Next Article in Journal / Special Issue
Novel TiO2-Supported Gold Nanoflowers for Efficient Photocatalytic NOx Abatement
Previous Article in Journal
Effects of Selective and Mixed-Action Kappa and Delta Opioid Receptor Agonists on Pain-Related Behavioral Depression in Mice
Previous Article in Special Issue
Fluorescence of Half-Twisted 10-Acyl-1-methyltetrahydrobenzoquinolines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pentafluorophenyl Copper–Biarylsulfoxide Complexes: Synthesis and Photoreactivity

1
Laboratoire Hétérochimie Fondamentale et Appliquée (UMR 5069), Université de Toulouse, CNRS, 118 Route de Narbonne, CEDEX 09, 31062 Toulouse, France
2
Institut de Chimie de Toulouse (UAR 2599), 118 Route de Narbonne, CEDEX 09, 31062 Toulouse, France
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(14), 3332; https://doi.org/10.3390/molecules29143332 (registering DOI)
Submission received: 18 June 2024 / Revised: 5 July 2024 / Accepted: 12 July 2024 / Published: 16 July 2024
(This article belongs to the Special Issue Feature Papers in Photochemistry and Photocatalysis)

Abstract

:
Pentafluorophenyl copper(I)–biarylsulfoxide complexes, existing as [Cu(C6F5)]4L2, both in solution and in the solid state, were prepared and thoroughly characterized. Subsequently, the photochemistry of the complexes was explored, showing inherent photoreactivity of the biarylsulfoxide moiety within the coordination sphere of the copper. Photoinduced cross-coupling reactions between the anthryl moiety of bis-anthracenylsulfoxide and pentafluorobenzene, and synthesis of Cu2O (cuprite), were demonstrated.

1. Introduction

The sulfoxide function is ubiquitous nowadays in chemistry, as illustrated per the well-known dimethylsulfoxide, which is widely used in medicine as a bio-compatible solvent for biologically relevant molecules [1,2,3,4], or in synthetic chemistry as a solvent presenting advantageous properties such as high polarity, high boiling point, and low toxicity.
Interestingly, some peculiar biarylsulfoxides scaffolds exhibit photoreactivity. For example, it has been known since 1973 that dibenzothiophene-S-oxide (DBTO 1) and its derivatives undergo S–O bond cleavage under UV-light irradiation, releasing the parent photostable dibenzothiophene (DBT 2) and atomic oxygen (see Scheme 1a) [5,6,7]. Remarkably, some biarylsulfoxides, such as bis-anthracenylsulfoxide (Anthra2SO 3), can undergo a fundamentally different type of photoreactivity, exhibiting extrusion of SO forming bianthryl 4 quantitatively (see Scheme 1b) under UV light [8].
On the other hand, biarylsulfoxides are often disregarded as ligands. Indeed, their weak coordination behavior often implies using a multidentate ligand to ensure that the sulfoxide enters the coordination sphere of the metal. However, those ligands have proved to coordinate via η1-S or η1-O bonding with numerous metals [9,10], including palladium, ruthenium, and rhodium, and have found applications in different catalytic processes, mostly exploiting their stereogenicity for the synthesis of chiral compounds [11,12].
Organocopper(I) compounds are known to form aggregates in solution, from species containing two to eight members up to a polymeric structure [13,14]. Interestingly, the addition of bulky and/or chelating ligands allowed breaking the clusters, leading to smaller, well-defined aggregates [15,16], ultimately providing monomeric species [17,18]. In the present study, pentafluorophenyl copper Cu(C6F5) 5 was chosen as an interesting candidate for the synthesis of copper(I)–biarylsulfoxide complexes because of its versatility. Indeed, this species, existing as a tetramer both in solution and in the solid state [19], has been reported to present different coordination behaviors depending on the nature and stoichiometry of the ligand used [20].
Thus, starting from the versatility of pentafluorophenyl copper Cu(C6F5), from the properties of sulfoxides in coordination chemistry, and from the photoreactivity of bis–arylsulfoxides, we describe in this study the combination of these complementary parameters through the synthesis of new bis-arylsulfoxide–copper(I) complexes and their inherent reactivity under irradiation.

