Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2025 Mar 11;147(12):10073–10077. doi: 10.1021/jacs.5c02179

Methane Beryllation Catalyzed by a Base Metal Complex

Josef T Boronski †,*, Agamemnon E Crumpton , Job J C Struijs , Simon Aldridge ‡,*
PMCID: PMC11951138  PMID: 40068010

Abstract

graphic file with name ja5c02179_0005.jpg

The homogeneous catalytic functionalization of methane is extremely challenging due to the relative nonpolarity and high C–H bond strength of this hydrocarbon. Here, using catalytic quantities (10 mol %) of CpMn(CO)3 or Cp*Re(CO)3, the conversion of methane and benzene C–H bonds to C–Be and H–Be bonds by CpBeBeCp has been achieved under photochemical conditions. Possible intermediates in the beryllation reactions—trans-bis(beryllyl)-manganese and -rhenium complexes—were also isolated. Quantum chemical calculations indicate that the inherent properties of the beryllyl ligands—which are powerfully σ-donating and feature highly Lewis acidic beryllium centers—are decisive in enabling methane functionalization by these systems.


The borylation of hydrocarbon C–H bonds has been a topic of intensive investigation for several decades,16 in part driven by the synthetic utility of the boryl moiety (BX2), which can be straightforwardly converted to a range of alternate functional groups.79 The first report of catalytic regiospecific alkane borylation detailed the use of cyclopentadienyl manganese- and rhenium-tricarbonyl derivatives, CpxM(CO)3 (1′: CpX = C5H4Me, M = Mn; 2: CpX = C5Me5, M = Re) under photochemical conditions, harnessing B2Pin2 (Pin = pinacolate) as the boron source.10 The corresponding metal–(bis)boryl complexes are key intermediates in these catalytic reactions.1113 While subsequent progress in the field of C–H borylation has been rapid, homogeneous catalysts—particularly those featuring base metals—for the functionalization of the simplest alkane, methane, are limited.1420 Indeed, although methane borylation using a diborane(4) reagent as the boron source is calculated to be thermodynamically downhill, the hydrocarbon’s insolubility, nonpolarity, and high C–H bond strength (104 kcal mol–1) make this reaction very challenging in practice.3

Beryllium and boron are abutting elements within the second period.21,22 However, while the reactivity of diborane(4) derivatives has been thoroughly developed, that of “diberyllanes”—molecules with a Be–Be bond—is almost entirely unknown.2328 We have recently shown that the Be–Be bond of CpBeBeCp (3) adds to low-oxidation state metal centers similarly to B–B bonds.29 As such, and with the aforementioned reactivity of 1′ and 2 in mind, we were interested to examine the competency of manganese- and rhenium-beryllyl complexes in C–H functionalization chemistry.10

Here, we report hydrocarbon metalation catalyzed by manganese—a base metal—and rhenium under photochemical conditions and at atmospheric pressures. Specifically, using catalytic quantities of either CpMn(CO)3 (1) or 2, and employing 3 as the beryllium source, the conversion of the unactivated C–H bonds of alkanes (methane) and arenes (benzene) to C–Be and H–Be bonds occurs rapidly. Proposed reaction intermediates, trans-CpMn(CO)2(BeCp)2 (4) and trans-Cp*Re(CO)2(BeCp)2 (5), were both isolated. Quantum chemical methods have been used to compare the mechanisms of methane beryllation and borylation by 1. These calculations indicate that the properties of the beryllyl ligands—which are highly electron releasing and Lewis acidic—are key to methane functionalization. Hence, our work provides insights that may inform the future design of functional homogeneous systems for C–H elementation.

