Next Article in Journal
First-Principles View on Photoelectrochemistry: Water-Splitting as Case Study
Next Article in Special Issue
[Bis(Trimethylsilyl)Methyl]Lithium and -Sodium: Solubility in Alkanes and Complexes with O- and N- Donor Ligands
Previous Article in Journal
Product Selectivity in Homogeneous Artificial Photosynthesis Using [(bpy)Rh(Cp*)X]n+-Based Catalysts
Previous Article in Special Issue
Hetero- and Homoleptic Magnesium Triazenides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Symmetric Assembly of a Sterically Encumbered Allyl Complex: Mechanochemical and Solution Synthesis of the Tris(allyl)beryllate, K[BeA′3] (A′ = 1,3-(SiMe3)2C3H3)

1
Department of Chemistry, Vanderbilt University, Nashville, TN 37235, USA
2
Department of Chemistry, University of Rochester, Rochester, NY 14627, USA
*
Author to whom correspondence should be addressed.
Inorganics 2017, 5(2), 36; https://doi.org/10.3390/inorganics5020036
Submission received: 23 April 2017 / Revised: 22 May 2017 / Accepted: 23 May 2017 / Published: 27 May 2017
(This article belongs to the Special Issue s-Block Metal Complexes)

Abstract

:
The ball milling of beryllium chloride with two equivalents of the potassium salt of bis(1,3-trimethylsilyl)allyl anion, K[A′] (A′ = [1,3-(SiMe3)2C3H3]), produces the tris(allyl)beryllate K[BeA’3] (1) rather than the expected neutral BeA’2. The same product is obtained from reaction in hexanes; in contrast, although a similar reaction conducted in Et2O was previously shown to produce the solvated species BeA’2(OEt2), it can produce 1 if the reaction time is extended (16 h). The tris(allyl)beryllate is fluxional in solution, and displays the strongly downfield 9Be NMR shift expected for a three-coordinate Be center (δ22.8 ppm). A single crystal X-ray structure reveals that the three allyl ligands are bound to beryllium in an arrangement with approximate C3 symmetry (Be–C (avg) = 1.805(10) Å), with the potassium cation engaging in cation–π interactions with the double bonds of the allyl ligands. Similar structures have previously been found in complexes of zinc and tin, i.e., M[M′A′3L] (M′ = Zn, M = Li, Na, K; M′ = Sn, M = K; L = thf). Density functional theory (DFT) calculations indicate that the observed C3-symmetric framework of the isolated anion ([BeA′3]) is 20 kJ·mol−1 higher in energy than a C1 arrangement; the K+ counterion evidently plays a critical role in templating the final conformation.

Graphical Abstract

1. Introduction

The physical and chemical properties of first-row elements often differ appreciably from their second-row and heavier counterparts; for the group 2 metals, the outlier (”black sheep” [1]) designation belongs to beryllium. To a considerably greater extent than its heavier congeners, even magnesium, the small size of the Be2+ cation (0.27 Å for CN = 4; cf. 0.57 Å for Mg2+) [2] and its corresponding high charge/size ratio ensures its bonds will be strongly polarized and possess substantial covalent character. Not surprisingly, beryllium compounds with the same ligand sets commonly have different structures from those of the other, more electropositive alkaline earth (Ae) metals. The bis(trimethylsilyl)amides of Mg–Ba, for example, have a common dimeric bridged structure, [Ae(N(SiMe3)(µ-N(SiMe3)2]2 [3], whereas that of beryllium is a two-coordinate monomer [4]. Similarly, the bis(cyclopentadienyl) complex Cp2Be has an η15-Cp structure [5] that is unlike that of the heavier metallocenes [6]. Investigation of these differences, and indeed research with all beryllium compounds, has traditionally been limited because of concerns about toxicity [7], but that has not prevented its compounds from serving as useful benchmarks of the steric and electronic consequences of crowded metal environments [8,9,10].
One of these consequences is the relative stability of η1- vs. η3-bonded allyl ligands in compounds of highly electropositive metals. We found some time ago that the bulky allyl [A′] (A′ = [1,3-(SiMe3)2C3H3]) can be used to form the ether adduct BeA’2∙OEt2, which displays η1-bonded A’ ligands in the solid state [11]. The compound is fluxional in solution, and exhibits symmetric, “π-type” bonding in its NMR spectra (e.g., only one peak is observed for the SiMe3 groups). Density functional theory (DFT) calculations suggested that a base-free Be(C3H3E2)2 (E = H, SiH3) complex would be more slightly more stable with delocalized, π-type allyls than with monodentate, sigma-bonded ligands (Scheme 1). If so, beryllium allyls would join those of magnesium, in which monodentate allyl ligands are uniformly found in complexes that are ether-solvated [12], but that in the absence of ethers, cation–π interactions with the metal can create “slipped-π” bonding [13].
The coordinated ether in BeA’2∙OEt2 proved impossible to remove without destroying the complex [11], and thus we investigated mechanochemical methods of synthesis as a means to bypass the use of ethereal solvents [14]. As detailed below, an unsolvated neutral complex was not isolated via this route, and the beryllate anion that was produced instead has structural parallels with previously described -ate complexes of Zn [15] and Sn [16]. In all of these species, the alkali metal counterion, usually K+ but sometimes Na+ and Li+, appears to play a critical role in the assembly of the symmetric complexes.