2. Results and Discussion

2.1. Synthesis and Photoreactivity of [Cu(C6F5)]4(DBTO)2

The highly electron-deficient metal core of 5 coordinates to arenes such as toluene, yielding the tetrameric copper cluster [Cu(C6F5)]4(toluene)2 [21]. Moreover, in 2012, Jäkle et al. demonstrated the versatility of the coordination of 5 with pyridine, obtaining at will the [Cu(C6F5)]4(pyridine)2 tetramer and the Cu(C6F5)(pyridine) monomer [17]. Interestingly, this study is supported by a systematic NMR investigation of the chemical shifts of both the fluorine and the protons of the organocopper complexes, which provide a strong basis for the present work. The present studies were thus started by mixing equimolar quantities of DBTO 1 and Cu(C6F5) 5. Pleasingly, only one set of 19F and 1H NMR signals was observed, with both showing a shift from starting materials, indicative of the formation of a new species. Crystals were then grown by slow evaporation and revealed by X-ray diffraction the tetrameric copper aggregate [Cu(C6F5)]4(DBTO)2 6, coordinated solely by two DBTO ligands (see Figure 1). The copper tetramer is nearly perfectly planar, with an elongated structure characteristic of tetranuclear copper species coordinated with two donor ligands. Pentafluorophenyl groups present interesting behavior: two are π-bonding with the DBTO moiety and lie in the plane of the Cu4 ring, whereas the other two are located above and under the Cu4 plane.
The use of DMSO as a ligand has been known to yield the monomeric complex Cu(C6F5)(DMSO) [21], with the observed metal/ligand ratio in its crystal state rather unexpected. This result prompted us to investigate its stoichiometry in solution. Similarly, to the case of Jäkle and coworkers with pyridine ligands, only one set of signals is observed by NMR, regardless of the equivalents of DBTO used (0.25 to 2 equiv. per CuC6F5), clearly stating the presence of a dynamic coordination equilibrium [17]. Moreover, both proton and fluorine NMR presented a noticeable shift depending on the equivalents of ligand used (see Figure 2a,b). A titration was thus carried out, plotting both the difference between the 19F shift of the fluorine para and the meta of the C6F5 moiety, Δδ(19F)m,p-C6F5, characteristic of the aggregation state of the copper complex [17,21] and the 1H shift of the proton Ha of DBTO, δ(1Ha) DBTO (see Figure 2c).
Δδ(19F)m,p-C6F5 quickly diminishes upon adding up to 0.5 equivalents of DBTO, clearly showing the formation of 6, and evolves only sparingly when more ligand is added. Known values of Δδ(19F)m,p-C6F5 above 10 ppm reported in the literature are consistent with [Cu(C6F5)]4L2 type complexes in solution, thus suggesting that 6 is preferentially formed in solution. More importantly, the compound does not undergo breakup of the tetramer, even upon adding excess ligand. On the other hand, the data obtained upon analyzing the 1H shifts also revealed a clear tendency, providing a regular shift upfield upon adding up to 2 equivalents of Cu(C6F5) 5, upon which 6 is formed, whereas further additions led to a plateau, suggesting that the monoligated complex [Cu(C6F5)]4(DBTO) formation is unfavored in these conditions. It is noteworthy that the shift towards the upfield region is not consistent with what was observed in the case of pyridine derivatives, but can be rationalized thanks to the X-ray diffraction structure, showing that Ha lie in the shielding region of a C6F5 substituent.
Finally, a variable temperature NMR was recorded between −30 and 50 °C using a Cu(C6F5)/DBTO ratio of 1:1 (see S2 Supplementary Information). Interestingly, lowering the temperature provided the same Δδ(19F)m,p-C6F5 value as that obtained when mixing Cu(C6F5)/DBTO in a ratio of 4:2, whereas heating the solution led to lower values. These data thus suggest again that the complex stoichiometry exhibits a dynamic equilibrium with a preferential ratio of 4:2 and that the equilibrium can shift slightly towards the incorporation of additional ligands upon warming in solution.