Photolysis of an equimolar quantity of either CpMn(CO)3 (1) or Cp*Re(CO)3 (2) with CpBeBeCp (3) in cyclohexane led to the quantitative formation of trans-CpMn(CO)2(BeCp)2 (4) or trans-Cp*Re(CO)2(BeCp)2 (5), respectively (Scheme 1).10,28,30 In the case of complex 1, the reaction to form 4 takes only 3 h, while the conversion of 2 to 5 takes longer (ca. 16 h).30 Both of the bis(beryllyl) complexes are trivalent, 18-electron, diamagnetic species, and are colorless.

Scheme 1. Synthesis of Bis(beryllyl) Complexes 4 and 5.

Scheme 1

Single crystals of both 4 and 5 could be obtained from concentrated cyclohexane solutions (Figures 1 and S9), allowing for the unambiguous determination of the solid-state structure of each. Both complexes exhibit a four-legged piano-stool geometry, with trans-BeCp ligands. The Mn–Be distances in 4 are 2.169(3) and 2.172(3) Å and the Re–Be distances in 5 are 2.298(17) and 2.316(18) Å. Both sets of bond lengths are comparable to the sum of the covalent radii of the respective elements (Mn–Be, 2.21 Å; Re–Be 2.33 Å).31

Figure 1.

Figure 1

Molecular structure of 4 in the solid state, as determined by X-ray crystallography. Thermal ellipsoids were set at 50% probability.

Cyclohexane-D12 solutions of 4 and 5 were also studied by multinuclear NMR spectroscopy. Of greatest interest are the 9Be NMR spectra of the complexes, which feature resonances at −12.5 ppm and −10.5 ppm, respectively.32,33 These signals are downfield shifted compared to previously disclosed transition metal–beryllyl complexes (e.g., cis-Fe(BeCp)2(CO)4 (−18.0 ppm) and Ni(BeCp)6 (−16.7 ppm)).28,29,34

In order to better understand the electronic structure of 4 and 5, we performed quantum chemical calculations on both, as well as their bis(boryl)-analogues, trans-CpMn(CO)2(BPin)2 (6) and trans-Cp*Re(CO)2(BPin)2 (7) (ωB97X-D4/ZORA-Def2-TZVPP).10 The HOMO of both 4 and 5 corresponds to M–CO π-backdonation, combined with a minor M–Be π-bonding component (Figures S10 and S13). The HOMO–1 of both complexes involves a M–CO π-bonding/M–Be σ-bonding combination (Figures S11 and S14).

Natural Bond Orbital (NBO) and Natural Population Analysis (NPA) calculations were performed on complexes 47. First-order NBO calculations do not find covalent Mn– or Re–Be interactions. This contrasts with 6 and 7, for which the M–B bonds are calculated to be composed of approximately 80:20 Re:B and 64:36 Mn:B character, respectively. Second-order NBO calculations indicate that there is substantial π-back-donation from filled Re d-orbitals to Be 2p-orbitals and B–O σ*-antibonding orbitals in 5 and 7, respectively. However, analogous Mn–Be/B π-back-donation is insignificant in complexes 4 and 6. According to the Pauling scale of electronegativity (χ: Be, 1.57; Mn, 1.55; Re, 1.90), addition of the Be–Be bond of 3 at manganese is formally an oxidative process, whereas the same reaction should be termed reductive at rhenium. On this basis, complex 4 could be formulated as a manganese(III) species and 5 as a rhenium(−I) complex. However, NPA charges are negative for the d-block metal centers of 47 (−1.08, −0.68, −0.81, and −0.79, respectively). Thus, the physical oxidation state of manganese in complex 4 is not well represented by the MnIII formalism derived from Pauling electronegativities.29 NPA charges at beryllium are similar in both 5 (av. +1.47) and 6 (av. +1.42), and markedly higher than the charges at the boron centers of 6 (av. +1.07) and 7 (av. +1.16).

