclization. J. Am. Chem. Soc. 2010, 132, 8569–8571. [Google Scholar] [CrossRef] [PubMed]
  71. Yeung, C.S.; Zhao, X.; Borduas, N.; Dong, V.M. Pd-catalyzed ortho-arylation of phenylacetamides, benzamides, and anilides with simple arenes using sodium persulfate. Chem. Sci. 2010, 1, 331–336. [Google Scholar] [CrossRef]
  72. Zhao, X.; Yeung, C.S.; Dong, V.M. Palladium-Catalyzed Ortho-Arylation of O-Phenylcarbamates with Simple Arenes and Sodium Persulfate. J. Am. Chem. Soc. 2010, 132, 5837–5844. [Google Scholar] [CrossRef] [PubMed]
  73. Zhou, D.-B.; Wang, G.-W. Synthesis of [60]Fullerene-Fused Tetralones via Palladium- Catalyzed Ketone-Directed sp2 C-H Activation and sp3 C-H Functionalization. Adv. Synth. Catal. 2016, 358, 1548–1554. [Google Scholar] [CrossRef]
  74. Tóth, B.L.; Kovács, S.; Sályi, G.; Novák, Z. Mild and Efficient Palladium-Catalyzed Direct Trifluoroethylation of Aromatic Systems by C–H Activation. Angew. Chem. Int. Ed. 2016, 55, 1988–1992. [Google Scholar] [CrossRef]
  75. Guner, V.A.; Houk, K.N.; Davies, I.W. Computational Studies on the Electrocyclizations of 1-Amino-1,3,5-hexatrienes. J. Org. Chem. 2004, 69, 8024–8028. [Google Scholar] [CrossRef] [PubMed]
  76. Yu, T.-Q.; Fu, Y.; Liu, L.; Guo, Q.-X. How to Promote Sluggish Electrocyclization of 1,3,5-Hexatrienes by Captodative Substitution. J. Org. Chem. 2006, 71, 6157–6164. [Google Scholar] [CrossRef]
  77. Xue, L.; Su, W.; Lin, Z. Mechanism of silver- and copper-catalyzed decarboxylation reactions of aryl carboxylic acids. Dalton Trans. 2011, 40, 11926–11936. [Google Scholar] [CrossRef]
  78. Ryabov, A.D. Cyclopalladated Complexes in Organic Synthesis. Synthesis 1985, 233–252. [Google Scholar] [CrossRef]
Scheme 1. Corresponding pKa of different acids in ortho–selective C–H bond functionalization.
Scheme 1. Corresponding pKa of different acids in ortho–selective C–H bond functionalization.
Molecules 26 04083 sch001
Scheme 2. Corresponding reaction yields of different acids in ortho–selective C–H bond functionalization.
Scheme 2. Corresponding reaction yields of different acids in ortho–selective C–H bond functionalization.
Molecules 26 04083 sch002
Scheme 3. Ligand exchange process.
Scheme 3. Ligand exchange process.
Molecules 26 04083 sch003
Scheme 4. The SEAr and CMD mechanism for Pd–catalyzed directed C–H activation.
Scheme 4. The SEAr and CMD mechanism for Pd–catalyzed directed C–H activation.
Molecules 26 04083 sch004
Figure 1. Calculated activation free energies by CMD mechanism (ΔGTS, kcal/mol) for the five substrates with three different Pd catalysts. The coordination atoms are marked in red.
Figure 1. Calculated activation free energies by CMD mechanism (ΔGTS, kcal/mol) for the five substrates with three different Pd catalysts. The coordination atoms are marked in red.
Molecules 26 04083 g001
Scheme 5. ΔGTS can be decomposed into two parts: ΔG1 and ΔG2. They represent coordinating energy and C–H activation energy barrier respectively.
Scheme 5. ΔGTS can be decomposed into two parts: ΔG1 and ΔG2. They represent coordinating energy and C–H activation energy barrier respectively.
Molecules 26 04083 sch005
Figure 2. (a) ΔG1 (kcal/mol) for the 5 substrates with three different Pd catalysts; (b) ΔG2 (kcal/mol) for the 5 substrates with three different Pd catalysts.
Figure 2. (a) ΔG1 (kcal/mol) for the 5 substrates with three different Pd catalysts; (b) ΔG2 (kcal/mol) for the 5 substrates with three different Pd catalysts.
Molecules 26 04083 g002
Scheme 6. The electrophilicity of the metal center and the basicity of the acid ligand in the (a) intramolecular model and (b) intermolecular model.
Scheme 6. The electrophilicity of the metal center and the basicity of the acid ligand in the (a) intramolecular model and (b) intermolecular model.
Molecules 26 04083 sch006
Figure 3. The relative free activation energies (kcal/mol) for the intermolecular models of sub5 with three different Pd catalysts and three different anionic external acid ligands (OAc, TFA, TfO).
Figure 3. The relative free activation energies (kcal/mol) for the intermolecular models of sub5 with three different Pd catalysts and three different anionic external acid ligands (OAc, TFA, TfO).
Molecules 26 04083 g003
Figure 4. Calculated relative activation free energies (kcal/mol) of intermolecular and intramolecular CMD mechanism. Intermolecular models were chosen as the reference point. (a) NMP−assistant process; (b) DMSO−assistant process.
Figure 4. Calculated relative activation free energies (kcal/mol) of intermolecular and intramolecular CMD mechanism. Intermolecular models were chosen as the reference point. (a) NMP−assistant process; (b) DMSO−assistant process.
Molecules 26 04083 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jiang, H.; Sun, T.-Y. The Activating Effect of Strong Acid for Pd-Catalyzed Directed C–H Activation by Concerted Metalation-Deprotonation Mechanism. Molecules 2021, 26, 4083. https://doi.org/10.3390/molecules26134083

AMA Style

Jiang H, Sun T-Y. The Activating Effect of Strong Acid for Pd-Catalyzed Directed C–H Activation by Concerted Metalation-Deprotonation Mechanism. Molecules. 2021; 26(13):4083. https://doi.org/10.3390/molecules26134083

Chicago/Turabian Style

Jiang, Heming, and Tian-Yu Sun. 2021. "The Activating Effect of Strong Acid for Pd-Catalyzed Directed C–H Activation by Concerted Metalation-Deprotonation Mechanism" Molecules 26, no. 13: 4083. https://doi.org/10.3390/molecules26134083

Article Metrics

Back to TopTop