Next Article in Journal
Cadaveric Adipose-Derived Stem Cells for Regenerative Medicine and Research
Next Article in Special Issue
Unveiling the Nature and Strength of Selenium-Centered Chalcogen Bonds in Binary Complexes of SeO2 with Oxygen-/Sulfur-Containing Lewis Bases: Insights from Theoretical Calculations
Previous Article in Journal
Review of Eukaryote Cellular Membrane Lipid Composition, with Special Attention to the Fatty Acids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Spectral Physics of Stable Cu(III) Produced by Oxidative Addition of an Alkyl Halide

1
School of Mathematics and Physics, University of Science and Technology Beijing, Beijing 100083, China
2
Research Institute for Electronic Science, Hokkaido University, Sapporo 001-0021, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(21), 15694; https://doi.org/10.3390/ijms242115694
Submission received: 5 October 2023 / Revised: 23 October 2023 / Accepted: 26 October 2023 / Published: 28 October 2023

Abstract

:
In this paper, we theoretically investigated spectral physics on Cu(III) complexes formed by the oxidative addition of α-haloacetonitrile to ionic and neutral Cu(I) complexes, stimulated by recent experimental reports. Firstly, the electronic structures of reactants of α-haloacetonitrile and neutral Cu(I) and two kinds of products of Cu(III) complexes are visualized with the density of state (DOS) and orbital energy levels of HOMO and LUMO. The visually manifested static and dynamic polarizability as well as the first hyperpolarizability are employed to reveal the vibrational modes of the normal and resonance Raman spectra of two Cu(III) complexes. The nuclear magnetic resonance (NMR) spectra are not only used to identify the reactants and products but also to distinguish between two Cu(III) complexes. The charge difference density (CDD) reveals intramolecular charge transfer in electronic transitions in optical absorption spectra. The CDDs in fluorescence visually reveal electron–hole recombination. Our results promote a deeper understanding of the physical mechanism of stable Cu(III) produced by the oxidative addition of an alkyl halide.

1. Introduction

As a cheap and plentiful metal, Cu can occur in oxidation states from 0 to 4+, although +1 (cuprous) and +2 (cupric) oxidation states are the most common, while Cu with 0 and +4 oxidation states are very rare. High-valent organo-Cu(III) species were mostly considered the main intermediates in many aerobic oxidation reactions, enzyme-catalyzed oxidation reactions, as well as Cu-catalyzed cross-coupling reactions [1,2,3,4,5,6,7]. Reductive elimination of Cu(III) intermediates is frequently projected as an important step in various Cu-catalyzed or copper-mediated formation of C-heteroatom or C-C bonds. Organo-Cu(III) species are produced in several ways [7]: (1) two-electron oxidative addition of Cu(I) species; (2) oxidation of Cu(II) intermediate via a single-electron transfer pathway or via halide atom transfer [8,9]; (3) σ-bond metathesis through a four-centre intermediate; and (4) p-complexation of Cu(I) on ArX. In (1) and (2), the Cu species changes its oxidation state during the catalytic cycle, whereas in (3) and (4), the Cu species preserves the same oxidation state through the whole cycle. Usually, there are four types of well-defined Cu(III) complexes among the d8 Cu(III) complexes: (1) Cu(III) complexes without Cu-Carbon bond [10,11,12,13,14,15,16,17,18,19,20]; (2) organo-Cu(III) complexes supported by macrocyclic ligand and with Cu-C bond [21,22,23,24,25,26,27,28,29,30,31,32]; (3) fluoroalkylated organo-Cu(III) complexes [33,34,35,36,37,38,39,40,41,42,43,44]; and (4) organo-Cu(III) complexes containing a Cu(III)-C bond [45,46,47,48,49,50,51,52,53].
Contrary to lots of other high-valent transition metal complexes, Cu(III) complexes are usually thermally unstable and hard to isolate. (CF3)2CuS2CNE, as an organo-Cu(III) specie, was isolated and characterized (vide infra) in 1989 [54]. Before 2000, few well-defined Cu(III) complexes were obtained and well-characterized because of their low thermal stabilities [55]. In the early 2000s, with rapid-injection nuclear magnetic resonance (RI-NMR) spectroscopy, Cu(III) specie was demonstrated [22,23,24,56]. It is urgent to study the organo-Cu(III) complexes with thermal stability and their elementary bond-forming reductive elimination because these investigations can reveal the fundamentals of Cu catalysis and promote the discovery of new reactions. Recently, Shen et al. experimentally reported the oxidative addition of an alkyl halide to form a stable Cu(III) product (see Figure 1) [57], and this oxidative addition of a C(sp3) − X bond to Cu(I) can help promote more effectual Cu-catalyzed cross-coupling reactions of alkyl electrophiles.
In this paper, we theoretically investigate the spectral physics of stable Cu(III) produced by the oxidative addition of an alkyl halide, stimulated by the recent experimental report in ref. [57]. The electronic structures of reactants of α-haloacetonitrile and neutral Cu(I) and two kinds of products of Cu(III) complexes are analyzed with DOS and by measuring the orbital energy levels. The static and dynamic polarizability as well as the first hyperpolarizability reveal the vibrational modes of the normal and resonance Raman spectra of two Cu(III) complexes. The NMR spectra are not only used to identify the reactants and products but also to distinguish between two Cu(III) complexes. Absorption and fluorescence spectra and charge transfer in electronic transitions are manifested. Our results promote a deeper understanding of the physical and chemical properties of Cu(III) complexes.