Quantum Theory of Atoms in Molecules (QTAIM) and Electron Localization Function (ELF) calculations were also performed on complexes 47. The calculated Bader charge distributions within these complexes contrast somewhat with those indicated by NPA. The metal centers in 4 and 5 bare charges close to zero (Mn, −0.059; Re, −0.091), while the Mn and Re centers of boryl complexes 6 and 7 are positively charged (Mn, +0.72; Re, +0.92). Bader charges of the Be centers of 4 (av. 1.60) and 5 (av. 1.58), and the B centers of 6 (av. 1.64) and 7 (av. 1.58), are all similar. The parentage of the ELF basins associated with the Mn–Be (75:20, Mn:Be) and Mn–B (30:66, Mn:B) interactions in 4 and 6, respectively, suggests that these bonds are of opposite polarity to one another. Thus, the computational data indicate that the metal centers in 4 and 5 are more electron rich than those in 6 and 7, and that the M–Be interactions are polarized to a much greater extent than the analogous M–B bonds (Table S3).

The intermediacy of complex 7 in alkane borylation has been established.10 Hence, we were interested to examine whether metal–beryllyl complexes might prove competent in C–H beryllation reactions (Scheme 2).35 Accordingly, photolysis of solutions of 3 in C6D6 with 10 mol % of 1, 2, 4, or 5 forms CpBeD and CpBe(C6D5), as indicated by multinuclear NMR monitoring (Table S1).36,37 Reactions occur at a much faster rate for manganese complexes 1 and 4 (and with lower energy irradiation), presumably due to the greater lability of the carbonyl ligands of these complexes.30 The novel complex CpBePh (along with CpBeI) was independently prepared by reaction of 3 with PhI. Notably, the [Pd(PCy3)2]-catalyzed magnesiation of benzene with [Mg(NacNacMes)]2—a complex that features a Mg–Mg bond—has been previously described.3840

Scheme 2. Beryllation of Hydrocarbons with 3 Catalyzed by Group 7 Complexes.

Scheme 2

Complexes 4 and 5 do not react with the secondary C–H bonds of cyclohexane, similarly to 7.10 Thus, we turned our attention to an alkane comprising only primary C–H bonds: methane. The [Pd(PCy3)2]-catalyzed C–H magnesiation of methane with [Mg(NacNacMes)]2 has not been demonstrated experimentally; the Gibbs free energy of this reaction is calculated to be very close to zero.38 However, photolysis of 3 under CH4 (1 atm.) in a cyclohexane-D12 solution containing 10 mol % of 1 or 2 led to the formation of CpBeMe and CpBeH.36,37,41,42 Moreover, 4 and 5 (10 mol %) also conduct this transformation in >75% yield under analogous conditions (Table S1).

Methane borylation catalyzed by complexes 1 and 2 has not been reported.10,18,20,43 Moreover, the mechanisms of alkane borylation mediated by these complexes have not been explored in-depth computationally. Thus, we employed quantum chemical calculations to probe the mechanisms of methane borylation (by B2Pin2) and beryllation (by 3) catalyzed by 1, with the aim of discerning and rationalizing the differences between the two reactions. A simplified catalytic cycle for these processes is displayed in Figure 2, with a full reaction profile illustrated in Figure S22.

Figure 2.

Figure 2

Simplified catalytic cycle for methane functionalization by CpMn(CO)3 with either B2Pin2 or CpBeBeCp. Values in italics correspond to the free energy changes in kcal mol–1.

Both mechanisms follow a similar path, beginning with the photoexcitation of 1, followed by the loss of CO, forming CpMn(CO)2. This particular reaction step has been extensively studied by experimental and theoretical methods.30,4447 Subsequent binding of 3 to CpMn(CO)2 is essentially barrierless, generating σ-complex I1Be, but an activation energy of +8.0 kcal mol–1 is calculated for formation of the analogous B2Pin2 complex, I1B. The addition of the Be–Be bond of 3 and B–B bond of B2Pin2 to Mn occurs in an unusual manner; in both cases a carbonyl ligand assists the process in a fashion that resembles a migratory insertion (T3).48 The activation energy barrier for this step is much smaller for the generation of I2Be (complex 4) than that for I2B (complex 6) (+21.6 and 31.3 kcal mol–1, respectively). Furthermore, while the formation of I2Be is exergonic with respect to that of I1 (−9.5 kcal mol–1), that of I2B is marginally uphill (+0.6 kcal mol–1). We ascribe this to the nature of beryllyl ligands, which are more reducing and Lewis acidic than their boryl analogues, as indicated by QTAIM, NPA, and ELF calculations (vide supra).29 This is also consistent with experimental work; in our hands complex 6 could not be isolated via photochemical reaction of 1 with B2Pin2 in cyclohexane.10

The next stage of the reaction involves photoinduced CO ejection from I2, generating a three-legged piano-stool complex, I3, which subsequently binds methane as a C–H σ-complex (I4). The next transition state, T6, is key and resembles a manganese-assisted concerted double σ-bond metathesis; the C–H bond is broken as E–C and E–H bonds are simultaneously formed.49,50 Attempts to optimize the products of methane C–H oxidative addition at Mn were unsuccessful at all relevant points along the reaction coordinates of the borylation and beryllation mechanisms. This is consistent with previous theoretical studies of C–H borylation with iron, rhodium, and tungsten complexes, which suggest that reaction pathways involving C–H oxidative can be much higher in energy than those involving σ-bond metathesis.1113 The activation energy barrier for the formation of I5Be from I4Be (+5.1 kcal mol–1) is notably smaller than that for the analogous formation of I5B from I4B (+15.3 kcal mol–1). Furthermore, T6B is located 22.2 kcal mol–1 above I3B, compared with the +8.4 kcal mol–1 energy difference between I3Be and T6Be. The magnitudes of these energetic barriers provide further evidence for the important role of the X-type boryl or beryllyl ligands in the respective C–H elementation reactions. Indeed, T6Be is stabilized by the electron-rich Mn center—a result of the powerful σ-donor properties of the beryllyl ligands—as well as the highly Lewis acidic nature of the beryllium atoms themselves, which polarize the methane C–H bond.11 The higher energy of T6B is a result of the more covalent Mn–B bonding (Table S3), which renders the boron centers of this transition state less Lewis acidic than the corresponding beryllium sites in T6Be, and the Mn atom less electron rich than in the analogous beryllyl transition state.

Finally, the newly formed element-hydride dissociates (T7) and CO recoordinates to Mn (T8), before dissociation of the E–CH3 species from the metal center, regenerating CpMn(CO)2 (T9).51 Overall, the catalytic cycle for methane borylation by 1 is slightly less exergonic than beryllation (−8.9 versus −11.6 kcal mol–1).

In conclusion, under photochemical conditions, the catalytic beryllation of methane and benzene by CpBeBeCp (3) has been achieved using CpMn(CO)3 (1) or Cp*Re(CO)3 (2). Viable intermediates in hydrocarbon beryllation—trans-CpMn(CO)2(BeCp)2 (4) and trans-Cp*Re(CO)2(BeCp)2 (5)—have been isolated. Inspection of the mechanisms of methane borylation and berylation by 1 reveals that the properties of the beryllyl ligands facilitate this novel C–H elementation reaction to an extent that is not possible with boron analogues.

Acknowledgments

J.T.B. thanks the Royal Society of Chemistry (Research Fund R23-3176939355) for financial support. We thank the EPSRC Centre for Doctoral Training in Inorganic Chemistry for Future Manufacturing (OxICFM, EP/S023828/1 studentships for A.E.C. and J.J.C.S).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.5c02179.

  • Materials and methods, experimental procedures, spectroscopic data, crystallographic details, and quantum chemical data (PDF)

Author Contributions

All authors have given approval to the final version of the manuscript.

The authors declare no competing financial interest.

