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

Synthesis of perhalogenated silylboranes (X = Cl, I) and their application in regiodivergent alkene silaboration

Jan Heller a, Christoph D. Bucha, Alexander V. Virovetsa, Eugenia Peresypkinaa, Hans-Wolfram Lernera, Felipe Fantuzzib and Matthias Wagner*a
aInstitut für Anorganische und Analytische Chemie, Goethe-Universität Frankfurt, Max-von-Laue-Straße 7, D-60438 Frankfurt am Main, Germany. E-mail: matthias.wagner@chemie.uni-frankfurt.de
bSchool of Chemistry and Forensic Science, University of Kent, Park Wood Rd, Canterbury CT2 7NH, UK

Received 15th August 2025 , Accepted 9th September 2025

First published on 10th September 2025


Abstract

Silaboration of olefins is a synthetically valuable and atom-economic mode of functionalization; however, it typically requires transition-metal catalysis. We have overcome this requirement by using highly reactive perhalogenated silylboranes, X2B–SiX3 (X = Cl, I), for which we herein report a straightforward synthesis, a full characterization, and their key properties. Access to this compound class was enabled by substantial improvement in the synthesis protocol for our previously published compound [Et4N][I3B–SiI3], now available on a 40 g scale via only two steps. Cation exchange with Li[Al(OC(CF3)3)4] affords the mixture Li[I3B–SiI3]/I2B–SiI3/LiI, serving as a synthetic equivalent of the elusive pure I2B–SiI3. Its chlorine analogue Cl2B–SiCl3 is accessible as a distillable liquid via treatment of [Et4N][I3B–SiI3] with GaCl3. For both perhalogenated silylboranes, various Lewis base adducts Do·X2B–SiX3 were obtained in excellent yields and structurally characterized by X-ray diffraction (Do = SMe2, Py, PPh3, IDipp; IDipp = 1,3-bis(2,6-diisopropylphenyl)-1,3-dihydro-2H-imidazol-2-ylidene). We demonstrated that Me2S·I2B–SiI3 undergoes efficient 1,2-silaboration of the challenging, non-activated substrate ethylene at rt with 0.1 eq. BI3 as promoter. In contrast, Li[I3B–SiI3]/I2B–SiI3/LiI effects a quantitative, unprecedented 1,1-silaboration of cyclohexene at rt. This remarkable reactivity switch was elucidated by experimental and quantum-chemical studies of the underlying steric and electronic factors.


Introduction

Once considered exotic and of limited utility, perhalogenated diborane(4) and disilane compounds (I, III; Fig. 1a) have recently emerged as valuable building blocks for purposes ranging from organic synthesis to materials development.1–6 The direct bond between two Lewis-acidic sites in I and III, each bearing good leaving groups, presents both challenges and opportunities: on the one hand, this unique combination promotes spontaneous disproportionation and vigorous decomposition upon exposure to air and moisture.7–10 On the other hand, it enables uncatalyzed diboration reactions using I,11–17 the in situ generation of versatile [SiX3] nucleophiles from III upon simple halide addition,18,19 and extensive late-stage derivatization at the B–X and Si–X bonds of the primary products. Thus, in contrast to the abundant bis(pinacolato)diboron (pinB–Bpin), whose B atoms are electronically tamed by O[double bond, length as m-dash]B π-donation and serve primarily as transmetallation partners in Suzuki–Miyaura cross-couplings,20–22 type-I halogenoboranes are tailored for applications where the B atoms are to remain as property-determining functional units in the final molecule.23–29 Likewise, the Si2X6/X trichlorosilylation system and the controlled disproportionation of Si2X6 with NR3 (ref. 30–32) have proven valuable for the synthesis of extensively derivatizable organosilanes,33–36 oligotetrelanes,37–42 and silicon clusters.43–50 In contrast to I, III undergoes no spontaneous 1,2-additions to unsaturated organic substrates, and theoretical studies predict a prohibitively high activation barrier without a catalyst.51,52
image file: d5sc06234a-f1.tif
Fig. 1 (a) Perhalogenated diboranes(4) (I), disilanes (III), and the perhalogenated silylboranes (II) studied in this work. (b) Previously studied silylboranes (IV) capable of undergoing uncatalyzed silaboration reactions (C atoms marked with asterisks bear tBu substituents); computed transition state (TS) for the silaboration of ethylene with IVH (tBu groups have been omitted in the calculations). (c) Silylborates H[V] and VI bearing some structural similarity with II.

Given the indispensable role of borylated53,54 and silylated55 building blocks in synthesis, it is desirable to combine both types of functional groups within a single building block, for which silylboranes of the type R2B–SiR3 are the most obvious candidates.56–60 Electronically stabilized representatives such as the prominent pinB–SiMe2Ph typically require activation by (precious) metal complexes prior to addition across C[double bond, length as m-dash]C double61–63 or C[triple bond, length as m-dash]C triple bonds.64–72 In only a few cases, the addition of a (Lewis) base (KOtBu,73,74 KN(SiMe3)2,75 PR3,76 pyridines77,78) has been sufficient to replace the transition metal catalyst in silaboration reactions. Yet, a significant proportion of base-catalyzed silylborane transformations results in incorporation of either the boryl74,79–84 or the silyl85–91 group,92 while the respective counterpart is discarded. So far, a single uncatalyzed silaboration reaction has been reported, employing compounds IVH and IVCl in THF (Fig. 1b).93,94 Key to this transformation is the incorporation of both the B and Si atoms of IV into planar heterofluorene scaffolds, which—compared to pinB–SiMe2Ph—enhances their exposure to the unsaturated substrate while reducing π-electron donation into the vacant B(pz) orbital (quantum-chemical calculations exclude a promoting effect of the THF ligand on B–Si-bond cleavage; cf. the transition state TS of olefin silaboration shown in Fig. 1b). Based on this background and the high reactivity of I and III, we reasoned that the perhalogenated silylborane II (Fig. 1a) as a silaboration reagent should uniquely combine a strong tendency towards B–Si heterolysis and diverse opportunities for subsequent derivatization. Herein, we demonstrate that type-II compounds with X = Cl, I can indeed be readily synthesized on a multigram scale. We provide a full characterization of their B-adducts with various Lewis bases and show that the Cl derivative Cl2B–SiCl3 can even be isolated in its free form as a distillable liquid. Notably, we disclose that both uncatalyzed 1,2- and rare 1,1-addition reactions to alkenes have been achieved. Only a few previously reported compounds share structural or electronic features with II. Among them are the borate H[V] and the nido cluster VI (Fig. 1c).95,96 Furthermore, the molecular structure of the anion [Cl3B–SiCl3] has been determined through single crystal X-ray structure analysis of the salt [(TMS2N)SiCl2–B(η5-C5Me5)][Cl3B–SiCl3] (TMS = Me3Si).97

Results and discussion

The synthesis of B2X4 (I) dates back to 1925, but for decades remained the domain of specialists capable of mastering the technically challenging gas-phase protocols of the time.98–100 A major breakthrough came in 1981, when Nöth et al. obtained B2Br4 in good yields by converting B2(OMe)4 with BBr3 through a convenient solution-phase synthesis.101 In 2017, Braunschweig et al. extended this approach to the other perhalogenated diboranes(4) via solution-phase reactions of B2Br4 with SbF3, GaCl3, and BI3.5

Si2Cl6, a side product of several large-scale processes in the silicon industry,102 is commercially available; quantitative Cl/F exchange with SbF3 affords Si2F6.103 The perbromo- and periododisilanes are accessible from Si2Ph6 by Ph/X exchange with MeC(O)X/AlX3 (X = Br, I).104

Analogous to how B2Br4 and Si2Cl6 grant access to their respective compound classes, the salt [Et4N][I3B–SiI3] ([Et4N][1]; Scheme 1) serves as a key starting material for developing perhalogenated silylboranes. Several years ago, we first reported [Et4N][1], primarily to demonstrate the in situ formation of [SiCl3] as the reactive intermediate in the Si2Cl6/Cl trichlorosilylation system via Lewis-adduct formation with BX3.19 Our study revealed that (excess) BI3 is a more effective trapping reagent than BCl3, because it is the stronger Lewis acid and outcompetes coexisting Si2Cl6 for coordination with [SiCl3], thus suppressing the formation of unwanted oligosilane side products.18 By thoroughly optimizing the original protocol, the yield of [Et4N][1] was increased from ≈45% to ≈70%, and the synthesis was scaled to ≈40 g (Scheme 1). A key improvement is the addition of a second portion of BI3 (0.1 eq.) toward the end of the reaction, following the initial addition of 2 eq. BI3. This prevents contamination of [Et4N][1] with [Et4N][I2ClB–SiI3], previously described as an ‘unknown side product’; its identity has now been unequivocally confirmed by X-ray crystallography (Fig. S102). This finding laid the foundation for a systematic exploration of perhalogenated silylboranes.


image file: d5sc06234a-s1.tif
Scheme 1 Optimized synthesis of [Et4N][1], enabling multigram-scale access to this key starting material. Conversion of the poorly soluble salt [Et4N][1] into moderately soluble neutral donor adducts Do (Py: pyridine; IDipp: 1,3-bis(2,6-diisopropylphenyl)-1,3-dihydro-2H-imidazol-2-ylidene). Formation of the liquid perchlorinated silylborane 3 and its adducts Do. Reaction of 3 with BI3 does not furnish 2, but yields the five-membered ring compound 4, characterized by X-ray crystallography. Reagents and conditions: (i) 1.25 eq. I2, n-heptane, 80 °C, min. 10 h, 48% yield; (ii) 1 eq. [Et4N]Cl, 0.5 eq. Si2Cl6, 0.05 eq. BI3, CH2Cl2, rt, 24 h, 72% yield; (iii) 1.1 eq. Li[Al(OC(CF3)3)4], oDFB, rt, 24 h; (iv) 1.1 eq. Do: SMe2, Py, PPh3, IDipp, CH2Cl2, rt, 24 h, SMe2 = 91%, Py = 83%, PPh3 = 87%, IDipp = 71% yield; (v) Method A: 2.1 eq. GaCl3, 80 °C, 1 h, 96% yield; Method B: 2.1 eq. GaCl3, oDFB, rt, 15 min; (vi) 1.0 eq. Do: SMe2, Py, PPh3, IDipp, oDFB, rt, 15 min, SMe2 = 94%, Py = 95%, PPh3 = 91%, IDipp = 89% yield; (vii) 2 eq. BI3, oDFB, rt, 15 min, 95% yield; (viii) 1 eq. [Et4N]Cl, oDFB, rt, 15 min, 92% yield.