2. Results and Discussion

Figure 2 shows the orbital energy level and DOS of trans-4, cis-4, and (bpy) being far away from Cu(CF3)2(CH2CN) before complex. Without interaction between (bpy) and Cu(CF3)2(CH2CN), the DOS of HOMO and LUMO are localized on (bpy) and Cu (CF3)2(CH2CN), respectively. When the complexes are formed, the DOS of HOMO are localized on the Cu CF3)2(CH2CN), and DOS are not only localized on Cu(CF3)2(CH2CN), but also on the (bpy), which directly reflects the interaction between (bpy) and Cu(CF3)2(CH2CN), especially for cis-4, and this interaction results in an increase in the orbital energy level of LUMO for the complexes.
Figure 3 demonstrates the noncovalent interactions (NCI) plotetween (bpy) and Cu(CF3)2(CH2CN) for trans-4 and cis-4, respectively, with top and side views. It was found that there is a strong attraction between (bpy) and Cu(CF3)2(CH2CN), which is the key factor in forming stable complexes. The visual interaction also clearly demonstrates the influence of the (CH2CN) position on the difference in interaction between trans-4 and cis-4.
Trans-4 is more stable than cis-4 since the ground energy of trans-4 is lower (0.1864 eV) than that of cis-4. The ground state permanent dipole moment energy refers to the permanent dipole moment possessed by a molecule in the ground state. It is usually represented by μ, which can be used to measure the interaction ability between molecules and external electric fields. Polarization rate is a physical quantity that describes the degree of polarization of a molecule or material under the effect of an external electric field. It describes the ability to react to external electric fields. The symbol for polarizability is usually α; the larger its value, the easier the material is to polarize under external electric fields, and vice versa. Hyperpolarizability is a physical quantity that describes the response of a material to an external electric field and can reflect its nonlinear optical properties. The symbol for hyperpolarizability is usually β. The data on ground permanent dipole moment, polarizability, and the first hyperpolarizability are listed in Table 1 and Table 2. It is found that there are similar ground dipole moments and polarizability between trans-4 and cis-4 (see Table 1), and the visualization of polarizability supports the above conclusion in Figure 4, but there is a large difference in the first hyperpolarizability in Table 2 and Figure 5. In Figure 4 and Figure 5, the blue, white, and red colors represent the polarizability from small to large. The arrows on the spherical surface in the figure represent the change in the dipole moment of the molecule when the center of the molecule applies an external electric field of the same strength in all directions. The green, pink, and light blue long arrows represent the general trend of the polarizability of the system along three different orientations.
Figure 6 shows the normal Raman spectra of trans-4 and cis-4. It is found that their profiles are almost the same because of their similar permanent polyaxiality, which can be seen in Figure 4, since Raman intensity (I) is estimated with I = α2E4, where α is permanent polarizability. The green, pink, and light blue long arrows in Figure 4 represent the general trend of the polarizability of the system along three different orientations, which determines the molecular vibrational modes along the direction of the vector arrow’s direction of polarizability. The insert in Figure 6 demonstrates the vibrational modes of trans-4 and cis-4, respectively. Due to the different orientations of (CH2CN), their Raman intensities are significantly different.
NMR spectroscopy is a technique that provides information about the quantities and types of atoms of a particular element in a molecule. Figure 7a–e are the NMR spectra of Cu, C, F, N, and H atoms of (bpy) being far away from Cu(CF3)2(CH2CN) before complex, respectively. Figure 7a’–e’ are the NMR spectra of Cu, C, F, N, and H atoms of trans-4 and cis-4, respectively, which can distinguish trans-4 from cis-4. Figure 7f,g are the trans-4 and cis-4 with labels, respectively. By comparing the NMR spectra of each atom before and after reactions, we can see that the NMR spectra of Cu, F, N, and H can well characterize the formation of trans-4 and cis-4 because of the large shielding. For example, the NMR of the Cu spectrum demonstrates that shielding is 63.46 ppm, while the shielding of trans-4 and cis-4 is 16.5 ppm and 91.1 ppm; thus, NMR can not only identify whether the C(III) complex formed but can also distinguish trans-4 from cis-4. Also, NMR spectra of F demonstrate that there is a lot of shielding, especially the shielding of atom #15, which is 94 ppm for cis-4, as can be seen in Figure 7c’, while the shielding of atom #15 of Cu(CF3)2(CH2CN) is about 30.9 ppm. The shielding of degenerated N#3 and N#8 of (bpy) is 66.8 ppm, which can be seen in Figure 7d, which are separated to −5 ppm and 28 ppm, respectively. The shielding of N#20 of CH2CN is 54 ppm in Figure 7d, which is separated to −7 ppm and −9 ppm, respectively, for trans-4 and cis-4. By comparing Figure 7e,e’, one can examine whether the trans-4 and cis-4 are formed. Furthermore, the NMR spectra of the H atom for trans-4 and cis-4 show a great difference, which can also be used to distinguish between trans-4 and cis-4 in Figure 7e’.
Electronic transitions corresponding to optical absorption and fluorescence spectra can be seen in Figure 8, where the inserts are the CDDs. Figure 8a,b shows the absorption spectra of trans-4 and cis-4, respectively. CDDs demonstrate electron transfer from Cu (CF3)2 to (bpy) and (CH2CN) for trans-4 at the absorption peak around 300 nm, while for cis-4, they demonstrate electron transfer from Cu(CF3)(CH2CN) to (CF3) and (bpy). While for the peak around 250 nm, they both exhibit intramolecular charge transfers, but within (bpy) and Cu(CF3)2(CH2CN), respectively. Figure 8c manifests that the fluorescence is around 410 nm and electron transfer is from (bpy) to Cu(CF3)2 in electron–hole recombination in fluorescence. The excited geometry optimization of cis-4 from the structure in Figure 8d was performed, but it did not converge and stopped at the structure in Figure 8e, which demonstrates that the structure of cis-4 was broken, where (CH2CN) moved to one side of (bpy). The intramolecular charge transfer between the organic component and Cu(III) can further promote stability by optical excitation, which reflects the stability of the sample explored in light atmospheric environments.
It can be seen that the normal Raman spectra of trans-4 and cis-4 in Figure 6 are similar and have weak Raman intensities due to the similar static permanent polarizability at the ground state, as can also be seen in Figure 7. To significantly increase the Raman intensities, as can be seen in Figure 6, the resonance Raman spectra of trans-4 and cis-4 are calculated; see Figure 9. Figure 9a–c shows the resonance Raman spectra of trans-4, excited by different resonant electronic state excitations. Dynamic polarization rate is one of the types of polarization rates that is mainly used to describe the polarization response of molecules or materials under high-frequency electric fields. It is found that resonance Raman spectra in Figure 9a–c are significantly selectively enhanced by charge transfer enhancement, resulting from the different dynamic polarizability at different incident lights; see the visualization with top and side views in Figure 10a. There are similar results for the resonance Raman spectra of cis-4 shown in Figure 9d–f, which are significantly enhanced by dynamic polarizability on different resonant electronic transitions in Figure 10b.