2.2. Synthesis and Photoreactivity of Complexes [Cu(C6F5)]4(Tol2SO)2 and [Cu(C6F5)]4(Anthra2SO)2

With the aim to further study the photoreactivity of biarylsulfoxide–CuC6F5, two new complexes, [Cu(C6F5)]4(Tol2SO)2 7 and [Cu(C6F5)]4(Anthra2SO)2 8, were synthesized, using, respectively, bis-p-tolylsulfoxide (Tol2SO) and the photoreactive bis-anthracenylsulfoxide 3. Pleasingly, X-ray diffraction analysis shown that even the strongly sterically demanding bis-anthracenylsulfoxide ligand yielded similar coordination behavior to DBTO and Tol2SO (see Figure 3).
In order to start probing the photoreactivity of the biarylsulfoxides within the coordination sphere of the copper, a quick survey of the UV-vis absorbance of the corresponding complexes was undertaken, showing a bathochromic shift up to the visible region of the main absorption features of the ligands (see S3 Supporting Information), thus encouraging us to study the photoreactivity using blue light (460 nm).
Pleasingly, the DBTO complex 6, prepared in situ in a CuC6F5/DBTO 4:2 stoichiometry, exhibited photoreactivity (see Scheme 2a). Indeed, upon irradiating 6 with a blue light LED for 5 h, DBT 2 was cleanly furnished in 92% conversion, alongside 65% of (C6F5)2 9 obtained as the sole product observed in 19F NMR, and Cu2O (see S5 Supporting Information). The deoxygenation process is in line with what is known for DBTO upon irradiation with UV light but can be promoted at a much energetically weaker wavelength thanks to the copper coordination. It is noteworthy that DBTO is unreactive in the same reaction conditions, whereas 5 undergoes slow photolysis yielding 9 in only 6% conversion. Importantly, when the solution containing a 1:1 stoichiometry was irradiated, 48% of DBT was obtained while 43% DBTO remained in solution, clearly stating that 6 is the photoreactive species of the reaction media. Encouraged by these results and hoping to extend the deoxygenation process to another biarylsulfoxide, the photoreactivity of 7 was investigated. However, no deoxygenated product was observed using blue light, and only the slow formation of 9 was observed (see Scheme 2b).
Finally, a solution of complex 8 was irradiated under blue light and yielded an unexpected range of products, with a full conversion obtained within 15 min of irradiation (see Scheme 2c). Firstly, as expected, bianthryl 4 and decafluorobiphenyl 9 were obtained in 26% and 21% NMR yields, respectively. Surprisingly, two other products arising from a cross coupling between C6F5 and both the anthryl and thioanthryl moiety, identified in 1H and 19F NMR as 10 and 11, were obtained in 28% and 15% NMR yields, respectively. Furthermore, 12, arising from an unexpected sulfur transfer of the bis-anthracenylsulfoxide to the C6F5 moiety, was formed in a 5% yield.

3. Materials and Methods

3.1. Reagents and Solvents

Unless otherwise noted, reagents were purchased from commercial suppliers and used directly without further purification. Unless indicated, technical grade solvents were purchased from commercial suppliers and used without further purification. CDCl3 and CD2Cl2 were dried and kept over activated 4 Å molecular sieves and degassed by the freeze–pump–thaw technique. All water was deionized before use. Unless stated, all reactions were carried out in Schlenk glassware under an inert atmosphere using either a standard Schlenk-line technique or an argon-filled glovebox. “Room temperature” can vary between 18 °C and 25 °C.
9,9′-Dianthryl sulfoxide [8] and 9-(anthracen-9-yldisulfanyl)anthracene [22] were prepared according to reported procedures.

3.2. Analysis and Characterization

Analytical thin layer chromatography (TLC) was performed on Merck aluminum-backed silica gel 60 F254 plates. Developed TLC plates were visualized by ultraviolet (UV) irradiation (254 nm). Column chromatography was carried out using Merk silica gel 60 Å, 220–440 mesh. Fourier transform infrared spectrometry (FTIR) was carried out using a Cary 630 FTIR using an attenuated total reflection (ATR) attachment and peaks were reported in terms of the frequency of absorption (cm–1). 1H, 13C, and 19F NMR spectra were recorded on Brucker Avance II 300 MHz, Avance III HD 400 MHz, and Avance I and II 500 MHz spectrometers (Brucker, Karlsruhe, Germany). Chemical shifts were expressed in parts per million (ppm), with residual solvent signals as an internal reference (1H and 13C{1H}). 19F NMR chemical shifts were reported in ppm relative to CFCl3. Coupling constants (J) were given in Hertz (Hz). The 1H NMR spectra were reported as follows: δ (multiplicity, coupling constant J, number of protons). Single-crystal X-ray data were collected at low temperature (193(2)K) on a Bruker APEX II Quazar diffractometer equipped with a 30 W air-cooled microfocus source (6) or on a Bruker D8 VENTURE diffractometer equipped with a PHOTON III detector (7 and 8), using MoKα radiation (λ = 0.71037 Å). The structures were solved by intrinsic phasing method [23] and refined by the full-matrix least-squares method on F2 [24]. All non-H atoms were refined with anisotropic displacement parameters and all the hydrogen atoms were refined isotropically at calculated positions using a riding model.

