Skip to main content
Science Advances logoLink to Science Advances
. 2023 Jun 9;9(23):eadg8130. doi: 10.1126/sciadv.adg8130

An osmium(II) methane complex: Elucidation of the methane coordination mode

Peter J Sempsrott 1, Brian B Trinh 1, Charity Flener Lovitt 1,2, Nicolas E Capra 1, Gregory S Girolami 1,*
PMCID: PMC10256148  PMID: 37294762

Abstract

The activation of inert C─H bonds by transition metals is of considerable industrial and academic interest, but important gaps remain in our understanding of this reaction. We report the first experimental determination of the structure of the simplest hydrocarbon, methane, when bound as a ligand to a homogenous transition metal species. We find that methane binds to the metal center in this system through a single M···H-C bridge; changes in the 1JCH coupling constants indicate clearly that the structure of the methane ligand is significantly perturbed relative to the free molecule. These results are relevant to the development of better C─H functionalization catalysts.


A detailed experimental determination shows how the simplest hydrocarbon, methane, binds to a homogenous transition metal species.

INTRODUCTION

The efficient and scalable conversion of hydrocarbon feedstocks (such as petroleum and natural gas) to higher-value commodity or specialty chemicals is an important goal of modern catalysis, and in part for this reason, the activation of C─H bonds by molecular transition metal complexes has been the subject of significant and growing interest since the 1980s (113). Particularly impressive has been the use of directing groups to control the functionalization of specific C─H bonds (1416). Despite these advances, the development of mild and selective C─H activation methods that do not rely on directing groups remains an important challenge (2, 5, 1721).

One mechanism by which alkane C─H bonds can be activated by a transition metal complex involves, as a first step, coordination of the alkane to the metal center. Such compounds, which are known as σ-alkane complexes, were initially studied by matrix isolation methods (22, 23) and have subsequently been studied by a variety of other techniques (2431). Despite these advances, there is little detailed experimental information about exactly how the alkane binds to the metal, especially in solution. One significant challenge in studying σ-alkane complexes in solution has been their low stability; the metal-alkane interaction is very weak, usually on the order of ~10 kcal mol−1. Consequently, solution information on the smallest and most weakly coordinating alkanes, particularly methane, is scarce, and in no case has the coordination mode of the methane ligand been determined experimentally (3234).

In 1998, we reported that the low-temperature protonation of the osmium(II) methyl complex (C5Me5)Os(dmpm)(CH3), where dmpm is the chelating diphosphine Me2PCH2PMe2, affords the methyl/hydride complex [(C5Me5)Os(dmpm)(CH3)H]+ (35). This compound was the first example of an alkyl/hydride complex that, on the nuclear magnetic resonance (NMR) time scale, is in dynamic equilibrium with its alkane tautomer: Interconversion of the methyl/hydride complex with [(C5Me5)Os(dmpm)(CH4)]+ occurs at a rate of 170 s−1 at −100°C. Density functional theory (DFT) calculations indicated that this σ-methane tautomer is about 5 kcal mol−1 higher in energy than the methyl/hydride complex and that variation of the ancillary ligands can reverse this energy difference so that the σ-methane tautomer is the predominant species in solution (3638).

Guided by those DFT calculations, we report here the synthesis of a new osmium(II) σ-methane complex and its characterization by low-temperature NMR spectroscopy. Isotopic labeling studies have enabled us to obtain detailed experimental information about how the methane ligand binds to the metal center and to what extent this interaction changes its structure.

RESULTS AND DISCUSSION

In our earlier work, we found that [(C5Me5)Os(dmpm)(CH3)H]+ is in equilibrium with the σ-methane tautomer [(C5Me5) Os(dmpm)(CH4)]+ in solution but that the equilibrium concentration of the latter is too low to detect directly (35). This result suggested that modifying the ancillary ligands to decrease the electron richness of the osmium center should lower the energy of the σ-methane tautomer by disfavoring oxidative addition of the C─H bond. We used benchmarked DFT calculations to evaluate the relative energies of [(C5R5)Os(diphosphine)(CH3)H]+ and [(C5R5)Os(diphosphine)(CH4)]+ complexes bearing a variety of substituted cyclopentadienyl and diphosphine ligands (38). The calculations, which confirmed our initial expectations, predicted that replacement of the dmpm ligand with the more electron-withdrawing phosphine (CF3)2PCH2P(CF3)2 (dfmpm) should render the σ-methane complex more stable than the methyl/hydride tautomer by 1.7 kcal mol−1, while keeping the barrier for methane dissociation large enough (12.1 kcal mol−1) to observe the complex at low temperatures. The calculations suggested that the stabilizing effect of dfmpm on the methane ligand is a result of a higher ratio of CH σ → Os d donation to Os d → CH σ* backdonation (38).

Although dfmpm has been synthesized previously (39), the reported method is not easily scalable to quantities greater than ~1 mmol. Accordingly, we developed an alternative preparation (see the Supplementary Materials) that provides multigram quantities of dfmpm from the commercially available diphosphine Cl2PCH2PCl2; a key improvement of our protocol is the use of the Ruppert-Prakash reagent (CF3SiMe3/CsF) as a CF3 source rather than (CF3)2PI (40).

The osmium methyl complex (C5Me5)Os(dfmpm)(CH3) was prepared (see the Supplementary Materials) in 54% overall yield in two steps from (C5Me5)2Os2Br4 (Fig. 1) (41, 42); the second step, alkylation of (C5Me5)Os(dfmpm)Br, required the use of dimethylzinc because methyllithium and methylmagnesium reagents deprotonate the CH2 group in the backbone of the dfmpm ligand, and trimethylaluminum activates the C─F bonds of the phosphine. The Os-CH3 resonance in the 1H NMR spectrum is a triplet (JHP = 7.5 Hz) of septets (JHF = 2 Hz); the septet splitting shows that there is a remarkably strong five-bond coupling between the methyl protons and 6 of the 12 fluorine atoms of the dfmpm ligand. The 1JCH coupling constant for the Os-CH3 group is 131 Hz. The monomeric nature of the complex is confirmed by its single-crystal x-ray structure of (C5Me5)Os(dfmpm)(CH3), which features an Os─C bond length of 2.177 (3) Å and a diphosphine bite angle of 70.14 (2)° (Fig. 2).

Fig. 1. Synthesis of (C5Me5)Os(dfmpm)(CH3) and the isolated yields for each step.

Fig. 1.

We have previously reported a preparation for the (C5Me5)2Os2Br4 starting material (41).

Fig. 2. An ORTEP view of the x-ray crystal structure of (C5Me5)Os(dfmpm)(CH3).

Fig. 2.

All atoms are shown as 30% probability ellipsoids except for hydrogen atoms, which are shown as arbitrarily sized spheres.

Protonation of (C5Me5)Os(dfmpm)(CH3) with either trifluoromethanesulfonic acid or bis(trifluoromethanesulfonyl)amine at −110°C in CDCl2F (43) gives a product 1 that exhibits a singlet at the unusual chemical shift of δ −1.94 in the 1H NMR spectrum. This value lies between the 1H NMR chemical shift for the parent methyl resonance (δ 0.36) and the 1H NMR chemical shifts of related osmium hydrides (δ −10 to −20) (42). Protonation of the 13C-labeled isotopolog (C5Me5)Os(dfmpm)(13CH3) causes the 1H NMR resonance at δ −1.94 to split into a doublet with JCH = 126 Hz; the 13C NMR spectrum of the same species features a binomial pentet at δ −45.11 with the same coupling constant (Fig. 3). The size of the average C─H coupling constant, which is similar to the values of 124 and 127 Hz seen for Rh(PONOP)(CH4)+ and (C5H5)Os(CO)2(CH4), respectively (3234), rules out the possibility that the protonation product 1 is a classical osmium cis-methyl/hydride complex in which the hydride and methyl hydrogen atoms are exchanging rapidly on the NMR time scale. If this were the case, the three 1JCH coupling constants within the Os-CH3 group would be approximately equal to the 131 Hz value seen in (C5Me5)Os(dfmpm)(13CH3), whereas the 2JCH coupling constant between the methyl carbon and the Os─H group would be close to zero: Exchange averaging would produce a JCH coupling constant of at most ~100 Hz, which is inconsistent with the observed JCH of 1. We therefore conclude that 1 is the methane coordination complex [(C5Me5)Os(dfmpm)(13CH4)]+.

Fig. 3. 1H and 13C NMR resonances for the coordinated methane ligand in [(C5Me5)Os(dfmpm)13CH4]+ at −110°C in CDCl2F.

Fig. 3.

A small quartet at δ −45.4 in the 13C NMR spectrum is due to residual (unprotonated) (C5Me5)Os(dfmpm)(13CH3).

Above −110°C, [(C5Me5)Os(dfmpm)(CH4)]+ releases methane (figs. S1 and S2) by a process that is first order in 1. An Eyring analysis (fig. S3) indicated that ΔH = 16 ± 2 kcal mol−1 and ΔS = 18 ± 12 cal mol−1 K−1 and that ΔG = 12.8 ± 0.1 kcal mol−1 at −100°C (44, 45). The positive entropy of activation is consistent with a dissociative mechanism. The experimental value for ΔG of 1 is in good agreement with the calculated binding energies of the model complex [(C5H5)Os(dfmpm)(CH4)]+ (12.1 kcal mol−1) and other σ-methane complexes (3234). Having determined the identity and stability of the protonated complex 1, we then turned to isotopic labeling studies to determine its structure.

There are four principal coordination modes by which methane could bind to a single metal center (Fig. 4): through one (κ1), two (κ2), or three (κ3) bridging hydrogen atoms or through both the hydrogen and carbon atom of a single C─H bond (η2) (24, 46). Because the bridging and terminal H nuclei exchange rapidly even at −130°C, it is not possible to determine directly which of these four structures is present, but the structure can be determined indirectly by an isotopic perturbation of equilibrium (IPE) (47, 48) study of the 13CH4, 13CH3D, 13CH2D2, and 13CHD3 isotopologs. Nearly 50 years ago, Calvert and Shapley (49) elegantly demonstrated the utility of IPE for determining the static structure of a triosmium complex containing an agostic methyl group, and more recently, Ball and coworkers (50, 51) have used IPE to study the structures of the alkane complexes (C5H4Pri)Re(CO)2(pentane) and (C5H5)Re(CO)2(cyclopentane). The IPE method relies on the fact that, if hydrogen and deuterium nuclei equilibrate between two sites with different force constants, the H and D nuclei will preferentially occupy the site(s) with the weaker and the stronger force constants, respectively, because this arrangement minimizes the zero point contribution to the total energy. In a methane ligand, the C─H bond(s) that bridge to the metal should have lower C─H stretching frequencies and weaker force constants; consequently, the bridging site(s) should preferentially be occupied by H.

Fig. 4. The four possible coordination modes of a σ-methane ligand.

Fig. 4.

The nonstandard descriptors proposed in 1996 (24) are listed in (A), but for reasons given in (46), here, we will use the descriptors given in (B).

At a given temperature, the exchange-averaged 1H NMR chemical shift of each isotopolog of [(C5Me5)Os(dfmpm)(CH4)]+ can be expressed as a function of three parameters (see the Supplementary Materials): the terminal and bridging 1H NMR chemical shifts for a static structure [δH(t) and δH(b)] and the zero point energy change that arises upon swapping an H/D pair between the two sites (ΔE). Similar equations apply for the exchange-averaged JCH coupling constant of each isotopolog, which depends on the parameters JCH(t) and JCH(b) and the same energy term ΔE. Because each methane binding mode has a different ratio of bridging and terminal hydrogen atoms (except κ1 versus η2, which we discuss later), each binding mode also has a unique set of equations for the exchange-averaged chemical shifts and JCH constants of the 13CH4, 13CH3D, 13CH2D2, and 13CHD3 isotopologs. Because these systems of equations (52) each contain three variables, measuring the chemical shifts and coupling constants of at least three isotopologs allows us to determine the binding mode, chemical shift, and coupling constants at the slow exchange limit.

The IPE equations predict that the exchange-averaged chemical shifts of the 13CH4, 13CH3D, 13CH2D2, and 13CHD3 methane adducts will show the following trend: For the κ1 (and η2) binding modes, which contain more terminal sites than bridging sites, each successive D for H substitution will cause a larger upfield change in δ than the previous substitution. For the κ2 binding mode, which contains an equal number of terminal and bridging sites, each successive D for H substitution will cause approximately the same upfield change. Last, for the κ3 binding mode, which contains fewer terminal sites than bridging sites, each successive D for H substitution will cause a smaller upfield change in δ than the previous substitution. The exchange-averaged C─H coupling constant will follow the same pattern: Each successive D for H substitution will produce a change in the averaged coupling constant that is unique to the ratio of terminal and bridging hydrogen atoms.

To obtain the isotopologs needed for the IPE study, we treated a 1:1 mixture of (C5Me5)Os(dfmpm)(13CH3) and (C5Me5)Os(dfmpm)(13CHD2) with a 1:2 mixture of protio- and deutero-acid (53) in CDCl2F at −110°C. At this temperature, four doublets are observed in the upfield region of the 1H NMR spectrum (Fig. 5), one for each methane isotopolog: δ −1.94 (Os–13CH4), −2.33 (Os–13CH3D), −2.86 (Os–13CH2D2), and −3.63 (Os–13CHD3). The chemical shift differences seen for each successive D for H substitution are thus 0.39, 0.53, and 0.77 parts per million (ppm). The C─H coupling constants of the four doublets are 126.5, 123.7, 120.0, and 114.6 Hz, respectively, so the differences are 2.8, 3.7, and 5.4 Hz. The same trends in δ and 1JCH are seen at −120°C and −130°C, but the differences seen for each substitution are larger, as expected from the Boltzmann statistics. The increasingly large changes in the chemical shifts and coupling constants seen for each successive D for H substitution is definitive evidence that only one hydrogen atom of the coordinated methane ligand bridges to the osmium center in the ground state structure and thereby proves that the binding mode of methane must be κ1 (or η2).

Fig. 5. 1H NMR resonances for the coordinated methane ligands in a mixture of the isotopologs [(C5Me5)Os(dfmpm)(13CH4-xDx)]+ at −110°C in CDCl2F.

Fig. 5.

Solving the system of equations for the κ12 configuration also enables us to determine the following NMR parameters of the terminal and bridging sites: δH(t) = 0.39 (5), δH(b) = −8.92 (17), and ΔE = 0.264 (5) kcal mol−1. The chemical shift of the terminal hydrogen atoms δH(t) is close to the chemical shift of free alkanes, and the chemical shift of the bridging hydrogen atom δH(b) is similar to the chemical shifts of related osmium(II) hydrides of the form (C5Me5)Os(diphosphine)H (42). The zero point energy difference ΔE between having H or D in the bridging position is about twice the 0.13(1) kcal mol−1 value reported by Calvert and Shapley (49) in their agostic methyl complex but is very similar to the 0.23(3) kcal mol−1 value reported by Ball and coworkers (51) for (C5H4Pri)Re(CO)2(pentane).

When the IPE equations are solved for the C─H coupling constants, we find that JCH(t) = 141 ± 3 Hz and JCH(b) = 83 ± 11 Hz; the difference in 1JCH between the bridging and terminal C─H bonds is large (almost 60 Hz) and shows how strongly the methane molecule is desymmetrized by its interaction with the osmium center (54). Comparing JCH(t) and JCH(b) to the value seen for free methane (125 Hz) (32) shows that the one bridging C─H bond is significantly weakened, whereas the three terminal C─H bonds are significantly strengthened (or have considerably more s-character) (55).

Having determined that one hydrogen atom is bound to the metal center, we turned to the question whether the binding mode is better described as κ1-H or η2-C,H; i.e., is the carbon atom also bonded to the metal (56)? The IPE data for complex 2 is consistent with both of these coordination modes, but DFT calculations on the model complex [(C5H5)Os(dfmpm)(CH4)]+ suggest that there is little or no direct overlap between osmium and carbon, as evidenced by the large Os─H─C angle of 115° and by the long Os-Cmethane distance (2.571 Å, c.f. the sum of the covalent radii ~2.20 Å) (38, 57).

Last, we wish to address why the 13C NMR shift of the σ-methane ligand in 2 is shielded by ~41 ppm relative to free methane. For (C5H4iPr)Re(CO)2(n-pentane), a similar strong shielding of the 13C NMR resonance for the bound σ-pentane ligand was proposed to indicate the presence of an η2-C,H interaction (50). We wish to propose an alternative possibility: that the observed shielding of the 13C NMR shift of the σ-methane ligand in 2 is a direct consequence of the weakness of the metal-alkane binding. As a result, the d6 osmium center is effectively five-coordinate rather than six-coordinate, which lowers the energy of one empty, σ*-character d-orbital; this lowering generates low-lying paramagnetic excited states, which mix into the ground state and perturb the observed NMR chemical shifts. This phenomenon, which is known as a diamagnetic anisotropy effect, is responsible for the strongly shielded 1H NMR chemical shifts of certain five-coordinate d6 iridium hydrides (58) and probably also for the shielding of weakly bound ligands in other d6 pseudo-octahedral compounds, such as (C5H4Pri)Re(CO)(PF3)Xe (59). Relativistic effects may also contribute to the shielding terms for third-row transition metal complexes (58, 6062).

Here, we have prepared a new methane coordination complex [(C5Me5)Os(dfmpm)(CH4)]+ and provided the first experimental evidence for how methane binds to a transition metal complex: In this system, the binding is through one hydrogen atom. The methane ligand is considerably distorted by its interaction with the metal, as indicated by a difference of almost 60 Hz between the 1JCH coupling constants of the bridging and terminal C─H bonds. The differences in 1JCH between the bridging and terminal C─H bonds in the frozen structure are too large to be due solely to dispersion forces (bond polarization or van der Waals interactions), so there must be considerable covalent d-σ mixing between osmium and the bridging C─H bond. We propose that the shielding of the 1H and 13C NMR resonances for the methane ligand may be a result of diamagnetic anisotropy from the pseudo–five-coordinate metal fragment (possibly augmented by relativistic effects), rather than an indication of direct osmium-carbon overlap. These results provide insights into the solution-phase structure and stability of σ-alkane species, which are important intermediates in the oxidative addition of saturated alkanes to transition metal centers.

MATERIALS AND METHODS

Unless stated otherwise, all manipulations were performed under argon or vacuum using standard Schlenk line or glove box techniques. Glassware was oven- or flame-dried and allowed to cool under vacuum or argon. Solvents were distilled under nitrogen from sodium/benzophenone (pentane and diethyl ether), magnesium (methanol), or sodium (toluene) and sparged with argon for 30 to 60 s immediately before use. Benzene-d6 was purchased from Sigma-Aldrich or Cambridge Isotope Laboratories in 1-ml ampoules and used without purification. The following starting materials were obtained from commercial sources and used as received unless stated otherwise: dimethylzinc (1.2 M solution in toluene; Strem), iodomethane-13C (Sigma-Aldrich), iodomethane-13C,d2 (Sigma-Aldrich), lithium powder (Sigma-Aldrich), bis(trifluoromethanesulfonyl)amine (dried over MgSO4 and sublimed; Matrix Scientific), bis(trifluoromethanesulfonyl)amine lithium salt (Oakwood Chemicals), sulfuric acid–d2 (Alfa Aesar), aluminum powder (-40 +325 mesh; Alfa Aesar), iodine (Alfa Aesar), dibromomethane (Sigma-Aldrich), hydrochloric acid (2 M in diethyl ether; Sigma-Aldrich), phosphorus(III) chloride (Sigma-Aldrich), phosphorus(V) oxychloride (Sigma-Aldrich), CF3SiMe3 (Oakwood Chemicals), cesium fluoride (Oakwood Chemicals), and trifluoromethanesulfonic acid (SynQuest Laboratories). The compounds (C5Me5)2Os2Br4 (42) and CDCl2F (43) were prepared according to published procedures. The CDCl2F was distilled and stored at −20°C in a sealed glass vessel over Linde 3 Å or 4 Å molecular sieves.

NMR spectra were acquired on Varian spectrometers (Unity 400, Unity Inova 400, Unity 500, VXR 500, and Unity Inova 600) or Bruker (500 and 600 MHz) spectrometers at room temperature, unless specified otherwise. Low-temperature NMR spectra were acquired on a Varian Unity Inova 600 spectrometer. Positive chemical shifts indicate shifts to higher frequency relative to a chemical shift standard. 1H and 13C spectra were referenced to SiMe4 by setting appropriate shifts to residual solvent signals for benzene-d5 (1H δ 7.16 and 13C δ 128.06) or dichlorofluoromethane (1H δ 7.47 and 13C δ 104.2). 19F and 31P spectra were referenced externally to 15% CFCl3 in CDCl3 and 85% H3PO4 in H2O, respectively. For spectra collected below room temperature, the NMR probe temperature was calibrated with neat methanol (63). NMR spectra were processed with the MestReNova NMR software package. Fourier transform infrared spectra were acquired on a Thermo Nicolet IR200 spectrometer as Nujol mulls between KBr plates and were processed using the OMNIC software package with automatic baseline corrections. Melting points were acquired on a Thomas-Hoover Uni-Melt apparatus in sealed capillaries under argon. Elemental analyses were performed by the University of Illinois Microanalysis Laboratory and are reported as the average of two replicates. X-ray crystallographic data were collected by the University of Illinois George L. Clark X-Ray Facility and 3M Materials Laboratory.

Acknowledgments

We thank L. Zhu and D. Olson for assistance with NMR experiments and L. McElwee-White for a helpful discussion about diamagnetic anisotropy effects.

Funding: This research was supported in part by the NSF through TeraGrid resources provided by the National Center for Supercomputing Applications, in part by the William and Janet Lycan fund of the University of Illinois to G.S.G., and in part by the Robert C. and Carolyn J. Springborn Fellowship fund to N.E.C. In addition, C.F.L. would like to acknowledge the support of the German-American Fulbright Commission and the Central European Summer Research Institute.

Author contributions: P.J.S. performed experiments and wrote the initial manuscript. B.B.T. performed experiments. C.F.L. performed theoretical modeling. N.E.C. validated results and finished the manuscript. G.S.G. conceptualized, funded, and supervised the experiments.

Competing interests: The authors declare that they have no competing interests.

Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.

Supplementary Materials

This PDF file includes:

Supplementary Text

Figs. S1 to S35

References

REFERENCES AND NOTES

  • 1.K. I. Goldberg, A. S. Goldman, Large-scale selective functionalization of alkanes. Acc. Chem. Res. 50, 620–626 (2017). [DOI] [PubMed] [Google Scholar]
  • 2.R. H. Crabtree, Alkane C–H activation and functionalization with homogeneous transition metal catalysts: A century of progress—A new millennium in prospect. J. Chem. Soc. Dalton Trans., 2437–2450 (2001). [Google Scholar]
  • 3.R. H. Crabtree, Organometallic alkane CH activation. J. Organomet. Chem. 689, 4083–4091 (2004). [Google Scholar]
  • 4.A. S. Goldman, K. I. Goldberg, Organometallic C—H bond activation: An introduction, in Activation and Functionalization of C—H Bonds, K. I. Goldberg, A. S. Goldman, Eds. (American Chemical Society Symposium Series, American Chemical Society, 2004), vol. 885, chap. 1, pp. 1–43. [Google Scholar]
  • 5.S. R. Golisz, T. Brent Gunnoe, W. A. Goddard, J. T. Groves, R. A. Periana, Chemistry in the center for catalytic hydrocarbon functionalization: An Energy Frontier Research Center. Catal. Lett. 141, 213–221 (2011). [Google Scholar]
  • 6.J. A. Labinger, J. E. Bercaw, Understanding and exploiting C–H bond activation. Nature 417, 507–514 (2002). [DOI] [PubMed] [Google Scholar]
  • 7.M. C. White, C–H Bond functionalization & synthesis in the 21st century: A brief history and prospectus. Synlett 23, 2746–2748 (2012). [Google Scholar]
  • 8.A. Sharma, J. F. Hartwig, Metal-catalysed azidation of tertiary C–H bonds suitable for late-stage functionalization. Nature 517, 600–604 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.J. E. Lyons, Selective oxidation of hydrocarbons via C-H bond activation by soluble and supported palladium catalysts. Catal. Today 3, 245–258 (1988). [Google Scholar]
  • 10.J. C. Lewis, R. G. Bergman, J. A. Ellman, Direct functionalization of nitrogen heterocycles via Rh-catalyzed C−H bond activation. Acc. Chem. Res. 41, 1013–1025 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.B. Li, P. H. Dixneuf, sp2 C–H bond activation in water and catalytic cross-coupling reactions. Chem. Soc. Rev. 42, 5744–5767 (2013). [DOI] [PubMed] [Google Scholar]
  • 12.H. Tsurugi, K. Yamamoto, H. Nagae, H. Kaneko, K. Mashima, Direct functionalization of unactivated C–H bonds catalyzed by group 3–5 metal alkyl complexes. Dalton Trans. 43, 2331–2343 (2014). [DOI] [PubMed] [Google Scholar]
  • 13.E. Nakamura, N. Yoshikai, Low-valent iron-catalyzed C−C bond formation⁠—addition, substitution, and C−H bond activation. J. Org. Chem. 75, 6061–6067 (2010). [DOI] [PubMed] [Google Scholar]
  • 14.D. A. Colby, A. S. Tsai, R. G. Bergman, J. A. Ellman, Rhodium catalyzed chelation-assisted C–H bond functionalization reactions. Acc. Chem. Res. 45, 814–825 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Y.-J. Liu, H. Xu, W.-J. Kong, M. Shang, H.-X. Dai, J.-Q. Yu, Overcoming the limitations of directed C–H functionalizations of heterocycles. Nature 515, 389–393 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.L. V. Desai, K. L. Hull, M. S. Sanford, Palladium-catalyzed oxygenation of unactivated sp3 C−H bonds. J. Am. Chem. Soc. 126, 9542–9543 (2004). [DOI] [PubMed] [Google Scholar]
  • 17.R. N. Saicic, Protecting group-free syntheses of natural products and biologically active compounds. Tetrahedron 70, 8183–8218 (2014). [Google Scholar]
  • 18.R. W. Hoffmann, Protecting-group-free synthesis. Synthesis 2006, 3531–3541 (2006). [Google Scholar]
  • 19.C. Hui, F. Chen, F. Pu, J. Xu, Innovation in protecting-group-free natural product synthesis. Nat. Rev. Chem. 3, 85–107 (2019). [Google Scholar]
  • 20.I. S. Young, P. S. Baran, Protecting-group-free synthesis as an opportunity for invention. Nat. Chem. 1, 193–205 (2009). [DOI] [PubMed] [Google Scholar]
  • 21.P. S. Baran, T. J. Maimone, J. M. Richter, Total synthesis of marine natural products without using protecting groups. Nature 446, 404–408 (2007). [DOI] [PubMed] [Google Scholar]
  • 22.R. N. Perutz, J. J. Turner, Photochemistry of the group 6 hexacarbonyls in low-temperature matrices. III. Interaction of the pentacarbonyls with noble gases and other matrices. J. Am. Chem. Soc. 97, 4791–4800 (1975). [Google Scholar]
  • 23.J. B. Turner, J. K. Burdett, R. N. Perutz, M. Poliakoff, Matrix photochemistry of metal carbonyls. Pure Appl. Chem. 49, 271–285 (1977). [Google Scholar]
  • 24.C. Hall, R. N. Perutz, Transition metal alkane complexes. Chem. Rev. 96, 3125–3146 (1996). [DOI] [PubMed] [Google Scholar]
  • 25.A. J. Cowan, M. W. George, Formation and reactivity of organometallic alkane complexes. Coord. Chem. Rev. 252, 2504–2511 (2008). [Google Scholar]
  • 26.R. D. Young, Characterisation of alkane σ-complexes. Chem. A Eur. J. 20, 12704–12718 (2014). [DOI] [PubMed] [Google Scholar]
  • 27.A. A. Shteinman, Coordinational activation of methane and other alkanes by metal complexes. Russ. Chem. Bull. 65, 1930–1944 (2016). [Google Scholar]
  • 28.A. S. Weller, F. M. Chadwick, A. I. McKay, Transition metal alkane-sigma complexes. Adv. Organomet. Chem. 66, 223–276 (2016). [Google Scholar]
  • 29.R. H. Crabtree, Sigma bonds as ligand donor groups in transition metal complexes, in The Chemical Bond III: 100 Years Old and Getting Stronger, D. M. P. Mingos, Ed. (Springer International Publishing, 2017), pp. 63–77. [Google Scholar]
  • 30.R. H. Crabtree, Alkane complexation, in Alkane Functionalization, A. J. L. Pombeiro, M. F. C. Guedes da Silva, Eds. (John Wiley & Sons, 2019), pp. 557–566. [Google Scholar]
  • 31.G. E. Ball, In situ photochemistry with NMR detection of organometallic complexes. Spectrosc. Prop. Inorg. Organomet. Compd. 41, 262–287 (2010). [Google Scholar]
  • 32.W. H. Bernskoetter, C. K. Schauer, K. I. Goldberg, M. Brookhart, Characterization of a rhodium(I) σ-methane complex in solution. Science 326, 553–556 (2009). [DOI] [PubMed] [Google Scholar]
  • 33.J. D. Watson, L. D. Field, G. E. Ball, Binding methane to a metal centre. Nat. Chem. 14, 801–804 (2022). [DOI] [PubMed] [Google Scholar]
  • 34.J. D. Watson, L. D. Field, G. E. Ball, [Fp(CH4)]+, [η5-CpRu(CO)2(CH4)]+, and [η5-CpOs(CO)2(CH4)]+: A complete set of group 8 metal–methane complexes. J. Am. Chem. Soc. 144, 17622–17629 (2022). [DOI] [PubMed] [Google Scholar]
  • 35.C. L. Gross, G. S. Girolami, Metal−alkane complexes. Rapid exchange of hydrogen atoms between hydride and methyl ligands in [(C5Me5)Os(dmpm)(CH3)H+]. J. Am. Chem. Soc. 120, 6605–6606 (1998). [Google Scholar]
  • 36.R. L. Martin, Hydrogen exchange between hydride and methyl ligands in [Cp*Os(dmpm)(CH3)H+]. J. Am. Chem. Soc. 121, 9459–9460 (1999). [Google Scholar]
  • 37.R. K. Szilagyi, D. G. Musaev, K. Morokuma, Hydrogen scrambling in [(C5Me5)Os(dmpm)(CH3)H]+. A density functional and ONIOM study. Organometallics 21, 555–564 (2002). [Google Scholar]
  • 38.C. F. Lovitt, N. E. Capra, R. J. Lastowski, G. S. Girolami, Steric and electronic analyses of ligand effects on the stability of σ-methane coordination complexes: A DFT study. Organometallics 41, 3834–3844 (2022). [Google Scholar]
  • 39.I. G. Phillips, R. G. Ball, R. G. Cavell, Synthesis and coordination chemistry (with platinum(ii) and molybdenum(0)) of new bis(bis(trifluoromethyl)phosphano)alkanes. Structure of a new bis(phosphano) methanide complex. Inorg. Chem. 27, 4038–4045 (1988). [Google Scholar]
  • 40.M. B. Murphy-Jolly, L. C. Lewis, A. J. M. Caffyn, The synthesis of tris(perfluoroalkyl)phosphines. Chem. Commun. 4479-4480, 4479 (2005). [DOI] [PubMed] [Google Scholar]
  • 41.C. L. Gross, J. L. Brumaghim, J. M. Jefferis, P. W. Dickinson, G. S. Girolami, C. W. Gribble, T. D. Tilley, Mono(η5-pentamethylcyclopentadienyl) complexes of osmium. Inorg. Synth. 36, 74–78 (2014). [Google Scholar]
  • 42.C. L. Gross, G. S. Girolami, Synthesis and characterization of osmium(II) compounds of stoichiometry (C5Me5)OsL2Br, (C5Me5)OsL2H, and (C5Me5)Os(NO)Br2. Organometallics 15, 5359–5367 (1996). [Google Scholar]
  • 43.J. S. Siegel, F. A. L. Anet, Dichlorofluoromethane-d: a versatile solvent for VT-NMR experiments. J. Org. Chem. 53, 2629–2630 (1988). [Google Scholar]
  • 44.P. M. Morse, M. D. Spencer, S. R. Wilson, G. S. Girolami, A static agostic ⍺-CH•••M interaction observable by NMR spectroscopy: Synthesis of the chromium(II) alkyl [Cr2(CH2SiMe3)6]2 and its conversion to the unusual "windowpane" bis(metallacycle) complex [Cr(κ2-C,C’-CH2SiMe2CH2)2]2. Organometallics 13, 1646–1655 (1994). [Google Scholar]
  • 45.Q. D. Shelby, W. Lin, G. S. Girolami, Metal-to-metal silyl migration and silicon−carbon bond cleavage/re-formation processes in the methylene/silyl complexes Cp*2Ru2(μ-CH2)(SiR3)(μ-Cl). Organometallics 18, 1904–1910 (1999). [Google Scholar]
  • 46.We prefer this nomenclature over that proposed in 1996 (24) because hapticity (η) refers to binding of contiguous atoms to a metal, whereas denticity (κ) refers to binding of non-contiguous atoms.
  • 47.R. H. Crabtree, The Organometallic Chemistry of the Transition Metals (Wiley, ed. 5, 2009), pp. 505. [Google Scholar]
  • 48.J. F. Hartwig, Organotransition Metal Chemistry: From Bonding to Catalysis (University Science Books, ed. 1, 2010), pp. 1127. [Google Scholar]
  • 49.R. B. Calvert, J. R. Shapley, ·Decacarbonyl(methyl)hydrotriosmium: NMR Evidence for a C••H••Os interaction. J. Am. Chem. Soc. 100, 7726–7727 (1978). [Google Scholar]
  • 50.S. Geftakis, G. E. Ball, Direct observation of a transition metal alkane complex, CpRe(CO)2(cyclopentane), using NMR spectroscopy. J. Am. Chem. Soc. 120, 9953–9954 (1998). [Google Scholar]
  • 51.D. J. Lawes, S. Geftakis, G. E. Ball, Insight into binding of alkanes to transition metals from NMR spectroscopy of isomeric pentane and isotopically labeled alkane complexes. J. Am. Chem. Soc. 127, 4134–4135 (2005). [DOI] [PubMed] [Google Scholar]
  • 52.Full derivations are provided in the Supplementary Materials.
  • 53.D. D. Desmarteau, M. Witz, N-Fluoro-bis(trifluoromethanesulfonyl)imide. An improved synthesis. J. Fluorine Chem. 52, 7–12 (1991). [Google Scholar]
  • 54.From the 1JCH values, we calculate that ΔE = 0.27(6) kcal mol−1, which is consistent with the more precise result calculated from the chemical shift data.
  • 55.M. Brookhart, M. L. H. Green, Carbon-hydrogen-transition metal bonds. J. Organomet. Chem. 250, 395–408 (1983). [Google Scholar]
  • 56.M. L. H. Green, G. Parkin, The covalent bond classification method and its application to compounds that feature 3-center 2-electron bonds, in The Chemical Bond III: 100 Years Old and Getting Stronger, D. M. P. Mingos, Ed. (Springer International Publishing, 2016), pp. 79–139. [Google Scholar]
  • 57.B. Cordero, V. Gómez, A. E. Platero-Prats, M. Revés, J. Echeverría, E. Cremades, F. Barragán, S. Alvarez, Covalent radii revisited. Dalton Trans., 2832–2838 (2008). [DOI] [PubMed] [Google Scholar]
  • 58.P. Garbacz, V. V. Terskikh, M. J. Ferguson, G. M. Bernard, M. Kędziorek, R. E. Wasylishen, Experimental characterization of the hydride 1H shielding tensors for HIrX2(PR3)2 and HRhCl2(PR3)2: Extremely shielded hydride protons with unusually large magnetic shielding anisotropies. J. Phys. Chem. A 118, 1203–1212 (2014). [DOI] [PubMed] [Google Scholar]
  • 59.G. E. Ball, T. A. Darwish, S. Geftakis, M. W. George, D. J. Lawes, P. Portius, J. P. Rourke, Characterization of an organometallic xenon complex using NMR and IR spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 102, 1853–1858 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.P. Hrobárik, V. Hrobáriková, F. Meier, M. Repiský, S. Komorovský, M. Kaupp, Relativistic four-component DFT calculations of 1H NMR chemical shifts in transition-metal hydride complexes: Unusual high-field shifts beyond the Buckingham–Stephens model. J. Phys. Chem. A 115, 5654–5659 (2011). [DOI] [PubMed] [Google Scholar]
  • 61.J. Vícha, J. Novotný, S. Komorovsky, M. Straka, M. Kaupp, R. Marek, Relativistic heavy-neighbor-atom effects on NMR shifts: Concepts and trends across the periodic table. Chem. Rev. 120, 7065–7103 (2020). [DOI] [PubMed] [Google Scholar]
  • 62.A. Bagno, G. Saielli, Relativistic DFT calculations of the NMR properties and reactivity of transition metal methane σ-complexes: Insights on C–H bond activation. Phys. Chem. Chem. Phys. 13, 4285–4291 (2011). [DOI] [PubMed] [Google Scholar]
  • 63.A. L. Van Geet, Calibration of methanol nuclear magnetic resonance thermometer at low temperature. Anal. Chem. 42, 679–680 (1970). [Google Scholar]
  • 64.Z. S. Novikova, A. A. Prishchenko, I. F. Lutsenko, Synthesis of tetrachloromethylenediphosphine and its derivatives. Zh. Obshch. Khim. 47, 775–781 (1977). [Google Scholar]
  • 65.L. Manojlović-Muir, I. R. Jobe, B. J. Maya, R. J. Puddephatt, Platinum(II) complexes with ligands (RO)2PCH2P(OR)2 (R = Me, Et, Ph, or C6H4Me-4); crystal structure of cis,cis-[Pt2Me4{μ-(EtO)2PCH2P(OEt)2}2]. J. Chem. Soc. Dalton Trans., 2117–2124 (1987). [Google Scholar]
  • 66.P. J. Sempsrott, “Low-temperature protonation studies of an electron-poor osmium methyl complex, and molecular precursors for the construction of graphene nanostructures with atomically precise edge structures,” thesis, University of Illinois at Urbana-Champaign (2018). [Google Scholar]
  • 67.H. H. Karsch, Alkylsubstituierte diphosphinomethane: Gemischte Cl/t-Bu- and Me/CH2SiMe3-derivate. Z. Naturforsch. 38b, 1027–1030 (1983). [Google Scholar]
  • 68.L. E. Mills, Method of Preparing Anhydrous Alkali Metal Phenoxides, U.S. Patent, 1955080 (1934).
  • 69.C. J. Smit, N. C. Van Der Koppel, Process for the Preparation of Di-alkyl Compounds of Group 2B Metals, GB Patent, 2268489A (1994).
  • 70.R. A. Bernheim, B. J. Lavery, Isotope shift in the proton magnetic resonance of CH4, CH3D, CH2D2, and CHD3. J. Chem. Phys. 42, 1464 (1965). [Google Scholar]
  • 71.B. Bennett, W. T. Raynes, C. W. Anderson, Temperature dependences of J(C,H) and J(C,D) in 13CH4 and some of its deuterated isotopomers. Spectrochim. Acta A Mol. Biomol. Spectrosc. 45, 821–827 (1989). [Google Scholar]

Associated Data

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

Supplementary Materials

Supplementary Text

Figs. S1 to S35

References


Articles from Science Advances are provided here courtesy of American Association for the Advancement of Science

RESOURCES