3. Materials and Methods

All calculations were performed with Gaussian 16 software [58] based on density functional theory (DFT) [59]. The geometries of trans-4 and cis-4 were optimized at ground state. B3LYP functional [60] and 6-31G(d) basis sets were used for C, N, F, and H, and the LANL2DZ basis set was used for Cu. With the optimized geometries, the electronic structures, the noncovalent interactions (NCI) [61], (hyper) polarizability, normal Raman spectra, and NMR spectra were calculated at the same level of the theory. Electronic transitions in absorption and fluorescence and resonance Raman spectra were calculated with time-dependent DFT (TD-DFT) [62], CAM-B3LYP functional [63], and 6-31G(d) basis sets for C, N, F, and H, and the LANL2DZ basis set for Cu. The geometries of the S1 excited state for trans-4 and cis-4 were optimized, with TD-DFT, B3LYP functional, and 6-31G(d) basis sets being used for C, N, F, and H, and the LANL2DZ basis set being used for Cu. With the optimized geometries of the S1 excited state, the fluorescence of trans-4 and cis-4 was calculated with TD-DFT and CAM-B3LYP functional, with the 6-31G(d) basis set being used for C, N, F, and H, and the LANL2DZ basis set being used for Cu. The visualizations of polarizability were performed according to refs. [64,65,66].

4. Conclusions

Stable Cu(III) produced by the oxidative addition of an alkyl halide was theoretically investigated using the spectral analysis method. NMR spectra can not only reveal the dynamic processes of forming trans-4 and cis-4 but also distinguish trans-4 from cis-4. Resonance Raman spectra are selectively and significantly enhanced by chemical enhancement, which can also distinguish trans-4 from cis-4. CDDs of electronic transitions reveal the electron–hole separation in optical absorption processes and can manifest the electron–hole recombination in fluorescence. Our theoretical analysis provides an understanding of the physical mechanism of stable Cu(III) produced by the oxidative addition of an alkyl halide.