Syntheses of new compounds

The quantitative Si–Cl/Si–I exchange during the formation of [Et4N][1] is advantageous, as iodinated products crystallize more readily in pure form than their chlorinated congeners. However, in combination with the salt-like nature of [Et4N][1], this results in extremely low solubility, posing challenges for subsequent transformations. As initial derivatizations, we consequently replaced one I ligand in [Et4N][1] with neutral donor ligands (Do) bearing solubilizing substituents. To this end, suspensions of [Et4N][1] and Krossing's salt (Li[Al(OC(CF3)3)4])105 in CH2Cl2 were treated with the respective ligand at rt [Do: SMe2, pyridine (Py), PPh3, 1,3-bis(2,6-diisopropylphenyl)-1,3-dihydro-2H-imidazol-2-ylidene (IDipp); Scheme 1]. After filtration, colorless crystals of the corresponding adducts Do readily grew from the filtrate (yields: SMe2 = 91%; Py = 83%; PPh3 = 87%; IDipp = 71%). In stark contrast to BI3, which is sensitive to sunlight, Do exhibit remarkable photostability, with no signs of decomposition upon light exposure.

Cation-exchange using Krossing's salt precipitated LiI as an insoluble byproduct instead of releasing soluble [Et4N]I, thereby driving the quantitative I/Do substitution and facilitating the isolation of pure Do. Most of the second byproduct, [Et4N][Al(OC(CF3)3)4], remained in the mother liquor; residual traces adhering to the crystals of Do were removed by rinsing with ortho-difluorobenzene (oDFB). To characterize the primary [Et4N]+/Li+ cation-exchange product, an equimolar mixture of [Et4N][1] and Li[Al(OC(CF3)3)4] was stirred in oDFB. The resulting solid, which proved insoluble in all common inert solvents, was analyzed by solid-state 11B and 29Si NMR spectroscopy. The data indicated the presence of Li[1] along with free 2 (and already eliminated LiI; see the SI for more details). Given that the insolubility of the free silylborane I2B–SiI3 (2) precluded its isolation and characterization in pure form, we next turned our attention to the synthesis of its perchlorinated congener Cl2B–SiCl3 (3; Scheme 1).

The targeted full I/Cl substitution was straightforwardly achieved by stirring [Et4N][1] with 2 eq. GaCl3—either as a solid mixture that gradually liquefied upon intermittent heating to 80 °C (Method A), or in oDFB (Method B). Neat 3 (Method A) or its calibrated oDFB solution (Method B) was obtained by gas-phase transfer of all volatiles under static vacuum at rt into a glass vessel cooled with liquid N2.106 The colorless donor adducts Do were harvested in crystalline form after stirring 3 and Do in oDFB for 15 min at rt (Scheme 1; yields: SMe2 = 94%; Py = 95%; PPh3 = 91%; IDipp = 89%).

As noted above, the perchlorinated analogue [Et4N][Cl3B–SiCl3] of [Et4N][1] is not accessible through trapping of in situ-generated [SiCl3] with BCl3. With the free silylborane 3 in hand, we have now demonstrated that its reaction with [Et4N]Cl in oDFB affords [Et4N][Cl3B–SiCl3] in >90% yield (Scheme 1). This confirms that the BCl3-based trapping experiment has failed not due to an inherent instability of [Cl3B–SiCl3], but rather because of interfering side reactions that dominate in the mixture BCl3/Si2Cl6/[Et4N]Cl.

As a final approach, we attempted to access pure 2 via Cl/I exchange at 3 using 2 eq. BI3 in oDFB. Instead of the target compound, we obtained the five-membered ring species 4 in good yields (Scheme 1). Its solid-state structure provides valuable insight to rationalize fundamental reactivity patterns of perhalogenated silylboranes (see below).

NMR-spectroscopic, mass-spectrometric, and X-ray-crystallographic characterization of new compounds107

Liquid-phase NMR spectra were recorded at rt in CD2Cl2 and on a sample of neat 3.

The free perchlorinated silylborane 3 gives rise to a singlet at 63.7 ppm in the 11B NMR spectrum and to a 1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 quartet at −8.2 ppm in the 29Si NMR spectrum (1J(B,Si) ≈ 200 Hz; Fig. S24, S25).

Tetracoordinate species typically show 11B NMR signals in the high-field region of the spectrum. The specific chemical shift values are governed by two main factors: (i) the electron density at the 11B nucleus, which reflects the donor strength of the coordinating ligand, and (ii) shielding effects arising from the magnetic anisotropy of the electron shells of the donor atoms, which are especially pronounced for heavier donors.108 To sidestep a comparative evaluation of magnetic anisotropy effects, we restrict our analysis to Py vs. IDipp (2nd-row donor atoms; δ(11B) = −24.8 vs. −37.1) and SMe2 vs. PPh3 (3rd-row donor atoms; δ(11B) = −31.8 vs. −40.6). The observed trends align with the expectation that the more stable adducts are formed with IDipp and PPh3, respectively. The chlorinated compounds Py vs. IDipp (δ(11B) = 3.7 vs. −4.4) and SMe2 vs. PPh3 (δ(11B) = 1.4 vs. −3.4) exhibit the same chemical-shift trends within each pair as observed for the corresponding Do adducts. However, all four signals appear at markedly lower field, which we attribute to decreased magnetic anisotropy shielding when going from the BI2(Do) to the BCl2(Do) fragments. The 29Si NMR resonances of Do and Do were not detectable in the solution spectra, owing to unresolved 1J(B,Si) coupling and broadening induced by the quadrupolar 11B nucleus (S = 3/2).108 The 31P NMR spectra of PPh3 and PPh3 are characterized by multiplet resonances at −7.2 and 2.1 ppm, respectively.

Electron ionization (EI) mass spectra were recorded for the full series of donor adducts Do and Do (introduced as solids). In most cases, we observed the molecular ion peak [(Do)X2B–SiX3+ and/or the peak corresponding to the donor-free silylborane [X2B–SiX3+, typically with low intensity (X = Cl, I; see the SI for details). Most adducts appear to release their neutral Do ligand under the measurement conditions—except for IDipp, which resists elimination. Prominent fragmentation products included [(Do)BSiX4]+/[BSiX4]+ and [(Do)BX2]+. The latter may arise either by [SiX3]˙ loss from the parent ion or via a concerted pathway: [SiX2] extrusion from [(Do)X2B–SiX3+, followed by X˙ elimination from the resulting [(Do)BX3+ intermediate. This, in turn, raises the question—relevant for later reactivity studies—of whether neutral Do and Do might undergo thermal [SiX2] extrusion. To probe this, IDipp was heated with the silylene-trapping reagent 2,3-dimethyl-1,3-butadiene (DMB; 10 eq.) in oDFB at 100 °C for 10 days in a flame-sealed NMR tube. [SiI2] was subsequently identified by GC-MS as its cycloadduct, 1,1-diiodo-3,4-dimethyl-1-silacyclopent-3-ene (Fig. S2).109 Consistently, the reaction mixture showed an 11B NMR signal corresponding to the byproduct BI3·IDipp formed through reductive elimination at the Si(IV) center of IDipp (−77.3 ppm; confirmed by comparison with an authentic sample and X-ray crystal structure analysis of a single crystal grown in the NMR tube).

All eight adducts Do and Do were structurally characterized by X-ray crystallography (Fig. S103–S106 and S109–S112).110 Given the particular relevance of SMe2 to silaboration reactions (see below), the molecular structures of this compound and its perchlorinated congener SMe2 are shown as representative examples in Fig. 2a. All B–Si-bond lengths in Do/Do fall within a narrow range of 2.005(3) to 2.043(4) Å, indicating that this parameter is not significantly influenced by either the nature of Do or the halogen ligand. In contrast, the B–Do bond lengths and the degree of pyramidalization at the B atom in the X2BSi fragments support the a priori expectations that (i) SMe2 is the weakest among the four donors Do, and (ii) the iodinated compound 2 is more Lewis acidic than its chlorinated analogue 3.111


image file: d5sc06234a-f2.tif
Fig. 2 (a) Solid-state structures of SMe2 (β-polymorph; left) and SMe2 (right); (b) solid-state structure of the (S,S)-diastereomer of 4 (left) and its corresponding structural formula, with one BSi and one BSi2 moiety highlighted in red (right). H atoms are omitted for clarity. Color code: B: green, C: black, Si: blue, S: yellow, Cl: yellow-green, I: violet.

Each molecule of 4 contains two chiral B atoms, giving rise to four possible diastereomers (Fig. 2b, left). In the lattice of the examined single crystal, two diastereomers [(S,S)/(S,R)] occupy the same crystallographic site in a 73[thin space (1/6-em)]:[thin space (1/6-em)]27 ratio, which leads to partial disorder involving the B(2)–Si(3) unit. This disorder, together with the comparatively weak scattering contribution of the light B atoms relative to the multiple heavy I atoms, limits the precision with which the B-atom positions and associated structural parameters can be determined. The five-membered ring in 4 features bridging I atoms (B–μ(I)–B and B–μ(I)–Si), resulting in tetracoordinate rather than tricoordinate, electron-deficient B sites (Fig. 2b, left).112 This feature prompts speculation that the extremely insoluble species 2 may adopt a polymeric or dimeric structure in the solid state, possibly featuring B2I2Si five-membered rings, with a single I atom replacing the Si(1)I3 substituent. Moreover, the presence of both a BSi and a BSi2 moiety in 4 (indicated by red-colored bonds in Fig. 2b, right) suggests that our methodology may provide access not only to perhalogenated silylboranes but also to disilylboranes.

Silaboration reactions with Do and Do

One of the primary motivations for developing Do and Do was to create highly reactive silaboration reagents that allow for the simultaneous introduction of both derivatizable functional groups, ideally under catalyst-free conditions. Ethylene was selected as the initial model substrate for two main reasons: (i) its gaseous nature and lack of activating substituents make its silaboration particularly challenging, and (ii) the expected products are highly symmetric molecules with low molecular weight, which facilitates analysis by NMR spectroscopy and mass spectrometry.61,94

The reactions were carried out by heating an excess of ethylene with Do or Do in CD2Cl2 or oDFB in sealed NMR tubes (Table 1 and Scheme 2a). oDFB was used when high temperatures and/or prolonged reaction times posed a risk of I/Cl exchange with CD2Cl2; for NMR measurements, oDFB was replaced with CD2Cl2 prior to measurement.