2. Results and Discussion

2.1. Solid-State Synthesis

The reaction of BeCl2 and K[A’] was conducted mechanochemically with a planetary ball mill, followed by an extraction with hexanes. Initial investigations used 2:1 molar ratios of BeCl2 and K[A’], based on the assumption that the product formed would be BeA’2 (Equation (1)). Although the reagents are off-white (K[A’]) and white (BeCl2), the ground reaction mixture (15 min/600 rpm) is orange. Extraction with hexanes, followed by filtration, yielded an orange filtrate and ultimately a dark orange, highly air-sensitive solid (1) on drying.
Inorganics 05 00036 i002
A single crystal analysis (described below) revealed that 1 is the potassium tris(allyl)beryllate, K[BeA’3]. This forms in spite of the fact that the 2:1 ratio of reagents used is not optimum for its production. As detailed below, conducting the reaction with 1:1 and 3:1 molar ratios of K[A’] and BeCl2 still yields 1 as the sole hexane-extractable product. It is possible that the excess halide is captured in the form of polyhalide anions such as [BeCl4]2− or [Be2Cl6]2− [17], although these have not been definitively identified.

2.2. Synthesis in Solution

The reaction of K[A’] and BeCl2 was also examined in solution, using diethyl ether and hexanes. These results are summarized in Table 1. Previous reactions with diethyl ether involved stirring for 2 h at room temperature, which formed BeA’2∙OEt2 from a 2:1 reaction (#5); Schlenk equilibrium was observed in a 1:1 mixture that was allowed to react for one hour (#4). When the 2:1 reaction in Et2O is allowed to proceed for 16 h, however, the formation of 1 is observed (#6) exclusively. Reaction in hexanes mimics the solid-state reactions, in that 1 is the exclusively detected organoberyllium product from a 1:1 reaction after 1 h (#7). Longer reactions and a higher ratio of K[A’] to BeCl2 (e.g., 3:1) do not change this outcome.

2.3. NMR Spectroscopy

The 1H NMR spectrum of 1 displays resonances typical of a “π-bound” A’ ligand, with a triplet representing H(β), a broad resonance (ν1/2 = 39 Hz, presumably an unresolved doublet) representing the equivalent H(α) and H(γ), and a singlet for the two equivalent trimethylsilyl groups (Figure 1). The appearance of such a symmetric spectrum even when σ-bound ligands are expected is consistent with a high degree of fluxionality, as was also observed in the σ-bound complex BeA’2∙(Et2O) [11]. The triplet resonance of the allyl ligands, at δ6.97, is shifted downfield from that of BeA’2∙(Et2O) (δ6.53); the resonance at δ3.19 is slightly upfield (cf. δ3.33 in BeA’2∙(Et2O)). The NMR chemical shifts for 1 are in line with those observed for other M[M’A’3] complexes (Table 2). In particular, the NMR shifts of the allyl ligands are sensitive both to the identity of the central divalent metal and to that of alkali metal counterion, evidence that the compounds exist as contact ion pairs in solution. Compound 1 and K[ZnA’3] share the greatest similarities, which may reflect their having the same counterion (K+) and central metals of similar electronegativity (χ Be (1.57); Zn (1.65)) [18].
John and co-workers have demonstrated that 9Be NMR chemical shift values can be diagnostic for coordination numbers in solution [19]. Typically, organoberyllium complexes with low formal coordination numbers, such as BeMe2∙Et2O (coordination number 3, δ20.8 ppm in Et2O), are observed well downfield of 0 ppm. BeA’2∙(Et2O) has a 9Be chemical shift of δ18.2 ppm, which is consistent with a three-coordinate geometry in solution [11]. It should be noted, however, that the correlation between coordination number and 9Be chemical shift is not exact, and can be strongly influenced by the electronic properties of the ligands. The 2-coordinate complex beryllium bis(N,N´-bis(2,6-diisopropylphenyl)-1,3,2-diazaborolyl), for example, has an extreme downfield shift of δ44 ppm [20], whereas the 2-ccordinate Be(N(SiMe3)2 displays a 9Be NMR shift at δ12.3 ppm [4]. Nevertheless, the 9Be of 1 is at δ22.8 ppm, which to our knowledge is the most positive shift yet reported for a three-coordinate species. DFT methods were used to predict the 9Be chemical shift value of 1 (B3LYP-D3/6-311+G(2d,p)//B3LYP-D3/6-31G(d)). It was calculated at δ25.9 ppm, in reasonable agreement with the observed value (referenced to [Be(OH2)4]2+ with an isotropic shielding constant of 108.98 ppm).