Author Contributions

Conceptualization, M.S.; methodology, M.S.; software, E.C.; validation, M.S. and E.C.; formal analysis, M.S. and E.C.; investigation, M.S. and E.C.; resources, M.S.; data curation, M.S. and E.C.; writing—original draft preparation, M.S.; writing—review and editing, M.S. and E.C.; visualization, M.S. and E.C.; supervision, M.S.; project administration, M.S.; funding acquisition, M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Foundation of China (Grant Nos. 11874407, 91436102, and 11374353) and the Fundamental Research Funds for the Central Universities (Grant No. 06500067).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available on request from authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, H.; Shen, Q. Well-defined organometallic Copper(III) complexes: Preparation, characterization and reactivity. Coord. Chem. Rev. 2021, 439, 213923. [Google Scholar] [CrossRef]
  2. Hickman, A.J.; Sanford, M.S. High-valent organometallic copper and palladium in catalysis. Nature 2012, 484, 177–185. [Google Scholar] [CrossRef] [PubMed]
  3. Hruszkewycz, D.; McCann, S.; Stahl, S.S. Cu-Catalyzed Aerobic Oxidation: Overview and New Developments; Wiley-VCH: Weinheim, Germany, 2016. [Google Scholar]
  4. Cho, S.H.; Kim, J.Y.; Kwak, J.; Chang, S. Recent advances in the transition metal-catalyzed twofold oxidative C-H bond activation strategy for C-C and C-N bond formation. Chem. Soc. Rev. 2011, 40, 5068–5083. [Google Scholar] [CrossRef]
  5. Que, L., Jr.; Tolman, W.B. Bis(μ-oxo)dimetal “Diamond” Cores in Copper and Iron Complexes Relevant to Biocatalysis. Angew. Chem. Int. Ed. 2002, 41, 1114–1137. [Google Scholar]
  6. Trammell, R.; Rajabimoghadam, K.; Garcia-Bosch, I. Copper-Promoted Functionalization of Organic Molecules: From Biologically Relevant Cu/O2 Model System to Organometallic Transformations. Chem. Rev. 2019, 119, 2954–3031. [Google Scholar] [CrossRef] [PubMed]
  7. Sperotto, E.; van Klink, G.P.M.; van Koten, G.; de Vries, J.G. The mechanism of the modified Ullmann reaction. Dalton Trans. 2010, 39, 10338–10351. [Google Scholar] [CrossRef] [PubMed]
  8. Casitas, A.; Ribas, X. The role of organometallic copper (III) complexes in homogeneous catalysis. Chem. Sci. 2013, 4, 2301–2318. [Google Scholar] [CrossRef]
  9. Li, S.-J.; Lan, Y. Is Cu(III) a necessary intermediate in Cu-mediated coupling reactions? A mechanistic point of view. Chem. Commun. 2020, 56, 6609–6619. [Google Scholar] [CrossRef]
  10. Cole, A.P.; Root, D.E.; Mukherjee, P.; Solomon, E.I.; Stack, T.D.P. A Trinuclear Intermediate in the Copper-Mediated Reduction of O2: Four Electrons from Three Copper. Science 1996, 273, 1848–1850. [Google Scholar] [CrossRef]
  11. Sinha, W.; Sommer, M.G.; Deibel, N.; Ehret, F.; Bauer, M.; Sarkar, B.; Kar, S. Experimental and Theoretical Investigations of the Existence of CuII, CuIII, and CuIV in Copper Corrolato Complexes. Angew. Chem. Int. Ed. 2015, 54, 13769–13774. [Google Scholar] [CrossRef]
  12. Srnec, M.; Navratil, R.; Andris, E.; Jasik, J.; Rouithova, J. Experimentally calibrated analysis of the electronic structure of CuO+: Implications for reactivity. Angew. Chem. Int. Ed. 2018, 52, 17053–17054. [Google Scholar] [CrossRef] [PubMed]
  13. Lemon, C.M.; Huynh, M.; Maher, A.G.; Anderson, B.L.; Bloch, E.D.; Powers, D.C.; Nocera, D.G. Electronic structure of copper corroles. Angew. Chem. Int. Ed. 2016, 55, 2176–2180. [Google Scholar] [CrossRef] [PubMed]
  14. Spaeth, A.D.; Gagnon, N.L.; Dhar, D.; Yee, G.M.; Tolman, W.B. Determination of the Cu(III)-OH Bond Distance by Resonance Raman Spectroscopy Using a Normalized Version of Badger’s Rule. J. Am. Chem. Soc. 2017, 139, 4477–4485. [Google Scholar] [CrossRef] [PubMed]
  15. Neisen, B.D.; Gagnon, N.L.; Dhar, D.; Spaeth, A.D.; Tolman, W.B. Formally Copper(III)-Alkylperoxo Complexes as Models of Possible Intermediates in Monooxygenase Enzymes. J. Am. Chem. Soc. 2017, 139, 10220–10223. [Google Scholar] [CrossRef] [PubMed]
  16. Dhar, D.; Yee, G.M.; Markle, T.F.; Mayer, J.M.; Tolman, W.B. Reactivity of the copper (III)-hydroxide unit with phenols. Chem. Sci. 2017, 8, 1075–1085. [Google Scholar] [CrossRef]