3.3. Synthesis of Copper–Biarylsulfoxide Complexes

[Cu(C6F5)]4(DBTO)2 6:
Molecules 29 03332 i001
Dibenzothiophene-S-oxide (10.1 mg, 0.05 mmol) and pentafluorophenyl copper (23.0 mg, 0.1 mmol) were weighted in an amber-glass vial before adding CDCl3 (1 mL) and stirring for 5 min, providing the pure complex in quantitative yield. Slow evaporation of chloroform at RT provided crystals suitable for X-ray diffraction.
1H NMR (CDCl3, 600 MHz) δH 7.74 (d, J = 7.8 Hz, 4H, Cb-H), 7.71 (d, J = 7.7 Hz, 4H, Ce-H), 7.64 (td, J = 7.5, 1.1 Hz, 4H, Cd-H), 7.49 (td, J = 7.5, 1.1 Hz, 4H, Cc-H). 13C{1H} NMR (CDCl3, 151 MHz) δC 152.4 (dd, JC-F = 233.9, 23.8 Hz, Ch), 142.4 (dd, JC-F = 253.8, 14.3 Hz, Cj), 142.1 (Ca), 136.7 (Cf), 136.3 (dddd, JC-F = 256.6, 25.0, 12.7, 5.9 Hz, Ci), 133.5 (Cd), 130.1 (Cc), 127.4 (Cb), 122.1 (Ce), 103.7 (t, JC-F = 57.2 Hz, Cg). 19F NMR (CDCl3, 282 MHz) δF −105.73 (dd, J = 30.4, 11.1 Hz), −148.18 (t, J = 20.5 Hz), −159.41 (td, J = 20.6, 20.2, 10.2 Hz). FTIR (neat) νmax/cm−1 3051, 2956, 2920, 2851, 1627, 1498, 1444, 1431, 1331, 1256, 1126, 1066, 1053, 1023, 988, 954, 753, 712.
[Cu(C6F5)]4(Tol2SO)2 7:
Molecules 29 03332 i002
Bis-p-tolyl sulfoxide (11.5 mg, 0.05 mmol) and pentafluorophenyl copper (23.0 mg, 0.1 mmol) were weighted in an amber-glass vial before adding CDCl3 (1 mL) and stirring for 5 min, providing the pure complex in quantitative yield. Layering the chloroform solution with pentane yielded crystals suitable for X-ray diffraction upon standing.
1H NMR (CDCl3, 500 MHz) δH 7.21 (d, J = 7.8 Hz, 8H, Cb-H), 7.15 (d, J = 8.4 Hz, 8H, Cc-H), 2.36 (s, 12H, Ce-H). 13C{1H} NMR (CDCl3, 126 MHz) δC 152.6 (dd, JC-F = 233.9, 24.2 Hz, Cg), 143.0 (Ca), 142.5 (d, JC-F = 253.9 Hz, Ci), 139.0 (Cd), 136.6 (ddd, JC-F = 256.9, 30.8, 12.9 Hz, Ch), 130.4 (Cc), 124.8 (Cb), 104.4 (t, JC-F = 58.7 Hz, Cf), 21.5 (Ce). 19F NMR (CDCl3, 282 MHz) δF −105.42 (d, J = 20.2 Hz, Cg-F), −149.06 (t, J = 19.9 Hz, Ci-F), −160.03 (ddt, J = 28.2, 18.0, 9.0 Hz, Ch-F). FTIR (neat) νmax/cm−1 2961, 2928, 2853, 1631, 1599, 1496, 1448, 1431, 1332, 1258, 1068, 1053, 982, 954, 805, 756, 704.
[Cu(C6F5)]4(Anthra2SO)2 8:
Molecules 29 03332 i003
9,9′-Dianthryl sulfoxide (20.1 mg, 0.05 mmol) and pentafluorophenyl copper (23.0 mg, 0.1 mmol) were weighted in an amber-glass vial before adding CD2Cl2 (2 mL) and stirring for 5 min, providing the pure complex in quantitative yield. Applying this exact protocol in CDCl3 led to precipitation of the complex, whereas using C6D6 provided crystals suitable for X-ray diffraction upon standing.
1H NMR (CD2Cl2, 600 MHz) δH 8.80 (dd, J = 9.0, 0.5 Hz, 8H, Cc-H), 8.55 (s, 4H, Ch-H), 7.99 (ddd, J = 8.4, 0.8, 0.8 Hz, 8H, Cf-H), 7.45 (ddd, J = 8.2, 6.6, 1.0 Hz, 8H, Ce-H), 7.40 (ddd, J = 8.8, 6.6, 1.4 Hz, 8H, Cd-H). 13C{1H} NMR (CD2Cl2, 151 MHz) δC 152.8 (dd, JC-F = 234.6, 24.1 Hz, Cj), 142.5 (d, JC-F = 252.1 Hz, Cl), 136.7 (ddd, JC-F = 255.5, 28.5, 11.7 Hz, Ck), 134.1, 131.5, 130.9, 130.1 (C + Ca), 128.7, 126.0, 122.6, 104.5 (broad, Ci). 19F NMR (CD2Cl2, 282 MHz) δF −106.08 (d, J = 20.5 Hz, Cj-F), −149.42 (t, J = 17.7 Hz, Cl-F), −160.44 (ddt, J = 27.3, 16.8, 8.6 Hz, Ck-F). FTIR (neat) νmax/cm−1 3082, 3051, 2956, 2924, 2853, 1625, 1497, 1446, 1431, 1331, 1258, 1068, 1053, 990, 954, 898, 775, 729.