2.4. Solid State Structure

The structure of 1 was determined from single crystal X-ray diffraction. In the solid state, 1 exhibits approximate C3-symmetry, with σ-bound A’ ligands and a potassium cation engaging in cation–π interactions with the three double bonds of the allyls. It is isostructural with the previously reported M[ZnA’3] (M = Li, Na, K) and K(thf)[SnA’3] complexes [15,16]. The beryllium center is in a nearly planar trigonal environment (sum of C–Be–C´ angles = 357.7°) (Figure 2).
The average Be–C distance of 1.805(10) Å has few direct points of comparison with other molecules, as 1 is only the second crystallographically characterized [BeR3] complex, the other being lithium tri-tert-butylberyllate [21]. The latter’s Be center, like that in 1, is in a nearly perfectly planar trigonal environment (sum of C–Be–C angles = 359.9°). In the solid state, however, tri-tert-butylberyllate is a dimer, [Li{Be(t-C4H9)3}]2, with some corresponding distortions in the Be–C bond lengths; Be–C distances range from 1.812(4) Å to 1.864(4) Å, averaging to 1.843(6) Å. The Be–C length in 1 is indistinguishable from the Be–Ccarbene length of 1.807(4) Å in the [Ph2Be(IPr)] (IPr = 1,3-bis(2,6-di-isopropylphenyl)-imidazol-2-ylidene) complex, which also has a three-coordinate Be center [22]. The anionic methyl groups in [Ph2Be(IPr)] are at a noticeably shorter distance, however (1.751(6) Å, ave.). A similar relationship between the Be–Ccarbene and Be–CH3 bond lengths exists in the related [Me2Be(IPr)] [23] and [Me2Be(IMes)] (IMes = N,N’-bis(2,4,6-trimethylphenyl)imidazol-1-ylidene) complexes [1]. A comparison of the Be–C length in 1 could also be made with the Be–C distance of 1.84 Å in lithium tetramethylberyllate, Li2[BeMe4], although the bond distance would be expected to be slightly longer in the latter owing to the higher coordination number of beryllium and the greater negative charge [24].
The C–C and C=C bonds in the alkyl groups in 1 are localized at 1.475(5) Å and 1.353(9) Å, respectively. The K+···C(olefin) contacts average 3.153(7) Å and 2.940(7) Å to the carbon atoms β (C2, C11, and C20) and γ (C3, C12, and C21) to the beryllium atom, respectively. These distances are comparable to, but slightly shorter than, the range of K+···C contacts found in the related zincate structure (3.205(3) Å and 2.945(3) Å, respectively), which reflects the shorter M–C(α) bonds in 1. The distance between Be and K (3.59 Å) is long enough to rule out significant metal-metal interactions.

2.5. Computational Investigations

It has previously been suggested that the occurrence of C3-symmetric M[M’A’3L] (M’ = Zn, M = Li, Na, K; M’ = Sn, M = K; L = thf) complexes is the result of a templating effect of the associated alkali metal counterion [25]. The rationale for this proposal is that the neutral MA’3 (M = As, Sb, Bi) complexes always occur in two diastereomeric forms, with R,R,R (equivalently, S,S,S) and R,R,S (or S,S,R) arrangements of the allyl ligands around the central element. The anionic [MA3´] complexes, in contrast, are always found in the C3-symmetric R,R,R (or S,S,S) configuration, and it is not unreasonable to assume that the counterion is responsible for the difference.
A DFT investigation was undertaken to explore the possible origins of this effect. The geometry of the free [BeA’3] anion was optimized with calculations employing the dispersion-corrected APF-D functional [26]. Three confirmations were examined: the C3-symmetric form (S,S,S) found in the X-ray crystal structure of 1, a related S,S,S form with one A’ ligand rotated antiparallel to the other two (C1 symmetry), and a R,R,S form, also with one ligand antiparallel to the other two, derived from the structure of the neutral AlA’3 complex (Figure 3) [27].
Not surprisingly, the calculated structures possess similar average Be–C bond lengths, ranging from 1.782 Å (the C3-symmetric form (Figure 3a)) to 1.788 Å (for the rotated S,S,S form (Figure 3b)). Energetically, the R,R,S form is the most stable; the rotated S,S,S form is 10.1 kJ·mol−1 higher in energy (∆G°), and the C3-symmetric form is higher still (20.3 kJ·mol−1 in ∆G°). The origin of these energy differences is not immediately obvious, but it may be related to the relative amounts of interligand congestion present. The low energy R,R,S form, for example, has no Me···Me´ contacts less than 4.0 Å, the sum of the van der Waals radii [18]. In contrast, the C3 symmetric form has multiple contacts between methyl groups of less than 4.0 Å, including two as short as 3.76 Å. At this level of theory, the energetics of the free anions do not provide a rationale for the exclusive formation of the S,S,S form.
Not surprisingly, incorporation of the K+ ion into the complex alters the relative stability of the species. The optimized geometries of the C3-symmetric K[BeA’3] found in the X-ray crystal structure of 1 and a related S,S,R form were calculated similarly to the isolated anions, and are depicted in Figure 4.
The C3-symmetric form is 6.1 kJ·mol−1 more stable than the S,S,R form. This is not a consequence of closer K+···(C=C) distances, which are nearly the same (avg. 3.91 Å in the C3 form; 2.86 Å in the S,S,R arrangement). The asymmetric arrangement of the ligands in the S,S,R form does lead to closer interligand C···C contacts in the allyl frameworks, however, as small as 3.37 Å, whereas there are no similar contacts less than 3.78 Å in the C3 form. The somewhat greater stability of the C3 form, possibly coupled with greater ease of crystal packing, may contribute to the exclusive appearance of that form in the crystal structure. It is likely that a similar analysis holds for the isostructural Zn and Sn complexes.
The failure to produce an unsolvated BeA’2 in the absence of a coordinating solvent (i.e., either mechanochemically or in hexanes) was also examined computationally with the aid of the Solid-G program [28]. Both BeA’2∙Et2O and 1 are found to have coordination sphere coverage (Gcomplex) above 90% (i.e., 97.0% (Figure 5a) and 92.6% (Figure 5b), respectively). Although the coverage of the metal center in the hypothetical BeA’2 varies somewhat with the angle between the ligands, the minimum energy position depicted in Figure 5c (C2 symmetry) has only 78.7% coverage. It is not unreasonable to assume that a monomeric BeA’2 may be too coordinately unsaturated to be readily isolable, and will bind an ethereal solvent molecule during synthesis, or, if that is not available, an additional A’ ligand, counterbalanced with a K+ ion.