  17. Bailey, W.D.; Dhar, D.; Cramblitt, A.C.; Tolman, W.B. Mechanistic dichotomy in proton-coupled electron-transfer reactions of phenols with a copper superoxide complex. Am. Chem. Soc. 2019, 141, 5470–5480. [Google Scholar] [CrossRef] [PubMed]
  18. Mandal, M.; Elwell, C.E.; Bouchey, C.J.; Zerk, T.J.; Tolman, W.B.; Cramer, C. Mechanisms for hydrogen-atom abstraction by mononuclear copper (III) cores: Hydrogen-atom transfer or concerted proton-coupled electron transfer? J. Am. Chem. Soc. 2019, 141, 17236–17244. [Google Scholar] [CrossRef]
  19. Wu, T.; MacMillan, S.N.; Rajabimoghadam, K.; Siegler, M.A.; Lancaster, K.M.; Garcia-Bosch, I. Structure, Spectroscopy, and Reactivity of a Mononuclear Copper Hydroxide Complex in Three Molecular Oxidation States. J. Am. Chem. Soc. 2020, 142, 12265–12276. [Google Scholar] [CrossRef]
  20. Bower, J.K.; Cypcar, A.D.; Henriquez, B.; Stieber, S.C.E.; Zhang, S. C (sp3)–H fluorination with a copper (II)/(III) redox couple. J. Am. Chem. Soc. 2020, 142, 8514–8521. [Google Scholar] [CrossRef]
  21. Hu, H.; Snyder, J.P. Organocuprate conjugate addition: The square-planar “CuIII” intermediate. Am. Chem. Soc. 2007, 129, 7210–7211. [Google Scholar] [CrossRef]
  22. Bertz, S.H.; Cope, S.; Murphy, M.; Ogle, C.A.; Taylor, B.J. Rapid injection NMR in mechanistic organocopper chemistry. Preparation of the elusive copper (III) intermediate. J. Am. Chem. Soc. 2007, 129, 7208–7209. [Google Scholar] [CrossRef] [PubMed]
  23. Gärtner, T.; Henze, W.; Gschwind, R.M. NMR-detection of Cu(III) intermediates in substitution reactions of alkyl halides with Gilman cuprates. Am. Chem. Soc. 2007, 129, 11362–11363. [Google Scholar] [CrossRef] [PubMed]
  24. Bertz, S.H.; Cope, S.; Dorton, D.; Murphy, M.; Ogle, C.A. Organocuprate cross-coupling: The central role of the copper (III) intermediate and the importance of the copper (I) precursor. Angew. Chem. Int. Ed. 2007, 46, 7082–7085. [Google Scholar] [CrossRef]
  25. Bartholomew, E.R.; Bertz, S.H.; Cope, S.; Murphy, M.; Ogle, C.A. Preparation of σ-and π-Allylcopper (III) Intermediates in SN2 and SN2′ Reactions of Organocuprate (I) Reagents with Allylic Substrates. J. Am. Chem. Soc. 2008, 130, 11244–11245. [Google Scholar] [CrossRef] [PubMed]
  26. Bertz, S.H.; Hardin, R.A.; Murphy, M.D.; Ogle, C.A.; Richter, J.D.; Thomas, A.A. Rapid Injection NMR Reveals η3 ‘π-Allyl’CuIII Intermediates in Addition Reactions of Organocuprate Reagents. J. Am. Chem. Soc. 2012, 134, 9557–9560. [Google Scholar] [CrossRef] [PubMed]
  27. Maurya, Y.K.; Noda, K.; Yamasumi, K.; Mori, S.; Uchiyama, T.; Kamitani, K.; Hirai, T.; Ninomiya, K.; Nishibori, M.; Hori, Y.; et al. Ground-state copper (III) stabilized by N-confused/N-linked corroles: Synthesis, characterization, and redox reactivity. J. Am. Chem. Soc. 2018, 140, 6883–6892. [Google Scholar] [CrossRef]
  28. Adinarayana, B.; Thomas, A.P.; Suresh, C.H.; Srinivasan, A. A 6, 11, 16-Triarylbiphenylcorrole with an adj-CCNN Core: Stabilization of an Organocopper (III) Complex. Angew. Chem. Int. Ed. 2015, 54, 10478–10482. [Google Scholar] [CrossRef]
  29. Ke, X.-S.; Hong, Y.; Tu, P.; He, Q.; Lynch, V.M.; Kim, D.; Sessler, J.L. Hetero Cu(III)–Pd (II) complex of a dibenzo [g, p] chrysene-fused bis-dicarbacorrole with stable organic radical character. J. Am. Chem. Soc. 2017, 139, 15232–15238. [Google Scholar] [CrossRef]
  30. Ribas, X.; Jackson, D.A.; Donnadieu, B.; Mahía, J.; Parella, T.; Xifra, R.; Hedman, B.; Hodgson, K.O.; Llobet, A.; Stack, T.D.P. Aryl C-H Activation by CuII To Form an Organometallic Aryl–CuIII Species: A Novel Twist on Copper Disproportionation. Angew. Chem. Int. Ed. 2002, 41, 2991–2994. [Google Scholar] [CrossRef]
  31. Zhang, Q.; Liu, Y.; Wang, T.; Zhang, X.; Long, C.; Wu, Y.-D.; Wang, M.-X. Mechanistic study on cu (II)-catalyzed oxidative cross-coupling reaction between arenes and boronic acids under aerobic conditions. J. Am. Chem. Soc. 2018, 140, 5579–5587. [Google Scholar] [CrossRef]