3.4. Photochemistry

The LEDs used were high-power Vision-EL (5W, λ = 460 ± 10 nm, 410 lm). Photochemistry experiments were carried out on an NMR scale. NMR conversions were obtained using 4,4′-difluoro-1,1′-biphenyl as internal standard allowing for quantitative 1H et 19F NMR, using in both cases a relaxation delay (d1) of 20 seconds.
Prior to the photochemistry studies, it was checked that the internal standard 4,4′-difluoro-1,1′-biphenyl was not interfering with the coordination behavior of the biarylsulfoxide copper complexes. Indeed, it was shown by Jäkle et al. that CuC6F5 can coordinate to aromatic hydrocarbons [21]. The NMR shifts of the complexes with and without the internal standard were compared, showing no difference. We thus assumed that 4,4′-difluoro-1,1′-biphenyl could be used. Moreover, the photoreactivity was investigated with and without 4,4′-difluoro-1,1′-biphenyl, showing comparable results and thus excluding an impact of the internal standard on the photoreactivity.

3.4.1. Irradiation of [Cu(C6F5)]4(DBTO)2 6

A freshly prepared 0.1 M solution of [Cu(C6F5)]4(DBTO)2 6 in CDCl3 was placed in a Young tap NMR tube in an argon-filled glovebox. After adding 4,4′-difluoro-1,1′-biphenyl (19.0 mg, 0.1 mmol, 1.0 equiv.), the tube was sealed and removed from the glovebox before recording a t = 0 NMR data point. The NMR tube was then irradiated with a blue LED for 6 h, at which point full consumption of the complex was observed. Analysis of the NMR data showed 65% of (C6F5)2 9 as the sole product in 19F NMR and 92% of DBT as the sole product in 1H NMR.

3.4.2. Irradiation of [Cu(C6F5)]4(Tol2SO)2 7

Irradiation of a solution of [Cu(C6F5)]4(Tol2SO)2 7 in CDCl3 led only to a slow degradation of the CuC6F5 moiety, furnishing 6% of (C6F5)2 9 after 5 h of irradiation and no noticeable change in 1H NMR.