3. Materials and Methods

General Considerations: All syntheses were conducted under rigorous exclusion of air and moisture using Schlenk line and glovebox techniques. (NOTE: Beryllium salts are toxic and should be handled with appropriate protective equipment.) After grinding was completed, the jars were opened according to glovebox procedures to protect the compounds and to prevent exposure to dust [10]. Proton (1H) and carbon (13C) spectra were obtained on Bruker DRX-500 or DRX-400 spectrometers (Karlsruhe, Germany), and were referenced to residual resonances of C6D6. Beryllium (9Be) spectra were obtained on a Bruker DRX-500 at 70.2 MHz, and were referenced to BeSO4(aq). Combustion analysis was performed by ALS Environmental, Tucson, AZ, USA. Beryllium chloride was purchased from Strem, stored under an N2 atmosphere and used as received. The K[A’] (A’ = 1,3-(SiMe3)2C3H3) reagent was synthesized as previously described [29,30]. Toluene, hexanes, and diethyl ether were distilled under nitrogen from potassium benzophenone ketyl [31]. Deuterated benzene (C6D6) was distilled from Na/K (22/78) alloy prior to use. Stainless steel (440 grade) ball bearings (6 mm,) were thoroughly cleaned with hexanes and acetone prior to use. Planetary milling was performed with a Retsch model PM100 mill (Haan, Germany), 50 mL stainless steel grinding jar type C, and safety clamp for air-sensitive grinding.

3.1. Mechanochemical Synthesis of K[BeA’3] (1)

Solid BeCl2 (56.7 mg, 0.71 mmol) and K[A’] (319 mg, 1.42 mmol) were added to a 50 mL stainless steel grinding jar (type C). The jar was charged with stainless steel ball bearings (6 mm dia, 50 count) and closed tightly with the appropriate safety closer device under an N2 atmosphere. The reagents were milled for 15 min at 600 rpm, resulting in a light orange solid. The product was extracted under an inert atmosphere with minimal hexanes (<100 mL) and filtered through a medium porosity ground glass frit, providing a dark orange filtrate. Drying under vacuum yielded a dark orange solid (61.5 mg, 21% yield of K[BeA’3]) which was recrystallized by the slow evaporation of toluene over one month to provide dark orange-brown crystals of 1 suitable for single crystal X-ray diffraction. For a 3:1 K[A’]:BeCl2 reaction, 812 mg (3.62 mmol) K[A’] and 95.2 mg (1.19 mmol) BeCl2 were added to a grinding jar. After extraction, 183 mg (25% yield) of orange solid was collected. Anal. Calcd. (%) for C27H63BeKSi6: C, 53.65; H, 10.51; Be, 1.49. Found: C, 52.09; H, 9.79; Be, 1.04. The values are somewhat low, possibly from the high air-sensitivity of the compound, but the C:H molar ratio is 2.34:1.00, close to the expected 2.33:1.00. 1H NMR (500 MHz, C6D6, 298K): δ 0.20 (s, 54H, SiMe3); 3.19 (br s (ν1/2 = 39 Hz), 6H, H(α,γ)); 6.97 (t, 3H, J1 = 16 Hz, H(β)). 13C NMR (100 MHz, C6D6, 298K): δ1.02 (s, SiMe3); 70.71 (s, C(α,γ)); 166.09 (s, C(β)). 9Be NMR (70.2 MHz, C6D6, 298K); δ22.8 (s) (ν1/2 = 360 Hz).

3.2. General Procedures for Synthesis of K[BeA’3] (1) with Solvents

Reactions were performed for either 1 or 16 h, and were run under inert atmosphere at room temperature. The ratio of K[A’] and BeCl2 was varied such that the reactions of emphasis were 1:1, 2:1, and 3:1. A general reaction involved dissolving the beryllium chloride (ca. 0.1 g) in the solvent of choice (Et2O or hexanes); to this solution solid K[A’] was added slowly and solvent was used to quantitatively transfer all material. Upon mixing, the solution was allowed to stir for the given time. In the case of Et2O, the solvent was removed in vacuo, and the resulting material was extracted with hexanes, filtered through a medium porosity glass frit, and then dried in vacuo. In the case of reaction in hexanes, the reaction mixture was filtered through a medium porosity fritted glass filter, and the hexane was removed in vacuo. The resulting material in all cases was then analyzed with 1H and 9Be NMR.

3.3. Procedures for X-ray Crystallography

A crystal (0.20 × 0.20 × 0.08 mm3) was placed onto the tip of a thin glass optical fiber and mounted on a Bruker SMART APEX II CCD platform diffractometer (Karlsruhe, Germany) for a data collection at 100.0(5) K [32]. The structure was solved using SIR2011 [33] and refined using SHELXL-2014/7 [34]. The space group P1 was determined based on intensity statistics. A direct-methods solution was calculated that provided most non-hydrogen atoms from the E-map. Full-matrix least squares/difference Fourier cycles were performed which located the remaining non-hydrogen atoms. All non-hydrogen atoms were refined with anisotropic displacement parameters. The allylic hydrogen atoms were found from the difference Fourier map and refined freely. All other hydrogen atoms were placed in ideal positions and refined as riding atoms with relative isotropic displacement parameters.

3.4. General Procedures for Calculations

All calculations were performed with the Gaussian 09W suite of programs [35]; an ultrafine grid was used for all cases (Gaussian keyword: int = ultrafine). Each conformation of the [BeA’3] complexes was studied with the APF-D functional, a global hybrid with 23% exact exchange [26]. The 6-31+G(d) basis set was used for C,H,Si; the 6-311+G(2d) basis was used for Be. For the neutral K[BeA’3] conformations, the APF-D functional was used with the 6-31G(d) basis set for C,H,Si; 6-311G(2d) was used for Be and K. The nature of the stationary points was determined with analytical frequency calculations; all of these optimized geometries were found to be minima (Nimag = 0). For the Solid-G calculations, the structures were preoptimized with the APF-D/6-311G(2d) (Be,K); 6-31G(d) (C,H,Si) protocol.