  32. Liu, Y.; Resch, S.G.; Klawitter, I.; Cutsai, G.E., III; Demeshko, S.; Dechert, S.; Kühn, F.E.; DeBeer, S.; Meyer, F. An Adaptable N-Heterocyclic Carbene Macrocycle Hosting Copper in Three Oxidation States. Angew. Chem. Int. Ed. 2020, 59, 5696–5705. [Google Scholar] [CrossRef]
  33. Brothers, P.J.; Roper, W.R. Transition-metal dihalocarbene complexes. Chem. Rev. 1988, 88, 1293–1326. [Google Scholar] [CrossRef]
  34. Romine, A.M.; Nebra, N.; Konovalov, A.I.; Martin, E.; Benet-Buchholz, J.; Grushin, V.V. Easy access to the copper (III) anion [Cu (CF3) 4]; Angew. Chem. Int. Ed. 2015, 54, 2745–2749. [Google Scholar] [CrossRef]
  35. Shen, H.; Liu, Z.; Zhang, P.; Tan, X.; Zhang, Z.; Li, C. Trifluoromethylation of alkyl radicals in aqueous solution. J. Am. Chem. Soc. 2017, 139, 9843–9846. [Google Scholar] [CrossRef] [PubMed]
  36. Tan, X.; Liu, Z.; Shen, H.; Zhang, P.; Zhang, Z.; Li, C. Silver-catalyzed decarboxylative trifluoromethylation of aliphatic carboxylic acids. J. Am. Chem. Soc. 2017, 139, 12430–12433. [Google Scholar] [CrossRef]
  37. Guo, S.; AbuSalim, D.I.; Cook, S.P. Aqueous benzylic C–H trifluoromethylation for late-stage functionalization. J. Am. Chem. Soc. 2018, 140, 12378–12382. [Google Scholar] [CrossRef] [PubMed]
  38. Choi, G.; Lee, G.S.; Park, B.; Kim, D.; Hong, S.H. Direct C (sp3)−H trifluoromethylation of unactivated alkanes enabled by multifunctional trifluoromethyl copper complexes. Angew. Chem. Int. Ed. 2021, 60, 5467–5474. [Google Scholar] [CrossRef]
  39. Guo, S.; AbuSalim, D.I.; Cook, S.P. 1, 2-(Bis) trifluoromethylation of Alkynes: A One-Step Reaction to Install an Underutilized Functional Group. Angew. Chem. Int. Ed. 2019, 58, 11704–11708. [Google Scholar] [CrossRef]
  40. Le, C.; Chen, T.Q.; Liang, T.; Zhang, P.; MacMillan, D.W.C. A radical approach to the copper oxidative addition problem: Trifluoromethylation of bromoarenes. Science 2018, 360, 1010–1014. [Google Scholar] [CrossRef]
  41. Kornfilt, D.J.P.; MacMillan, D.W.C. Copper-catalyzed trifluoromethylation of alkyl bromides. J. Am. Chem. Soc. 2019, 141, 6853–6858. [Google Scholar] [CrossRef]
  42. Lu, Z.; Liu, H.; Liu, S.; Leng, X.; Lan, Y.; Shen, Q. A Key Intermediate in Copper-Mediated Arene Trifluoromethylation,[nBu4N][Cu (Ar)(CF3) 3]: Synthesis, Characterization, and C (sp2)−CF3 Reductive Elimination. Angew. Chem. Int. Ed. 2019, 58, 8510–8514. [Google Scholar] [CrossRef] [PubMed]
  43. Paeth, M.; Tyndall, S.B.; Chen, L.-Y.; Hong, J.-C.; Carson, W.P.; Liu, X.; Sun, X.; Liu, J.; Yang, K.; Hale, E.M.; et al. Csp3–Csp3 bond-forming reductive elimination from well-defined copper (III) complexes. J. Am. Chem. Soc. 2019, 141, 3153–3159. [Google Scholar] [CrossRef]
  44. Liu, S.; Liu, H.; Liu, S.; Lu, Z.; Lu, C.; Leng, X.; Lan, Y.; Shen, Q. C (sp3)-CF3 Reductive elimination from a five-coordinate neutral copper (III) complex. J. Am. Chem. Soc. 2020, 142, 9785–9791. [Google Scholar] [CrossRef]
  45. Wing, R.M. Carborane analogs of. pi.-allyls. The crystal and molecular structure of triphenylmethylphosphonium bis (3)-1, 2-dicarboryl) cuprate (III). J. Am. Chem. Soc. 1968, 90, 4828–4834. [Google Scholar] [CrossRef]
  46. Varadarajan, A.; Johnson, S.E.; Gomez, F.A.; Chakrabarti, S.; Knobler, C.B.; Hawthorne, M.F. Synthesis and structural characterization of pyrazole-bridged metalla-bis (dicarbollide) derivatives of cobalt, nickel, copper, and iron: Models for venus flytrap cluster reagents. J. Am. Chem. Soc 1992, 114, 9003–9011. [Google Scholar] [CrossRef]
  47. García-López, J.; Yañez-Rodríguez, V.; Roces, L.; García-Granda, S.; Martínez, A.; Guevara-García, A.; Castro, G.R.; Jiménez-Villacorta, F.; Iglesias, M.J.; Ortiz, F.L. Synthesis and characterization of a coupled binuclear CuI/CuIII complex. J. Am. Chem. Soc. 2010, 132, 10665–10667. [Google Scholar] [CrossRef]
  48. Liu, L.; Zhu, M.; Yu, H.-T.; Zhang, W.-X.; Xi, Z. Organocopper (III) spiro complexes: Synthesis, structural characterization, and redox transformation. J. Am. Chem. Soc. 2017, 139, 13688–13691. [Google Scholar] [CrossRef] [PubMed]
  49. Santo, R.; Miyamoto, R.; Tanaka, R.; Nishioka, T.; Sato, K.; Toyota, K.; Obata, M.; Yano, S.; Kinoshita, I.; Ichimura, A.; et al. Diamagnetic–Paramagnetic Conversion of Tris (2-pyridylthio) methylcopper (III) through a Structural Change from Trigonal Bipyramidal to Octahedral. Angew. Chem. Int. Ed. 2006, 45, 7611–7614. [Google Scholar] [CrossRef]