Table 1 Conditions and product distributions for the reactions of the adducts Do or Do with an excess of ethylene in sealed NMR tubes
Adduct Conditions Product(s)
a After the initial heating period, heating was continued at 140 °C for 1 day and at 160 °C for 1 day.b The reaction mixture contained unconsumed starting material.
SMe2 CD2Cl2, 6 d, 80 °C SMe2, 98%
SMe2/0.1 BI3 CD2Cl2, 12 h, rt SMe2, 85%
Py oDFB, 20 d, 120 °C Py, 97%
PPh3 oDFB, 20 d, 120 °C PPh3, BI3·PPh3b
IDipp oDFB, 6 d, 100 °C BI3·IDippb
SMe2 CD2Cl2, 31 d, 80 °C SMe2, BCl3·SMe2b
Py oDFB, 7 da, 120 °Ca BCl3·Pyb
PPh3 oDFB, 17 d, 120 °C BCl3·PPh3b
IDipp oDFB, 7 da, 120 °Ca BCl3·IDippb



image file: d5sc06234a-s2.tif
Scheme 2 (a) 1,2-Silaborations of ethylene with SMe2 or Py afford SMe2 or Py. (b) 1,1-Silaboration of cyclohexene with [Et4N]1/Li[Al(OC(CF3)3)4] furnishes 7. (c) Reaction of phenylacetylene with SMe2 leads to the addition of the 3/SMe2 Lewis pair across the C[triple bond, length as m-dash]C bond to give 8. (d) Solid-state structures of Py, 7, and 8; H atoms are omitted for clarity. Color code: B: green, C: black, N: pale blue, Si: blue, S: yellow, Cl: yellow-green; I: violet. Reagents and conditions: (i) exc. ethylene, CD2Cl2, 80 °C, 6 d, 98% yield; (ii) exc. ethylene, 0.1 eq. BI3, CD2Cl2, rt, 12 h, 85% yield; (iii) exc. ethylene, oDFB, 120 °C, 20 d, 97% yield; (iv) 10 eq. cyclohexene, 1 eq. Li[Al(OC(CF3)3)4], oDFB, rt, 15 min, 99% yield; (v) 5 eq. phenylacetylene, CH2Cl2, rt, 1 d, 37% yield.

To evaluate general reactivity trends, we employed the pure silylborane adducts without added promoters. Under these conditions, SMe2 underwent quantitative conversion with ethylene to afford the 1,2-silaboration product Me2S·I2B–C2H4–SiI3 (SMe2; Table 1 and Scheme 2a). This transformation proceeded to completion at 80 °C within 6 days. Likewise, Py gave similarly high yields, albeit under even harsher conditions (120 °C, 20 d). At similar temperatures and reaction times, PPh3 was only partially consumed; the fraction that reacted generated both the silaboration product PPh3 and the thermolysis product BI3·PPh3. Among the perchlorinated analogues, only SMe2 produced a notable amount of the corresponding 1,2-silaboration product. However, this transformation took five times longer than the reaction of SMe2 and furnished Me2S·Cl2B–C2H4–SiCl3 (SMe2; Table 1) contaminated with residual starting material and the side product BCl3·SMe2. No silaboration was observed for the other adducts Do and Do; instead, they formed varying amounts of BX3·Do, likely due to thermally induced [SiX2] extrusion, as discussed above and corroborated by our previous trapping experiments with DMB. To promote the reaction between SMe2 and ethylene, BI3 (0.1 eq.) was added to the mixture. Now, silaboration proceeded smoothly at rt within 12 h, affording SMe2 in high yields (85%; entry 2 in Table 1). This final result of our systematic screening thus offers a practical and efficient access route to this promising functionalized building block. Notably, neither the Li[1]/2/LiI mixture nor free 3 provided further improvement, as both led to pronounced side reactions, presumably including ethylene polymerization. We further emphasize that haloboration did not compete with silaboration under any of the tested conditions.

Based on these experimental findings, two key principles emerge to guide further synthetic applications: (i) the highly reactive free species 2 and 3 must be tamed by adduct formation with a suitable donor to prevent non-selective transformations. In this regard, the soft ligand SMe2 performs best in terms of product selectivity while still allowing for reasonable temperatures and reaction times—especially when 0.1 eq. of BI3 is added as a promoter, which likely generates small concentrations of the free Lewis acid 2 in situ. (ii) The iodinated adducts are more effective in silaborations than their chlorinated congeners. From that, we offer the following mechanistic interpretations: (i) silaborations with Do and Do are apparently not initiated by donor-induced B–Si-bond cleavage. Instead, displacement of Do by π-bonded ethylene must precede the 1,2-addition step (comparable B···olefin complexes have been structurally characterized by Yamaguchi et al.113). (ii) To maximize the interaction between the vacant B(pz) orbital and the π-electron cloud of ethylene, competing π-backbonding from X to B must be minimized, which accounts for the superior suitability of X = I (2-type compounds) over Cl (3-type compounds). A comprehensive quantum-chemical assessment of the overall reaction mechanism is provided below.

In a second reactivity test, a mixture of SMe2 and the internal olefin cyclohexene in oDFB was heated to 120 °C for 24 days. Subsequent 11B NMR analysis of the sample showed essentially one signal at −11.6 ppm, indicating quantitative and selective conversion. As such forcing reaction conditions lack practical relevance, efforts were directed toward significantly accelerating the reaction prior to detailed product analysis. In this instance, the addition of BI3 as a promoter did not prove beneficial. However, a successful outcome was ultimately achieved using an equimolar mixture of [Et4N][1] and Li[Al(OC(CF3)3)4] in oDFB, which effected complete conversion within only 15 min at rt. It is evident that the increased kinetic protection of the C[double bond, length as m-dash]C double bond in this case suppresses unwanted side reactions, even in the absence of any donor ligand apart from the residual I ions. More remarkably, olefin internalization exerts a decisive effect on regioselectivity: the reaction with cyclohexene selectively afforded the 1,1-silaboration product 7 rather than the 1,2-isomer (Scheme 2).114 Such a transformation is unprecedented—not only in silaboration but also in the related diboration or disilylation of olefins.115

In a final test experiment, phenylacetylene was chosen as the third representative substrate. Since the iodinated SMe2 led to complex product mixtures, we turned to the chlorinated analogue SMe2, which underwent complete conversion at rt after 1 day. From the reaction mixture, the zwitterionic species 8 crystallized in 37% yield (Scheme 2). Unlike Do and 7 (Do = SMe2, Py), 8 is not generated via silaboration but instead represents the typical outcome of a concerted reaction between a free thioether Lewis base and a free borane Lewis acid acting on the same C[triple bond, length as m-dash]C triple bond.116–122 This finding thus supports our earlier assumption that replacement of the B-bonded donor Do with the unsaturated substrate constitutes the initial step in the reactions of Do and Do. In the case of olefin substrates, both a boryl and a silyl group are introduced into the molecule. Yet, with phenylacetylene, the B–Si bond remains intact, and SMe2 is instead transferred to the substrate. In 8, the B atom is attached to the terminal position of the resulting olefin, while the SMe2 substituent resides near the phenyl ring. This can be explained by the fact that the positive charge accumulated on the carbon framework during electrophilic borylation is better stabilized by resonance at the α-position relative to the phenyl ring.

NMR-spectroscopic and X-ray crystallographic characterization of Do, 7, and 8 (ref. 107)

The 11B NMR spectra of the 1,2-silaboration products, SMe2 and Py, exhibit resonances at −18.9 and −14.0 ppm, respectively, consistent with the presence of tetracoordinate B nuclei.108 In contrast, the 11B NMR signal of the 1,1-silaboration product 7 appears at 53.5 ppm, indicative of a tricoordinate B center.108 The 29Si NMR shifts of SMe2, Py, and 7 are very similar with values of −115.1, −114.8, and −122.8 ppm, respectively. Furthermore, all three compounds give rise to signals exclusively in the aliphatic region of their 1H NMR spectra, confirming complete consumption of the C[double bond, length as m-dash]C double bonds present in the starting materials. The resonances of the axial and equatorial H atoms within the cyclohexyl moiety of compound 7 are distinctly resolved, indicating that bulky substituents on the saturated ring act as effective conformational locks on the NMR timescale.123,124 The 11B NMR spectrum of 8 is characterized by a resonance at −2.6 ppm. As in the cases of Do and Do, the 29Si NMR signal of the B-bonded Si atom is broadened beyond detection. A singlet at δ(1H) = 7.35, together with a corresponding broad resonance at δ(13C) = 163.1, is consistent with the presence of an olefinic fragment in 8.

The molecular structures of SMe2 (Fig. S115), Py, and 7 (Scheme 2d), confirm their proposed identities as 1,2- and 1,1-silaboration products, respectively. The C–C-bond length in Py falls within the typical single-bond range (1.533(4) Å), as do all C–C bonds in 7. As expected,123,124 the bulkier SiI3 substituent occupies an equatorial position, whereas the less bulky BI2 group adopts an axial orientation in the cyclohexane ring of 7. In contrast to compound 4 (Fig. 2), there is no B–μ(I)–Si bridge in 7; rather, the boryl group remains trigonal-planar coordinated. Nonetheless, the vacant B(pz) orbital may acquire some electron density from the occupied Si–C σ orbital, reminiscent of the well-known stabilization of carbenium ions bearing β-positioned silyl groups (see below).125 The C(1)–C(2) distance in compound 8 is 1.335(5) Å, characteristic of a double bond (Scheme 2d). The S and B atoms adopt a mutual E configuration, with the sterically demanding (silyl)boryl substituent located at the terminal position of the styrene core.

Quantum-chemical calculations rationalizing the 1,2- vs. 1,1-silaboration of ethylene vs. cyclohexene to give SMe2 vs. 7

For the reactions of SMe2 and [Et4N][1]/Li[Al(OC(CF3)3)4] with the olefins, two scenarios were examined: 1,2-silaboration and 1,1-silaboration. Potentially competing haloboration pathways126 as well as, for SMe2, the hypothetical addition of the Me2S/I2B–SiI3 Lewis pair to ethylene, were also considered but found to be irrelevant (see the SI for corresponding reaction pathways, activation barriers, and reaction energies). Fig. 3 shows the plausible silaboration sequences for ethylene (a) and cyclohexene (b). As a first important result, the experimentally observed products correspond to pathways that are both kinetically and thermodynamically favored (highlighted in red).
image file: d5sc06234a-f3.tif
Fig. 3 Computed reaction mechanisms for (a) the observed 1,2-silaboration of ethylene with SMe2/0.1 BI3 (red) vs. the not observed 1,1-silaboration pathway (black) and (b) the observed 1,1-silaboration of cyclohexene with 2 (red) vs. the not observed 1,2-silaboration pathway (black). Color code: H: white, B: green, C: grey, Si: blue, I: violet. The Gibbs free energy changes (ΔG) were computed at the SMD(DCM)/PBE0-D3(BJ)/def2-QZVPPD level of theory, using geometries optimized at the SMD(DCM)/PBE0-D3(BJ)/def2-SVPD level. Note: compounds 9–11 appear in the SI as part of theoretically examined but energetically unfavorable alternative mechanisms.