4. Conclusions

The generation of products from reagents that are not in the optimum stoichiometric ratio is a known feature of some Group 2 reactions [36,37], a testament to the role that kinetic factors play in s-block chemistry. It is perhaps not surprising that when mechanochemical activation is used with alkaline earth reagents, a nonstoichiometric product such as the organoberyllate 1 is formed, as grinding and milling environments are often far from equilibrium [38,39,40,41]. However, the fact that 1 is also generated in hexanes indicates how non-ethereal synthesis can reveal features of reactions that are obscured when they are conducted in coordinating solvents. It is now apparent that the production of the previously described BeA’2∙Et2O, which was the expected complex from a 2:1 reaction of K[A’] and BeCl2 in diethyl ether [11], actually depends critically on the presence of the solvent to prevent further reaction of the beryllium center with an additional A’ ligand. (In a preliminary study, the reaction of K[A′] and BeCl2 in a 2:1 molar ratio in THF (1 h) was found not to produce K[BeA′3]. A species with a 9Be NMR shift of δ16.6 ppm was present instead, tentatively identified as BeA′2(thf). If correct, this indicates that THF, like Et2O, can block the formation of the tris(allyl) anion with Be). Without such ethereal solvent support, whether conducted mechanochemically or in hexanes, the reaction between K[A’] and BeCl2 rapidly forms the kinetic product 1.
Parallels of the beryllium chemistry to the related tris(allyl) -ate complexes of Zn and Sn are instructive, although they cannot be pushed too far. All the [MA3´] species possess approximate C3 symmetry, and it is likely that the associated alkali metal cation is intimately involved in templating their constructions. The formation of the zinc species K[ZnA’3] is also similar to that of 1 in that it is formed from the reaction of 2 equiv. of K[A’] and ZnCl2, i.e., in a non-stoichiometric reaction [15]. However, both it and K(thf)[SnA’3] are synthesized in THF, so it is clear that the driving force for -ate formation compared to that for the neutral (Zn,Sn)A’2 species is greater than that for 1. This may reflect the somewhat lesser covalency of Be–C versus Zn–C and Sn–C bonds, and the greater robustness of M2+⟵:OR2 interactions with beryllium.

Supplementary Materials

The following are available online at https://www.mdpi.com/2304-6740/5/2/36/s1: CIF and checkCIF file, and fractional coordinates of geometry-optimized structures.

Acknowledgments

Financial support by the National Science Foundation (CHE-1112181), The American Chemical Society–Petroleum Research Fund (56027-ND3), and a Discovery Grant of Vanderbilt University is gratefully acknowledged.