  50. Snyder, J.P. Elusiveness of CuIII complexation. preference for trifluoromethyl oxidation in the formation of [CuI (CF3)4]−salts. Angew. Chem. Int. Ed. Engl. 1995, 34, 80–81. [Google Scholar] [CrossRef]
  51. Walroth, R.C.; Lukens, J.T.; MacMillan, S.N.; Finkelstein, K.D.; Lancaster, K.M. Spectroscopic Evidence for a 3d10 Ground State Electronic Configuration and Ligand Field Inversion in [Cu (CF3)4] 1–. J. Am. Chem. Soc. 2016, 138, 1922–1931. [Google Scholar] [CrossRef]
  52. Dimucci, I.M.; Lukens, J.T.; Chatterjee, S.; Carsch, K.M.; Titus, C.J.; Lee, S.J.; Nordlund, D.; Betley, T.A.; MacMillan, S.N.; Lancaster, K.M. The myth of d8 copper (III). J. Am. Chem. Soc. 2019, 141, 18508–18520. [Google Scholar] [CrossRef]
  53. Hoffmann, R.; Alvarez, S.; Mealli, C.; Falceto, A.; Cahill, T.J., III; Zeng, T.; Manca, G. From widely accepted concepts in coordination chemistry to inverted ligand fields. Chem. Rev. 2016, 116, 8173–8192. [Google Scholar] [CrossRef] [PubMed]
  54. van Koten, G. Organocopper Compounds: From Elusive to Isolable Species, from Early Supramolecular Chemistry with RCuI Building Blocks to Mononuclear R2–nCuII and R3–mCuIII Compounds. A Personal View. Organometallics 2012, 31, 7634–7764. [Google Scholar] [CrossRef]
  55. Melník, M.; Kabešová, M. Copper (III) coordination compounds: Classification and analysis of crystallographic and structural data. Coord. Chem. 2000, 50, 323–338. [Google Scholar] [CrossRef]
  56. Bartholomew, E.R.; Bertz, S.H.; Cope, S.; Dorton, D.C.; Murphy, M.; Ogle, C.A. Neutral organocopper (III) complexes. Chem. Commun. 2008, 10, 1176–1177. [Google Scholar] [CrossRef] [PubMed]
  57. Luo, Y.; Li, Y.; Wu, J.; Xue, X.; Hartwig, J.F.; Shen, Q. Oxidative addition of an alkyl halide to form a stable Cu(III) product. Science 2023, 381, 1072–1079. [Google Scholar] [CrossRef]
  58. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.P.; Petersson, G.A.; Nakatsuji, H.R.; et al. Gaussian 16, Revision A. 03; Gaussian. Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  59. Kohn, W.; Sham, L.J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  60. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef] [PubMed]
  61. García, J.C.; Johnson, E.R.; Keinan, S.; Chaudret, R.; Piquemal, J.; Beratan, D.N.; Yang, W.J. NCIPLOT: A program for plotting noncovalent interaction regions. Chem. Theory Comput. 2011, 7, 625–632. [Google Scholar] [CrossRef]
  62. Gross, E.K.U.; Kohn, W. Local density-functional theory of frequency-dependent linear response. Phys. Rev. Lett. 1985, 55, 2850–2852. [Google Scholar] [CrossRef] [PubMed]
  63. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef]
  64. Lu, T.; Chen, F.W. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  65. Tuer, A.; Krouglov, S.; Cisek, R.; Tokarz, D.; Barzda, V.J. Three-dimensional visualization of the first hyperpolarizability tensor. Comput. Chem. 2011, 32, 1128–1134. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, Z.; Lu, T.; Yuan, A.; Wang, X.; Chen, Q.; Yan, X. Remarkable Size Effect on Photophysical and Nonlinear Optical Properties of All-Carboatomic Rings, Cyclo [18] carbon and Its Analogues. Chem. Asian J. 2021, 16, 2267–2271. [Google Scholar] [CrossRef]
Figure 1. The trans-[(bpy)Cu(CF3)2(CH2CN)] (trans-4) and cis-[(bpy)Cu(CF3)2(CH2CN)] (cis-4). This figure was taken from ref. [57].
Figure 1. The trans-[(bpy)Cu(CF3)2(CH2CN)] (trans-4) and cis-[(bpy)Cu(CF3)2(CH2CN)] (cis-4). This figure was taken from ref. [57].
Ijms 24 15694 g001
Figure 2. The orbital energy level and DOS of trans-4, cis-4, and (bpy) being far away from Cu(CF3)2(CH2CN) before complex.
Figure 2. The orbital energy level and DOS of trans-4, cis-4, and (bpy) being far away from Cu(CF3)2(CH2CN) before complex.
Ijms 24 15694 g002
Figure 3. Interaction between (bpy) and Cu(CF3)2(CH2CN) for trans-4 and cis-4 with top and side views, (ac) for trans-4, and (df) for cis-4.
Figure 3. Interaction between (bpy) and Cu(CF3)2(CH2CN) for trans-4 and cis-4 with top and side views, (ac) for trans-4, and (df) for cis-4.