3.4.3. Irradiation of [Cu(C6F5)]4(Anthra2SO)2 8

A freshly prepared 0.05M solution of [Cu(C6F5)]4(Anthra2SO)2 8 in CD2Cl2 was placed in a Young tap NMR tube in an argon-filled glovebox. After adding 4,4′-difluoro-1,1′-biphenyl (19.0 mg, 0.1 mmol, 1.0 equiv.), the tube was sealed and removed from the glovebox before recording a t = 0 NMR data point. The NMR tube was then irradiated with a blue LED for 15 min, at which point full consumption of the complex was observed. Analysis of the 19F NMR data showed 21% of (C6F5)2 9 [25], 28% of 10, 15% of 11, and 5% of 12. The 1H NMR data revealed the products 10 and 11, as well as 26% of the expected bianthryl 4.

3.5. X-ray Data

CCDC 2360088 (6), CCDC 2360089 (7), and CCDC 2360090 (8) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif (accessed on 11 July 2024).

4. Conclusions

In summary, three new biarylsulfoxide–copper(I) complexes are presented. These complexes present a dynamic coordination behavior, of which [Cu(C6F5)]4L2 is the main species both in solution and in the solid state. Finally, the photochemical properties of the biarylsulfoxides within the coordination sphere of the copper have been evaluated, allowing (1) the S–O bond cleavage of DBTO under visible light and (2) access to unprecedented cross-coupling reactions between the C6F5 moiety of the organocopper and the photoreactive Anthra2SO. Finally, we want to emphasize that the present method allows for the controlled photoinduced synthesis of Cu2O, which we believe may find applications in material science, as this compound is well known to exhibit interesting optical and electrical properties [26,27].

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules29143332/s1: S1 Stoichiometry titration study data. S2 Variable-temperature 19F NMR of a solution containing a 1:1 Cu(C6F5)/DBTO ratio. S3 UV-vis absorbance of complexes 6–8. S4 Photolysis of complex 8. S5 X-ray diffraction studies. S6 NMR Spectra. S7 IR Spectra.

Author Contributions

Experiments design, V.M. and D.M.; chemical reactions, V.M. and R.L.; X-ray diffraction analysis, S.M.-L.; writing—original draft preparation, V.M.; writing—review and editing, V.M., R.L., S.M.-L., E.M., and D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Centre National de la Recherche Scientifique (CNRS) and the Université de Toulouse (UPS). We thank the ANR funding agency for its financial support of the CROSS project (ANR-21-CE09-002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available on request from the authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. David, N.A. The Pharmacology of Dimethyl Sulfoxide. Annu. Rev. Pharmacol. Toxicol. 1972, 12, 353–374. [Google Scholar] [CrossRef]
  2. Rubin, L.F. Toxicity of Dimethyl Sulfoxide, alone and in combination. Ann. N. Y. Acad. Sci. 1975, 243, 98–103. [Google Scholar] [CrossRef] [PubMed]
  3. Brayton, C.F. Dimethyl Sulfoxide (DMSO): A Review. Cornell Vet. 1986, 76, 61–90. [Google Scholar] [PubMed]
  4. Dimethylsulfoxide. In Meyler’s Side Effects of Drugs, 16th ed.; Elsevier: Amsterdam, The Netherlands, 2016; pp. 992–993. [CrossRef]
  5. Gurria, G.M.; Posner, G.H. Photochemical Deoxygenation of Aryl Sulfoxides. J. Org. Chem. 1973, 38, 2419–2420. [Google Scholar] [CrossRef]
  6. Gregory, D.D.; Wan, Z.; Jenks, W.S. Photodeoxygenation of Dibenzothiophene Sulfoxide: Evidence for a Unimolecular S-O Cleavage Mechanism. J. Am. Chem. Soc. 1997, 119, 94–102. [Google Scholar] [CrossRef]
  7. Omlid, S.M.; Dergunov, S.A.; Isor, A.; Sulkowski, K.L.; Petroff, J.T.; Pinkhassik, E.; McCulla, R.D. Evidence for Diffusing Atomic Oxygen Uncovered by Separating Reactants with a Semi-Permeable Nanocapsule Barrier. Chem. Commun. 2019, 55, 1706–1709. [Google Scholar] [CrossRef]
  8. Christensen, P.R.; Patrick, B.O.; Caron, É.; Wolf, M.O. Oxidation-State-Dependent Photochemistry of Sulfur-Bridged Anthracenes. Angew. Chem. Int. Ed. 2013, 52, 12946–12950. [Google Scholar] [CrossRef] [PubMed]
  9. Calligaris, M.; Carugo, O. Structure and Bonding in Metal Sulfoxide Complexes. Coord. Chem. Rev. 1996, 153, 83–154. [Google Scholar] [CrossRef]
  10. Calligaris, M. Structure and Bonding in Metal Sulfoxide Complexes: An Update. Coord. Chem. Rev. 2004, 248, 351–375. [Google Scholar] [CrossRef]
  11. Mellah, M.; Voituriez, A.; Schulz, E. Chiral Sulfur Ligands for Asymmetric Catalysis. Chem. Rev. 2007, 107, 5133–5209. [Google Scholar] [CrossRef]
  12. Sipos, G.; Drinkel, E.E.; Dorta, R. The Emergence of Sulfoxides as Efficient Ligands in Transition Metal Catalysis. Chem. Soc. Rev. 2015, 44, 3834–3860. [Google Scholar] [CrossRef]
  13. Jastrzebski, J.T.B.H.; van Koten, G. Structures and Reactivities of Organocopper Compounds. In Modern Organocopper Chemistry; Wiley: Hoboken, NJ, USA, 2002; Volume 34, pp. 1–44. [Google Scholar] [CrossRef]
  14. Liu, L.; Chen, H.; Yang, Z.; Wei, J.; Xi, Z. C,C- and C,N-Chelated Organocopper Compounds. Molecules 2021, 26, 5806. [Google Scholar] [CrossRef] [PubMed]
  15. Guss, J.M.; Mason, R.; Søtofte, I.; Van Koten, G.; Noltes, J.G. A Tetranuclear Cluster Complex of Copper(I) with Bridging Aryl Ligands: The Crystal Structure of (4-Methyl-2-Cupriobenzyl)Dimethylamine. J. Chem. Soc. Chem. Commun. 1972, 8, 446–447. [Google Scholar] [CrossRef]
  16. Liu, L.; Geng, W.; Yang, Q.; Zhang, W.X.; Xi, Z. Well-Defined Butadienyl Organocopper(I) Aggregates from Zirconacyclopentadienes and CuCl: Synthesis and Structural Characterization. Organometallics 2015, 34, 4198–4201. [Google Scholar] [CrossRef]
  17. Doshi, A.; Sundararaman, A.; Venkatasubbaiah, K.; Zakharov, L.N.; Rheingold, A.L.; Myahkostupov, M.; Piotrowiak, P.; Jäkle, F. Pentafluorophenyl Copper-Pyridine Complexes: Synthesis, Supramolecular Structures via Cuprophilic and π-Stacking Interactions, and Solid-State Luminescence. Organometallics 2012, 31, 1546–1558. [Google Scholar] [CrossRef]
  18. Gambarotta, S.; Strologo, S.; Floriani, C.; Chlesl-Vllla, A.; Guastini, C. Synthesis and Structure of a Mononuclear Copper(I) Complex Containing the Copper(I) σ-Phenyl Functionality. Organometallics 1984, 3, 1444–1445. [Google Scholar] [CrossRef]