Author Contributions

The project concept was devised equally by all authors. Nicholas C. Boyde and Nicholas R. Rightmire performed the experiments and analyzed the resulting data. William W. Brennessel obtained and solved the single crystal X-ray data. The manuscript was written by Timothy P. Hanusa, with contributions from all the co-authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Arrowsmith, M.; Hill, M.S.; Kociok-Köhn, G. Activation of N-Heterocyclic Carbenes by {BeH2} and {Be(H)(Me)} Fragments. Organometallics 2015, 34, 653–662. [Google Scholar] [CrossRef]
  2. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  3. Westerhausen, M. Synthesis and Spectroscopic Properties of Bis(trimethylsilyl)amides of the Alkaline-earth Metals Magnesium, Calcium, Strontium, and Barium. Inorg. Chem. 1991, 30, 96–101. [Google Scholar] [CrossRef]
  4. Naglav, D.; Neumann, A.; Blaser, D.; Wolper, C.; Haack, R.; Jansen, G.; Schulz, S. Bonding situation in Be[N(SiMe3)2]2—An experimental and computational study. Chem. Commun. 2015, 51, 3889–3891. [Google Scholar] [CrossRef] [PubMed]
  5. Nugent, K.W.; Beattie, J.K.; Hambley, T.W.; Snow, M.R. A precise Low-temperature crystal structure of bis(cyclopentadienyl)beryllium. Aust. J. Chem. 1984, 37, 1601–1606. [Google Scholar] [CrossRef]
  6. Burkey, D.J.; Hanusa, T.P. Structural Lessons from Main-Group Metallocenes. Comments Inorg. Chem. 1995, 17, 41–77. [Google Scholar] [CrossRef]
  7. Civic, T.M. Beryllium Metal Toxicology: A Current Perspective. In Encyclopedia of Inorganic and Bioinorganic Chemistry; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2011. [Google Scholar]
  8. Hanusa, T.P.; Bierschenk, E.J.; Engerer, L.K.; Martin, K.A.; Rightmire, N.R. 1.37—Alkaline Earth Chemistry: Synthesis and Structures. In Comprehensive Inorganic Chemistry II, 2nd ed.; Reedijk, J., Poeppelmeier, K., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 1133–1187. [Google Scholar]
  9. Iversen, K.J.; Couchman, S.A.; Wilson, D.J.D.; Dutton, J.L. Modern organometallic and coordination chemistry of beryllium. Coord. Chem. Rev. 2015, 297–298, 40–48. [Google Scholar] [CrossRef]
  10. Naglav, D.; Buchner, M.R.; Bendt, G.; Kraus, F.; Schulz, S. Off the Beaten Track—A Hitchhiker's Guide to Beryllium Chemistry. Angew. Chem. Int. Ed. 2016, 55, 10562–10576. [Google Scholar] [CrossRef] [PubMed]
  11. Chmely, S.C.; Hanusa, T.P.; Brennessel, W.W. Bis(1,3-trimethylsilylallyl)beryllium. Angew. Chem. Int. Ed. 2010, 49, 5870–5874. [Google Scholar] [CrossRef] [PubMed]
  12. Solomon, S.A.; Muryn, C.A.; Layfield, R.A. s-Block metal complexes of a bulky, donor-functionalized allyl ligand. Chem. Commun. 2008, 3142–3144. [Google Scholar] [CrossRef] [PubMed]
  13. Chmely, S.C.; Carlson, C.N.; Hanusa, T.P.; Rheingold, A.L. Classical versus Bridged Allyl Ligands in Magnesium Complexes: The Role of Solvent. J. Am. Chem. Soc. 2009, 131, 6344–6345. [Google Scholar] [CrossRef] [PubMed]
  14. Rightmire, N.R.; Hanusa, T.P. Advances in Organometallic Synthesis with Mechanochemical Methods. Dalton Trans. 2016, 45, 2352–2362. [Google Scholar] [CrossRef] [PubMed]
  15. Gren, C.K.; Hanusa, T.P.; Rheingold, A.L. Threefold Cation–π Bonding in Trimethylsilylated Allyl Complexes. Organometallics 2007, 26, 1643–1649. [Google Scholar] [CrossRef]
  16. Layfield, R.A.; Garcia, F.; Hannauer, J.; Humphrey, S.M. Ansa-tris(allyl) complexes of alkali metals: Tripodal analogues of cyclopentadienyl and ansa-metallocene ligands. Chem. Commun. 2007, 5081–5083. [Google Scholar] [CrossRef] [PubMed]
  17. Neumüller, B.; Weller, F.; Dehnicke, K. Synthese, Schwingungsspektren und Kristallstrukturen der Chloroberyllate (Ph4P)2[BeCl4] und (Ph4P)2[Be2Cl6]. Z. Anorg. Allg. Chem. 2003, 629, 2195–2199. [Google Scholar] [CrossRef]
  18. Allred, A.L. Electronegativity values from thermochemical data. J. Inorg. Nucl. Chem. 1961, 17, 215–221. [Google Scholar] [CrossRef]
  19. Plieger, P.G.; John, K.D.; Keizer, T.S.; McCleskey, T.M.; Burrell, A.K.; Martin, R.L. Predicting 9Be Nuclear Magnetic Resonance Chemical Shielding Tensors Utilizing Density Functional Theory. J. Am. Chem. Soc. 2004, 126, 14651–14658. [Google Scholar] [CrossRef] [PubMed]
  20. Arnold, T.; Braunschweig, H.; Ewing, W.C.; Kramer, T.; Mies, J.; Schuster, J.K. Beryllium bis(diazaborolyl): Old neighbors finally shake hands. Chem. Commun. 2015, 51, 737–740. [Google Scholar] [CrossRef] [PubMed]
  21. Wermer, J.R.; Gaines, D.F.; Harris, H.A. Synthesis and molecular structure of lithium tri-tert-butylberyllate, Li[Be(tert-C4H9)3]. Organometallics 1988, 7, 2421–2422. [Google Scholar] [CrossRef]
  22. Gottfriedsen, J.; Blaurock, S. The First Carbene Complex of a Diorganoberyllium: Synthesis and Structural Characterization of Ph2Be(i-Pr-carbene) and Ph2Be(n-Bu2O). Organometallics 2006, 25, 3784–3786. [Google Scholar] [CrossRef]
  23. Arrowsmith, M.; Hill, M.S.; Kociok-Köhn, G.; MacDougall, D.J.; Mahon, M.F. Beryllium-Induced C–N Bond Activation and Ring Opening of an N-Heterocyclic Carbene. Angew. Chem. Inter. Ed. 2012, 51, 2098–2100. [Google Scholar] [CrossRef] [PubMed]
  24. Weiss, E.; Wolfrum, R. Über metall–alkyl-verbindungen VIII. Die kristallstruktur des lithium-tetramethylberyllats. J. Organomet. Chem. 1968, 12, 257–262. [Google Scholar] [CrossRef]
  25. Rightmire, N.R.; Bruns, D.L.; Hanusa, T.P.; Brennessel, W.W. Mechanochemical Influence on the Stereoselectivity of Halide Metathesis: Synthesis of Group 15 Tris(allyl) Complexes. Organometallics 2016, 35, 1698–1706. [Google Scholar] [CrossRef]
  26. Austin, A.; Petersson, G.A.; Frisch, M.J.; Dobek, F.J.; Scalmani, G.; Throssell, K. A Density Functional with Spherical Atom Dispersion Terms. J. Chem. Theory Comp. 2012, 8, 4989–5007. [Google Scholar] [CrossRef] [PubMed]
  27. Rightmire, N.R.; Hanusa, T.P.; Rheingold, A.L. Mechanochemical Synthesis of [1,3-(SiMe3)2C3H3]3(Al,Sc), a Base-Free Tris(allyl)aluminum Complex and Its Scandium Analogue. Organometallics 2014, 33, 5952–5955. [Google Scholar] [CrossRef]
  28. Guzei, I.A.; Wendt, M. An improved method for the computation of ligand steric effects based on solid angles. Dalton Trans. 2006, 3991–3999. [Google Scholar] [CrossRef] [PubMed]
  29. Fraenkel, G.; Chow, A.; Winchester, W.R. Dynamics of solvated lithium(+) within exo,exo-[1,3-bis(trimethylsilyl)allyl]lithium N,N,N′,N′-tetramethylethylenediamine complex. J. Am. Chem. Soc. 1990, 112, 1382–1386. [Google Scholar] [CrossRef]
  30. Harvey, M.J.; Hanusa, T.P.; Young, V.G., Jr. Synthesis and Crystal Structure of a Bis(allyl) Complex of Calcium, [Ca{C3(SiMe3)2H3}2·(thf)2]. Angew. Chem. Int. Ed. 1999, 38, 217–219. [Google Scholar] [CrossRef]
  31. Perrin, D.D.; Armarego, W.L.F. Purification of Laboratory Chemicals, 3rd ed.; Pergamon: Oxford, UK, 1988. [Google Scholar]
  32. APEX2, Version 2014.11–0; Bruker AXS: Madison, WI, USA, 2013.