Ijms 24 15694 g003
Figure 4. The permanent polarizability of trans-4 and cis-4 with top and side views: (ac) top and side views of trans-4; (df) top and side views of cis-4.
Figure 4. The permanent polarizability of trans-4 and cis-4 with top and side views: (ac) top and side views of trans-4; (df) top and side views of cis-4.
Ijms 24 15694 g004
Figure 5. The permanent polarizability of trans-4 and cis-4 with top and side views: (ac) top and side views of trans-4; (df) top and side views of cis-4.
Figure 5. The permanent polarizability of trans-4 and cis-4 with top and side views: (ac) top and side views of trans-4; (df) top and side views of cis-4.
Ijms 24 15694 g005
Figure 6. Normal Raman spectra of trans-4 and cis-4: (a) rans-4; and (b) cis-4.
Figure 6. Normal Raman spectra of trans-4 and cis-4: (a) rans-4; and (b) cis-4.
Ijms 24 15694 g006
Figure 7. NMR spectra of Cu, C, F, N, and H atoms before and after they formed trans-4 and cis-4, which can also be used to distinguish trans-4 from cis-4. The parts (ae) are before complex; (a’e’) trans-4 and cis-4; and (f,g) trans-4 and cis-4 with labels, respectively.
Figure 7. NMR spectra of Cu, C, F, N, and H atoms before and after they formed trans-4 and cis-4, which can also be used to distinguish trans-4 from cis-4. The parts (ae) are before complex; (a’e’) trans-4 and cis-4; and (f,g) trans-4 and cis-4 with labels, respectively.
Ijms 24 15694 g007aIjms 24 15694 g007b
Figure 8. Absorption and fluorescence spectra of trans-4 and cis-4: (a,b) absorption spectra of trans-4 and cis-4, where green and red stand for hole and electron, respectively; and (c) fluorescence of absorption spectra of trans-4, and geometries before (d) and after (e) excited state optimization of cis-4.
Figure 8. Absorption and fluorescence spectra of trans-4 and cis-4: (a,b) absorption spectra of trans-4 and cis-4, where green and red stand for hole and electron, respectively; and (c) fluorescence of absorption spectra of trans-4, and geometries before (d) and after (e) excited state optimization of cis-4.
Ijms 24 15694 g008
Figure 9. Resonance Raman spectra: (ac) resonance Raman spectra of trans-4; and (df) resonance Raman spectra of cis-4.
Figure 9. Resonance Raman spectra: (ac) resonance Raman spectra of trans-4; and (df) resonance Raman spectra of cis-4.
Ijms 24 15694 g009
Figure 10. Dynamic polarizability: (a) dynamic polarizability of trans-4; and (b) dynamic polarizability of cis-4.
Figure 10. Dynamic polarizability: (a) dynamic polarizability of trans-4; and (b) dynamic polarizability of cis-4.
Ijms 24 15694 g010
Table 1. The permanent dipole moment (Debye) and polarizability (au) of trans-4 and cis-4 at ground state, where data are the total values.
Table 1. The permanent dipole moment (Debye) and polarizability (au) of trans-4 and cis-4 at ground state, where data are the total values.
μ (x) μ (y) μ (z) α ( x x ) α ( y y ) α ( z z )
trans-4−9.00.70.0237231128
cis-410.2−2.7−1.4215253139
Table 2. The permanent first hyperpolarizability (au) of trans-4 and cis-4 at ground state.
Table 2. The permanent first hyperpolarizability (au) of trans-4 and cis-4 at ground state.
β ( x x x ) β ( x x y ) β ( x x z ) β ( x y y ) β ( x z z ) β ( y y y ) β ( y y z ) β ( y z z ) β ( z z z )
trans-4340−260102−2151330−480
cis-4−610−1551−67110−41−709184
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

Cao, E.; Sun, M. Spectral Physics of Stable Cu(III) Produced by Oxidative Addition of an Alkyl Halide. Int. J. Mol. Sci. 2023, 24, 15694. https://doi.org/10.3390/ijms242115694

AMA Style

Cao E, Sun M. Spectral Physics of Stable Cu(III) Produced by Oxidative Addition of an Alkyl Halide. International Journal of Molecular Sciences. 2023; 24(21):15694. https://doi.org/10.3390/ijms242115694

Chicago/Turabian Style

Cao, En, and Mengtao Sun. 2023. "Spectral Physics of Stable Cu(III) Produced by Oxidative Addition of an Alkyl Halide" International Journal of Molecular Sciences 24, no. 21: 15694. https://doi.org/10.3390/ijms242115694

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