Next Article in Journal
Field-Induced Slow Magnetic Relaxation of Mono- and Dinuclear Dysprosium(III) Complexes Coordinated by a Chloranilate with Different Resonance Forms
Next Article in Special Issue
Hybrid Disila-Crown Ethers as Hosts for Ammonium Cations: The O–Si–Si–O Linkage as an Acceptor for Hydrogen Bonding
Previous Article in Journal
Pentacoordinated Chloro-Iron(III) Complexes with Unsymmetrically Substituted N2O2 Quadridentate Schiff-Base Ligands: Syntheses, Structures, Magnetic and Redox Properties
Previous Article in Special Issue
Modification of TiO2 Surface by Disilanylene Polymers and Application to Dye-Sensitized Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis of a α-Chlorosilyl Functionalized Donor-Stabilized Chlorogermylene

1
Tata Institute for Fundamental Research Hyderabad Gopanpally, Hyderabad-500107, Telangana, India
2
Krupp-Chair of General and Inorganic Chemistry, Saarland University, 66123 Saarbrücken, Germany
*
Authors to whom correspondence should be addressed.
Inorganics 2018, 6(1), 6; https://doi.org/10.3390/inorganics6010006
Submission received: 8 December 2017 / Revised: 23 December 2017 / Accepted: 25 December 2017 / Published: 29 December 2017
(This article belongs to the Special Issue Coordination Chemistry of Silicon)

Abstract

:
Peripherally functionalized low-valent main group species allow for the introduction/interconversion of functional groups without increasing the formal oxidation state of the main group center. Herein, we report a straightforward method for the incorporation of a α-chlorosilyl moiety adjacent to the NHC-coordinated germanium(II) center.

Graphical Abstract

1. Introduction

In recent years, the chemistry of the heavier analogues of carbenes (tetrylenes) has been expanded beyond mere synthetic curiosity [1,2,3] towards application in synthesis. Heavier carbene analogues are applied as a donor ligands in low-valent main group species [4,5] as well as in transition metal complexes [6,7]. An increasing number of examples show competitive catalytic activity in different organic transformations [8]. As the complexity of the tetrylenes increases with more intricate ligand architectures [9], functionalization protocols in the presence of uncompromised low-valent Group 14 centers conveniently allow for a comparatively straightforward diversification in the final stages of ligand synthesis. While the interconversion of functional groups is just beginning to emerge in the case of heavier multiple bonds [10], numerous examples have been reported for the heavier tetrylenes [11,12,13,14,15]. Recently, Scheschkewitz et al. have taken a similar approach with the synthesis of the multiply functional NHC-coordinated silagermenylidenes, I and II (Scheme 1) [16,17], which serve as precursors for cyclic NHC-coordinated germylenes of type III, IV, and V under consumption of the Si=Ge bond, but retention of the low-valent germanium center (Scheme 1) [16,18,19]. The leaving group characteristics of the peripheral chloro functionality of II can be exploited for the incorporation of different organic substituents by treatment with organolithium reagents in order to fine-tune the steric requirements of the ligand scaffold of cyclic germylenes of type IV [20,21]. Directly chloro-functionalized silylenes and germylenes are readily converted to a variety of novel low-valent group 14 compounds by functional group interconversion at the low-valent tetrel center [22].
An additional leaving group adjacent to a chloro-functionalized heavier carbene center would in principle provide a precursor for the synthesis of further examples of heavier vinylidene such as I and II. The synthesis of an NHC-coordinated (chlorogermyl)chlorogermylene, IV, from NHC-coordinated diaryl germylene and NHC- or 1,4-dioxane coordinated dichlorogermylene has been reported by the groups of Baines and Tobitah [23,24]. Herein, we now report the synthesis of NHC-stabilized (chlorosilyl)chlorogermylene, 1.