In the reaction of SMe2 with ethylene, the SMe2 donor must first dissociate to generate a vacant coordination site at the B atom for olefin binding. The dissociation requires an energy input of 14.6 kcal mol−1 (cf. Fig. S134). However, the presence of BI3 renders the in situ release of the active silaboration reagent 2 significantly less endergonic (SMe2 + BI32 + BI3·SMe2; ΔG = 4.8 kcal mol−1). Subsequent ethylene binding to free 2 is endergonic by an additional 5.4 kcal mol−1. The resulting intermediate, (C2H4), features a strongly pyramidalized B atom [∑(I–B–I/Si) = 320.2°]; the ethylene ligand remains essentially planar.127 The reaction proceeds via transition state TS1, characterized by B–Si-bond cleavage and the concerted formation of a C–Si bond. The 1,2-silaboration product 5 lies −21.7 kcal mol−1 below the starting materials, with an overall activation barrier of only 15.7 kcal mol−1. In the final step, 5 acquires an SMe2 ligand from SMe2 to afford SMe2 and free 2 with a similar endoergicity as observed in the case of SMe2/BI3, explaining why only minor amounts of BI3 are necessary to promote the silaboration at rt. The alternative 1,1-silaboration of ethylene to furnish 12 would have to proceed via the much higher–energy transition state TS2G = 25.2 kcal mol−1 relative to the starting materials) and is thus not observed.

Due to the modified protocol used for the silaboration of cyclohexene, dissociation of SMe2 is not an issue here. Instead, free 2 can directly interact with the added olefin. Formation of the primary olefin complex (C6H10) is somewhat more endergonic than in the case of ethylene (ΔG = 11.3 vs. 5.4 kcal mol−1), which can be attributed partly to steric factors and partly to a more pronounced reorganization energy: while the B atom in (C6H10) is comparably pyramidalized as in (C2H4), one B-bonded C atom now also deviates significantly from planarity [∑(C–C–C/H) = 348.2°].127 Starting from (C6H10), two subsequent transition states are most relevant: TS1′ leads, via an overall activation barrier of 27.6 kcal mol−1, to the (experimentally unobserved) syn-1,2-silaboration product 13G = −7.5 kcal mol−1). In contrast, TS2′, which lies 12.4 kcal mol−1 lower in energy than TS1′, corresponds to a 1,2-hydride shift leading to intermediate Int1. Subsequent 1,2-silyl migration via the low-lying TS3′ furnishes the experimentally obtained 1,1-silaboration product 7, with an overall reaction energy of ΔG = −10.9 kcal mol−1.

The differing regioselectivities observed in the silaborations of ethylene and cyclohexene arise as early as in intermediates (C2H4) and (C6H10): In (C2H4), the ethylene coordination is near symmetric with B–C distances of 1.837 and 1.868 Å; the C(2)–C(1)–B–Si torsion angle is 23.0°, which represents an ideal conformation for an ensuing 1,2-silyl shift (Fig. 4a, left). According to a Natural Bond Orbital (NBO) analysis,128 all three atoms—B, C(1), and C(2)—carry negative partial charges of −0.52, −0.43, and −0.49 e, respectively. Intermediate (C2H4) can thus be described as a σ-type donor–acceptor complex, in which charge is transferred from the occupied π-orbital of the olefin to the vacant orbital at the B atom, resulting in a substantial interaction energy of –313 kcal mol−1.129 Notably, an Intrinsic Bond Orbital (IBO)130 analysis even suggests the presence of a C–B–C two-electron–three-center (2e–3c) bond, with relative contributions of 29.4% (B), 35.4% (C(1)), and 34.7% (C(2); Fig. 4a, right). In (C6H10), olefin binding is markedly unsymmetric, likely due to the higher steric bulk of cyclohexene relative to ethylene (Fig. 4a, left):131 a short σ bond is found between the B center and the pyramidalized C(1) atom (1.829 Å), while the distance to the still planar C(2) atom is significantly longer (B⋯C = 2.356 Å). Concomitantly, the torsion angle C(2)–C(1)–B–Si is increased to 52.0°, thereby disfavoring a 1,2-silyl shift due to the longer Si⋯C(2) distance that would have to be traversed in the corresponding transition state. While the NBO charges on B and C(1) in (C6H10) remain comparably negative to those in (C2H4), C(2) now carries a positive charge of +0.09e. Cyclohexene can accommodate the steric constraints, as the carbenium ion at C(2) is stabilized by both the +I effect of the alkyl substituent and hyperconjugative interactions132–134 between its vacant pz orbital and the neighboring B–C and C–Hax σ bonds with contributions worth –92.7 and –19.5 kcal mol−1 (ref. 129; Fig. 4a, right; see section 6.3.6 in the SI for a comparison with (C6H10) where an NBO analysis reveals that 3 is coordinated primarily through a conventional, symmetric π→B interaction, most likely reflecting the lower Lewis acidity of 3). Similar hyperconjugative interactions as described for (C6H10) are also present in the rearrangement intermediate Int1—this time between the carbenium ion's pz orbital and the B–Si σ bond or two equivalent C–Hax σ bonds (relative energy contributions: –41.4 and 2 × –29.1 kcal mol−1, respectively; Fig. 4b, right). The former interaction corresponds to the well-known β-effect of a silyl group.125 The overall stabilizing influence of steric and electronic factors makes Int1 thermodynamically more favorable than (C6H10). In summary, the distinct regioselectivities in ethylene and cyclohexene silaboration originate from substrate-dependent binding geometries to 2: symmetric coordination of ethylene facilitates direct 1,2-silaboration, whereas the unsymmetric activation of cyclohexene favors a stepwise 1,1-pathway via a stabilized carbenium-ion intermediate. The computed energy profiles and bonding analyses offer a coherent explanation for the experimentally observed selectivities and underscore the critical influence of steric and electronic substrate effects in directing the specific silaboration pathway.


image file: d5sc06234a-f4.tif
Fig. 4 NBO and IBO analyses rationalizing the divergent silaboration pathways of ethylene and cyclohexene.129 Color code: H: white, B: green, C: grey, Si: blue, I: violet. (a) Left: optimized structures of 2·(C2H4) (top) and 2·(C6H10) (bottom) with selected NBO charges given in elementary charges (e); right: key orbital interactions with associated stabilization energies; top right: IBO representation of the two-electron–three-center (2e–3c) bond in (C2H4). (b) Left: Optimized structure of Int1 with selected NBO charges given in elementary charges (e); right: hyperconjugative interactions stabilizing the carbenium ion, including the β-silicon effect and C–Hax σ donation; SMD(DCM)/PBE0-D3(BJ)/def2-SVPD level of theory.

Conclusions

The addition of a reactant X–Y across a C[double bond, length as m-dash]C double bond is a perfectly atom-economic transformation. When employing versatile (orthogonal) functional groups for X and Y, the primary addition products can be made valuable platforms for a wide range of applications. This is particularly true for X–Y-type reactants featuring covalently bonded boryl and silyl groups: both substituents are not only among the most versatile handles for downstream functionalization, but also play key roles as property-defining units in organic functional materials. Consequently, there is a growing demand for the development of novel silaboration reactions and tailored silylborane reagents R2B–SiR3. We have now found a way to make perhalogenated derivatives (R = Cl, I) readily accessible on a multigram scale—both as free Lewis acids (e.g., Cl2B–SiCl3) and as Lewis base adducts (Do·R2B–SiR3; Do = SMe2, Py, PPh3, IDipp). These developments create a versatile platform with the following key features: (i) Me2S·I2B–SiI3 and the in situ-generated mixture Li[I3B–SiI3]/I2B–SiI3/LiI react directly with olefins in silaboration reactions without the need for a catalyst, which is virtually without precedent.135 (ii) Cyclohexene undergoes a 1,1-addition reaction—so far unobserved not only for silaborations, but also for diboration and disilylation reactions. Combined experimental and quantum-chemical studies revealed that the steric demand of cyclohexene renders symmetrical coordination of the olefin to the B site unfavorable and instead promotes the formation of a zwitterionic B(sp3)–C(sp3)–C(sp2) fragment as a key entry point for the 1,1-silaboration cascade. While such a motif is prohibitively high in energy for ethylene, the carbenium center in the zwitterionic cyclohexene intermediate is efficiently stabilized through a combination of positive inductive (+I) and hyperconjugative effects. (iii) The halide substituents on the introduced boryl and silyl units enable diverse late-stage derivatizations—an aspect of particular importance when these functional groups are not merely used for transmetallation purposes in C–C-coupling reactions, but are instead retained as property-defining elements in the final molecule. (iv) Bulk Cl2B–SiCl3 can be distilled without decomposition. Considering that Si2Cl6 has been successfully used for the gas-phase deposition of silicon thin films,136 and B2F4 for their boron doping,137,138 Cl2B–SiCl3 emerges as a promising single-source precursor for semiconductor fabrication. Taken together, these findings pave the way for the future utilization of perhalogenated silylboranes in both synthesis (i–iii) and materials science (iv).

Author contributions

J. H. performed the experimental studies and characterized all new compounds. C. D. B. performed the quantum-chemical calculations. A. V. V. and E. P. performed the X-ray crystal structure analyses of all compounds. H.-W. L., F. F. and M. W. supervised the project. The manuscript was written by J. H., C. D. B. and M. W. and edited by all co-authors.

Conflicts of interest

There are no conflicts to declare.

Data availability

CCDC 2470869, 2470870, 2470871, 2470872, 2470873, 2470874, 2470875, 2470876, 2470877, 2470878, 2470879, 2470880, 2470881, 2470882, 2470883, 2470884, 2470885, 2470886, 2470887, 2470888, and 2470889, contain the supplementary crystallographic data for this paper.139a–u

The data supporting this article have been included as part of the SI. Supplementary information is available. See DOI: https://doi.org/10.1039/d5sc06234a.

Acknowledgements

We thank the Center for Biomolecular Magnetic Resonance (BMRZ, Goethe University Frankfurt) for the solid-state NMR measurements. J. H. and C. D. B. thank Dr Jannik Gilmer for helpful discussions. We acknowledge the microanalytical laboratory Pascher for the elemental analyses.