  33. Burla, M.C.; Caliandro, R.; Camalli, M.; Carrozzini, B.; Cascarano, G.L.; Giacovazzo, C.; Mallamo, M.; Mazzone, A.; Polidori, G.; Spagna, R. SIR2011: A New Package for Crystal Structure Determination and Refinement, Version 1.0; Istituto di Cristallografia: Bari, Italy, 2012.
  34. Sheldrick, G.M. SHELXL-2014/7, University of Göttingen: Göttingen, Germany, 2014.
  35. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09W, Revision D.01; Gaussian, Inc.: Wallingford CT, USA, 2009.
  36. Glock, C.; Görls, H.; Westerhausen, M. N,N,N′,N′-Tetramethylethylendiamine adducts of amido calcium bases—Synthesis of monomeric [(tmeda)Ca{N(SiMe3)2}2], [(tmeda)Ca{NiPr2}2], and dimeric Hauser base-type [(tmeda)Ca(tmp)(μ-I)]2 (tmp = 2,2,6,6-tetramethylpiperidide). Inorg. Chim. Acta 2011, 374, 429–434. [Google Scholar] [CrossRef]
  37. Fitts, L.S.; Bierschenk, E.J.; Hanusa, T.P.; Rheingold, A.L.; Pink, M.; Young, V.G. Selective modification of the metal coordination environment in heavy alkaline-earth iodide complexes. New J. Chem. 2016, 40, 8229–8238. [Google Scholar] [CrossRef]
  38. Hernández, J.G.; Friščić, T. Metal-catalyzed organic reactions using mechanochemistry. Tetrahedron Lett. 2015, 56, 4253–4265. [Google Scholar] [CrossRef]
  39. Boldyreva, E. Mechanochemistry of inorganic and organic systems: What is similar, what is different? Chem. Soc. Rev. 2013, 42, 7719–7738. [Google Scholar] [CrossRef] [PubMed]
  40. Wang, G.-W. Mechanochemical organic synthesis. Chem. Soc. Rev. 2013, 42, 7668–7700. [Google Scholar] [CrossRef] [PubMed]
  41. Waddell, D.C.; Thiel, I.; Clark, T.D.; Marcum, S.T.; Mack, J. Making kinetic and thermodynamic enolates via solvent-free high speed ball milling. Green Chem. 2010, 12, 209–211. [Google Scholar] [CrossRef]
Scheme 1. Optimized geometries of Be(1,3-(SiH3)2C3H3)2. At the B3PW91/aug-cc-pVDZ level, the π-bound structure (a) is 4.0 kcal·mol−1 lower in energy (∆G°) than the σ-bound geometry (b) [11].
Scheme 1. Optimized geometries of Be(1,3-(SiH3)2C3H3)2. At the B3PW91/aug-cc-pVDZ level, the π-bound structure (a) is 4.0 kcal·mol−1 lower in energy (∆G°) than the σ-bound geometry (b) [11].
Inorganics 05 00036 sch001
Figure 1. 1H NMR spectrum (500 Mhz) of isolated 1, recorded in C6D6. The starred peaks represent impurities: δ7.15 (residual protons of C6D6); δ0.9 and δ1.3, residual hexanes.
Figure 1. 1H NMR spectrum (500 Mhz) of isolated 1, recorded in C6D6. The starred peaks represent impurities: δ7.15 (residual protons of C6D6); δ0.9 and δ1.3, residual hexanes.
Inorganics 05 00036 g001
Figure 2. Thermal ellipsoid plot of 1, illustrating the numbering scheme used in the text. Ellipsoids are drawn at the 50% level, and for clarity, hydrogen atoms have been removed from the trimethylsilyl groups. Selected bond distances (Å) and angles (deg): Be1–C1, 1.795(6); Be1–C10, 1.810(6); Be1–C19, 1.811(6); C(2)–C(3), 1.351(5); C(11)–C(12), 1.350(5); C(20)–C(21), 1.358(5); K(1)–C(2), 3.138(4); K(1)–C(3), 2.940(4); K(1)–C(11), 3.206(4); K(1)–C(12), 2.943(4); K(1)–C(20), 3.114(4); K(1)–C(21), 2.938(4); C(1)–Be(1)–C(10), 119.1(3); C(1)–Be(1)–C(19), 119.0(3); C(10)–Be(1)–C(19), 119.4(3).
Figure 2. Thermal ellipsoid plot of 1, illustrating the numbering scheme used in the text. Ellipsoids are drawn at the 50% level, and for clarity, hydrogen atoms have been removed from the trimethylsilyl groups. Selected bond distances (Å) and angles (deg): Be1–C1, 1.795(6); Be1–C10, 1.810(6); Be1–C19, 1.811(6); C(2)–C(3), 1.351(5); C(11)–C(12), 1.350(5); C(20)–C(21), 1.358(5); K(1)–C(2), 3.138(4); K(1)–C(3), 2.940(4); K(1)–C(11), 3.206(4); K(1)–C(12), 2.943(4); K(1)–C(20), 3.114(4); K(1)–C(21), 2.938(4); C(1)–Be(1)–C(10), 119.1(3); C(1)–Be(1)–C(19), 119.0(3); C(10)–Be(1)–C(19), 119.4(3).
Inorganics 05 00036 g002
Figure 3. Geometry optimized structures of [BeA’3] anions: (a) as found in the crystal structure of 1; (b) related S,S,S form with one A’ ligand rotated (C1 symmetry); and (c) R,R,S form derived from the structure of AlA’3.
Figure 3. Geometry optimized structures of [BeA’3] anions: (a) as found in the crystal structure of 1; (b) related S,S,S form with one A’ ligand rotated (C1 symmetry); and (c) R,R,S form derived from the structure of AlA’3.
Inorganics 05 00036 g003
Figure 4. Geometry optimized structures of the K[BeA’3] complex: (a) as found in the crystal structure of 1; and (b) related S,S,R form.
Figure 4. Geometry optimized structures of the K[BeA’3] complex: (a) as found in the crystal structure of 1; and (b) related S,S,R form.
Inorganics 05 00036 g004
Figure 5. Visualization of the extent of coordination sphere coverage (Gcomplex) of: (a) BeA’2∙Et2O (the coverage from the two allyls are assigned blue and green; that from the ether is in red); (b) 1 (all three allyls are in blue); and (c) BeA’2, using optimized coordinates (APF-D/6-311G(2d) (Be); 6-31G(d) (other atoms)) and the program Solid-G [28]. The Gcomplex value takes into account the net coverage; regions of the coordination sphere where the projections of the ligands overlap are counted only once.
Figure 5. Visualization of the extent of coordination sphere coverage (Gcomplex) of: (a) BeA’2∙Et2O (the coverage from the two allyls are assigned blue and green; that from the ether is in red); (b) 1 (all three allyls are in blue); and (c) BeA’2, using optimized coordinates (APF-D/6-311G(2d) (Be); 6-31G(d) (other atoms)) and the program Solid-G [28]. The Gcomplex value takes into account the net coverage; regions of the coordination sphere where the projections of the ligands overlap are counted only once.
Inorganics 05 00036 g005
Table 1. Summary of K[A’] and BeCl2 reactions; amounts of reagents given as molar ratios.
Table 1. Summary of K[A’] and BeCl2 reactions; amounts of reagents given as molar ratios.
No.K[A’]:BeCl2Medium aTimeOrganoberyllium Product(s)Yield (%) b
11:1 Inorganics 05 00036 i00115 minK[BeA’3]97
22:1 Inorganics 05 00036 i00115 minK[BeA’3]21
33:1 Inorganics 05 00036 i00115 minK[BeA’3]25
41:1Et2O1 h2A’BeCl ⇄ BeA’2 + BeCln/a c,d
52:1Et2O2 hBeA’2∙OEt277 c
62:1Et2O16 hK[BeA’3]98
71:1hexanes1 hK[BeA’3]24
a Inorganics 05 00036 i001 = ball milling at 600 rpm. The symbol for mechanical milling has been proposed in ref. [14]; b Unrecrystallized; limiting reagent taken into account; c Ref. [11]; d Products were observed with 9Be NMR, and were not isolated.
Table 2. 1H NMR shifts (ppm) and bond distances in M[M’A’3] complexes.
Table 2. 1H NMR shifts (ppm) and bond distances in M[M’A’3] complexes.
Complexδ H(α)/H(γ)δ H(β)δ SiMe3M–C (σ) ÅM’···C(olefin) ÅRef.
Li[ZnA’3]6.463.500.152.117(3) b2.745(4), 2.268(3) b[15]
Na[ZnA’3]7.594.000.162.103(3)2.857(3), 2.567(3)[15]
K[BeA’3]6.973.190.201.805(10)3.153(7), 2.940(7)this work
K[ZnA’3]7.053.420.232.068(4)3.205(3), 2.945(3)[15]
K(thf)[SnA’3]6.434.420.42, 0.23 a2.344(7)3.201(7), 3.164(8), 3.065(8)[16]
a Two resonances are observed for the SiMe3 groups, as the A’ ligands are not fluxional; b Distance(s) affected by crystallographic disorder.