Supplementary Material

ja5c02179_si_001.pdf (1.1MB, pdf)

References

  1. Hartwig J. F. Borylation and Silylation of C-H Bonds: A Platform for Diverse C-H Bond Functionalizations. Acc. Chem. Res. 2012, 45 (6), 864–873. 10.1021/ar200206a. [DOI] [PubMed] [Google Scholar]
  2. Hartwig J. F.; Larsen M. A. Undirected, Homogeneous C-H Bond Functionalization: Challenges and Opportunities. ACS Cent. Sci. 2016, 2 (5), 281–292. 10.1021/acscentsci.6b00032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Mkhalid I. A. I.; Barnard J. H.; Marder T. B.; Murphy J. M.; Hartwig J. F. C-H Activation for the Construction of C-B Bonds. Chem. Rev. 2010, 110 (2), 890–931. 10.1021/cr900206p. [DOI] [PubMed] [Google Scholar]
  4. Yu I. F.; Wilson J. W.; Hartwig J. F. Transition-Metal-Catalyzed Silylation and Borylation of C-H Bonds for the Synthesis and Functionalization of Complex Molecules. Chem. Rev. 2023, 123, 11619–11663. 10.1021/acs.chemrev.3c00207. [DOI] [PubMed] [Google Scholar]
  5. Hartwig J. F. Regioselectivity of the Borylation of Alkanes and Arenes. Chem. Soc. Rev. 2011, 40 (4), 1992–2002. 10.1039/c0cs00156b. [DOI] [PubMed] [Google Scholar]
  6. Ishiyama T.; Miyaura N. Transition Metal-Catalyzed Borylation of Alkanes and Arenes via C-H Activation. J. Organomet. Chem. 2003, 680, 3–11. 10.1016/S0022-328X(03)00176-1. [DOI] [Google Scholar]
  7. Miyaura N.; Suzuki A. Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem. Rev. 1995, 95 (7), 2457–2483. 10.1021/cr00039a007. [DOI] [Google Scholar]
  8. Fyfe J. W. B.; Watson A. J. B. Recent Developments in Organoboron Chemistry: Old Dogs, New Tricks. Chem. 2017, 3, 31–55. 10.1016/j.chempr.2017.05.008. [DOI] [Google Scholar]
  9. Boronic Acids: Preparation and Applications in Organic Synthesis, Medicine and Materials, Second ed.; Hall D. G., Ed.; Wiley: Weinheim, 2011. 10.1002/9783527639328 [DOI] [Google Scholar]
  10. Chen H.; Hartwig J. F. Catalytic, Regiospecific End-Functionalization of Alkanes: Rhenium-J. Am. Chem. Soc.d Borylation under Photochemical Conditions. Angew. Chem., Int. Ed. 1999, 38 (22), 3391–3393. . [DOI] [PubMed] [Google Scholar]
  11. Webster C. E.; Fan Y.; Hall M. B.; Kunz D.; Hartwig J. F. Experimental and Computational Evidence for a Boron-Assisted, σ-Bond Metathesis Pathway for Alkane Borylation. J. Am. Chem. Soc. 2003, 125 (4), 858–859. 10.1021/ja028394c. [DOI] [PubMed] [Google Scholar]
  12. Boller T. M.; Murphy J. M.; Hapke M.; Ishiyama T.; Miyaura N.; Hartwig J. F. Mechanism of the Mild Functionalization of Arenes by Diboron Reagents Catalyzed by Iridium Complexes. Intermediacy and Chemistry of Bipyridine-Ligated Iridium Trisboryl Complexes. J. Am. Chem. Soc. 2005, 127 (41), 14263–14278. 10.1021/ja053433g. [DOI] [PubMed] [Google Scholar]
  13. Hartwig J. F.; Cook K. S.; Hapke M.; Incarvito C. D.; Fan Y.; Webster C. E.; Hall M. B. Rhodium Boryl Complexes in the Catalytic, Terminal Functionalization of Alkanes. J. Am. Chem. Soc. 2005, 127 (8), 2538–2552. 10.1021/ja045090c. [DOI] [PubMed] [Google Scholar]
  14. Crabtree R. H. The Organometallic Chemistry of Alkanes. Chem. Rev. 1985, 85 (4), 245–269. 10.1021/cr00068a002. [DOI] [Google Scholar]
  15. Fujisaki H.; Kojima T. Functionalization of Methane Using Molecular Metal Complexes as Catalysts. Catal. Sci. Technol. 2023, 13, 4270–4284. 10.1039/D3CY00647F. [DOI] [Google Scholar]
  16. Labinger J. A. Methane Activation in Homogeneous Systems. Fuel Proc. Technol. 1995, 42, 325–338. 10.1016/0378-3820(94)00107-5. [DOI] [Google Scholar]
  17. Cavaliere V. N.; Mindiola D. J. Methane: A New Frontier in Organometallic Chemistry. Chem. Sci. 2012, 3, 3356–3365. 10.1039/c2sc20530k. [DOI] [Google Scholar]
  18. Smith K. T.; Berritt S.; González-Moreiras M.; Ahn S.; Smith M. R.; Baik M.-H.; Mindiola D. J. Catalytic Borylation of Methane. Science 2016, 351 (6280), 1424–1427. 10.1126/science.aad9730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Pulcinella A.; Mazzarella D.; Noël T. Homogeneous Catalytic C(sp3)-H Functionalization of Gaseous Alkanes. Chem. Commun. 2021, 57 (78), 9956–9967. 10.1039/D1CC04073A. [DOI] [PubMed] [Google Scholar]
  20. Cook A. K.; Schimler S. D.; Matzger A. J.; Sanford M. S. Catalyst-Controlled Selectivity in the C–H Borylation of Methane and Ethane. Science 2016, 351 (6280), 1421–1424. 10.1126/science.aad9289. [DOI] [PubMed] [Google Scholar]
  21. Buchner M. R. Recent Contributions to the Coordination Chemistry of Beryllium. Chem. Eur. J. 2019, 25, 12018–12036. 10.1002/chem.201901766. [DOI] [PubMed] [Google Scholar]
  22. 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 (36), 10562–10576. 10.1002/anie.201601809. [DOI] [PubMed] [Google Scholar]
  23. Neeve E. C.; Geier S. J.; Mkhalid I. A. I.; Westcott S. A.; Marder T. B. Diboron(4) Compounds: From Structural Curiosity to Synthetic Workhorse. Chem. Rev. 2016, 116, 9091–9161. 10.1021/acs.chemrev.6b00193. [DOI] [PubMed] [Google Scholar]
  24. Dewhurst R. D.; Neeve E. C.; Braunschweig H.; Marder T. B. sp2-sp3 Diboranes: Astounding Structural Variability and Mild Sources of Nucleophilic Boron for Organic Synthesis. Chem. Commun. 2015, 51 (47), 9594–9607. 10.1039/C5CC02316E. [DOI] [PubMed] [Google Scholar]
  25. Boronski J. T. Alkaline Earth Metals: Homometallic Bonding. Dalton Trans. 2023, 53 (1), 33–39. 10.1039/D3DT03550F. [DOI] [PubMed] [Google Scholar]
  26. Boronski J. T.; Thomas-Hargreaves L. R.; Ellwanger M. A.; Crumpton A. E.; Hicks J.; Bekiş D. F.; Aldridge S.; Buchner M. R. Inducing Nucleophilic Reactivity at Beryllium with an Aluminyl Ligand. J. Am. Chem. Soc. 2023, 145 (8), 4408–4413. 10.1021/jacs.3c00480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Boronski J. T.; Crumpton A. E.; Roper A. F.; Aldridge S. A Nucleophilic Beryllyl Complex via Metathesis at [Be–Be]2+. Nat. Chem. 2024, 16 (8), 1295–1300. 10.1038/s41557-024-01534-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Boronski J. T.; Crumpton A. E.; Wales L. L.; Aldridge S. Diberyllocene, a Stable Compound of Be(I) with a Be–Be Bond. Science 2023, 380 (6650), 1147–1149. 10.1126/science.adh4419. [DOI] [PubMed] [Google Scholar]
  29. Boronski J. T.; Crumpton A. E.; Aldridge S. A Crystalline NiX6 Complex. J. Am. Chem. Soc. 2024, 146, 35208–35215. 10.1021/jacs.4c12125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Turner J. J.; George M. W.; Poliakoff M.; Perutz R. N. Photochemistry of Transition Metal Carbonyls. Chem. Soc. Rev. 2022, 51, 5300–5329. 10.1039/D1CS00826A. [DOI] [PubMed] [Google Scholar]
  31. Pyykkö P. Additive Covalent Radii for Single-, Double-, and Triple-Bonded Molecules and Tetrahedrally Bonded Crystals: A Summary. J. Phys. Chem. A 2015, 119 (11), 2326–2337. 10.1021/jp5065819. [DOI] [PubMed] [Google Scholar]
  32. Buchanan J. K.; Plieger P. G. 9Be Nuclear Magnetic Resonance Spectroscopy Trends in Discrete Complexes: An Update. Zeitschrift fur Naturforschung - Section B Journal of Chemical Sciences. 2020, 75, 459–472. 10.1515/znb-2020-0007. [DOI] [Google Scholar]
  33. 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 (44), 14651–14658. 10.1021/ja046712x. [DOI] [PubMed] [Google Scholar]
  34. Boronski J. T.; Griffin L. P.; Conder C.; Crumpton A. E.; Wales L. L.; Aldridge S. On the Nature and Limits of Alkaline Earth-Triel Bonding. Chem. Sci. 2024, 15, 15377–15384. 10.1039/D4SC03832K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Field-Theodore T. E.; Couchman S. A.; Wilson D. J. D.; Dutton J. L. Carboberyllation: Addition of Organoberyllium Species to Alkenes and Alkynes. A Comparison with Carboboration. Dalton Trans. 2017, 46 (24), 7686–7692. 10.1039/C7DT01370A. [DOI] [PubMed] [Google Scholar]
  36. Bartke T. C.; Bjorseth A.; Haaland A.; Marstokk K.-M.; Mo̷llendal H. Microwave Spectrum, Structure and Dipole Moment of Cyclopentadienylberyllium Hydride. J. Organomet. Chem. 1975, 85 (3), 271–277. 10.1016/S0022-328X(00)80300-9. [DOI] [Google Scholar]
  37. Schmidbaur H.Be: Organoberyllium Compounds; Springer Berlin Heidelberg: Berlin, Heidelberg, 1987. 10.1007/978-3-662-06024-7 [DOI] [Google Scholar]
  38. Garçon M.; White A. J. P.; Crimmin M. R. Palladium-Catalysed Magnesiation of Benzene. Chem. Commun. 2018, 54 (87), 12326–12328. 10.1039/C8CC06392C. [DOI] [PubMed] [Google Scholar]
  39. Green S. P.; Jones C.; Stasch A. Stable Magnesium(I) Compounds with Mg-Mg Bonds. Science 2007, 318 (5857), 1754–1757. 10.1126/science.1150856. [DOI] [PubMed] [Google Scholar]
  40. Jones C. Dimeric Magnesium(I) β-Diketiminates: A New Class of Quasi-Universal Reducing Agent. Nat. Rev. Chem. 2017, 1 (8), 0059. 10.1038/s41570-017-0059. [DOI] [Google Scholar]
  41. Saulys D. A.; Powell D. R. Synthesis, Experimental/Theoretical Characterization, and Thermolysis Chemistry of CpBe(SiMe3), a Molecule Containing an Unprecedented Beryllium-Silicon Bond. Organometallics 2003, 22 (3), 407–413. 10.1021/om0201720. [DOI] [Google Scholar]
  42. Drew D. A.; Morgan G. L. Synthesis and Spectroscopic Properties of Cyclopentadienyl(Methyl)Beryllium and Cyclopentadienylberyllium Halide Complexes. Inorg. Chem. 1977, 16 (7), 1704–1708. 10.1021/ic50173a027. [DOI] [Google Scholar]
  43. Gunsalus N. J.; Koppaka A.; Park S. H.; Bischof S. M.; Hashiguchi B. G.; Periana R. A. Homogeneous Functionalization of Methane. Chem. Rev. 2017, 117, 8521–8573. 10.1021/acs.chemrev.6b00739. [DOI] [PubMed] [Google Scholar]
  44. Wu X.; Liu Z.; Murphy T. S.; Sun X. Z.; Hanson-Heine M. W. D.; Towrie M.; Harvey J. N.; George M. W. The Effect of Coordination of Alkanes, Xe and CO21-OCO) on Changes in Spin State and Reactivity in Organometallic Chemistry: A Combined Experimental and Theoretical Study of the Photochemistry of CpMn(CO)3. Faraday Discuss. 2019, 220, 86–104. 10.1039/C9FD00067D. [DOI] [PubMed] [Google Scholar]
  45. Yang H.; Asplund M. C.; Kotz K. T.; Wilkens M. J.; Frei H.; Harris C. B. Reaction Mechanism of Silicon–Hydrogen Bond Activation Studied Using Femtosecond to Nanosecond IR Spectroscopy and Ab Initio Methods. J. Am. Chem. Soc. 1998, 120 (39), 10154–10165. 10.1021/ja980692f. [DOI] [Google Scholar]
  46. Kemnitz C. R.; Ball E. S.; McMahon R. J. Photochemistry of CpMn(CO)3 and Related Derivatives: Spectroscopic Observation of Singlet and Triplet CpMn(CO)2. Organometallics 2012, 31 (1), 70–84. 10.1021/om200555e. [DOI] [Google Scholar]
  47. Zhang Z. F.; Su M. Der. The Mechanistic Investigations of Photochemical Carbonyl Elimination and Oxidative Addition Reactions of (η5-C5H5)M(CO)3, (M = Mn and Re) Complexes. RSC Adv. 2018, 8 (20), 10987–10998. 10.1039/C8RA01118D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Frank R.; Howell J.; Tirfoin R.; Dange D.; Jones C.; Mingos D. M. P.; Aldridge S. Circumventing Redox Chemistry: Synthesis of Transition Metal Boryl Complexes from a Boryl Nucleophile by Decarbonylation. J. Am. Chem. Soc. 2014, 136 (44), 15730–15741. 10.1021/ja508979x. [DOI] [PubMed] [Google Scholar]
  49. Perutz R. N.; Sabo-Etienne S. The σ-CAM Mechanism: σ Complexes as the Basis of σ-Bond Metathesis at Late-Transition-Metal Centers. Angew. Chem., Int. Ed. 2007, 46, 2578–2592. 10.1002/anie.200603224. [DOI] [PubMed] [Google Scholar]
  50. Perutz R. N.; Sabo-Etienne S.; Weller A. S. Metathesis by Partner Interchange in σ-Bond Ligands: Expanding Applications of the σ-CAM Mechanism. Angew. Chem., Int. Ed. 2022, 61, e202111462 10.1002/anie.202111462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Schlecht S.; Hartwig J. F. σ-Borane Complexes of Manganese and Rhenium. J. Am. Chem. Soc. 2000, 122 (39), 9435–9443. 10.1021/ja001546o. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ja5c02179_si_001.pdf (1.1MB, pdf)

Articles from Journal of the American Chemical Society are provided here courtesy of American Chemical Society

RESOURCES