Notes and references

  1. S. A. Westcott and E. Fernández, in Adv. Organometal. Chem., Academic Press, Cambridge, 2015, vol. 63, pp. 39–89 Search PubMed.
  2. H. Braunschweig and R. D. Dewhurst, Single, Double, Triple Bonds and Chains: The Formation of Electron-Precise B–B Bonds, Angew. Chem. Int. Ed., 2013, 52, 3574–3583 CrossRef CAS PubMed.
  3. R. D. Dewhurst, E. C. Neeve, H. Braunschweig and T. B. Marder, sp2-sp3 diboranes: astounding structural variability and mild sources of nucleophilic boron for organic synthesis, Chem. Commun., 2015, 51, 9594–9607 RSC.
  4. M. Arrowsmith, H. Braunschweig and T. E. Stennett, Formation and Reactivity of Electron-Precise B-B Single and Multiple Bonds, Angew. Chem., Int. Ed., 2017, 56, 96–115 CrossRef CAS.
  5. M. Arrowsmith, J. Böhnke, H. Braunschweig, A. Deißenberger, R. D. Dewhurst, W. C. Ewing, C. Hörl, J. Mies and J. H. Muessig, Simple solution-phase syntheses of tetrahalodiboranes(4) and their labile dimethylsulfide adducts, Chem. Commun., 2017, 53, 8265–8267 RSC.
  6. J. Teichmann and M. Wagner, Silicon chemistry in zero to three dimensions: from dichlorosilylene to silafullerane, Chem. Commun., 2018, 54, 1397–1412 RSC.
  7. J. A. Morrison, Chemistry of the Polyhedral Boron Halides and the Diboron Tetrahalides, Chem. Rev., 1991, 91, 35–48 CrossRef CAS.
  8. T. B. Marder and N. C. Norman, Transition metal catalysed diboration, Top. Catal., 1998, 5, 63–73 CrossRef CAS.
  9. Y. J. Lin, C. H. Liu, M. G. Chin, C. C. Wang, S. H. Wang, H. Y. Tsai, J. R. Chen, E. Y. Ngai and R. Ramachandran, Characterization of Shock-Sensitive Deposits from the Hydrolysis of Hexachlorodisilane, ACS Omega, 2019, 4, 1416–1424 CrossRef CAS PubMed.
  10. Y. J. Lin, T. T. Nguyen, M. G. Chin, C. C. Wang, C. H. Liu, H. Y. Tsai, J. R. Chen, E. Y. Ngai and R. Ramachandran, Disposal of hexachlorodisilane and its hydrolyzed deposits, J. Loss Prev. Proc. Ind., 2020, 65, 104136 CrossRef CAS.
  11. G. Urry, J. Kerrigan, T. D. Parsons and H. I. Schlesinger, Diboron Tetrachloride, B2Cl4, as a Reagent for the Synthesis of Organo-boron Compounds. I. The Reaction of Diboron Tetrachloride with Ethylene, J. Am. Chem. Soc., 1954, 76, 5299–5301 CrossRef CAS.
  12. P. Ceron, A. Finch, J. Frey, J. Kerrigan, T. Parsons, G. Urry and H. I. Schlesinger, Diboron Tetrachloride and Tetrafluoride as Reagents for the Synthesis of Organoboron Compounds. II. The Behavior of the Diboron Tetrahalides toward Unsaturated Organic Compounds, J. Am. Chem. Soc., 1959, 81, 6368–6371 CrossRef CAS.
  13. T. Wartik and W. B. Fox, Reaction of Diboron Tetrachloride with Aromatic Substances, J. Am. Chem. Soc., 1961, 83, 498–499 CrossRef.
  14. J. Feeney, A. K. Holliday and F. J. Marsden, Diboron Tetrachloride–Olefin Compounds. Part III. The Reaction of Diboron Tetrachloride with Trichloroethylene, Isobutene, and cis- and trans-But-2-ene, J. Chem. Soc., 1961, 356–360 RSC.
  15. M. Zeldin and A. Rosen, The Chemistry of Tetrachlorodiborane(4): I. Reactions with Cyclic Olefins, J. Organomet. Chem., 1971, 31, 319–328 CrossRef CAS.
  16. M. Zeldin and A. Rosen, The Chemistry of Tetrachlorodiborane(4): II. Reactions with Saturated Ring Hydrocarbons, J. Organomet. Chem., 1972, 34, 259–268 CrossRef CAS.
  17. W. Siebert, M. Hildenbrand, P. Hornbach, G. Karger and H. Pritzkow, 1,2- und 1,1-Diborylalkene, Z. Naturforsch. B, 1989, 44, 1179–1186 CrossRef CAS.
  18. J. Tillmann, L. Meyer, J. I. Schweizer, M. Bolte, H.-W. Lerner, M. Wagner and M. C. Holthausen, Chloride-Induced Aufbau of Perchlorinated Cyclohexasilanes from Si2Cl6: A Mechanistic Scenario, Chem. Eur. J., 2014, 20, 9234–9239 CrossRef CAS.
  19. J. Teichmann, M. Bursch, B. Köstler, M. Bolte, H.-W. Lerner, S. Grimme and M. Wagner, Trapping Experiments on a Trichlorosilanide Anion: a Key Intermediate of Halogenosilane Chemistry, Inorg. Chem., 2017, 56, 8683–8688 CrossRef CAS.
  20. E. C. Neeve, S. J. Geier, I. A. I. Mkhalid, S. A. Westcott and T. B. Marder, Diboron(4) Compounds: From Structural Curiosity to Synthetic Workhorse, Chem. Rev., 2016, 116, 9091–9161 CrossRef CAS.
  21. J. Takaya and N. Iwasawa, Catalytic, Direct Synthesis of Bis(boronate) Compounds, ACS Catal., 2012, 2, 1993–2006 CrossRef CAS.
  22. I. A. I. Mkhalid, J. H. Barnard, T. B. Marder, J. M. Murphy and J. F. Hartwig, C–H Activation for the Construction of C–B Bonds, Chem. Rev., 2010, 110, 890–931 CrossRef CAS.
  23. M. Hildenbrand, H. Pritzkow, U. Zenneck and W. Siebert, Synthesis and Structure of a 1,3-Dihydro-1,3-diborete, Angew. Chem. Int. Ed. Engl., 1984, 23, 371–372 CrossRef.
  24. P. Hornbach, M. Hildenbrand, H. Pritzkow and W. Siebert, A Puckered and a Planar 1,3-Diboretane, Angew. Chem. Int. Ed. Engl., 1986, 25, 1112–1114 CrossRef.
  25. H. Braunschweig, R. D. Dewhurst, K. Hammond, J. Mies, K. Radacki and A. Vargas, Ambient-Temperature Isolation of a Compound with a Boron-Boron Triple Bond, Science, 2012, 336, 1420–1422 CrossRef CAS PubMed.
  26. J. Böhnke, H. Braunschweig, W. C. Ewing, C. Hörl, T. Kramer, I. Krummenacher, J. Mies and A. Vargas, Diborabutatriene: An Electron-Deficient Cumulene, Angew. Chem. Int. Ed., 2014, 53, 9082–9085 CrossRef.
  27. J. Böhnke, H. Braunschweig, T. Dellermann, W. C. Ewing, K. Hammond, J. O. C. Jimenez-Halla, T. Kramer and J. Mies, The Synthesis of B2(SIDip)2 and its Reactivity Between the Diboracumulenic and Diborynic Extremes, Angew. Chem. Int. Ed., 2015, 54, 13801–13805 CrossRef PubMed.
  28. H. Braunschweig, S. Demeshko, W. C. Ewing, I. Krummenacher, B. B. Macha, J. D. Mattock, F. Meyer, J. Mies, M. Schäfer and A. Vargas, A Binuclear 1,1’-Bis(boratabenzene) Complex: Unprecedented Intramolecular Metal–Metal Communication through a B–B Bond, Angew. Chem. Int. Ed., 2016, 55, 7708–7711 CrossRef CAS.
  29. W. Lu, Y. Li, R. Ganguly and R. Kinjo, Alkene–Carbene Isomerization induced by Borane: Access to an Asymmetrical Diborene, J. Am. Chem. Soc., 2017, 139, 5047–5050 CrossRef CAS PubMed.
  30. G. Urry, Systematic Synthesis in the Polysilane Series, Acc. Chem. Res., 1970, 3, 306–312 CrossRef CAS.
  31. J. I. Schweizer, M. G. Scheibel, M. Diefenbach, F. Neumeyer, C. Würtele, N. Kulminskaya, R. Linser, N. Auner, S. Schneider and M. C. Holthausen, A Disilene Base Adduct with a Dative Si–Si Single Bond, Angew. Chem. Int. Ed., 2016, 55, 1782–1786 CrossRef CAS PubMed.
  32. C. Kunkel, M. Bolte, H.-W. Lerner, P. Albert and M. Wagner, Subvalent mixed SixGey oligomers: (Cl3Si)4Ge and Cl2(Me2EtN)SiGe(SiCl3)2, Chem. Commun., 2021, 57, 12028–12031 RSC.
  33. I. Georg, J. Teichmann, M. Bursch, J. Tillmann, B. Endeward, M. Bolte, H.-W. Lerner, S. Grimme and M. Wagner, Exhaustively Trichlorosilylated C1 and C2 Building Blocks: Beyond the Müller–Rochow Direct Process, J. Am. Chem. Soc., 2018, 140, 9696–9708 CrossRef CAS.
  34. I. Georg, M. Bursch, J. B. Stückrath, E. Alig, M. Bolte, H.-W. Lerner, S. Grimme and M. Wagner, Building up Strain in One Step: Synthesis of an Edge-Fused Double Silacyclobutene from an Extensively Trichlorosilylated Butadiene Dianion, Angew. Chem. Int. Ed., 2020, 59, 16181–16187 CrossRef CAS PubMed.
  35. I. Georg, M. Bursch, B. Endeward, M. Bolte, H.-W. Lerner, S. Grimme and M. Wagner, The power of trichlorosilylation: isolable trisilylated allyl anions, allyl radicals, and allenyl anions, Chem. Sci., 2021, 12, 12419–12428 RSC.
  36. M. Schmidt, J. Gilmer, A. Virovets, M. Bolte, H.-W. Lerner and M. Wagner, Adjusting the Number of Functional Groups in Vicinal Bis(trichlorosilylated) Benzenes, Chem. Eur. J., 2024, 30, e202402998 CrossRef CAS.
  37. J. Tillmann, M. Moxter, M. Bolte, H.-W. Lerner and M. Wagner, Lewis Acidity of Si6Cl12 and Its Role as Convenient SiCl2 Source, Inorg. Chem., 2015, 54, 9611–9618 CrossRef CAS PubMed.
  38. J. Teichmann, B. Köstler, J. Tillmann, M. Moxter, R. Kupec, M. Bolte, H.-W. Lerner and M. Wagner, Halide-Ion Diadducts of Perhalogenated Cyclopenta- and Cyclohexasilanes, Z. Anorg. Allg. Chem., 2018, 644, 956–962 CrossRef CAS.
  39. J. Teichmann, C. Kunkel, I. Georg, M. Moxter, T. Santowski, M. Bolte, H.-W. Lerner, S. Bade and M. Wagner, Tris(trichlorosilyl)tetrelide Anions and a Comparative Study of Their Donor Qualities, Chem. Eur. J., 2019, 25, 2740–2744 CrossRef CAS.
  40. B. Köstler, H. Bae, J. Gilmer, A. Virovets, H.-W. Lerner, P. Albert, F. Fantuzzi and M. Wagner, Dope it with germanium: selective access to functionalized Si5Ge heterocycles, Chem. Commun., 2023, 59, 716–719 RSC.
  41. B. Köstler, F. Jungwirth, L. Achenbach, M. Sistani, M. Bolte, H.-W. Lerner, P. Albert, M. Wagner and S. Barth, Mixed-Substituted Single-Source Precursors for Si1–xGex Thin Film Deposition, Inorg. Chem., 2022, 61, 17248–17255 CrossRef.
  42. R. Behrle, V. Krause, M. S. Seifner, B. Köstler, K. A. Dick, M. Wagner, M. Sistani and S. Barth, Electrical and Structural Properties of Si1−xGex Nanowires Prepared from a Single-Source Precursor, Nanomaterials, 2023, 13, 627 CrossRef CAS PubMed.
  43. J. Tillmann, J. H. Wender, U. Bahr, M. Bolte, H.-W. Lerner, M. C. Holthausen and M. Wagner, One-Step Synthesis of a [20]Silafullerane with an Endohedral Chloride Ion, Angew. Chem. Int. Ed., 2015, 54, 5429–5433 CrossRef CAS.
  44. B. Köstler, M. Bolte, H.-W. Lerner and M. Wagner, Selective One-Pot Syntheses of Mixed Silicon-Germanium Heteroadamantane Clusters, Chem. Eur. J., 2021, 27, 14401–14404 CrossRef.
  45. M. Bamberg, M. Bursch, A. Hansen, M. Brandl, G. Sentis, L. Kunze, M. Bolte, H.-W. Lerner, S. Grimme and M. Wagner, [Cl@Si20H20]: Parent Siladodecahedrane with Endohedral Chloride Ion, J. Am. Chem. Soc., 2021, 143, 10865–10871 CrossRef CAS.
  46. M. Bamberg, T. Gasevic, M. Bolte, A. Virovets, H.-W. Lerner, S. Grimme, M. Bursch and M. Wagner, Brominated [20]silafulleranes: pushing the limits of steric loading, Chem. Commun., 2023, 59, 7459–7462 RSC.
  47. M. Bamberg, T. Gasevic, M. Bolte, A. Virovets, H.-W. Lerner, S. Grimme, M. Bursch and M. Wagner, Regioselective Derivatization of Silylated [20]Silafulleranes, J. Am. Chem. Soc., 2023, 145, 11440–11448 CrossRef CAS PubMed.
  48. B. Köstler, J. Gilmer, M. Bolte, A. Virovets, H.-W. Lerner, P. Albert, F. Fantuzzi and M. Wagner, Group IV heteroadamantanes: synthesis of Si6Sn4 and site-selective derivatization of Si6Ge4, Chem. Commun., 2023, 59, 2295–2298 RSC.
  49. S. Kühn, B. Köstler, C. True, L. Albers, M. Wagner, T. Müller and C. Marschner, Selective synthesis of germasila-adamantanes through germanium–silicon shift processes, Chem. Sci., 2023, 14, 8956–8961 RSC.
  50. T. Gasevic, M. Bamberg, J. Wicke, M. Bolte, A. Virovets, H.-W. Lerner, S. Grimme, A. Hansen, M. Wagner and M. Bursch, Confined Lewis Pairs: Investigation of the X→Si20 Interaction in Halogen-Encapsulating Silafulleranes, Angew. Chem. Int. Ed., 2024, 63, e202314238 CrossRef CAS PubMed.
  51. F. Raaii and M. S. Gordon, Potential Energy Surfaces for the Bis-Silylation of Ethylene, J. Phys. Chem. A, 1998, 102, 4666–4668 CrossRef CAS.
  52. Y. Alexeev and M. S. Gordon, Theoretical Study of the Bis-Silylation Reaction of Ethylene Catalyzed by Titanium Dichloride, Organometallics, 2003, 22, 4111–4117 CrossRef CAS.
  53. D. G. Hall, Boronic Acids: Preparation and Applications in Organic Synthesis, Medicine and Materials, Wiley-VCH, Weinheim, 2011 Search PubMed.
  54. J. W. B. Fyfe and A. J. B. Watson, Recent Developments in Organoboron Chemistry: Old Dogs, New Tricks, Chem, 2017, 3, 31–55 CAS.
  55. T. Hiyama and M. Oestreich, Organosilicon Chemistry: Novel Approaches and Reactions, Wiley-VCH, Weinheim, 2019 Search PubMed.
  56. J. J. Feng, W. Mao, L. Zhang and M. Oestreich, Activation of the Si–B interelement bond related to catalysis, Chem. Soc. Rev., 2021, 50, 2010–2073 RSC.
  57. M. Oestreich, E. Hartmann and M. Mewald, Activation of the Si–B Interelement Bond: Mechanism, Catalysis, and Synthesis, Chem. Rev., 2013, 113, 402–441 CrossRef CAS.
  58. T. Ohmura and M. Suginome, Silylboranes as New Tools in Organic Synthesis, Bull. Chem. Soc. Jpn., 2009, 82, 29–49 CrossRef CAS.
  59. R. Koyanagi, M. Tanaka, Y. Nonaka, K. Mori, S. Morisako, Y. Yamamoto and A. Kawachi, Preparation and Reactions of (Hydrosilyl)diarylborane, Eur. J. Inorg. Chem., 2025, 28, e202500068 CrossRef CAS.
  60. R. Takahashi, J. Jiang, S. Maeda and H. Ito, Introducing Steric Bulk into Silylboranes: Enhanced Bench Stability and Novel Chemical Reactivity, Angew. Chem. Int. Ed., 2025, e202506194 CAS.
  61. M. Suginome, H. Nakamura and Y. Ito, Platinum-Catalyzed Regioselective Silaboration of Alkenes, Angew. Chem. Int. Ed. Engl., 1997, 36, 2516–2518 CrossRef CAS.
  62. M. Suginome, T. Matsuda, T. Yoshimoto and Y. Ito, Stereoselective 1,4-silaboration of 1,3-dienes catalyzed by nickel complexes, Org. Lett., 1999, 1, 1567–1569 CrossRef CAS.
  63. G. Durieux, M. Gerdin, C. Moberg and A. Jutand, Rate and Mechanism of the Oxidative Addition of a Silylborane to Pt0 Complexes – Mechanism for the Pt-Catalyzed Silaboration of 1,3-Cyclohexadiene, Eur. J. Inorg. Chem., 2008, 4236–4241 CrossRef CAS.
  64. M. Suginome, H. Nakamura and Y. Ito, Regio- and stereo-selective silaboration of alkynes catalysed by palladium and platinum complexes, Chem. Commun., 1996, 2777–2778 RSC.
  65. S.-y. Onozawa, Y. Hatanaka and M. Tanaka, Palladium-catalysed borylsilylation of alkynes and borylsilylative carbocyclization of diynes and an enyne compound, Chem. Commun., 1997, 1229–1230 RSC.
  66. M. Suginome, T. Matsuda and Y. Ito, Nickel-Catalyzed Silaborative Dimerization of Alkynes, Organometallics, 1998, 17, 5233–5235 CrossRef CAS.
  67. M. Suginome, T. Matsuda, H. Nakamura and Y. Ito, Regio-and Stereoselective Synthesis of (Z)-β-Silylalkenylboranes by Silaboration of Alkynes Catalyzed by Palladium and Platinum Complexes, Tetrahedron, 1999, 55, 8787–8800 CrossRef CAS.
  68. T. Sagawa, Y. Asano and F. Ozawa, Synthesis and Reactions of cis-Silyl(boryl)platinum(II) Complexes, Organometallics, 2002, 21, 5879–5886 CrossRef CAS.
  69. T. Ohmura, K. Oshima and M. Suginome, Palladium-catalysed cis- and trans-silaboration of terminal alkynes: complementary access to stereo-defined trisubstituted alkenes, Chem. Commun., 2008, 1416–1418 Search PubMed.
  70. T. Ohmura, K. Oshima, H. Taniguchi and M. Suginome, Switch of Regioselectivity in Palladium-Catalyzed Silaboration of Terminal Alkynes by Ligand-Dependent Control of Reductive Elimination, J. Am. Chem. Soc., 2010, 132, 12194–12196 CrossRef CAS PubMed.
  71. M. B. Ansell, J. Spencer and O. Navarro, (N-Heterocyclic Carbene)2-Pd(0)-Catalyzed Silaboration of Internal and Terminal Alkynes: Scope and Mechanistic Studies, ACS Catal., 2016, 6, 2192–2196 CrossRef CAS.
  72. M. Zhao, C.-C. Shan, Z.-L. Wang, C. Yang, Y. Fu and Y.-H. Xu, Ligand-Dependent-Controlled Copper-Catalyzed Regio- and Stereoselective Silaboration of Alkynes, Org. Lett., 2019, 21, 6016–6020 CrossRef CAS.
  73. H. Ito, Y. Horita and E. Yamamoto, Potassium tert-butoxide-mediated regioselective silaboration of aromatic alkenes, Chem. Commun., 2012, 48, 8006–8008 RSC.
  74. E. Yamamoto, R. Shishido, T. Seki and H. Ito, Tris(trimethylsilyl)silylboronate Esters: Novel Bulky, Air- and Moisture-Stable Silylboronate Ester Reagents for Boryl Substitution and Silaboration Reactions, Organometallics, 2017, 36, 3019–3022 CrossRef CAS.
  75. Y. Gu, Y. Duan, Y. Shen and R. Martin, Stereoselective Base-Catalyzed 1,1-Silaboration of Terminal Alkynes, Angew. Chem. Int. Ed., 2020, 59, 2061–2065 CrossRef CAS PubMed.
  76. K. Nagao, H. Ohmiya and M. Sawamura, Anti-Selective Vicinal Silaboration and Diboration of Alkynoates through Phosphine Organocatalysis, Org. Lett., 2015, 17, 1304–1307 CrossRef CAS PubMed.
  77. K. Oshima, K. Kurotobi, M. Suginome, Y. Takano, T. Umeyama and H. Imahori, Dearomatizing conversion of pyrazines to 1,4-dihydropyrazine derivatives via transition-metal-free diboration, silaboration, and hydroboration, Chem. Commun., 2012, 48, 8571–8573 RSC.
  78. Y. Morimasa, K. Kabasawa, T. Ohmura and M. Suginome, Pyridine-Based Organocatalysts for Regioselective syn-1,2-Silaboration of Terminal Alkynes and Allenes, Asian J. Org. Chem., 2019, 8, 1092–1096 CrossRef CAS.
  79. E. Yamamoto, K. Izumi, Y. Horita and H. Ito, Anomalous Reactivity of Silylborane: Transition-Metal-Free Boryl Substitution of Aryl, Alkenyl, and Alkyl Halides with Silylborane/Alkoxy Base Systems, J. Am. Chem. Soc., 2012, 134, 19997–20000 CrossRef CAS PubMed.
  80. E. Yamamoto, K. Izumi, Y. Horita, S. Ukigai and H. Ito, Formal Nucleophilic Boryl Substitution of Organic Halides with Silylborane/Alkoxy Base System, Top. Catal., 2014, 57, 940–945 CrossRef CAS.
  81. E. Yamamoto, S. Ukigai and H. Ito, Boryl substitution of functionalized aryl-, heteroaryl- and alkenyl halides with silylborane and an alkoxy base: expanded scope and mechanistic studies, Chem. Sci., 2015, 6, 2943–2951 RSC.
  82. R. Uematsu, E. Yamamoto, S. Maeda, H. Ito and T. Taketsugu, Reaction Mechanism of the Anomalous Formal Nucleophilic Borylation of Organic Halides with Silylborane: Combined Theoretical and Experimental Studies, J. Am. Chem. Soc., 2015, 137, 4090–4099 CrossRef CAS.
  83. E. Yamamoto, K. Izumi, R. Shishido, T. Seki, N. Tokodai and H. Ito, Direct Introduction of a Dimesitylboryl Group Using Base-Mediated Substitution of Aryl Halides with Silyldimesitylborane, Chem. Eur. J., 2016, 22, 17547–17551 CrossRef CAS.
  84. E. Yamamoto, S. Maeda, T. Taketsugu and H. Ito, Transition-Metal-Free Boryl Substitution Using Silylboranes and Alkoxy Bases, Synlett, 2017, 28, 1258–1267 CrossRef CAS.
  85. J. M. O'Brien and A. H. Hoveyda, Metal-Free Catalytic C–Si Bond Formation in an Aqueous Medium. Enantioselective NHC-catalyzed Silyl Conjugate Additions to Cyclic and Acyclic α,β-Unsaturated Carbonyls, J. Am. Chem. Soc., 2011, 133, 7712–7715 CrossRef.
  86. C. Kleeberg and C. Borner, On the Reactivity of Silylboranes toward Lewis Bases: Heterolytic B–Si Cleavage vs. Adduct Formation, Eur. J. Inorg. Chem., 2013, 2799–2806 CrossRef CAS.
  87. C. Kleeberg, On the structural diversity of [K(18-crown-6)EPh3] complexes (E = C, Si, Ge, Sn, Pb): Synthesis, crystal structures and NOESY NMR study, Dalton Trans., 2013, 42, 8276–8287 RSC.
  88. J. Plotzitzka and C. Kleeberg, [(NHC)CuI–ER3] Complexes (ER3 = SiMe2Ph, SiPh3, SnMe3): From Linear, Mononuclear Complexes to Polynuclear Complexes with Ultrashort CuI···CuI Distances, Inorg. Chem., 2016, 55, 4813–4823 CrossRef CAS PubMed.
  89. J. Plotzitzka and C. Kleeberg, [(18-C-6)K][(N≡C)CuI–SiMe2Ph], a Potassium Silylcyanocuprate as a Catalyst Model for Silylation Reactions with Silylboranes: Syntheses, Structures, and Catalytic Properties, Inorg. Chem., 2017, 56, 6671–6680 CrossRef CAS PubMed.
  90. P. Gao, G. Wang, L. Xi, M. Wang, S. Li and Z. Shi, Transition-Metal-Free Defluorosilylation of Fluoroalkenes with Silylboronates, Chin. J. Chem., 2019, 37, 1009–1014 CrossRef CAS.
  91. R. Shishido, M. Uesugi, R. Takahashi, T. Mita, T. Ishiyama, K. Kubota and H. Ito, General Synthesis of Trialkyl- and Dialkylarylsilylboranes: Versatile Silicon Nucleophiles in Organic Synthesis, J. Am. Chem. Soc., 2020, 142, 14125–14133 CrossRef CAS PubMed.
  92. A. B. Cuenca, R. Shishido, H. Ito and E. Fernández, Transition-metal-free B–B and B–interelement reactions with organic molecules, Chem. Soc. Rev., 2017, 46, 415–430 RSC.
  93. J. Gilmer, M. Bolte, A. Virovets, H.-W. Lerner, F. Fantuzzi and M. Wagner, A Hydride-Substituted Homoleptic Silylborate: How Similar is it to its Diborane(6)-Dianion Isostere?, Chem. Eur. J., 2023, 29, e202203119 CrossRef CAS PubMed.
  94. J. Gilmer, T. Trageser, L. Čaić, A. Virovets, M. Bolte, H.-W. Lerner, F. Fantuzzi and M. Wagner, Catalyst-free diboration and silaboration of alkenes and alkynes using bis(9-heterofluorenyl)s, Chem. Sci., 2023, 14, 4589–4596 RSC.
  95. S. Bochmann, U. Böhme, E. Brendler, M. Friebel, M. Gerwig, F. Gründler, B. Günther, E. Kroke, R. Lehnert and L. Ruppel, Unexpected Formation of the Highly Symmetric Borate Ion [B(SiCl3)4], Eur. J. Inorg. Chem., 2021, 2583–2594 CrossRef CAS.
  96. P. Greiwe, A. Bethäuser, H. Pritzkow, T. Kühler, P. Jutzi and W. Siebert, Borane-stabilized Boranediyls (Borylenes): Neutral nido-1-Borane-2,3,4,5,6-pentamethyl-2,3,4,5,6-pentacarbahexaboranes(6), Eur. J. Inorg. Chem., 2000, 1927–1929 CrossRef CAS.
  97. N. Sen, N. Parvin, S. Tothadi and S. Khan, Reactivity of (TMS)2N(η1-Cp*)Si[double bond, length as m-dash]Si(η1-Cp*)N(TMS)2 toward the Halides of Groups 13–15, Organometallics, 2021, 40, 1874–1883 CrossRef CAS.
  98. A. Stock, A. Brandt and H. Fischer, Der Zink-Lichtbogen als Reduktionsmittel, Ber. Dtsch. Chem. Ges. B, 1925, 58, 643–657 CrossRef.
  99. T. Wartik, R. Moore and H. I. Schlesinger, Derivatives of Diborine, J. Am. Chem. Soc., 1949, 71, 3265–3266 CrossRef CAS.
  100. G. Urry, T. Wartik, R. E. Moore and H. I. Schlesinger, The Preparation and Some of the Properties of Diboron Tetrachloride, B2Cl4, J. Am. Chem. Soc., 1954, 76, 5293–5298 CrossRef CAS.
  101. H. Nöth and H. Pommerening, Eine einfache Synthese von Dibortetrabromid, Chem. Ber., 1981, 114, 398–399 CrossRef.
  102. X. Zhou, M. A. Wanous, X. Wang, D. V. Eldred and T. L. Sanders, Study on the Shock Sensitivity of the Hydrolysis Products of Hexachlorodisilane, Ind. Eng. Chem. Res., 2018, 57, 10354–10364 CrossRef CAS.
  103. S. Isomura and K. Takeuchi, Preparation of hexafluorodisilane, J. Fluorine Chem., 1997, 83, 89–91 CrossRef CAS.
  104. M. Berger, N. Auner and M. Bolte, Hexabromo- and hexaiododisilane: small and simple molecules showing completely different crystal structures, Acta Crystallogr. Sect. C, 2014, 70, 1088–1091 CrossRef CAS PubMed.
  105. I. Krossing, The Facile Preparation of Weakly Coordinating Anions: Structure and Characterisation of Silverpolyfluoroalkoxyaluminates AgAl(ORF)4, Calculation of the Alkoxide Ion Affinity, Chem. Eur. J., 2001, 7, 490–502 CrossRef CAS.
  106. Alternative syntheses of Cl2B–SiCl3 have been reported, but are of limited practical use due to the demanding apparatus and low yields: (a) A. G. Massey and D. S. Urch, Proc. Chem. Soc., 1964, 273–312 Search PubMed : Cl2B–SiCl3 formed in trace amounts during a mercury discharge of BCl3 vapor, likely via reaction with SiCl4 generated from etching of a quartz discharge tube; (b) P. L. Timms, Inorg. Chem., 1968, 7, 387–389 CrossRef CAS : SiCl2, generated from Si and SiCl4 at high temperature, was cocondensed with BCl3 (1[thin space (1/6-em)]:[thin space (1/6-em)]1) at –196 °C, yielding a blue solid that, upon warming, released Cl2B–SiCl3; (c) R. W. Kirk, D. L. Smith, W. Airey and P. L. Timms, J. Chem. Soc., Dalton Trans., 1972, 13, 1392–1396 RSC : Condensation of SiCl2 with B2Cl4 at –196 °C, followed by warming under reduced pressure, produced a complex, unstable mixture from which only Cl2B–SiCl3 could be isolated and identified; (d) M. Zeldin, D. Solan and B. Dickman, J. Inorg. Nucl. Chem., 1975, 37, 25–28 CrossRef CAS : Using a SiCl4/BCl3 gas mixture under electric discharge, Cl2B–SiCl3 was prepared and isolated in low yield (≈0.2%) by fractional condensation and distillation..
  107. Deposition Numbers 2470869 (for 2·IDipp), 2470870 (for 2·PPh3 (α-)), 2470871 (for 2·PPh3 (β-)), 2470872 (for 2·Py), 2470873 (for 2·SMe2 (α-)), 2470874 (for 2·SMe2 (β-)), 2470875 (for 3·IDipp), 2470876 (for 3·PPh3), 2470877 (for 3·Py), 2470878 (for 3·SMe2), 2470879 (for 4), 2470880 (for 5·SMe2), 2470881 (for 5·Py), 2470882 (for 7 (α-)), 2470883 (for 7 (β-)), 2470884 (for 8), 2470885 (for [Et4N][Cl3B–SiCl3]), 2470886 (for [Et4N][(I2.03/Cl0.97)B–SiI3]), 2470887 (for BI3·IDipp), 2470888 (for BI3·PPh3), and 2470889 (for Me2S·I2B–C2H4–I) contain the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service.
  108. H. Nöth and B. Wrackmeyer, Nuclear Magnetic Resonance Spectroscopy of Boron Compounds, in NMR Basic Principles and Progress, ed. P. Diehl, E. Fluck and R. Kosfeld, Springer Verlag, Berlin, Heidelberg, New York, 1978 Search PubMed.
  109. Replacement of oDFB with CH2Cl2 led to I/Cl exchange, giving rise to CD2ICl, CD2I2, and 1,1-dichloro-3,4-dimethyl-1-silacyclopent-3-ene (GC-MS: Fig. S1), along with IDipp·BCl3 (11B NMR spectroscopy; Fig. S62).
  110. 2·SMe2 is dimorphic. To confirm the phase purity of a freshly prepared sample of 2·SMe2, measured powder diffraction data must be compared with simulated patterns of both polymorphs.
  111. The B–Do bonds within each 2·Do/3·Do pair featuring the same ligand tend to be slightly shorter in 2·Do, e.g., 2·SMe2 (B–S = 1.927[2] Å)a vs. 3·SMe2 (B–S = 1.959(2) Å). Pyramidalization at boron was assessed by comparing the sum of the three bond angles in each SiBX2 fragment to that of three equivalent angles in an ideal tetrahedron (∑ = 328.5°). Almost all B sites in 2·Do/3·Do show equal or even greater pyramidalization [exceptions: β-2·SMe2 (∑ = 330.3(6)°) and 3·SMe2 (∑ = 333.1(4)°)]. Adducts 2·Do are generally more pyramidalized than 3·Do, with the strongest donor IDipp producing the most pronounced effect: 2·IDipp (∑ = 308.6(5)°), 3·IDipp (∑ = 315.1[4]°).b (a) The average value with standard deviations in square brackets was calculated from the following individual bond lengths: B–S = 1.925(6), 1.927(6) (α-polymorph; two crystallographically unique molecules); 1.930(5) Å (β-polymorph); (b) The average value with standard deviations in square brackets was calculated from the bond angles of three crystallographically unique molecules: ∑ = 314.75(54), 314.95(54), 315.65(53)°..
  112. The Si(2)–I bond involving a bridging I atom is clearly longer (2.611(4) Å) than the terminal Si(2)–I bonds (2.428(4)/2.403(4) Å).
  113. R. Oshimizu, N. Ando and S. Yamaguchi, Olefin–Borane Interactions in Donor–π–Acceptor Fluorophores that Undergo Frustrated-Lewis-Pair-Type Reactions, Angew. Chem. Int. Ed., 2022, 61, e202209394 CrossRef CAS.