Share and Cite

MDPI and ACS Style

Boyde, N.C.; Rightmire, N.R.; Hanusa, T.P.; Brennessel, W.W. Symmetric Assembly of a Sterically Encumbered Allyl Complex: Mechanochemical and Solution Synthesis of the Tris(allyl)beryllate, K[BeA′3] (A′ = 1,3-(SiMe3)2C3H3). Inorganics 2017, 5, 36. https://doi.org/10.3390/inorganics5020036

AMA Style

Boyde NC, Rightmire NR, Hanusa TP, Brennessel WW. Symmetric Assembly of a Sterically Encumbered Allyl Complex: Mechanochemical and Solution Synthesis of the Tris(allyl)beryllate, K[BeA′3] (A′ = 1,3-(SiMe3)2C3H3). Inorganics. 2017; 5(2):36. https://doi.org/10.3390/inorganics5020036

Chicago/Turabian Style

Boyde, Nicholas C., Nicholas R. Rightmire, Timothy P. Hanusa, and William W. Brennessel. 2017. "Symmetric Assembly of a Sterically Encumbered Allyl Complex: Mechanochemical and Solution Synthesis of the Tris(allyl)beryllate, K[BeA′3] (A′ = 1,3-(SiMe3)2C3H3)" Inorganics 5, no. 2: 36. https://doi.org/10.3390/inorganics5020036

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop