Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Multiple ethylene activation by heteroleptic L(Cl)Ga-substituted germylenes

Anna Bücker a, Christoph Wölper a, Hannah Siera b, Gebhard Haberhauer b and Stephan Schulz *ac
aInstitute of Inorganic Chemistry, University of Duisburg-Essen, 45117 Essen, Germany. E-mail: stephan.schulz@uni-due.de
bInstitute of Organic Chemistry, University of Duisburg-Essen, 45117 Essen, Germany
cCenter for Nanointegration Duisburg-Essen (CENIDE), University of Duisburg-Essen, 47057 Duisburg, Germany

Received 24th November 2023 , Accepted 4th December 2023

First published on 5th December 2023


Abstract

Ethylene insertion into the Ga–Ge bond of the L(Cl)Ga-substituted germylene LGa(μ-Cl)GeDMP 1 (L = HC(C(Me)NAr)2, Ar = 2,6-iPr2C6H3; DMP = 2,6-Mes2C6H3, Mes = 2,4,6-Me3C6H2) at ambient temperature is followed by dimerization of the as-formed germylene to give the digermene 3, which further reacted with ethylene in a [2 + 2] cycloaddition to give the 1,2-digermacyclobutane 4. In marked contrast, the amino-substituted germylene L(Cl)GaGeN(SiMe3)Ar 2 reacted directly to the 1,2-digermacyclobutane 5. Quantum chemical calculations confirmed the assumed reaction mechanism, hence demonstrating the crucial role of the substituent on the reaction mechanism.


Introduction

Ethylene belongs to the most important industrial feedstock chemicals, with a current production of more than 150 million tons.1 It is catalytically converted into fine chemicals including polyethylene, ethylene oxide, alcohols, and aldehydes,2 typically by using transition metal-based catalysts. More recently, the potential of low valent s- and p-block compounds for the activation of industrially relevant small molecules, i.e., H2, CO, NH3 and ethylene, has been demonstrated, and even catalytic reactions have been reported.3,4 Such compounds contain a stereochemically active lone pair (donor) and an empty π-type orbital (acceptor) and thereby resemble the electronic configuration of transition-metal compounds.

Low valent group 14 compounds were found to react with ethylene in different ways.5 Monovalent tetrylynes typically reacted in [2 + 2]6 and [2 + 2 + 2] cycloaddition reactions7 as was exemplarily shown in reactions of digermyne B[thin space (1/6-em)]6b to compounds II and III, respectively, and of distannyne C[thin space (1/6-em)]7c to compound V. Interestingly, compound II reacted reversibly under kinetic control to bisgermacyclopropane IV.6b In contrast, the analogous bissilacyclopropane I was formed exclusively with disilyne A,8a whereas hydridotetrylenes were reported to react with alkenes with hydrometallation,8b,8c respectively. In addition, the digermyne TbbGeGeTbb (Tbb = 2,6-(CH(SiMe3)2-4-tBu-C6H2)) served as a precatalyst in the catalytic cyclotrimerisation of terminal alkynes (Reppe reaction).9 In contrast, divalent tetrylenes, which have an electron lone pair in a sp2-hybridised orbital in the plane of the molecule and a vacant p-type orbital perpendicular to the molecular plane as a result of their typically large singlet–triplet energy gap, reacted mainly with ethylene in (sometimes reversible) [2 + 1] cycloaddition reactions.10 For example, the acyclic silylene D reacted to silirane VI (Scheme 1), which underwent 1,2-insertion of ethylene at 60 °C to finally give silirane VII as was confirmed by in situ1H NMR spectroscopy with C2D4.10c In contrast, Lappert's germylene E directly reacted to the 1,2-digermacyclobutane IX, most likely via the intermediate formation of germirane VIII,10g while the sterically demanding diarylgermylene (Ar′)2Ge (Ar′ = DMP, 2,6-Ar2C6H3) reacted reversibly with ethylene to the corresponding germirane.10h In contrast, the reaction of diarylstannylenes (Ar*)2Sn F (Ar* = 2,6-Ar2C6H3, 2,6-Trip2C6H3, Trip = 2,4,6-iPr3-C6H2) with C2H4 at 60 °C occurred with insertion into a Sn–C bond to yield stannylene X.11


image file: d3dt03944g-s1.tif
Scheme 1 Ethylene activation reactions by tetrylynes and tetrylenes; R = CH(SiMe3)2, Ar* = 2,6-Dipp2C6H3, 2,6-Trip2C6H3 (Trip = 2,4,6-iPr3-C6H2), Bbt = 2,6-[CH(SiMe3)2]2-4-C(SiMe3)3-C6H2, Ar′ = 2,6-Dipp2-C6H3, 2,6-Trip2-3,5-iPr2-C6H.6b,7c,8a,10c,g,11

We are particularly interested in electron-rich tetrylenes and recently reported the synthesis of room temperature-stable silylene carbonyl complexes (L(X)Ga)2SiCO (L = HC[C(Me)N(Ar)]2; X = Br, I, Me),12 which showed enhanced π-backbonding to the CO group.13 In addition, heteroleptic metallasilylenes L(Cl)MSi[PhC(N-tBu)2] (M = Al, Ga) were found to activate CO2, N2O and azides, respectively, and also reacted with white phosphorus (P4) in unprecedented [2 + 1 + 1] fragmentation reactions and with cyclo-E5 rings of [Cp*Fe(η5-E5)] (Cp* = η5-C5Me5; E = P, As).14 To explore the reactivity of heteroleptic tetrylenes and to investigate the influence of the group 14 element on their reactivity in more detail, we recently synthesised the L(Cl)Ga-substituted germylenes LGa(μ-Cl)GeDMP 1, which represents an intermediate of the oxidative addition reaction of DMPGeCl and LGa, in which the bridging Cl atom stabilises the empty p orbital at the Ge centre, and L(Cl)GaGeN(SiMe3)Ar 2.15 LGa(μ-Cl)GeDMP 1 was found to react with CO2 with decarbonylation and with isocyanates and carbodimides with insertion into the Ga–Ge bond, demonstrating its potential for small molecule activation reactions.16 We here report on (multiple) ethene activation reactions of germylenes 1 and 2 including detailed mechanistic studies (DFT calculations).

Results and discussion

Germylene LGa(μ-Cl)GeDMP 1 reacted at ambient temperature under one atmosphere of ethylene within minutes to give the digermene 3, which precipitated as a yellow solid in benzene during the reaction. Digermene 3 was found to further react with ethylene in 1,2-difluorobenzene in a [2 + 2] cycloaddition reaction at elevated temperature (60 °C) to give the corresponding 1,2-digermacyclobutane 4. In marked contrast, the amido-substituted germylene L(Cl)GaGeN(SiMe3)Ar 2 reacted with ethylene at ambient temperature directly to the 1,2-digermacyclobutane 5 (Scheme 2).
image file: d3dt03944g-s2.tif
Scheme 2 Syntheses of digermene (3) and 1,2-digermacyclobutanes (4, 5); Ar = 2,6-iPr2C6H3.

Digermene 3 is only slightly soluble in fluorobenzene and 1,2-diflourobenzene, insoluble in benzene, toluene, THF, acetonitrile, n-hexane and n-pentane and decomposes in dichloromethane-d2, bromobenzene-d5 and 1,2-dichlorobenzene-d4 within minutes at room temperature. The 1H NMR spectrum of 3 was obtained from a solution in toluene-d8 and 1,2-difluorobenzene at 90 °C. Unfortunately, 3 was found to decompose to yet unidentified products under these conditions (Fig. S9), hence no meaningful 13C NMR spectrum was obtained. In contrast, 1H and 13C NMR spectra of 4 and 5 were recorded at ambient temperature in benzene-d6, showing the expected resonances of the β-diketiminate and C2H4 units as well as the terphenyl (4) and amine ligands (5).

Single crystals were obtained from solutions in hot fluorobenzene (3) and pentane (5) after storage at room temperature and from benzene after storage at 6 °C (4), respectively. Digermene 3 crystallised as bright yellow blocks in the monoclinic space group C2/c, 1,2-digermacyclobutane 4 as colourless plates in the triclinic space group P[1 with combining macron], and 1,2-digermacyclobutane 5 as colourless blocks in the monoclinic space group P21/c (Fig. 1). The Ge–Ge bond length in digermene 3 (2.3307(5) Å) is within the reported range for Ge–Ge double bonds (2.212–2.535[thin space (1/6-em)]17), and the substituents adopt a trans-bent orientation, which is typical for heavier dimetallenes.5f,18 However, the Ge–Ge–C bond angles (Ge–Ge–CTer 108.72(6)°; Ge–Ge–CH2 119.48(6)°) largely differ. The [2 + 2] cycloaddition of ethylene to the Ge–Ge double bond in digermene 3 causes a rearrangement of the substituents on the Ge–Ge unit from a transoid (3) to a cisoid orientation (4), and the Ge–Ge–CTer bond angle is widened by 5° (3 108.72(6)°; 4 113.77(3)°, 114.02(6)°). The Ge–Ge single bonds in the 1,2-digermacyclobutane rings (2.5440 Å 4, 2.4607 Å 5) are not only significantly shorter than the Ge–Ge double bond in 3, but also largely differ by almost 8 pm. In addition, the Ge–C bonds in the 1,2-digermacyclobutane rings in 5 (2.008(2) Å, 2.010(2) Å) are slightly elongated compared to those in 1,2-digermacyclobutane 4 (1.982(2) Å, 1.973(2) Å), and the endocyclic Ge–Ge–C bond angles of 5 (73.72(7)°, 73.64(6)°) are slightly wider than those of 4 (71.42(7)°, 71.27(7)°), while the CH2–CH2 bond lengths are almost identical (1.551(3) Å 4; 1.548(3) Å 5), respectively. These results are most likely due to the sterically more demanding terphenyl substituent compared to the amido substituent as is proven by comparing the calculated buried volumes Vbur[thin space (1/6-em)]19 of 2,6-Mes2C6H3 (36.8%) vs. N(SiMe3)(Ar) (34.5%).20 The exocyclic ethane units in 3–5 have identical CH2–CH2 (1.532(3)–1.537(3) Å), Ge–CH2 (1.982(2)–1.993(2) Å) and Ga–CH2 (1.963(2)–1.990(2) Å) bond lengths, as is expected from the comparable atomic radii of Ga (124 pm) and Ge (121 pm).21


image file: d3dt03944g-f1.tif
Fig. 1 Solid state structures of 3–5 with thermal ellipsoid (50%); H atoms omitted for clarity. Selected average bond lengths [Å] and bond angles [°]: compound 3: Ge1–Ge#1 2.3307(5), Ge1–C56 1.993(2), Ga1–C55 1.970(2), C55–C56 1.535(3), C30–Ge1–Ge#1 108.72(6), C56–Ge1–Ge#1 119.48(6), Ga1–C55–C56 115.52(14), Ge1–C56–C55 113.93(14); compound 4: Ge1–Ge2 2.5440(3), Ge1–C63 1.982(2), Ge1–C60 1.986(2), Ge2–C64 1.973(2), Ge2–C62 1.982(2), Ga1–C59 1.977(2), Ga2–C61 1.990(2), C59–C60 1.537(3), C61–C62 1.535(3), C63–C64 1.551(3), Ge1–Ge2–C64 71.42(7), Ge2–Ge1–C63 71.27(7), Ge2–Ge1–C60 132.38(7), Ge1–Ge2–C62 130.68(6), Ge1–Ge2–C010 114.02(6), Ge2–Ge1–C65 113.77(3), Ga1–C59–C60 112.29(14), Ga2–C61–C62 112.12(14), Ge1–C60–C59 115.47(15), Ge2–C62–C61 117.24(14); compound 5: Ge1–Ge2 2.4607(3), Ge1–C31 1.987(2), Ge1–C47 2.008(2), Ge2–C48 2.010(2), Ge2–C61 1.986(2), Ga1–C30 1.966(2), Ga2–C62 1.963(2), C30–C31 1.536(3), C47–C48 1.548(3), C61–C62 1.532(3), Ge1–Ge2–C48 73.77(7), Ge2–Ge1–C47 73.64(6), Ge2–Ge1–C31 121.99(7), Ge1–Ge2–C61 120.91(7), Ge1–Ge2–N4 122.51(6), Ge2–Ge1–N3 123.00(6), Ga1–C30–C31 116.37(15), Ga2–C62–C61 116.38(15), Ge1–C31–C30 113.69(15), Ge2–C61–C62 114.19(14).

Recently, regioselective and stereoselective insertion reactions of phenylacetylene into the tetryl-tetrylene linkage of germylstannylene, germylplumbylene and stannylplumbylene as well as the silylated germylene [(Me3Si)3Si]2Ge have been reported.22 In addition, 1,2-insertion reactions of alkynes into Ge–C bonds of dibromodigermenes were found to give bromovinylgermylenes,23 while silylenes containing a sterically demanding, electropositive hypersilyl (Si(SiMe3)3) ligand reacted with ethylene to the corresponding silirane.10c–e Moreover, trimethylstannyl-substituted silylenes reversibly inserted ethylene into the Si–Sn bond.24 However, to the best of our knowledge, we report here for the first time the multiple activation of ethylene using germylenes containing a bulky, electropositive main group metal-substituent.

The mechanism of the reaction was subsequently investigated by quantum chemical calculations. In addition to the reaction of 1 and 2 with ethylene (Fig. 2), the addition of C2H4 to reference systems with sterically less demanding substituents as well as substituents without a Ga centre was computed (Fig. S10–S12). While the barrier for cycloaddition of germylenes to the C–C double bond of ethylene has been found to be close to zero in gas-phase kinetic studies25 and quantum chemical calculations26 (see also Fig. S10–S12), activation barriers of 13.6 and 19.6 kcal mol−1 are found for germylenes 1 and 2 as a result of the sterically demanding substituents. The higher barrier for 2 agrees with previous quantum chemical calculations, which showed that electron-withdrawing groups on the germylene hinder the cycloaddition.27 As-formed germiranes Int-1 then undergo 1,2-migration of the C2H4 unit into the Ge–Ga bonds to give germylenes Int-2, which dimerise to digermenes as demonstrated by the isolation of 3 (Fig. 2). A comparison with the reference systems shows that the driving force of the 1,2-migration is essentially the reduction of steric interactions.


image file: d3dt03944g-f2.tif
Fig. 2 Gibbs energies for the reaction of germylenes 1 and 2 with ethylene yielding the 1,2-digermacyclobutanes 4 and 5 calculated by means of PBE0-D3BJ. The values are given in kcal mol−1. L = HC(C(Me)NAr)2, Ar = 2,6-iPr2C6H3; DMP = 2,6-Mes2C6H3, Mes = 2,4,6-Me3C6H2; NRR′ = N(SiMe3)(Ar).

Insertion of a C2H4 unit into a tetrel-ligand bond was previously observed by Power et al. in the reactions of sterically hindered stannylenes with C2H4,11 whereas, to the best of our knowledge, it is without precedent in germylene chemistry except for the C2H4 insertion into the Ge–Ge double bond of 1,2-cyclodigermene II, which occurred under kinetic control to give the bisgermacyclopropane IV.6b The digermenes 3 and Int-3b then react with another equivalent of ethylene in [2 + 2] cycloaddition to give the corresponding 1,2-digermacyclobutanes 4 and 5, respectively. The calculations reveal that the addition of ethylene to Int-3b is barrier free (Fig. S13), whereas an activation energy of 35.1 kcal mol−1 is found for the addition of C2H4 to 3, which is due to the sterically very demanding DMP ligands (Fig. 2 and S13). The calculations clearly explain why digermene 3 can be isolated in marked contrast to Int-3b.

Experimental

Synthetic procedures

All experiments were performed either in a glovebox and using standard Schlenk line techniques under argon atmosphere. Toluene and n-hexane were dried using a mBraun solvent purification system (SPS), degassed and stored in Schlenk flasks under argon atmosphere. Deuterated solvents were stored over molecular sieves (4 Å) and degassed prior to use. The anhydrous nature of the solvents was verified by Karl Fischer titration. LGa(μ-Cl)GeDMP (1) and L(Cl)GaGeN(SiMe3)Ar (2) were synthesized according to literature procedures.15 Microanalysis was performed at the Elemental Laboratory for Microanalysis of the University of Duisburg-Essen. The melting points were measured using a Thermo Scientific 9300 apparatus.

Spectroscopic methods

NMR spectra were either recorded using a Ascend™ 400 spectrometer (1H NMR, 400 MHz; 13C{1H} NMR, 100 MHz) or a Bruker Avance DPX-300 spectrometer (1H NMR, 300 MHz). The spectra were referenced to internal C6D5H (1H: δ = 7.16, C6D6) and toluene-d7 (1H: δ = 2.08, toluene-d8) or to natural-abundance carbon resonances C6D6 (13C: δ = 128.06, C6D6) and toluene-d8 (13C: δ = 20.43, toluene-d8). IR spectra were recorded with an ALPHA-T FT-IR spectrometer equipped with a single reflection ATR sampling module. The IR spectrometer was placed in a glovebox to guarantee measurements under inert gas conditions.

Crystallography

The crystals were mounted on nylon loops in inert oil. Data of 3 were collected on a Bruker AXS D8 Kappa diffractometer with APEX2 detector (monochromated Mo radiation, λ = 0.71073 Å) and data of 4 and 5 were collected on a Bruker AXS D8 Venture diffractometer with photon II detector (monochromated Cu radiation, λ = 1.54178 Å, microfocus source) at 100(2) K. Structures were solved by direct methods (SHELXS-2013)28 and refined anisotropically by full-matrix least-squares on F2 (SHELXL-2017).29,30 Absorption corrections were performed semi-empirically from equivalent reflections on the basis of multi-scans (Bruker AXS APEX3). Hydrogen atoms were refined using a riding model or rigid methyl groups.

Two solvent molecules are disordered in 3 over a two-fold rotational axis. The local symmetry was ignored in the refinement (negative PART) and all bond lengths and angles of the solvent molecules were restrained to be equal (SADI). In addition, the atoms were restrained to lie on a common plane (FLAT). RIGU and SIMU restraints were applied to the displacement parameters.

4 contains four highly disordered benzene molecules and 5 two highly disordered n-pentane molecules, respectively. The final refinements were done with solvent free datasets from a PLATON/SQUEEZE run.31 The molecules were included in the sum formulas for completeness.

The bond lengths and angles of the solvent molecule in 5 were restrained to be equal (SADI).

CCDC 2250957 (3), 2250958 (4) and, 2250959 (5) contain the supplementary crystallographic data for this paper.

Computations

All calculations were performed by using the program package Gaussian 16.32 The geometrical parameters of all stationary points were optimized by means of PBE0[thin space (1/6-em)]33 with additional dispersion correction with Becke–Johnson damping.34 The basis set def2-SVP35,36 was employed. For all stationary points no symmetry restriction was applied. Frequency calculations were carried out at each of the structures to verify the nature of the stationary point. It turned out that all transition states have exactly one imaginary frequency, whereas all other structures have none.

Furthermore, the energies of the stationary points were calculated using the density functionals PBE0-D3BJ and the basis set def2-TZVP.35,36 To take solvent effects into account, the solvent model SMD37 (benzene as solvent) was used for the single point calculations.

Synthesis of compound 3. A degassed red solution of germylene LGa(μ-Cl)GeDMP 1 (65 mg, 0.071 mmol) in 1 mL of 1,2-difluorobenzene under an ethene atmosphere (1 atm) was stirred for 5 minutes, upon which a yellow participate was formed. The ethene atmosphere was removed and replaced by an argon atmosphere. The solid was dissolved in hot 1,2-difluorobenzene (2 mL) and stored at ambient temperature to give yellow crystals of digermene 3. Yield: 39 mg (0.021 mmol, 59%), m.p. 128 °C decomp. Anal. Calcd for C110H140Cl2Ga2Ge2N4: C 70.50, N 2.99, H 7.53; found: C 69.30, N 2.70, H 7.24. 1H NMR (300 MHz, Tol-d8 with 1,2-F2-C6H4, 363 K): δ [ppm] = 7.30–6.10 (m, ArH, overlapping with solvent) 4.83 (s, 2H, γ-C[H with combining low line]), 3.49 (sept, 3JHH = 6.7 Hz, 4H, C[H with combining low line](CH3)2), 2.83 (sept, 3JHH = 6.7 Hz, 4H, C[H with combining low line](CH3)2), 2.33 (s, 10H, C[H with combining low line]3), 1.44 (s, 12H, C[H with combining low line]3), 1.37 (s, 6H, C[H with combining low line]3), 1.09 (m, 52H), 0.87 (d, 3JHH = 6.6 Hz,12H, CH(C[H with combining low line]3)2), 0.81 (m, 4H), 0.70 (m, 6H). It is not possible to assign every signal due to the unsatisfactory resolution of the spectrum. Due to the insolubility and instability of digermene 3 in common solvents it was not possible to measure 13C NMR spectra or 2D spectra. IR(ATR): ν [cm−1] = 2958, 2914, 2861, 1596, 1529, 1438, 1385, 1308, 1266, 1223, 1180, 1094, 1061, 1027, 998, 935, 869, 850, 802, 750, 688, 630, 592, 549, 501, 449.
Synthesis of compound 4. A degassed yellow suspension of digermene 3 (20 mg, 0.011 mmol) in 2 mL of 1,2-difluorobenzene under an ethene atmosphere (1 atm) was slowly warmed to 60 °C, giving a colourless solution after16 hours. All volatile components were removed in vacuo and the residue was dissolved in benzene (0.3 mL) and stored at 6 °C to give colourless crystals of 1,2-digermacyclobutane 4. Yield: 6 mg (0.003 mmol, 27%), m.p. 57–58 °C. Anal. Calcd for C112H144Cl2Ga2Ge2N4: C 70.73, N 2.95, H 7.63; found: C 70.20, N 3.13, H 7.54. 1H NMR (400 MHz, C6D6, 298 K): δ [ppm] = 7.28 (m, 4H, Ar[H with combining low line]), 7.10 (m, 9H, Ar[H with combining low line], overlapping with solvent resonances), 6.96 (m, 3H, Ar[H with combining low line]), 6.88 (m, 8H, Ar[H with combining low line]), 6.72 (m, 4H, Ar[H with combining low line]), 4.79 (s, 2H, γ-C[H with combining low line]), 3.73 (sept, 3JHH = 6.4 Hz, 4H, C[H with combining low line](CH3)2), 3.00 (sept, 3JHH = 6.7 Hz, 2H, C[H with combining low line](CH3)2), 2.83 (sept, 3JHH = 6.6 Hz, 2H, C[H with combining low line](CH3)2), 2.58 (s, 6H, C[H with combining low line]3(Mes)), 2.51 (s, 6H, C[H with combining low line]3(Mes)), 2.25 (s, 6H, C[H with combining low line]3(Mes)), 2.10 (s, 6H, C[H with combining low line]3(Mes)), 1.51 (d, 3JHH = 6.5 Hz, 6H, CH(C[H with combining low line]3)2), 1.47 (s, 12H, NCC[H with combining low line]3), 1.39 (s, 6H, C[H with combining low line]3(Mes)), 1.31 (d, 3JHH = 6.8 Hz, 6H, CH(C[H with combining low line]3)2), 1.27 (d, 3JHH = 6.8 Hz, 14H, CH(C[H with combining low line]3)2 overlapping with C[H with combining low line]2), 1.20 (d, 3JHH = 6.8 Hz, 6H, CH(C[H with combining low line]3)2), 1.02 (d, 3JHH = 6.7 Hz, 6H, CH(C[H with combining low line]3)2), 0.92 (d, 3JHH = 6.3 Hz, 12H, CH(C[H with combining low line]3)2 overlapping with C[H with combining low line]3(Mes)), 0.88 (d, 3JHH = 6.9 Hz, 6H, CH(C[H with combining low line]3)2), 0.72 (m, 2H, C[H with combining low line]2), 0.49 (m, 1H, C[H with combining low line]2), 0.19 (m, 2H, C[H with combining low line]2), −0.80 (m, 2H, C[H with combining low line]2). 13C NMR (101 MHz, C6D6, 298 K): δ [ppm] = 169.2 (N[C with combining low line]CH3), 146.5 (Ar), 145.7 (Ar), 143.2 (Ar), 143.2 (Ar), 142.6 (Ar), 141.5 (Ar), 141.2 (Ar), 141.0 (Ar), 141.0 (Ar), 138.5 (Ar), 138.2 (Ar), 135.6 (Ar), 135.5 (Ar), 135.3 (Ar), 135.1 (Ar), 135.0 (Ar), 129.5 (Ar), 129.3 (Ar), 128.9 (Ar), 127.8 (Ar), 126.2 (Ar), 125.9 (Ar), 124.3 (Ar), 124.0 (Ar), 97.6 (γ-[C with combining low line]H), 29.6 ([C with combining low line]H(CH3)2), 28.3 (CH([C with combining low line]H3)2), 28.1 (CH([C with combining low line]H3)2), 28.0 ([C with combining low line]H(CH3)2), 25.2 (CH([C with combining low line]H3)2), 25.1 (CH([C with combining low line]H3)2), 24.7 (CH([C with combining low line]H3)2), 24.5 (CH([C with combining low line]H3)2), 24.5 (CH([C with combining low line]H3)2), 24.4 (CH3), 23.9, 23.8 (NC[C with combining low line]H3), 22.5 ([C with combining low line]H2), 22.5 ([C with combining low line]H3), 22.0 (CH([C with combining low line]H3)2), 21.7 ([C with combining low line]H2), 21.5 ([C with combining low line]H3), 21.1 ([C with combining low line]H3), 21.0 (CH([C with combining low line]H3)2), 21.0 (CH([C with combining low line]H3)2), 16.8 ([C with combining low line]H2), 5.1 ([C with combining low line]H2), 1.7 ([C with combining low line]H2). IR(ATR): ν [cm−1] = 2972, 2924, 2862, 1538, 1438, 1390, 1314, 1261, 1175, 1103, 1022, 935, 869, 845, 802, 760, 745, 639, 558, 453.
Synthesis of compound 5. A degassed purple solution of germylene L(Cl)GaGeN(SiMe3)Ar 2 (45 mg, 0.053 mmol) in 1 mL of toluene was stirred for 5 minutes under an ethene atmosphere (1 atm) to give a colourless solution. After the removal of all volatile components in vacuo, the residue was dissolved in n-pentane (1 mL) and stored at room temperature. Colourless crystals of 1,2-digermacyclobutane 5 were formed within 12 hours. Yield: 30 mg (0.017 mmol, 61%), m.p. 220–221 °C. Anal. Calcd for C94H146Cl2Ga2Ge2N4Si4: C 63.71, N 4.74, H 8.31; found: C 64.20, N 4.52, H 8.59. 1H NMR (400 MHz, C6D6, 298 K): δ [ppm] = 7.16 (m, 8H, Ar[H with combining low line], overlapping with solvent resonances), 7.12–7.05 (m, 12H, Ar[H with combining low line]), 6.99 (dd, 3JHH = 7.0, 2.3 Hz, 2H, Ar[H with combining low line]), 4.85 (s, 2H, γ-C[H with combining low line]), 3.76 (sept, 3JHH = 6.7 Hz, 2H, C[H with combining low line](CH3)2), 3.56 (sept, 3JHH = 6.7 Hz, 4H, C[H with combining low line](CH3)2), 3.41 (sept, 3JHH = 6.8 Hz, 2H, C[H with combining low line](CH3)2), 3.11 (sept, 3JHH = 6.8 Hz, 2H, C[H with combining low line](CH3)2), 3.04 (sept, 3JHH = 6.7 Hz, 2H, C[H with combining low line](CH3)2), 1.68 (m, 2H, C[H with combining low line]2), 1.50 (s, 6H, NCC[H with combining low line]3), 1.46 (s, 6H, NCC[H with combining low line]3), 1.44–1.33 (m, 30H, CH(C[H with combining low line]3)2), 1.27 (d, 3JHH = 6.7 Hz, 6H, C[H with combining low line](CH3)2), 1.22 (d, 3JHH = 6.9 Hz, 6H, C[H with combining low line](CH3)2), 1.18 (d, 3JHH = 6.9 Hz, 6H, C[H with combining low line](CH3)2), 1.10 (d, 3JHH = 6.5 Hz, 6H, C[H with combining low line](CH3)2), 1.04 (d, 3JHH = 7.1 Hz, 6H, C[H with combining low line](CH3)2), 1.02 (d, 3JHH = 7.1 Hz, 6H, C[H with combining low line](CH3)2), 0.88 (t, 3JHH = 7.0 Hz, 6H, C[H with combining low line]2), 0.69 (d, 3JHH = 6.9 Hz, 6H, C[H with combining low line](CH3)2), 0.64–0.51 (m, 1H, C[H with combining low line]2), 0.25 (m, 2H, C[H with combining low line]2), −0.02 (s, 18H, Si(C[H with combining low line]3)3). 13C NMR (101 MHz, C6D6, 298 K): δ [ppm] = 169.27 (N[C with combining low line]CH3), 147.3 (Ar), 146.6 (Ar), 146.2 (Ar), 146.0 (Ar), 143.5 (Ar), 142.8 (Ar), 142.2 (Ar), 141.6 (Ar), 140.6 (Ar), 127.6 (Ar), 125.8 (Ar), 125.5 (Ar), 124.9 (Ar), 124.2 (Ar), 124.1 (Ar), 123.8 (Ar), 123.4 (Ar), 97.9 (Ar), 34.4 ([C with combining low line]H2), 29.41 ([C with combining low line]H(CH3)2), 29.3 ([C with combining low line]H(CH3)2), 28.8 (CH([C with combining low line]H3)2), 27.9 ([C with combining low line]H(CH3)2), 27.8 ([C with combining low line]H(CH3)2), 27.6 ([C with combining low line]H(CH3)2), 27.3 ([C with combining low line]H(CH3)2), 26.2 (CH([C with combining low line]H3)2), 25.8 (CH([C with combining low line]H3)2), 25.0 (CH([C with combining low line]H3)2), 25.0 (CH([C with combining low line]H3)2), 24.9 (CH([C with combining low line]H3)2), 24.8 ([C with combining low line]H2), 24.1 (CH([C with combining low line]H3)2), 24.1 (CH([C with combining low line]H3)2), 24.0 (CH([C with combining low line]H3)2), 23.8 (CH([C with combining low line]H3)2), 23.5 (NC[C with combining low line]H3), 23.5 (NC[C with combining low line]H3), 22.7 ([C with combining low line]H2), 17.3 ([C with combining low line]H2), 14.3 ([C with combining low line]H2), 13.6 (CH([C with combining low line]H3)2), 3.5 ([C with combining low line]H2), 2.3 (Si([C with combining low line]H3)3). IR(ATR): ν [cm−1] = 3066, 2962, 2932, 2866, 1524, 1462, 1433, 1380, 1318, 1237, 1180, 1098, 1046, 1012, 922, 864, 830, 793, 749, 692, 638, 573, 540, 444.

Conclusions

The distinct influence of aryl (DMP) and amino (N(SiMe3)Ar) substituents on the reaction pathway of L(Cl)Ga-substituted germylenes LGa(μ-Cl)GeDMP 1 and L(Cl)GaGeN(SiMe3)Ar 2 with ethylene is demonstrated. Based on quantum chemical calculations (DFT), both germylenes react with ethylene with [2 + 1] cycloaddition and formation of germiranes (Int-1), which undergo 1,2-migration of the C2H4 unit into the Ge–Ga bonds to give the corresponding germylenes (Int-2). Dimerisation of these germylenes to the corresponding digermenes is proven by the isolation and structural characterisation of digermene 3, which is stabilized by the sterically very demanding DMP substituent. In contrast, the amido-substituted digermene (Int-3b) directly reacts with one equivalent of C2H4via [2 + 2] cycloaddition to the corresponding 1,2-digermacyclobutane 5. DFT calculations showed that this reaction is barrier free, whereas an activation energy of 35.1 kcal mol−1 is found for the addition of C2H4 to digermene 3 due to the bulky DMP ligands. As a consequence, digermene 3 only reacts with ethylene at elevated temperature to the corresponding 1,2-digermacyclobutane 4.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

Generous financial support by the DFG (SCHU1069/26-1; INST 20876/282-1 FUGG) and the University of Duisburg-Essen is gratefully acknowledged. We acknowledge support by the Open Access Publication Fund of the University of Duisburg-Essen.

References

  1. Y. Gao, L. Neal, D. Ding, W. Wu, C. Baroi, A. M. Gaffney and F. Li, ACS Catal., 2019, 9, 8592 CrossRef.
  2. L. Kniel, O. Winter and K. Stork, in Ethylene, Keystone to the Petrochemical Industry, Marcel Dekker Inc., New York, 1st edn, 1980 Search PubMed.
  3. (a) K. Yuvaraj, I. Douair, L. Maron and C. Jones, Chem. – Eur. J., 2020, 26, 14665 CrossRef PubMed; (b) B. Rösch and S. Harder, Chem. Commun., 2021, 57, 9354 RSC; (c) J. M. Parr, A. Phanopoulos, A. Vickneswaran and M. R. Crimmin, Chem. Sci., 2023, 14, 1590 RSC.
  4. (a) P. Power, Nature, 2010, 463, 171 CrossRef PubMed; (b) C. Weetman and S. Inoue, ChemCatChem, 2018, 10, 4213 CrossRef CAS; (c) R. L. Melen, Science, 2019, 363, 479 CrossRef CAS PubMed; (d) C. Weetman, Chem. – Eur. J., 2021, 27, 1941 CrossRef CAS; (e) S. Inoue, R. L. Melen and S. Harder, Eur. J. Inorg. Chem., 2022, e202200414 CrossRef CAS.
  5. (a) V. Y. Lee, Molecules, 2023, 28, 1558 CrossRef CAS PubMed; (b) L. Wang, Y. Li, Z. Li and M. Kira, Coord. Chem. Rev., 2022, 457, 214413 CrossRef CAS; (c) S. Fujimori and S. Inoue, Comprehensive Organometallic Chemistry IV, 2022, vol. 10, p. 1 Search PubMed; (d) C. Shan, S. Yao and M. Driess, Chem. Soc. Rev., 2020, 49, 6733 RSC; (e) S. Fujimori and S. Inoue, Eur. J. Inorg. Chem., 2020, 3131 CrossRef CAS; (f) F. Hanusch, L. Groll and S. Inoue, Chem. Sci., 2021, 12, 2001 RSC; (g) Y. E. Türkmen, Comprehensive Organometallic Chemistry IV, 2022, vol. 10, p. 506 Search PubMed; (h) J.-D. Guo and T. Sasamori, Chem. – Asian J., 2018, 13, 3800 CrossRef CAS.
  6. Si: (a) N. Wiberg, S. K. Vasisht, G. Fischer and P. Mayer, Z. Anorg. Allg. Chem., 2004, 630, 1823 CrossRef CAS. Ge: (b) T. Sasamori, T. Sugahara, T. Agou, K. Sugamata, J.-D. Guo, S. Nagase and N. Tokitoh, Chem. Sci., 2015, 6, 5526 RSC.
  7. Ge: (a) T. J. Hadlington, J. Li, M. Hermann, A. Davey, G. Frening and C. Jones, Organometallics, 2015, 34, 3175 CrossRef CAS; (b) J. A. Kelly, M. Juckel, T. J. Hadlington, I. Fernandez, G. Frenking and C. Jones, Chem. – Eur. J., 2019, 25, 2773 CrossRef CAS PubMed. Sn: (c) Y. Peng, B. D. Ellis, X. Wang, J. C. Fettinger and P. P. Power, Science, 2009, 325, 1668 CrossRef CAS.
  8. (a) J. S. Han, T. Sasamori, Y. Mizuhata and N. Tokitoh, J. Am. Chem. Soc., 2010, 132, 2546 CrossRef PubMed; (b) T. J. Hadlington, M. Hermann, G. Frenking and C. Jones, Chem. Sci., 2015, 6, 7249 RSC; (c) S. Wang, M. L. McCrea-Hendrick, C. M. Weinstein, C. A. Caputo, E. Hoppe, J. C. Fettinger, M. M. Olmstead and P. P. Power, J. Am. Chem. Soc., 2017, 139, 6586 CrossRef.
  9. T. Sugahara, J.-D. Guo, T. Sasamori, S. Nagase and N. Tokitoh, Angew. Chem., Int. Ed., 2018, 57, 3499 CrossRef.
  10. Si: (a) R. Rodriguez, D. Gau, T. Kato, N. Saffon-Merceron, A. De Cozar, F. P. Cossio and A. Baceiredo, Angew. Chem., Int. Ed., 2011, 50, 10414 CrossRef; (b) F. Lips, J. C. Fettinger, A. Mansikkamäki, H. M. Tuononen and P. P. Power, J. Am. Chem. Soc., 2014, 136, 634 CrossRef PubMed; (c) D. Wendel, W. Eisenreich, C. Jandl, A. Pöthig and B. Rieger, Organometallics, 2016, 35, 1 CrossRef; (d) D. Wendel, A. Porzelt, F. A. D. Herz, D. Sarkar, C. Jandl, S. Inoue and B. Rieger, J. Am. Chem. Soc., 2017, 139, 8134 CrossRef PubMed; (e) K. I. Leszczyńska, P. Deglmann, C. Präsang, V. Huch, M. Zimmer, D. Schweinfurth and D. Scheschkewitz, Dalton Trans., 2020, 49, 13218 RSC; (f) R. Nougué, S. Takahashi, A. Baceiredo, N. Saffon-Merceron, V. Branchadell and T. Kato, Angew. Chem., Int. Ed., 2023, 62, e202215394 CrossRef. Ge: (g) H. Ohgaki, Y. Kabe and W. Ando, Organometallics, 1995, 14, 2139 CrossRef; (h) T. Y. Lai, K. L. Gullen, C.-Y. Chen, J. C. Fettinger and P. P. Power, Organometallics, 2019, 38, 1425 CrossRef.
  11. T. Y. Lai, J.-D. Guo, J. C. Fettinger, S. Nagase and P. P. Power, Chem. Commun., 2019, 55, 405 RSC.
  12. (a) C. Ganesamoorthy, J. Schoening, C. Wölper, L. Song, P. R. Schreiner and S. Schulz, Nat. Chem., 2020, 12, 608 CrossRef CAS; (b) J. Schoening, C. Ganesamoorthy, C. Wölper, E. Solel, P. R. Schreiner and S. Schulz, Dalton Trans., 2022, 51, 8249 RSC.
  13. T. Sergeieva, D. Mandal and D. M. Andrada, Chem. – Eur. J., 2021, 27, 10601 CrossRef CAS.
  14. (a) J. Schoening, C. Wölper and S. Schulz, Eur. J. Inorg. Chem., 2023, 26, e202200638 CrossRef CAS; (b) J. Schoening, A. Gehlhaar, C. Wölper and S. Schulz, Chem. – Eur. J., 2022, 28, e2022010 Search PubMed; (c) X. Sun, A. Hinz, S. Schulz, L. Zimmermann, M. Scheer and P. W. Roesky, Chem. Sci., 2023, 14, 4769 RSC.
  15. A. Bücker, C. Wölper, G. Haberhauer and S. Schulz, Chem. Commun., 2022, 58, 9758 RSC.
  16. A. Bücker, C. Wölper and S. Schulz, Polyhedron, 2024, 247, 116702 CrossRef.
  17. Search for Ge[double bond, length as m-dash]Ge bonds in CSD 5.43 gave 78 hits (85 bonds) (2.212–2.535 Å, mean 2.34(8) Å). Search for Ge–Ge double bonds (coord. no of Ge restricted to 3) in CSD 5.43 gave 66 hits (71 bonds) (2.212–2.535 Å, mean 2.34(8) Å). See: C. R. Groom, I. J. Bruno, M. P. Lightfoot and S. C. Ward, Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater., 2016, 72, 171 CrossRef CAS.
  18. (a) R. C. Fischer and P. P. Power, Chem. Rev., 2010, 110, 3877 CrossRef CAS; (b) J.-D. Guo, D. J. Liptrot, S. Nagase and P. P. Power, Chem. Sci., 2015, 6, 6235 RSC; (c) P. P. Power, Organometallics, 2020, 39, 4127 CrossRef; (d) T. Sasamori, V. Lee, N. Nagahora and S. Morisako, in Comprehensive Inorganic Chemistry III, ed. J. Reedijk and K. R. Poeppelmeier, Elsevier, Oxford, 3rd edn, 2023, p. 118 Search PubMed.
  19. L. Faliven, R. Credendino, A. Poater, A. Petta, L. Serra, R. Oliva, V. Scarano and L. Cavallo, Organometallics, 2016, 35, 2286 CrossRef.
  20. C. Helling, G. E. Cutsail III, H. Weinert, C. Wölper and S. Schulz, Angew. Chem., Int. Ed., 2020, 59, 7561 CrossRef PubMed.
  21. P. Pyykkö and M. Atsumi, Chem. – Eur. J., 2009, 15, 186 CrossRef.
  22. (a) J.-J. Maudrich, F. Diab, S. Weiß, M. Zweigart, K. Eichele, H. Schubert, R. Müller, M. Kaupp and L. Wesemann, Chem. – Eur. J., 2021, 27, 4691 CrossRef PubMed; (b) M. Walewska, J. Baumgartner and C. Marschner, Molecules, 2020, 25, 68 CrossRef.
  23. R. Rodriguez, Y. Contie, D. Gau, N. Saffon-Merceron, K. Miqueu, J.-M. Sotiropoulos, A. Baceiredo and T. Kato, Angew. Chem., Int. Ed., 2013, 52, 8437 CrossRef PubMed.
  24. T. Sugahara, A. E. Ferao, A. R. Planells, J.-D. Guo, S. Aoyama, K. Igawa, K. Tomooka, T. Sasamori, D. Hashizume, S. Nagase and N. Tokitoh, Dalton Trans., 2020, 49, 7189 RSC.
  25. R. Becerra, S. E. Boganov, M. P. Egorov, V. I. Faustov, V. M. Promyslov, O. M. Nefedov and R. Walsh, Phys. Chem. Chem. Phys., 2002, 4, 5079 RSC.
  26. A. A. Birukov, V. I. Faustov, M. P. Egorov and O. M. Nefedov, Russ. Chem. Bull., Int. Ed., 2005, 54, 2003 CrossRef.
  27. M.-D. Su and S.-Y. Chu, J. Am. Chem. Soc., 1999, 121, 11478 CrossRef.
  28. G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 1990, 46, 467 CrossRef.
  29. G. M. Sheldrick, SHELXL-2017, Program for the Refinement of Crystal Structures, University of Göttingen, Göttingen (Germany), 2017 Search PubMed. See also: G. M. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr., Sect. C: Struct. Chem., 2015, 71, 3 Search PubMed.
  30. shelXle, A Qt GUI for SHELXL, C. B. Hübschle, G. M. Sheldrick and B. Dittrich, J. Appl. Crystallogr., 2011, 44, 1281 CrossRef.
  31. P. van der Sluis and A. L. Spek, PLATON/SQUEEZE, Acta Crystallogr., Sect. A: Found. Crystallogr., 1990, 46, 194 CrossRef.
  32. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16, Revision A.03, Wallingford CT, 2016 Search PubMed.
  33. C. Adamo and V. Barone, J. Chem. Phys., 1999, 110, 6158 CrossRef CAS.
  34. S. Grimme, S. Ehrlich and L. Goerigk, J. Comput. Chem., 2011, 32, 1456 CrossRef CAS.
  35. F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005, 7, 3297 RSC.
  36. F. Weigend, Phys. Chem. Chem. Phys., 2006, 8, 1057 RSC.
  37. A. V. Marenich, C. J. Cramer and D. G. Truhlar, J. Phys. Chem., 2009, 113, 6378 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Experimental and analytical data (NMR, IR spectra, crystallographic data) and details from DFT calculations. CCDC 2250957 (3), 2250958 (4) and 2250959 (5). For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3dt03944g

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.