  114. The presence or absence of SMe2 in the reaction mixture does not appear to be decisive in this context, as the product obtained under SMe2-free conditions has the same NMR signature after addition of the ligand as the compound formed in the long-term reaction between 2·SMe2 and cyclohexene (see the SI for more details).
  115. Alternative synthesis routes to 1,1-diboryl- and 1-boryl-1-silylalkanes do exist and the particular synthetic value of such compounds has been emphasized. Selected examples: (a) H. Li, X. Shangguan, Z. Zhang, S. Huang, Y. Zhang and J. Wang, Org. Lett., 2014, 16, 448–451 CrossRef CAS PubMed; (b) A. Millán, P. D. Grigol Martinez and V. K. Aggarwal, Chem. Eur. J., 2018, 24, 730–735 CrossRef.
  116. J. Guo, M. Yan and D. W. Stephan, Frustrated Lewis pair chemistry of alkynes, Organic Chemistry Frontiers, 2024, 11, 2375–2396 RSC.
  117. M. A. Dureen, C. C. Brown and D. W. Stephan, Deprotonation and Addition Reactions of Frustrated Lewis Pairs with Alkynes, Organometallics, 2010, 29, 6594–6607 CrossRef CAS.
  118. C. A. Tanur and D. W. Stephan, The Thioether–Methyleneborane (PhSCH2B(C6F5)2)2: Synthesis and Reactivity with Donors and Alkynes, Organometallics, 2011, 30, 3652–3657 CrossRef CAS.
  119. A. Fukazawa, E. Yamaguchi, E. Ito, H. Yamada, J. Wang, S. Irle and S. Yamaguchi, Zwitterionic Ladder Stilbenes with Phosphonium and Borate Bridges: Intramolecular Cascade Cyclization and Structure–Photophysical Properties Relationship, Organometallics, 2011, 30, 3870–3879 CrossRef CAS.
  120. C. Eller, G. Kehr, C. G. Daniliuc, R. Fröhlich and G. Erker, Facile 1,1-Carboboration Reactions of Acetylenic Thioethers, Organometallics, 2013, 32, 384–386 CrossRef CAS.
  121. D. J. Faizi, A. J. Davis, F. B. Meany and S. A. Blum, Catalyst-Free Formal Thioboration to Synthesize Borylated Benzothiophenes and Dihydrothiophenes, Angew. Chem. Int. Ed., 2016, 55, 14286–14290 CrossRef CAS.
  122. A. J. Warner, A. Churn, J. S. McGough and M. J. Ingleson, BCl3-Induced Annulative Oxo- and Thioboration for the Formation of C3-Borylated Benzofurans and Benzothiophenes, Angew. Chem. Int. Ed., 2017, 56, 354–358 CrossRef CAS PubMed.
  123. K. B. Wiberg, J. D. Hammer, H. Castejon, W. F. Bailey, E. L. DeLeon and R. M. Jarret, Conformational Studies in the Cyclohexane Series. 1. Experimental and Computational Investigation of Methyl, Ethyl, Isopropyl, and tert-Butylcyclohexanes, J. Org. Chem., 1999, 64, 2085–2095 CrossRef CAS.
  124. H. M. Pickett and H. L. Strauss, Conformational Structure, Energy, and Inversion Rates of Cyclohexane and Some Related Oxanes, J. Am. Chem. Soc., 1970, 92, 7281–7290 CrossRef.
  125. J. B. Lambert, Y. Zhao, R. W. Emblidge, L. A. Salvador, X. Liu, J. H. So and E. C. Chelius, The β Effect of Silicon and Related Manifestations of σ Conjugation, Acc. Chem. Res., 1999, 32, 183–190 CrossRef CAS.
  126. We have confirmed that the model reactant BI3·SMe2 is competent in the iodoboration of ethylene (see the SI for details)..
  127. Compared to the computed C[double bond, length as m-dash]C double-bond lengths of ethylene and cyclohexene, the corresponding distances in the olefin complexes are considerably elongated: (i) 1.332 [C2H4] vs. 1.380 Å [2·(C2H4)]; (ii) 1.340 [C6H10] vs. 1.392 Å [2·(C6H10)].
  128. J. P. Foster and F. Weinhold, Natural Hybrid Orbitals, J. Am. Chem. Soc., 1980, 102, 7211–7218 CrossRef CAS.
  129. The second-order perturbation energy E(2) in Natural Bond Orbital (NBO) analysis quantifies the stabilization arising from donor–acceptor interactions between occupied (donor) and unoccupied (acceptor) NBOs. It reflects the energetic benefit of intramolecular electron delocalization but must not be interpreted as a bond dissociation energy or a direct measure of intrinsic bond strength. Rather, E(2) serves as a qualitative indicator of hyperconjugative interactions and other delocalization effects within the electronic structure. Although E(2) values are conventionally reported as positive, we present them as negative throughout the text and figures to emphasize their stabilizing effect.
  130. G. Knizia, Intrinsic Atomic Orbitals: An Unbiased Bridge Between Quantum Theory and Chemical Concepts, J. Chem. Theory Comput., 2013, 9, 4834–4843 CrossRef CAS.
  131. This assumption is supported, inter alia, by quantum-chemical calculations showing that the cyclohexene in the hypothetical complex 3·(C6H10), with iodine atoms replaced by smaller chlorine atoms, primarily coordinates via symmetric π-donation—as observed in 2·(C2H4); more details are provided in the SI.
  132. H. Cohn, E. D. Hughes, M. H. Jones and M. G. Peeling, Effects of Alkyl Groups in Electrophilic Additions and Substitutions, Nature, 1952, 169, 291 CrossRef CAS.
  133. I. Fernández and G. Frenking, Hyperconjugative Stabilization in Alkyl Carbocations: Direct Estimate of the β-Effect of Group-14 Elements, J. Phys. Chem. A, 2007, 111, 8028–8035 CrossRef.
  134. M. C. Elliott, C. E. Hughes, P. J. Knowles and B. D. Ward, Alkyl groups in organic molecules are NOT inductively electron-releasing, Org. Biomol. Chem., 2025, 23, 352–359 RSC.
  135. The only other example reported to date, which requires a unique type of silylborane, is described in ref. 94.
  136. Evonik Industries, Bau einer Spezialchemie-Anlage für Elektronikchips gestartet, https://publications.evonik.com/de/presse/pressemitteilungen/corporate/bau-einer-spezialchemie-anlage-fuer-elektronikchips-gestartet-105329.html.
  137. O. Byl, E. Jones, J. Sweeney and R. Kaim, Properties of Diboron Tetrafluoride (B2F4), a New Gas for Boron Ion Implantation, AIP Conf. Proc., 2011, 1321, 408–410 CrossRef.
  138. Y. Tang, O. Byl, A. Avila, J. Sweeney, R. Ray, J. Koo, M.-S. Jeon, T. Miller, S. Krause, W. Skinner and J. Mullin, High-efficiency, high-productivity boron doping implantation by diboron tetrafluoride (B2F4) gas on Applied Materials solar ion implanter, 20th International Conference on Ion Implantation Technology (IIT),  DOI:10.1109/IIT.2014.6939984.
  139. (a) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470869, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4g7; (b) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470870, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4h8; (c) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470871, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4j9; (d) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470872, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4kb; (e) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470873, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4lc; (f) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470874, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4md; (g) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470875, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4nf; (h) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470876, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4pg; (i) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470877, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4qh; (j) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470878, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4rj; (k) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470879, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4sk; (l) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470880, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4tl; (m) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470881, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4vm; (n) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470882, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4wn; (o) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470883, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4xp; (p) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470884, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4yq; (q) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470885, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny4zr; (r) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470886, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny50t; (s) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470887, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny51v; (t) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470888, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny52w; (u) J. Heller, C. D. Buch, A. V. Virovets, E. Peresypkina, H.-W. Lerner, F. Fantuzzi and M. Wagner, CCDC 2470889, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny53x.

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.