2. Results and Discussions

We anticipated that West’s N-heterocyclic silylene, 2 [25] would insert into the Ge–Cl bond of the NHC-germanium(II)dichloride adduct 3 [26] as it is well known for the oxidative addition of different types of bonds e.g., C–Cl [27] and Ge–N [28]. Indeed, the reaction of 2 and 3 in a 1:1 ratio in toluene at room temperature afforded the NHC-stabilized (chlorosilyl)chlorogermylene 1 which was isolated as a crystalline compound (Scheme 2). We did not obtain any indication for the formation of donor–acceptor adducts between 2 and 3 or rearrangement products as often described for reactions of silylenes and germylenes [29,30].
Compound 1 was characterized in solution state by NMR spectroscopy as well as in solid state by single crystal X-ray molecular structure determination. The insertion of the silylene into the Ge–Cl bond turns the germanium atom into a center of chirality. As a result the two diastereotopic C–H protons of C2N2Si-moiety give rise to two doublets at δ = 5.87 and 6.03 ppm (1J(H, H) = 3.92 Hz) in the 1H NMR. Similarly, the two tBu groups show 1H NMR resonances at δ = 1.29 and 1.65 ppm. Hindered rotation can be excluded as the explanation for the doubling of these resonances. Despite the increased congestion about the germanium center, the NHC retains the local rotational C2-symmetry in solution: the 1H NMR shows only a single septet for the two CH moieties of the isopropyl groups. In contrast, there are again two signals for the adjacent diastereotopic methyl groups of the N-isopropyl moiety. In 13C{1H} NMR, the carbenic carbon shows a resonance at δ = 171.01 ppm, which is similar to the chemical shifts observed for other NHC-coordinated Ge(II) compounds [18,19]. In 29Si{1H} NMR, the singlet at δ = −3.39 ppm is strongly highfield shifted compared with that of the free N-heterocyclic silylene (δ = +78.3 ppm) [25]. Notably, even repeated crystallization of 1 did not yield NMR spectra uncontaminated by residual 2, which led us to speculate about the reversibility of the oxidative Ge–Cl addition to the silylene. NMR at variable temperatures, however, did not show any temperature dependence of the sample composition.
Nonetheless, single crystals of 1 suitable for a X-ray diffraction study were obtained from saturated toluene solution at −20 °C after one day. Compound 1 crystallizes in the monoclinic P21/c space group. However in the obtained single crystal X-ray diffraction data we did not see any residual electron density for the cocrystalization of 2 along with 1. Analysis of molecular structure determination revels the presence of α-chlorosilyl moiety adjacent to the NHC-coordinated germanium(II) center; which was anticipated from the solution state structure (Figure 1). The Si–Ge bond length is 2.4969(7) Å which is close to reported Si(IV)–Ge(II) bond length [19]. The distance between carbenic carbon and germanium(II) center is 2.081(2) Å, which is slightly shorter than that of the corresponding NHC-coordinated germanium(II)dichloride (2.106(3) Å) [26].

3. Materials and Methods

3.1. General Information

All manipulation were carried out under an argon atmosphere using either a Schlenk line technique or inside a GloveBox. All solvents were dried by Innovative Technology solvent purification system. Compounds 2 [25] and 3 [26] were prepared according to literature procedures. Benzene-d6 was dried and distilled over potassium under argon. NMR spectra were recorded on a Bruker Avance III 300 MHz NanoBay NMR spectrometer (Bruker, Switzerland). 1H and 13C{1H} NMR spectra were referenced to the peaks of residual protons of the deuterated solvent (1H) or the deuterated solvent itself (13C{1H}). 29Si{1H} NMR spectra were referenced to external SiMe4.

3.2. Experimental Details

Synthesis of compound 1: 25-mL dry and degassed toluene were added to a Schlenk flask containing 2 (0.242 g, 1.23 mmol) and 3 (0.4 g, 1.23 mmol) at −78 °C. The mixture is brought to room temperature within one hour and stirred continuously for another two hours. Removal of the solvent in vacuum and washing of the solid residue with n-hexane was followed by extraction with 20 mL warm toluene. The resulting yellow solution was concentrated to about 15 mL and kept at −20 °C for one day to get bright yellow crystals of the desired compound, 1, suitable for single crystal X-ray diffraction study. Despite apparently uniform crystals, a pure sample of compound 1 without free silylene 2 could not be obtained. Yield: 0.360 g (56% which include 13% of compound 2). 1H NMR (300 MHz, C6D6, 298 K): δ = 1.13 (d, 1J(H, H) = 7.02 Hz, 6H, CH(CH3)2), 1.24 (d, 1J(H, H) = 6.9 Hz, 6H, CH(CH3)2), 1.29 (s, 9H, N(CH3)3), 1.51 (s, 6H, CCH3), 1.63 (s, 9H, N(CH3)3), 5.74 (sept, 1J(H, H) = 7.02 Hz, 2H, CH(CH3)2), 5.87 (d, 1J(H, H) = 3.92 Hz, 1H, CHCH) 6.03 (d, 1H, 1J(H, H) = 3.92 Hz, CHCH) (1.42 and 6.76 ppm refer to the resonances for compound 2) ppm. 13C{1H} NMR (75.4 MHz, C6D6, 298 K): δ = 10.35 (2C, CCH3), 22.06(4C, CH(CH3)2), 31.97 (3C, C(CH3)3), 32.15 (3C, C(CH3)3), 51.97 (1C, C(CH3)3), 52.49 (1C, C(CH3)3), 54.07 (2C, CH(CH3)2), 113.70 (1C, CHCH), 115.26 (1C, CHCH), 127.13 (2C, CCH3), 171.01 (1C, NCN). 29Si{1H} NMR (59.6 MHz, C6D6, 298 K): δ = −3.39 ppm.

3.3. X-ray Crystallographic Analysis

Single crystals of 1 were obtained from saturated toluene solution at −20 °C. Intensity data were collected on a Bruker SMART APEX CCD diffractometer (Bruker, Germany with a Mo Kα radiation (λ = 0.71073 Å) at T = 182(2) K. The structures were solved by a direct method (SHELXS [31]) and refined by a full-matrix least square method on F2 for all reflections (SHELXL-2014 [32]). All hydrogen atoms were placed using AFIX instructions, while all other atoms were refined anisotropically. Crystallographic data (Supplementary Materials) were deposited at the Cambridge Crystallographic Data Center (CCDC; under reference number: CCDC-1587144) and can be obtained free of charge via https: //www.ccdc.cam.ac.uk/structures/. X-ray crystallographic data for 1: M = 520.15, monoclinic, P21/c, a = 11.3853(3) Å, b = 13.1071(3) Å, c = 17.5390(5) Å, β = 96.2860(10)°, V = 2601.58(12) Å3, Z = 4, Dcalc. = 1.328 gcm−3, m = 1.444 mm−1, 2θmax = 54.20°, measd./unique refls. = 48784/5747 (Rint. = 0.0292), GOF = 1.075, R1 = 0.0408/0.0471 [I>2σ(I)/all data], wR2 = 0.1077/0.1112 [I>2σ(I)/all data], largest diff. peak and hole 3.152 and −0.604 e.Å−3.

4. Conclusions

We have demonstrated a proof of principle study for the straightforward incorporation of a α-chlorosilyl moiety adjacent to the donor-stabilized germanium(II) center. The resulting product features a 1,2-dicholoro functionality and should therefore in principle be suitable as precursor for the targeted synthesis of NHC-coordinated silagermenylidenes, the heavier analogues of vinylidenes. So far, attempts to eliminate the two chloro substituents reductively were not met with success.

Supplementary Materials

The following are available online at www.mdpi.com/2304-6740/6/1/6/s1, Cif, cif-checked, NMR spectra files.
Supplementary File 1

Acknowledgments

This work is supported by the Research Group Linkage Program of the Alexander von Humboldt Foundation, Germany; the CSIR (01(2863)/16/EMR-II), India; SERB-DST (EMR/2014/001237), India, the TIFR Center for Interdisciplinary Science Hyderabad, India; and Saarland University.

Author Contributions

David Scheschkewitz and Anukul Jana conceived and designed the experiments and wrote the paper; Debabrata Dhara performed the experiments; Volker Huch performed the XRD analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, S.B.; Sinsermsuksakul, P.; Hock, A.S.; Pike, R.D.; Gordon, R.G. Synthesis of N-heterocyclic stannylene (Sn(II)) and germylene (Ge(II)) and a Sn(II) amidinate and their application as precursors for atomic layer deposition. Chem. Mater. 2014, 26, 3065–3073. [Google Scholar] [CrossRef]
  2. Purkait, T.K.; Swarnakar, A.K.; De Los Reyes, G.B.; Hegmann, F.A.; Rivard, E.; Veinot, J.G.C. One-pot synthesis of functionalized germanium nanocrystals from a single source precursor. Nanoscale 2015, 7, 2241–2244. [Google Scholar] [CrossRef] [PubMed]
  3. Blom, B.; Said, A.; Szilvási, T.; Menezes, P.W.; Tan, G.; Baumgartner, J.; Driess, M. Alkaline-earth-metal-induced liberation of rare allotropes of elemental silicon and germanium from N-heterocyclic metallylenes. Inorg. Chem. 2015, 54, 8840–8848. [Google Scholar] [CrossRef] [PubMed]
  4. Shan, Y.-L.; Yim, W.-L.; So, C.-W. An N-heterocyclic silylene-stabilized digermanium(0) complex. Angew. Chem. Int. Ed. 2014, 53, 13155–13158. [Google Scholar] [CrossRef] [PubMed]
  5. Shan, Y.-L.; Leong, B.-X.; Xi, H.-W.; Ganguly, R.; Li, Y.; Limb, K.H.; So, C.-W. Reactivity of an amidinato silylene and germylene toward germanium(II), tin(II) and lead(II) halides. Dalton Trans. 2017, 46, 3642–3648. [Google Scholar] [CrossRef] [PubMed]
  6. Blom, B.; Stoelzel, M.; Driess, M. New vistas in N-heterocyclic silylene (NHSi) transition-metal coordination chemistry: Syntheses, structures and reactivity towards activation of small molecules. Chem. Eur. J. 2013, 19, 40–62. [Google Scholar] [CrossRef] [PubMed]
  7. Parvin, N.; Dasgupta, R.; Pal, S.; Sen, S.S.; Khan, S. Strikingly diverse reactivity of structurally identical silylene and stannylene. Dalton Trans. 2017, 46, 6528–6532. [Google Scholar] [CrossRef] [PubMed]
  8. Brck, A.; Gallego, D.; Wang, W.; Irran, E.; Driess, M.; Hartwig, J.F. Pushing the s-donor strength in iridium pincer complexes: Bis-(silylene) and bis(germylene) ligands are stronger donors than bis(phosphorus(III)) ligands. Angew. Chem. Int. Ed. 2012, 51, 11478–11482. [Google Scholar] [CrossRef] [PubMed]
  9. Driess, M.; Yao, S.; Brym, M.; van Wüllen, C.; Lentz, D. A new type of N-heterocyclic silylene with ambivalent reactivity. J. Am. Chem. Soc. 2006, 128, 9628–9629. [Google Scholar] [CrossRef] [PubMed]
  10. Präsang, C.; Scheschkewitz, D. Reactivity in the periphery of functionalised multiple bonds of heavier group 14 elements. Chem. Soc. Rev. 2016, 45, 900–921. [Google Scholar] [CrossRef] [PubMed]
  11. Azhakar, R.; Ghadwal, R.S.; Roesky, H.W.; Wolf, H.; Stalke, D. A début for base stabilized monoalkylsilylenes. Chem. Commun. 2012, 48, 4561–4563. [Google Scholar] [CrossRef] [PubMed]
  12. Gallego, D.; Brück, A.; Irran, E.; Meier, F.; Kaupp, M.; Driess, M.; Hartwig, J.F. From bis(silylene) and bis(germylene) pincer-type nickel(II) complexes to isolable intermediates of the nickel-catalyzed sonogashira cross-coupling reaction. J. Am. Chem. Soc. 2013, 135, 15617–15626. [Google Scholar] [CrossRef] [PubMed]
  13. Baus, J.A.; Poater, J.; Bickelhaupt, F.M.; Tacke, R. Silylene-induced reduction of [Mn2(CO)10]: Formation of a five-coordinate silicon(IV) complex with an O-bound [(OC)4Mn=Mn(CO)4]2− ligand. Eur. J. Inorg. Chem. 2017, 186–191. [Google Scholar] [CrossRef]
  14. Cabeza, J.A.; García-Álvarez, P.; Gobetto, R.; González-Álvarez, L.; Nervi, C.; Pérez-Carreño, E.; Polo, D. [MnBrL(CO)4] (L = Amidinatogermylene): Reductive dimerization, carbonyl substitution, and hydrolysis reactions. Organometallics 2016, 35, 1761–1770. [Google Scholar] [CrossRef]
  15. Rivard, E. Group 14 inorganic hydrocarbon analogues. Chem. Soc. Rev. 2016, 45, 989–1003. [Google Scholar] [CrossRef] [PubMed]
  16. Jana, A.; Huch, V.; Scheschkewitz, D. NHC-stabilized silagermenylidene: A heavier analogue of vinylidene. Angew. Chem. Int. Ed. 2013, 52, 12179–12182. [Google Scholar] [CrossRef] [PubMed]
  17. Jana, A.; Majumdar, M.; Huch, V.; Zimmer, M.; Scheschkewitz, D. NHC-coordinated silagermenylidene functionalized in allylic position and its behaviour as a ligand. Dalton Trans. 2014, 43, 5175–5181. [Google Scholar] [CrossRef] [PubMed]
  18. Jana, A.; Omlor, I.; Huch, V.; Rzepa, H.S.; Scheschkewitz, D. N-heterocyclic carbene coordinated neutral and cationic heavier cyclopropylidenes. Angew. Chem. Int. Ed. 2014, 53, 9953–9956. [Google Scholar] [CrossRef] [PubMed]
  19. Jana, A.; Huch, V.; Rzepa, H.S.; Scheschkewitz, D. A multiply functionalized base-coordinated GeII compound and its reversible dimerization to the digermene. Angew. Chem. Int. Ed. 2015, 54, 289–292. [Google Scholar] [CrossRef] [PubMed]
  20. Nieder, D.; Yildiz, C.B.; Jana, A.; Zimmer, M.; Huch, V.; Scheschkewitz, D. Dimerization of a marginally stable disilenyl germylene to tricyclic systems: Evidence for reversible NHC-coordination. Chem. Commun. 2016, 52, 2799–2802. [Google Scholar] [CrossRef] [PubMed]
  21. Nieder, D.; Huch, V.; Yildiz, C.B.; Scheschkewitz, D. Regiodiscriminating reactivity of isolable NHC-coordinated disilenyl germylene and its cyclic isomer. J. Am. Chem. Soc. 2016, 138, 13996–14005. [Google Scholar] [CrossRef] [PubMed]
  22. Nagendran, S.; Roesky, H.W. The chemistry of aluminum(I), silicon(II), and germanium(II). Organometallics 2008, 27, 457–492. [Google Scholar] [CrossRef]
  23. Rupar, P.A.; Jennings, M.C.; Baines, K.M. Synthesis and structure of N-heterocyclic carbene complexes of germanium(II). Organometallics 2008, 27, 5043–5051. [Google Scholar] [CrossRef]
  24. Tashita, S.; Watanabe, T.; Tobita, H. Synthesis of a base-stabilized (Chlorogermyl)metallogermylene and Its photochemical conversion to a (Chlorogermyl)germylyne complex. Chem. Lett. 2013, 42, 43–44. [Google Scholar] [CrossRef]
  25. Denk, M.; Lennon, J.R.; Hayashi, R.; West, R.; Belyakov, A.V.; Verne, H.P.; Haaland, A.; Wagner, M.; Metzler, N. Synthesis and structure of a stable silylene. J. Am. Chem. Soc. 1994, 116, 2691–2692. [Google Scholar] [CrossRef]
  26. Rupar, P.A.; Staroverov, V.N.; Ragogna, P.J.; Baines, K.M. A germanium(II)-centered dication. J. Am. Chem. Soc. 2007, 129, 15138–15139. [Google Scholar] [CrossRef] [PubMed]
  27. Moser, D.F.; Naka, A.; Guzei, I.A.; Müller, T.; West, R. Formation of disilanes in the reaction of stable silylenes with halocarbons. J. Am. Chem. Soc. 2005, 127, 14730–14738. [Google Scholar] [CrossRef] [PubMed]
  28. Gehrhus, B.; Hitchcock, P.B.; Lappert, M.F. New reactions of a silylene: Insertion into M–N bonds of M[N(SiMe3)2]2 (M = Ge, Sn, or Pb). Angew. Chem. Int. Ed. Engl. 1997, 36, 2514–2516. [Google Scholar] [CrossRef]
  29. Al-Rafia, S.M.I.; Momeni, M.R.; McDonald, R.; Ferguson, M.J.; Brown, A.; Rivard, E. Controlled growth of dichlorogermanium oligomers from Lewis basic hosts. Angew. Chem. Int. Ed. 2013, 52, 6390–6395. [Google Scholar] [CrossRef] [PubMed]
  30. Schafer, A.; Saak, W.; Weidenbruch, M.; Marsmann, H.; Henkel, G. Reactions of a silylene with a germylene and a stannylene: Formation of a digermene with an unusual arrangement of the substituents and of a stannane. Chem. Ber. 1997, 130, 1733–1737. [Google Scholar] [CrossRef]
  31. Sheldrick, G.M. A short history of SHELX. Acta Cryst. 2008, A64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  32. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef]
Scheme 1. Chemical structures of α-chlorosilyl-functionalized silagermenyledene I, and α-chloro-functionalized germylenes IIIV (R = Tip = 2,4,6-iPr3C6H2, NHCiPr2Me2 = 1,3-diisopropyl-4,5-dimethylimidazol-2-ylidene, NHCMe4 = 1,3,4,5-tetramethylimidazol-2-ylidene, Xyl = 2,6-Me2C6H3, and Mes = 2,6-Me2C6H3).
Scheme 1. Chemical structures of α-chlorosilyl-functionalized silagermenyledene I, and α-chloro-functionalized germylenes IIIV (R = Tip = 2,4,6-iPr3C6H2, NHCiPr2Me2 = 1,3-diisopropyl-4,5-dimethylimidazol-2-ylidene, NHCMe4 = 1,3,4,5-tetramethylimidazol-2-ylidene, Xyl = 2,6-Me2C6H3, and Mes = 2,6-Me2C6H3).
Inorganics 06 00006 sch001
Scheme 2. Synthesis of 1.
Scheme 2. Synthesis of 1.
Inorganics 06 00006 sch002
Figure 1. Molecular structure of 1 at 30% probability level, all H-atoms were deleted for clarity. Selected bond lengths (Å) and bond angles (deg.): Ge1–C1 2.081(2), Ge1–Cl1 2.2891(8), Ge1–Si1 2.4969(7); C1–Ge1–Cl1 92.58(7), C1–Ge1–Si1 102.56(7), Cl–Ge1–Si1 101.83(3).
Figure 1. Molecular structure of 1 at 30% probability level, all H-atoms were deleted for clarity. Selected bond lengths (Å) and bond angles (deg.): Ge1–C1 2.081(2), Ge1–Cl1 2.2891(8), Ge1–Si1 2.4969(7); C1–Ge1–Cl1 92.58(7), C1–Ge1–Si1 102.56(7), Cl–Ge1–Si1 101.83(3).
Inorganics 06 00006 g001

Share and Cite

MDPI and ACS Style

Dhara, D.; Huch, V.; Scheschkewitz, D.; Jana, A. Synthesis of a α-Chlorosilyl Functionalized Donor-Stabilized Chlorogermylene. Inorganics 2018, 6, 6. https://doi.org/10.3390/inorganics6010006

AMA Style

Dhara D, Huch V, Scheschkewitz D, Jana A. Synthesis of a α-Chlorosilyl Functionalized Donor-Stabilized Chlorogermylene. Inorganics. 2018; 6(1):6. https://doi.org/10.3390/inorganics6010006

Chicago/Turabian Style

Dhara, Debabrata, Volker Huch, David Scheschkewitz, and Anukul Jana. 2018. "Synthesis of a α-Chlorosilyl Functionalized Donor-Stabilized Chlorogermylene" Inorganics 6, no. 1: 6. https://doi.org/10.3390/inorganics6010006

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