  19. Cairncross, A.; Omura, H.; Sheppard, W.A. Organocopper Cluster Compounds. II. Pentafluorophenylcopper and o-(Trifluoromethyl)Phenylcopper Tetramers. J. Am. Chem. Soc. 1971, 93, 248–249. [Google Scholar] [CrossRef]
  20. Jäkle, F. Pentafluorophenyl Copper: Aggregation and Complexation Phenomena, Photoluminescence Properties, and Applications as Reagent in Organometallic Synthesis. Dalton Trans. 2007, 27, 2851–2858. [Google Scholar] [CrossRef] [PubMed]
  21. Sundararaman, A.; Lalancette, R.A.; Zakharov, L.N.; Rheingold, A.L.; Jäkle, F. Structural Diversity of Pentafluorophenylcopper Complexes. First Evidence of π-Coordination of Unsupported Arenes to Organocopper Aggregates. Organometallics 2003, 22, 3526–3532. [Google Scholar] [CrossRef]
  22. Akihiko, I.; Norio, N.; Yasunaka, T. 5,9b-dihydro-5,9b-banzonaphtho[1,2-b] Chalcogenophene Derivative, and Production Method of the Same. JP2012041295A, 1 March 2013. [Google Scholar]
  23. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A 2015, 71, 3–8. [Google Scholar] [CrossRef]
  24. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  25. Hofer, M.; Nevado, C. Cross-coupling of arene–gold(III) complexes. Tetrahedron 2013, 69, 5751–5757. [Google Scholar] [CrossRef]
  26. Meyer, B.K.; Polity, A.; Reppin, D.; Becker, M.; Hering, P.; Kramm, B.; Klar, P.J.; Sander, T.; Reindl, C.; Heiliger, C.; et al. The Physics of Copper Oxide (Cu2O). Semicond. Semimet. 2013, 88, 201–226. [Google Scholar] [CrossRef]
  27. Zoolfakar, A.S.; Rani, R.A.; Morfa, A.J.; O’Mullane, A.P.; Kalantar-Zadeh, K. Nanostructured Copper Oxide Semiconductors: A Perspective on Materials, Synthesis Methods and Applications. J. Mater. Chem. C 2014, 2, 5247–5270. [Google Scholar] [CrossRef]
Scheme 1. Known photoreactive biarylsulfoxide derivatives. (a) Deoxygenation of DBTO. (b) SO extrusion of Anthra2SO.
Scheme 1. Known photoreactive biarylsulfoxide derivatives. (a) Deoxygenation of DBTO. (b) SO extrusion of Anthra2SO.
Molecules 29 03332 sch001
Figure 1. (a) ORTEP view of X-ray structure of [Cu(C6F5)]4(DBTO)2 6 at 50% probability. (b) View of the Cu4 ring coordination structure.
Figure 1. (a) ORTEP view of X-ray structure of [Cu(C6F5)]4(DBTO)2 6 at 50% probability. (b) View of the Cu4 ring coordination structure.
Molecules 29 03332 g001
Figure 2. (a) 19F NMR spectra for the titration of the complexation of 5 with DBTO. (b) 1H NMR spectra for the titration of the complexation of DBTO with 5. (c) NMR data plot for the titration of the formation of the complex 6.
Figure 2. (a) 19F NMR spectra for the titration of the complexation of 5 with DBTO. (b) 1H NMR spectra for the titration of the complexation of DBTO with 5. (c) NMR data plot for the titration of the formation of the complex 6.
Molecules 29 03332 g002
Figure 3. ORTEP view of X-ray structure of (a) [Cu(C6F5)]4(Tol2SO)2 7 and (b) [Cu(C6F5)]4(Anthra2SO)2 8 at 50% probability, and their corresponding view of the Cu4 ring coordination structure. Three cocrystallized molecules of benzene were omitted for clarity in complex 8.
Figure 3. ORTEP view of X-ray structure of (a) [Cu(C6F5)]4(Tol2SO)2 7 and (b) [Cu(C6F5)]4(Anthra2SO)2 8 at 50% probability, and their corresponding view of the Cu4 ring coordination structure. Three cocrystallized molecules of benzene were omitted for clarity in complex 8.
Molecules 29 03332 g003
Scheme 2. (ac) Photoreactivity of biarylsulfoxide–copper(I) complexes 68.
Scheme 2. (ac) Photoreactivity of biarylsulfoxide–copper(I) complexes 68.
Molecules 29 03332 sch002
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

Magné, V.; Lenk, R.; Mallet-Ladeira, S.; Maerten, E.; Madec, D. Pentafluorophenyl Copper–Biarylsulfoxide Complexes: Synthesis and Photoreactivity. Molecules 2024, 29, 3332. https://doi.org/10.3390/molecules29143332

AMA Style

Magné V, Lenk R, Mallet-Ladeira S, Maerten E, Madec D. Pentafluorophenyl Copper–Biarylsulfoxide Complexes: Synthesis and Photoreactivity. Molecules. 2024; 29(14):3332. https://doi.org/10.3390/molecules29143332

Chicago/Turabian Style

Magné, Valentin, Romaric Lenk, Sonia Mallet-Ladeira, Eddy Maerten, and David Madec. 2024. "Pentafluorophenyl Copper–Biarylsulfoxide Complexes: Synthesis and Photoreactivity" Molecules 29, no. 14: 3332. https://doi.org/10.3390/molecules29143332

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop