Abstract
gem-Hydrogenation of an internal alkyne with the aid of [Cp*RuCl]4 as the precatalyst is a highly unorthodox transformation, in which one C atom of the triple bond is transformed into a methylene group, whereas the second C atom gets converted into a ruthenium carbene. In the case of 1,3-enynes bearing a propargylic steering substituent as the substrates, the reaction occurs regioselectively, giving rise to vinyl carbene complexes that adopt interconverting η1/η3-binding modes in solution; a prototypical example of such a reactive intermediate was characterized in detail by spectroscopic means. Although both forms are similarly stable, only the η3-vinyl carbene proved kinetically competent to insert into primary, secondary, or tertiary C–H bonds on the steering group itself or another suitably placed ether, acetal, orthoester, or (sulfon)amide substituent. The ensuing net hydrogenative C–H insertion reaction is highly enabling in that it gives ready access to spirocyclic as well as bridged ring systems of immediate relevance as building blocks for medicinal chemistry. Moreover, the reaction scales well and lends itself to the formation of partly or fully deuterated isotopologues. Labeling experiments in combination with PHIP NMR spectroscopy (PHIP = parahydrogen induced polarization) confirmed that the reactions are indeed triggered by gem-hydrogenation, whereas kinetic data provided valuable insights into the very nature of the turnover-limiting transition state of the actual C–H insertion step.
Introduction
The ability to transfer both H atoms of H2 to the same C atom of an internal alkyne is a fundamentally new reactivity mode that was discovered only recently (Scheme 1).1 The conversion of the receiving C atom into a methylene group is accompanied by the formation of a discrete metal carbene at the adjacent position. For the time being, such “gem-hydrogention” reactions have been accomplished with [CpXRuCl] (or closely related metal fragments)2,3 as well as with [NHC(η6-cymene)RuCl2]; this latter system is photochemically driven and opens an unconventional entry into second-generation Grubbs-type catalysts.4,5 The use of [CpXRuCl], in contrast, provides access to piano-stool ruthenium carbene complexes, the electrophilic character of which manifests itself in multifarious reactivity (Scheme 1):6 so far, it could be harnessed in the form of hydrogenative cyclopropanation,3,7 hydrogenative metathesis,7,8 hydrogenative heterocycle syntheses,3,9 hydrogenative ring expansion reactions,3 and hydrogenative rearrangements,10−12 all of which are conceptually new types of transformations; some of them are even counterintuitive when judged on the basis of conventional chemical logic.13
Scheme 1. Concept of gem-Hydrogenation.
It was during a recent application of hydrogenative metathesis to the total synthesis of the marine natural product sinularone F that C–H insertion was observed as yet another possibility for the transient carbene to evolve (Scheme 2A).8 However, this then undesired side reaction infringed only in cases such as 1, in which the derived carbene complex D carried a neighboring ketone group. Even a flanking ester or dimethylamide did not suffice to upregulate the electrophilicity of the intermediate to the necessary extent; these compounds simply got reduced to alkene and alkane.8,14
Scheme 2. Lead Findings.
Although these preliminary data spoke for a narrow window of opportunity, they sparked our interest. Additional motivation was drawn from a literature survey, which showed that certain piano-stool ruthenium carbenes generated by an entirely different route, namely, carbene/alkyne metathesis (CAM),15−17 are capable of inserting into secondary or tertiary C–H bonds of suitably disposed acetals or ethers in moderate to good yields (Scheme 2B);18 the reaction was carefully studied by computational means.19−25 If one were able to generate the presumed vinylcarbenes of type E by gem-hydrogenation, one could avoid the use of hazardous diazoalkanes altogether.26 At the same time, it might be possible to enlarge the scope of the reaction to a considerable extent, since, in practice, the CAM-based route had worked well only for the addition of trimethylsilyl diazomethane to terminal alkynes (Scheme 2B);18,27 a hydrogenative approach should not face such limitations.
At the outset of our project, however, the goal of establishing a reasonably general hydrogenative C–H insertion protocol seemed (over)ambitious: 1,3-enynes are known to bind very tightly to [Cp*Ru] fragments and had proven problematic in the past in various other reactions effected by such catalysts;28−30 it was therefore not clear whether they are amenable to gem-hydrogenation at all. Not only was this proved to be the case, but the ensuing C–H insertion reactions turned out to be truly enabling. Most notably, they open access to (spirocyclic) building blocks of immediate relevance for medicinal chemistry. In parallel, the gathered mechanistic information brings the understanding for this type of transformation to a new level.
Results and Discussion
Reaction Development and Control Experiments
All it took was to subject model compound 5 carrying a tertiary propargylic ether substituent to the conditions previously optimized for other gem-hydrogenation reactions in order to convert this substrate into the tetrahydrofuran derivative 7 in high yield (Scheme 3).3,6,8 [Cp*RuCl]4 proved to be the catalyst of choice;31,32 the cationic complexes [CpRu(MeCN)3]PF6 (43%) and [Cp*Ru(MeCN)3]PF6 (16%) turned out to be much less efficient and were not studied any further at this point.33 The reaction was best carried out under hydrogen atmosphere (1 bar) in 1,2-dichloroethane (DCE) as the solvent at 70 °C to ensure reasonable rates. This favorable result shows that (i) suitably functionalized 1,3-enynes are indeed amenable to gem-hydrogenation, (ii) the reaction proceeds regioselectively to generate the required vinylcarbene 6, and (iii) this presumed intermediate is capable of inserting even into the primary C–H bond of the steering methyl ether. This latter aspect was deemed particularly encouraging since successful insertions of metal carbenes in general into aliphatic primary C–H bonds are rare34−38 and had been called a “remaining major challenge”.39 The specific literature precedent based on CAM had not reported any such example.18 This latter fact, however, might have solely been an oversight since CAM and gem-hydrogenation should pass through the same carbene. The direct comparison shown in Scheme 3 confirms this notion in that both reactions likely pass through a common intermediate G which evolves into the C–H insertion product 10. At the same time, however, it reveals a first significant advantage of the novel hydrogenative approach: since product 10 derived from enyne 8 was obtained as a single compound, the sequence of gem-hydrogenation/insertion must have proceeded stereospecifically, whereas CAM converts substrate 9 into an inseparable mixture of double-bond isomers.18
Scheme 3. Proof-of-Concept and Control Experiments.
Prior to exploring the scope of the reaction in more detail, several control experiments were carried out. Specifically, the labeled substrate [D3]-5 furnished product [D3]-7 exclusively. Even more compelling is the hydrogenation of unlabeled 5 with D2, which led to the gem-dideuterated tetrahydrofuran [D2]-7 as the only detectable product. These results are in excellent agreement with a reaction sequence consisting of initial gem-hydrogenation followed by C–H insertion. The lack of scrambling speaks against any hidden mechanistic complexity along the reaction coordinate; therefore it is reasonable to assume that this method provides access to all possible isotopologues of 7 by proper combination of substrate (5 versus [D3]-5) and reagent (H2 versus D2).
In addition to the information gathered by analysis of the products, reactions performed with parahydrogen (p-H2) allowed the spectral fingerprints of the transient intermediates formed under catalytic conditions to be detected by virtue of the exceptional sensitivity of PHIP-enhanced 1H NMR spectroscopy (PHIP = p-H2-induced polarization) (Figure 1);40−42 the “only parahydrogen spectroscopy (OPSY)” pulse sequence proved particularly convenient.43 The appearance of the hyperpolarized antiphase signals characteristic of a diastereotopic methylene unit is contingent on pairwise delivery of both H atoms of H2 to the same C atom of the alkyne, which, in turn, implies a (largely concerted) gem-hydrogenation step (insert b). Stacking of the spectra of this transient intermediate with those of the carbene complex formed using stoichiometric amounts of [Cp*RuCl] proved identity.
Figure 1.

The “only parahydrogen spectrum (OPSY)” of the carbene intermediate 6 generated by gem-hydrogenaton of 5 using p-H2 shows massive signal enhancement indicative of pairwise delivery of the geminal hydrogen atoms to the same C atom. Inserts: (a) ordinary 1H NMR (methylene signals) of complex 6 formed using stoichiometric [Cp*RuCl] (see below); (b) PHIP hyperpolarized antiphase signal of the methylene group of 6 generated under catalytic conditions with p-H2; (c) signals of the OPSY spectrum of 6 for comparison.
Scope
A set of appropriate substrates was readily attained by one of two methods (Scheme 4A): (i) Sonogahira-type cross coupling of a propargyl alcohol derivative with a suitable alkenyl halide (sulfonate) or (ii) addition of a lithiated enyne to a ketone followed by (in situ) alkylation of the resulting alkoxide with the alkyl halide of choice (for details, see the Supporting Information).
Scheme 4. Hydrogenative C–H Insertion.
The newly formed C–C bond is highlighted in red. [a] The structure of this compound in the solid state is contained in the Supporting Information. [b] Single diastereomer of unknown stereochemistry at the propenyl branch.
Most of these enynes proved amenable to hydrogenative C–H insertion under standard conditions. It is important to note that many of the products shown in Scheme 4B would be difficult, if not even impossible, to obtain by the CAM-based route.18 The fact that preconfiguration of the alkene in the substrate allowed compounds such as 10 and 18 to be obtained in isomerically pure form has already been pointed out in the context of the control experiments discussed above. An even more significant advantage is the fact that products 12, 14–17 and 19–24 would require hazardous diazomethane as carbene precursor if one were to use CAM for their synthesis, whereas the new route is simple, safe, and convenient. For this very reason, the reaction scales well, as illustrated by the preparation of 20, which was obtained in virtually the same yield independent of whether 23 mg (70%) and 1.53 g (72%) of product were made. It is unnecessary to reiterate that this (and any other) compound can also be formed in partially or fully labeled format. Use of perdeuteromethyl iodide and D2 as two of the cheapest deuterium sources, for example, allowed us to make [D5]-20. Since labeled compounds are of eminent importance in medicinal chemistry and elsewhere, this facile, flexible, and if necessary, scalable entry is arguably significant.44,45
CAM struggles when it comes to using nonterminal alkynes, since delivery of the primary carbene derived from the diazo derivative is then typically regiounselective;15,17,18 once again, the new gem-hydrogenative approach has no problem in providing access to such products as amply illustrated by Scheme 4B. Moreover, compounds 7, 11, and 13 comprising a cyclic alkene moiety would be basically inaccessible via CAM.
Although the current study capitalized on insertions into the arguably most challenging primary C–H bonds of methyl ethers, (more activated) secondary and tertiary C–H bonds are also amenable to the reaction (see 11, 23, and Schemes 6 and 7). Particularly noteworthy in this context is the cyclization of compound 25, in which the −OMe group is shifted away from the triple bond and the steering effect hence weak as had been shown in previous mechanistic studies.28,29 This aspect notwithstanding, a remarkably clean formation of cyclopentane 26 by kinetically favored regioselective insertion into the −CH2O– rather than the −OMe group was observed (Scheme 4C). This result suggests that the new method extends to carbocyclic rings and hence encourages further study (for further evidence, see the bicyclic products in Scheme 6).46
Scheme 6. Structural Diversity by Site-Selective C–H Insertion.
After desilylation of the crude product with TBAF.
Scheme 7. Formation of Pyrrolidine Derivatives by gem-Hydrogenation.
Limitations
Limitations were encountered in case of substrates bearing allylic groups. Although the unsaturated acetal derivative 21 was obtained in excellent yield (Scheme 4), other compounds comprising an allylic silyl ether, ester, carbamate, or silane substituent reacted unselectively or even failed completely (Figure 2A). As functional groups of these types are compatible with the reaction otherwise, it must be the allylic placement that opens competitive pathways which consume the precatalyst and/or decompose the substrates.47,48 Unsurprisingly perhaps, insertion of a ruthenium carbene generated by gem-hydrogenation into the C–H bond of a methyl ether does not outcompete cyclopropanation of a suitably placed olefin (Figure 2B).
Figure 2.
Representative examples illustrating current limitations of hydrogenative C–H insertion:(a) incomplete conversion; (b) complex mixture; (c) substrate recovered unchanged; (d) inseparable mixture of C–H insertion, cyclopropanation, and −OMe elimination products derived thereof
Our previous investigations had shown that alkyne gem-hydrogenation is mechanistically linked to trans-hydrogenation as a particularly facile competing process. For gem-hydrogenation to prevail, a tertiary −OR group at the propargylic position of the substrate is usually necessary.1−3 In line with this notion, enynes with secondary or primary propargylic ether substituents furnished product mixtures or were subject to trans-reduction only (Figure 2C).
The massive bias toward five-membered ring formation is a well-known hallmark of aliphatic C–H insertion in general.49−55 Indeed, first attempts at closing larger cycles by gem-hydrogenation have so far met with failure (Figure 2D). The same is true for intermolecular reactions; in assessing this aspect, however, one has to keep in mind that effective intermolecular trapping of vinylcarbenes generated via CAM by any reaction partner is basically unknown.15−17
The Acetal and Orthoester Series
A noteworthy extension of gem-hydrogenation pertains to enynes in which the steering substituent is part of an acetal or orthoester rather than a simple ether. Scheme 5 shows different ways of how the compliance of such substrates can be harnessed. Of arguably highest significance is the ability to form spiroketals from lactones in two straightforward and high-yielding steps via enyne addition/alkylation followed by gem-hydrogenative C–H insertion (see compounds 31a,b and 36); we are unaware of any precedent for this approach. Importantly, however, the concept is not limited to the formation of spirocycles, since bridged arrays such as 30 and 35 are equally within reach. When applied to an orthoester, the reaction provides access to butanolide derivatives such as 33 upon hydrolysis of the ortholactone 32 primarily formed.
Scheme 5. Hydrogenative C–H Insertion of Acetals and Orthoesters.
The product is acid sensitive.
C–H Insertion Selectivity
These examples show that the insertion into primary, secondary, and tertiary C–H bonds proceeds with similar efficiency. It was therefore interesting to study whether substrates comprising more than one potentially reactive site are amenable to regioselective C–H insertion.
At first sight, the gem-hydrogenation with formation of product 31b seems to follow the conventional order of C–H insertions35−39 because the weaker tertiary C–H bond of the R = OiPr group rather than the endocyclic −OCH2– unit has reacted (Scheme 5). The analogous substrate 28a (R = OMe) leads to the same spirocyclic scaffold in 31a: it is important to note that this outcome mandates a drastic switch in selectivity in that exclusive insertion into the a priori less activated primary C–H bond of the methyl group must have taken place. This striking dichotomy suggests that the course of the reactions cannot be rationalized on thermodynamic grounds. In line with this notion, enyne 38 afforded spirocycle 15 as the only detectable product (Scheme 6), once again by exclusive insertion of the transient ruthenium carbene into the primary C–H bond of the steering −OMe substituent. The analogous reactive intermediate derived from 39, however, gave rise to the bridged bicycle 40 in similar yield. This result proves that there is no inherent problem in engaging the methylene group of the pre-existing heterocycle into bond formation.56
Taken together, these results suggest that the observed selectivities are largely kinetic in origin, which, in turn, implies that the actual C–H insertion step must have a strong steric component to it, likely as a result of the bulky ancillary Cp* ligand. The C–H bond of a “tangling” and hence freely rotatable −OR substituent will align faster with the reactive [C=Ru] unit than a slightly more rigid cyclic array. In any case, the ability to produce noticeably different skeletons from a single precursor solely by switching a protecting group, as manifested in the couples 15/40, 14/43, 41/44, and 42/45, is deemed a significant asset (Scheme 6). Scaffolds of these types are prominently featured in contemporary medicinal chemistry as manifested in innumerous patents; the examples shown in Figure 3 are representative.
Figure 3.
Spirocyclic scaffolds commonly used in medicinal chemistry.
Propargyl Amides
As expected, the scope of the reaction extends beyond propargyl ethers and acetals (Scheme 7). Specifically, tert-amide derivatives such as 46 proved well behaved, even though the very nature of the amide group does affect the yield of the resulting pyrrolidine derivative (compare 47/48).
The exclusive formation of the cis-configured compounds 49 and 50 is another remarkable feature; actually, product 49 is remotely related to kainic acid and related neuroexcitatory agents, the structures of which have been edited in countless ways.57 Hemiaminal 50 is a valuable N-sulfonyliminium ion surrogate. In the absence of external nucleophiles, catalytic HCl converts it into the functionalized 1,3-diene building block 51 in readiness for use in Diels–Alder cycloadditions. In line with the results outlined above, a Thorpe–Ingold effect in the substrate is also necessary in this series to ensure efficient C–H insertion.58
Exemplary Downstream Functionalization
During the past decade, small ring systems in general and spirocyclic scaffolds in particular gained prominence as novel types of building blocks for medicinal chemistry.59−63 Replacement of traditional flat (hetero)aromatic cores of drug candidates by three-dimensional and sp3-rich templates can be largely beneficial: if properly chosen, they ensure optimal display of attached functionality toward reciprocal groups in the binding site of the targeted biological receptor; moreover, they provide potential advantages with regard to metabolic stability, often lead to reduced lipophilicity as compared to (hetero)arenes, open uncommon or even uncharted chemical and pharmacological space, and hence provide many opportunities for innovation and therapeutic advances.59−63
The ease with which gem-hydrogenation brings such compounds into reach even on a larger scale encouraged us to briefly explore their downstream functionalization. Compound 20 was chosen as the model substrate since its isopropenyl substituent provides a versatile handle (Scheme 8). While the cleavage of the double bond by ozonolysis with formation of 52 is an obvious possibility, some other transformations are more involved.64,65 Specifically, a cobalt-catalyzed hydration furnished the tertiary alcohol derivative 53; rather than trapping the transient radical with oxygen, an intermediate of this type can also be engaged in 1,4-addition reactions to, for example, ethyl acrylate as illustrated by the formation of 54.66 The other entries illustrate the possibility of iron mediated dealkenylative oxidation with formation of either the valuable TEMPO-adduct 55 or ketone 56.67,68 Sulfone 57 further illustrates the structural and functional diversity accessible from a single such platform.69,70 In this context, it is of note that compound 56 is a commercially available yet expensive building block,71−73 which the new route is able to deliver on scale and, if desirable, in labeled format. Extrapolation of the chemistry shown in Scheme 8 to the other (spirocyclic) products containing isopropenyl (or related alkenyl) substituents described above should give access to a multitude of valuable scaffolds for medicinal chemistry and chemical biology.
Scheme 8. Downstream Functionalization.
Reagents and conditions: (a) (i) O3, CH2Cl2, −78 °C; (ii) Me2S, −78 °C → RT, 76%; (b) Co(acac)3 (25 mol %), PhSiH3, O2, THF, 69%; (c) Fe(acacc)3, PhSiH3, methyl acrylate, 1,2-dichloroethane, ethylene glycol, 60 °C, 74%; (d) (i) O3, MeOH, −78 °C; (ii) TEMPO, FeSO4·7H2O, −78 °C → RT, 63%; (e) (i) O3, MeOH, −78 °C; (ii) TEMPO, FeSO4·7H2O, magnesium bis(monoperoxyphthalate) hexahydrate (MMPP), −78 °C → RT, 62%; (f) (i) O3, MeOH, −78 °C; (ii) PhSSPh, −78 °C → 0 °C; FeSO4·7H2O; (g) mCPBA, CH2Cl2, 45% (over both steps).
Reactive Intermediates
As mentioned in the Introduction, gem-hydrogenation empowers various types of transformations including cyclopropanation, metathesis, skeletal rearrangements, and various heterocycle syntheses;1−12 C–H insertion was late to appear on scene.8 In view of this history, the ease of the reaction, as manifested in the examples outlined above, is all the more striking. It hence seemed possible that the ability to insert into C–H bonds might be a peculiarity of vinylcarbenes or α-oxocarbenes, which had led to the initial discovery (see Scheme 2).8
The comparison of substrates differing only in the degree of unsaturation of the flanking ring confirmed this notion (Scheme 9): only compound 5 which gives rise to the transient piano-stool ruthenium vinyl carbene intermediate 6 furnished the corresponding tetrahydrofuran derivative 7; the two other compounds pass through the “ordinary” piano-stool carbenes 59 and 61 that proved incompetent in C–H insertion; rather, they evolve into the corresponding trans-alkenes and partial over-reduction products (for details, see the Supporting Information).3
Scheme 9. C–H Insertion Downstream of gem-Hydrogenation Is Contingent on the Presence of a Vinylcarbene.
Reagents and conditions: (a) [Cp*RuCl]4 (2 mol %), H2 (1 bar), 1,2-dichloroethane, 70 °C.
To gain better understanding for why that is so, the gem-hydrogenation of 5 was repeated with a stoichiometric amount of [Cp*RuCl], as this might allow the structure and reactivity of the resulting vinylcarbene intermediate 6 to be studied in more detail. The reaction proceeded smoothly at 0 °C; the resulting crude material consisted of ∼85% of the expected complex as judged by NMR. The fact that the gem-hydrogenation occurs rapidly even at this low temperature whereas all catalytic reactions described above required gentle heating suggests that the turnover-limiting step must be later in the catalytic cycle (see below).
Complex 6 proved too unstable for isolation in crystalline form, but the spectral data are highly informative (Scheme 10). Specifically, two different isomers are present in solution in a ratio of ≈3:2; the barrier for interconversion is on the order of only ≈10.4 kcal·mol–1 as deduced by VT-1H NMR spectroscopy (for details, see the Supporting Information). At −80 °C, both forms are frozen out and all relevant signals well resolved to allow for an unambiguous assignment.
Scheme 10. Spectral Data of the Carbene Intermediate: Comparison with Literature Data.
The slightly preferred species is distinguished by η3-binding of the vinylcarbene unit to the ruthenium atom (η3-6). This coordination mode is manifested in the characteristic 13C NMR signals of the bound alkene (C6, C7) at δC = 90.7, 72.3 ppm; an upfield shift of almost 70 ppm relative to the alkene signals of the second isomer indicates a strong electronic communication between the π-bond and the metal center.74 The second isomer is an ordinary η1-vinylcarbene (η1-6), the double bond of which is slightly polarized but otherwise largely unperturbed by the neighboring carbene site as indicated by the resonances of C6 and C7 at δC = 159.0 and 139.3 ppm, respectively. Its ruthenium center likely reaches the 18e count by a supporting Lewis acid/base interaction with the ether moiety.75
The distinctly different shifts of the carbene centers of these two interconverting isomers are equally significant. With δC = 308.3 ppm, η1-6 falls in the typical range previously observed for other piano-stool ruthenium carbene complexes with lateral Ru···O bonding such as 64;2,3 no such derivative has been found competent in C–H insertion reactions. In sharp contrast, η3-6 has the carbene resonance at δC = 288.8 ppm, which speaks for a notably different electronic character. In this regard, it is quite similar to the so far only structurally characterized α-oxocarbene complex 63 (δC = 295.5 ppm).3 The furyl carbene species 62, the η3-binding mode of which has been crystallographically proven, shows an even stronger upfield shift (δC = 266.1 ppm);9 this complex is likely kept from undergoing intramolecular C–H insertion into the methyl ether substituent by the high strain of the four-membered ring that would ensue.
These experimental and spectroscopic data concur with the conclusions drawn by Saá and co-workers, who studied the fate of ruthenium vinylcarbene intermediates generated by CAM in silico.19,20 These authors suggested that the η3-bound isomer accounts for the downstream chemistry; it draws its higher reactivity from an out-of-plane distortion that causes an electronic perturbation and, at the same time, renders the site sterically more exposed.
The Actual C–H Insertion Step
The ability to form 6 in fairly high purity allowed us to study the fate of this reactive intermediate by NMR spectroscopy. In the temperature range from +15 °C to +50 °C, the decay follows a first order rate law (ΔG‡ (25 °C) = 23 kcal·mol–1), as expected for an intramolecular C–H insertion (for details, see the Supporting Information). The experiments were repeated with [D3]-6 carrying a perdeuterated methyl ether; comparison of the derived rate constants allowed the kinetic isotope effect (KIE) and its temperature-dependence to be accurately determined (Table 1; for the full data set, see the Supporting Information). With values on the order of 3.3–3.9, determined by measuring two separate rate constants, it is clear that C–H insertion occurs during the rate-determining step.76,77
Table 1. Kinetic Data for the Transformation of 6 or [D3]-6 into 7 or [D3]-7, Respectively: Putative Transition State Connecting Vinylcarbene 6 and Product 7.
Importantly, these data also allow valuable
information concerning
the actual transition state (TS) to be deduced. In an insightful study,
Kwart had analyzed four different extremes for three-center processes
in general and found that each TS geometry has a characteristic footprint
manifested in a data-triple consisting of the actual KIE (kH/kD, at 25 °C),
the difference in activation energy [ΔEA]
, and the quotient of the pre-exponential
terms of the Arrhenius equations (AH/AD).78 This formalism
applies to the interaction between a carbene center and an incoming
C–H bond. In the present case, the temperature-dependent KIE
and a characteristic data-triple (kH/kD (25 °C) = 3.72; [ΔEA]
= 0.82 kcal·mol–1; AH/AD =
0.94) are in excellent agreement with a linear, unsymmetrical H-transfer
process. In other words, the electrophilic metal carbene center is
attacked by the hydrogen atom, which develops hydridic character in
a transition state distinguished by a very obtuse (in the extreme:
linear) C–H–C angle (see the insert in Table 1).79,80 The recorded data rule out a more Dewar–Chatt–Duncanson-like
scenario, in which the metal concomitantly interacts with the σ/σ*
orbitals of the C–H bond, which would require a “bent”
geometry (in the extreme a coplanar orientation of the carbene and
the C–H bond to be broken); in such a case, the KIE is expected
to be basically temperature-independent.78
Once again, the conclusions drawn from our experimental data tally well with the computational results of Saá and co-workers, who suggested that the vinylcarbene species formed by CAM evolve via a “hydride transfer” mechanism.19,20 In addition, our own computations of the C–H insertion pathway populated by ruthenium α-oxocarbene complexes such as D (Scheme 2) showed a very obtuse angle between the carbene center and the incoming reaction partner.81
Conclusions
gem-Hydrogenation is a conceptually novel mode of H2 transfer to an organic substrate, which our group was able to discover after a century of intense research devoted to catalytic hydrogenation in innumerous academic as well as industrial laboratories. The present study showed that 1,3-enynes bearing an appropriate propargylic substituent are amenable to this process using [Cp*RuCl]4 as the catalyst. The resulting piano-stool ruthenium vinylcarbene intermediates adopt interconverting η1- and η3-binding modes, which are easily distinguished by virtue of their markedly different spectral fingerprints. Although these two forms are similarly stable, only the η3-isomer is competent to insert into primary, secondary, or tertiary C–H bonds of suitably disposed ethers, acetals, or N-alkylated (sulfon)amide derivatives. The ensuing reaction is highly enabling in preparative terms; most notably, it provides ready access to spirocyclic as well as bridged ring systems of immediate relevance as building blocks for medicinal chemistry and chemical biology. The method scales well and lends itself to the preparation of deuterated isotopologues. This novel hydrogenative C–H insertion process hence provides a notable addendum to the growing list of reactions exploiting gem-hydrogenation as a means to generate reactive intermediates and augurs well for further explorations of this field of research.
Acknowledgments
Generous financial support by the Max-Planck-Gesellschaft and Spanish Ministry of Universities—FPU Program (fellowship for A.G.G., Grant FPU17/00711) is gratefully acknowledged. We thank Dr. R. Weiss, G. Späth, and K. Radkowski for assistance, C. Wirtz for help with the analysis of various NMR spectra, and the analytical departments of our Institute for excellent support.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.1c13446.
Experimental Section containing supporting NMR data (PDF)
Accession Codes
CCDC 2106229 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
Open access funded by Max Planck Society.
The authors declare the following competing financial interest(s): Patent applied for.
Supplementary Material
References
- Fürstner A. trans-Hydrogenation, gem-Hydrogenation, and trans-Hydrometalation of Alkynes: An Interim Report on an Unorthodox Reactivity Paradigm. J. Am. Chem. Soc. 2019, 141 (1), 11–24. 10.1021/jacs.8b09782. [DOI] [PubMed] [Google Scholar]
- Leutzsch M.; Wolf L. M.; Gupta P.; Fuchs M.; Thiel W.; Farès C.; Fürstner A. Formation of Ruthenium Carbenes by gem-Hydrogen Transfer to Internal Alkynes: Implications for Alkyne trans-Hydrogenation. Angew. Chem., Int. Ed. 2015, 54 (42), 12431–12436. 10.1002/anie.201506075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guthertz A.; Leutzsch M.; Wolf L. M.; Gupta P.; Rummelt S. M.; Goddard R.; Farès C.; Thiel W.; Fürstner A. Half-Sandwich Ruthenium Carbene Complexes Link trans-Hydrogenation and gem-Hydrogenation of Internal Alkynes. J. Am. Chem. Soc. 2018, 140 (8), 3156–3169. 10.1021/jacs.8b00665. [DOI] [PubMed] [Google Scholar]
- Biberger T.; Zachmann R. J.; Fürstner A. Grubbs Metathesis Enabled by a Light-Driven gem-Hydrogenation of Internal Alkynes. Angew. Chem., Int. Ed. 2020, 59 (42), 18423–18429. 10.1002/anie.202007030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zachmann R. J.; Fürstner A. Light-Driven gem-Hydrogenation: An Orthogonal Entry into “Second-Generation” Ruthenium Carbene Catalysts for Olefin Metathesis. Chem.—Eur. J. 2021, 27 (28), 7663–7666. 10.1002/chem.202101176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Biberger T.; Gordon C. P.; Leutzsch M.; Peil S.; Guthertz A.; Copéret C.; Fürstner A. Alkyne gem-Hydrogenation: Formation of Pianostool Ruthenium Carbene Complexes and Analysis of Their Chemical Character. Angew. Chem., Int. Ed. 2019, 58, 8845–8850. 10.1002/anie.201904255. [DOI] [PubMed] [Google Scholar]
- Peil S.; Guthertz A.; Biberger T.; Fürstner A. Hydrogenative Cyclopropanation and Hydrogenative Metathesis. Angew. Chem., Int. Ed. 2019, 58, 8851–8856. 10.1002/anie.201904256. [DOI] [PubMed] [Google Scholar]
- Peil S.; Bistoni G.; Goddard R.; Fürstner A. Hydrogenative Metathesis of Enynes via Piano-Stool Ruthenium Carbene Complexes Formed by Alkyne gem-Hydrogenation. J. Am. Chem. Soc. 2020, 142 (43), 18541–18553. 10.1021/jacs.0c07808. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peil S.; Fürstner A. Mechanistic Divergence in the Hydrogenative Synthesis of Furans and Butenolides: Ruthenium Carbenes Formed by gem-Hydrogenation or via Carbophilic Activation of Alkynes. Angew. Chem., Int. Ed. 2019, 58, 18476–18481. 10.1002/anie.201912161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Song L.; Feng Q.; Wang Y.; Ding S.; Wu Y. D.; Zhang X.; Chung L. W.; Sun J. Ru-Catalyzed Migratory Geminal Semihydrogenation of Internal Alkynes to Terminal Olefins. J. Am. Chem. Soc. 2019, 141, 17441–17451. 10.1021/jacs.9b09658. [DOI] [PubMed] [Google Scholar]
- For closely related rearrangements by gem-hydroboration of internal alkynes, see the following:; Feng Q.; Wu H.; Li X.; Song L.; Chung L. W.; Wu Y.-D.; Sun J. Ru-Catalyzed Geminal Hydroboration of Silyl Alkynes via a New gem-Addition Mechanism. J. Am. Chem. Soc. 2020, 142 (32), 13867–13877. 10.1021/jacs.0c05334. [DOI] [PubMed] [Google Scholar]
- Biberger T.; Hess S. N.; Leutzsch M.; Fürstner A. Hydrogenative Cycloisomerization and Sigmatropic Rearrangement Reactions of Cationic Ruthenium Carbenes Formed by Catalytic Alkyne gem-Hydrogenation. Angew. Chem., Int. Ed. 2022, 61, e202113827. 10.1002/anie.202113827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cyclopropanes, for example, are usually cleaved but not formed by catalytic hydrogenolysis.
- An additional unsuccessful example of merging gem-hydrogenation with C–H insertion is reported in ref (3).
- Padin D.; Varela J. A.; Saá C. Recent Advances in Ruthenium-Catalyzed Carbene/Alkyne Metathesis (CAM) Transformations. Synlett 2020, 31, 1147–1157. 10.1055/s-0039-1690861. [DOI] [Google Scholar]
- Dey S.; De Sarkar S. Synthetic Applications of Vinyl Ruthenium Carbenes Derived from Diazoalkanes and Alkynes. Adv. Synth. Catal. 2017, 359, 2709. 10.1002/adsc.201700622. [DOI] [Google Scholar]
- Pei C.; Zhang C.; Qian Y.; Xu X. Catalytic Carbene/Alkyne Metathesis (CAM): A Versatile Strategy from Alkyne Bifunctionalization. Org. Biomol. Chem. 2018, 16, 8677. 10.1039/C8OB02420K. [DOI] [PubMed] [Google Scholar]
- Cambeiro F.; López S.; Varela J. A.; Saá C. Cyclization by Cationic Ruthenium Carbene Insertion into Csp3-H Bonds. Angew. Chem., Int. Ed. 2012, 51, 723–727. 10.1002/anie.201107344. [DOI] [PubMed] [Google Scholar]
- Cambeiro F.; Martínez-Núñez E.; Varela J. A.; Saá C. DFT and Kinetic Monte Carlo Study of TMS-Substituted Ruthenium Vinyl Carbenes: Key Intermediates for Stereoselective Cyclizations. ACS Catal. 2015, 5, 6255–6262. 10.1021/acscatal.5b01333. [DOI] [Google Scholar]
- Padin D.; Varela J. A.; Saá C. Cp*RuCl-Vinyl Carbenes: Two Faces and the Bifunctional Role in Catalytic Processes. Chem. - Eur. J. 2020, 26, 7470–7478. 10.1002/chem.202000391. [DOI] [PubMed] [Google Scholar]
- For yet other reactions of piano-stool ruthenium carbene complexes, see refs (22−25) and the following:; Monnier F.; Vovard-Le Bray C.; Castillo D.; Aubert V.; Derien S.; Dixneuf P. H.; Toupet L.; Ienco A.; Mealli C. Selective Ruthenium-Catalyzed Transformations of Enynes with Diazoalkanes into Alkenylbicyclo][3.1.0]hexanes. J. Am. Chem. Soc. 2007, 129, 6037–6049. 10.1021/ja0700146. [DOI] [PubMed] [Google Scholar]
- Le Paih J.; Vovard-Le Bray C.; Dérien S.; Dixneuf P. H. Ruthenium-Catalyzed Synthesis of Functional Conjugated Dienes via Addition of Two Carbene Units to Alkynes. J. Am. Chem. Soc. 2010, 132, 7391–7397. 10.1021/ja101064b. [DOI] [PubMed] [Google Scholar]
- Cambeiro F.; López S.; Varela J. A.; Saá C. Vinyl Dihydropyrans and Dihydrooxazines: Cyclizations of Catalytic Ruthenium Carbenes Derived from Alkynals and Alkynones. Angew. Chem., Int. Ed. 2014, 53 (23), 5959–5963. 10.1002/anie.201400675. [DOI] [PubMed] [Google Scholar]
- Tortoreto C.; Achard T.; Zeghida W.; Austeri M.; Guénée L.; Lacour J. Enol Acetal Synthesis through Carbenoid C-H Insertion into Tetrahydrofuran Catalyzed by CpRu Complexes. Angew. Chem., Int. Ed. 2012, 51, 5847. 10.1002/anie.201201541. [DOI] [PubMed] [Google Scholar]
- Tortoreto C.; Achard T.; Egger L.; Guénée L.; Lacour J. Synthesis of Spiro Ketals, Orthoesters, and Orthocarbonates by CpRu-Catalyzed Decomposition of α-Diazo-β-ketoesters. Org. Lett. 2016, 18 (2), 240–243. 10.1021/acs.orglett.5b03380. [DOI] [PubMed] [Google Scholar]
- For the use of tosylhydrazones as carbene precursors for CAM/C–H insertion cascades catalyzed by dirhodium paddlewheel complexes, see the following:; Le P. Q.; May J. A. Hydrazone-Initiated Carbene/Alkyne Cascades to Form Polycyclic Products: Ring-Fused Cyclopropenes as Mechanistic Intermediates. J. Am. Chem. Soc. 2015, 137 (38), 12219–12222. 10.1021/jacs.5b08157. [DOI] [PubMed] [Google Scholar]
- A single example of the use of ethyl diazaoacetate was also shown, which gave a modest yield and was not diastereoselective either.
- Rummelt S. M.; Radkowski K.; Roşca D.-A.; Fürstner A. Interligand Interactions Dictate the Regioselectivity of trans-Hydrometalations and Related Reactions Catalyzed by [Cp*RuCl]. Hydrogen Bonding to a Chloride Ligand as a Steering Principle in Catalysis. J. Am. Chem. Soc. 2015, 137 (16), 5506–5519. 10.1021/jacs.5b01475. [DOI] [PubMed] [Google Scholar]
- Roşca D.-A.; Radkowski K.; Wolf L. M.; Wagh M.; Goddard R.; Thiel W.; Fürstner A. Ruthenium-Catalyzed Alkyne trans-Hydrometalation: Mechanistic Insights and Preparative Implications. J. Am. Chem. Soc. 2017, 139 (6), 2443–2455. 10.1021/jacs.6b12517. [DOI] [PubMed] [Google Scholar]
- Schaubach S.; Michigami K.; Fürstner A. Hydroxyl-Assisted trans-Reduction of 1,3-Enynes: Application to the Formal Synthesis of (+)-Aspicilin. Synthesis 2016, 49, 202–208. 10.1055/s-0035-1562381. [DOI] [Google Scholar]
- [Cp*RuCl]4 readily disassembles in solution in the presence of an alkyne substrate to give monomeric [Cp*RuCl] as the actual catalytically active fragment. For the preparation of this complex, see the following:; Fagan P. J.; Ward M. D.; Calabrese J. C. Molecular Engineering of Solid-state Materials: Organometallic Building Blocks. J. Am. Chem. Soc. 1989, 111 (5), 1698–1719. 10.1021/ja00187a024. [DOI] [Google Scholar]
- Fagan P. J.; Mahoney W. S.; Calabrese J. C.; Williams I. D. Structure and Chemistry of the Complex Tetrakis(η5-pentamethylcyclopentadienyl)tetrakis(μ3-chloro)tetra-ruthenium(II): A Useful Pprecursor to (Pentamethylcyclopenta-dienyl)ruthenium(0), -(II), and -(IV) Complexes. Organometallics 1990, 9 (6), 1843–1852. 10.1021/om00156a025. [DOI] [Google Scholar]
- This fact currently prevents a systemtatic exploration of enantioselective C–H insertion via gem-hydrogenation. Very few structurally defined chloride-bridged chiral (and even achiral) complexes comprising substituted CpR rings are known, since the established synthesis routes usually give access to cationic complexes only; compare ref (6) and literature cited therein. Work trying to overcome this important limitation of ruthenium coordination chemistry is in progress.
- Lo V. K.-Y.; Guo Z.; Choi M. K.-W.; Yu W.-Y.; Huang J.-S.; Che C.-M. Highly Selective Intramolecular Carbene Insertion into Primary C-H Bond of α-Diazoacetamides Mediated by a (p-Cymene)ruthenium(II) Carboxylate Complex. J. Am. Chem. Soc. 2012, 134 (18), 7588–7591. 10.1021/ja3006989. [DOI] [PubMed] [Google Scholar]
- For general reviews, see refs (36−39) and the following:; Doyle M. P.; Duffy R.; Ratnikov M.; Zhou L. Catalytic Carbene Insertion into C-H Bonds. Chem. Rev. 2010, 110, 704–724. 10.1021/cr900239n. [DOI] [PubMed] [Google Scholar]
- Davies H. M. L.; Manning J. R. Catalytic C-i-H Functionalization by Metal Carbenoid and Nitrenoid Insertion. Nature 2008, 451 (7177), 417–424. 10.1038/nature06485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giri R.; Shi B.-F.; Engle K. M.; Maugel N.; Yu J.-Q. Transition Metal-catalyzed C-H Activation Reactions: Diastereoselectivity and Enantioselectivity. Chem. Soc. Rev. 2009, 38 (11), 3242–3272. 10.1039/b816707a. [DOI] [PubMed] [Google Scholar]
- Baudoin O. In Science of Synthesis: Catalytic Transformations via C—H Activation; Yu J.-Q., Ed.; Thieme Chemistry, 2015; Vol. 2, p 37. [Google Scholar]
- Cui X.; Zhang X. P. In Science of Synthesis: Catalytic Transformations via C—H Activation; Yu J.-Q., Ed.; Thieme Chemistry, 2015; Vol. 2, p 63. [Google Scholar]
- Bowers C. R.; Weitekamp D. P. Parahydrogen and Synthesis Allow Dramatically Enhanced Nuclear Alignment. J. Am. Chem. Soc. 1987, 109 (18), 5541–5542. 10.1021/ja00252a049. [DOI] [Google Scholar]
- Duckett S. B.; Mewis R. E. Application of Parahydrogen Induced Polarization Techniques in NMR Spectroscopy and Imaging. Acc. Chem. Res. 2012, 45 (8), 1247–1257. 10.1021/ar2003094. [DOI] [PubMed] [Google Scholar]
- Eisenberg R. Parahydrogen-induced Polarization: A New Spin on Reactions with Molecular Hydrogen. Acc. Chem. Res. 1991, 24 (4), 110–116. 10.1021/ar00004a004. [DOI] [Google Scholar]
- Aguilar J. A.; Elliott P. I. P.; López-Serrano J.; Adams R. W.; Duckett S. B. Only para-Hydrogen Spectroscopy (OPSY), a Technique for the Selective Observation of para-Hydrogen Enhanced NMR Signals. Chem. Commun. 2007, (11), 1183–1185. 10.1039/B616307F. [DOI] [PubMed] [Google Scholar]
- Atzrodt J.; Derdau V.; Kerr W. J.; Reid M. Deuterium- and Tritium-Labelled Compounds: Applications in the Life Sciences. Angew. Chem., Int. Ed. 2018, 57 (7), 1758–1784. 10.1002/anie.201704146. [DOI] [PubMed] [Google Scholar]
- Pirali T.; Serafini M.; Cargnin S.; Genazzani A. A. Applications of Deuterium in Medicinal Chemistry. J. Med. Chem. 2019, 62 (11), 5276–5297. 10.1021/acs.jmedchem.8b01808. [DOI] [PubMed] [Google Scholar]
- Preliminary results show that the Thorpe–Ingold effect exerted by the gem-dimethyl substituents in 25 is mandatory for success.
- Fourteen-electron [CpRRuX] fragments readily undergo oxidative insertion into various allylic derivatives. See ref (48) and the following:; Nagashima H.; Mukai K.; Shiota Y.; Yamaguchi K.; Ara K.; Fukahori T.; Suzuki H.; Akita M.; Morooka Y.; Itoh K. Oxidative Addition of Allylic Halides to Ruthenium(II) Compounds. Preparation, Reactions, and X-ray Crystallographic Structure of Ruthenium(IV)-Allyl Complexes. Organometallics 1990, 9 (3), 799–807. 10.1021/om00117a041. [DOI] [Google Scholar]
- Albers M. O.; Liles D. C.; Robinson D. J.; Shaver A.; Singleton E. Chemistry of Cyclopentadienylruthenium and -Osmium Complexes. 4. The Facile Synthesis of Cyclopentadienylruthenium(IV) and -Osmium(IV) Allyl Complexes and the Characterization of a Novel Ruthenium(IV) Intermediate in the Dehydrohalogenation and Dehydrogenation of 3-Bromocyclohexene. The X-ray Crystal Structure of [(η5-C5H5)Ru(η3-C4H4OMe)Cl2]. Organometallics 1987, 6 (11), 2347–2354. 10.1021/om00154a014. [DOI] [Google Scholar]
- The formation of six-membered rings by insertion into aliphatic C–H bonds is exceedingly rare even with conventional carbene sources. For pertinent reviews see refs (38) and (39) and the following:; Qiu D.; Wang J.. Recent Developments of Diazo Compounds in Organic Synthesis; World Scientific: London, 2021. [Google Scholar]
- The few exceptions use metals other than ruthenium and pass through electronically and/or conformationally highly biased carbene intermediates. For leading refernces, see refs (51−55) and the following:; John J. P.; Novikov A. V. Selective Formation of Six-Membered Cyclic Sulfones and Sulfonates by C-H Insertion. Org. Lett. 2007, 9 (1), 61–63. 10.1021/ol062592h. [DOI] [PubMed] [Google Scholar]
- Wolckenhauer S. A.; Devlin A. S.; Du Bois J. δ-Sultone Formation Through Rh-Catalyzed C-H Insertion. Org. Lett. 2007, 9 (21), 4363–4366. 10.1021/ol701950d. [DOI] [PubMed] [Google Scholar]
- Flynn C. J.; Elcoate C. J.; Lawrence S. E.; Maguire A. R. Highly Enantioselective Intramolecular Copper Catalyzed C-H Insertion Reactions of α-Diazosulfones. J. Am. Chem. Soc. 2010, 132 (4), 1184–1185. 10.1021/ja909713a. [DOI] [PubMed] [Google Scholar]
- Taber D. F.; Paquette C. M.; Gu P.; Tian W. Cyclohexanones by Rh-Mediated Intramolecular C-H Insertion. J. Org. Chem. 2013, 78 (19), 9772–9780. 10.1021/jo4014996. [DOI] [PubMed] [Google Scholar]
- McKervey M. A.; Ye T. Asymmetric Synthesis of Substituted Chromanones via C-H Insertion Reactions of α-Diazoketones Catalysed by Homochiral Rhodium(II) Carboxylates. J. Chem. Soc., Chem. Commun. 1992, (11), 823–824. 10.1039/C39920000823. [DOI] [Google Scholar]
- Ito M.; Kondo Y.; Nambu H.; Anada M.; Takeda K.; Hashimoto S. Diastereo- and Enantioselective Intramolecular 1,6-C-H Insertion Reactions of α-Diazo Esters Catalyzed by Chiral Dirhodium(II) Carboxylates. Tetrahedron Lett. 2015, 56 (11), 1397–1400. 10.1016/j.tetlet.2015.01.125. [DOI] [Google Scholar]
- For the formation of related bridged scaffolds by carbene C–H insertion starting from diazo derivatives, see the following:; Jansone-Popova S.; May J. A. Synthesis of Bridged Polycyclic Ring Systems via Carbene Cascades Terminating in C-H Bond Insertion. J. Am. Chem. Soc. 2012, 134 (43), 17877–17880. 10.1021/ja308305z. [DOI] [PubMed] [Google Scholar]
- Moloney M. G. Excitatory Amino Acids. Nat. Prod. Rep. 1998, 15 (2), 205–219. 10.1039/a815205y. [DOI] [PubMed] [Google Scholar]
- The primary propargyl amide analogous to 46 succumbed solely to trans-hydrogenation.
- Carreira E. M.; Fessard T. C. Four-Membered Ring-Containing Spirocycles: Synthetic Strategies and Opportunities. Chem. Rev. 2014, 114 (16), 8257–8322. 10.1021/cr500127b. [DOI] [PubMed] [Google Scholar]
- Zheng Y.; Tice C. M.; Singh S. B. The Use of Spirocyclic Scaffolds in Drug Discovery. Bioorg. Med. Chem. Lett. 2014, 24 (16), 3673–3682. 10.1016/j.bmcl.2014.06.081. [DOI] [PubMed] [Google Scholar]
- Hiesinger K.; Dar’in D.; Proschak E.; Krasavin M. Spirocyclic Scaffolds in Medicinal Chemistry. J. Med. Chem. 2021, 64 (1), 150–183. 10.1021/acs.jmedchem.0c01473. [DOI] [PubMed] [Google Scholar]
- Fominova K.; Diachuk T.; Granat D.; Savchuk T.; Vilchynskyi V.; Svitlychnyi O.; Meliantsev V.; Kovalchuk I.; Litskan E.; Levterov V. V.; Badlo V. R.; Vaskevych R. I.; Vaskevych A. I.; Bolbut A. V.; Semeno V. V.; Iminov R.; Shvydenko K.; Kuznetsova A. S.; Dmytriv Y. V.; Vysochyn D.; Ripenko V.; Tolmachev A. A.; Pavlova O.; Kuznietsova H.; Pishel I.; Borysko P.; Mykhailiuk P. K. Oxa-Spirocycles: Synthesis, Properties and Applications. Chem. Sci. 2021, 12 (34), 11294–11305. 10.1039/D1SC03615G. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Talele T. T. Opportunities for Tapping into Three-Dimensional Chemical Space through a Quaternary Carbon. J. Med. Chem. 2020, 63 (22), 13291–13315. 10.1021/acs.jmedchem.0c00829. [DOI] [PubMed] [Google Scholar]
- Crossley S. W. M.; Obradors C.; Martinez R. M.; Shenvi R. A. Mn-, Fe-, and Co-Catalyzed Radical Hydrofunctionalizations of Olefins. Chem. Rev. 2016, 116 (15), 8912–9000. 10.1021/acs.chemrev.6b00334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fürstner A. Iron Catalysis in Organic Synthesis: A Critical Assessment of What It Takes To Make This Base Metal a Multitasking Champion. ACS Cent. Sci. 2016, 2 (11), 778–789. 10.1021/acscentsci.6b00272. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lo J. C.; Kim D.; Pan C.-M.; Edwards J. T.; Yabe Y.; Gui J.; Qin T.; Gutiérrez S.; Giacoboni J.; Smith M. W.; Holland P. L.; Baran P. S. Fe-Catalyzed C-C Bond Construction from Olefins via Radicals. J. Am. Chem. Soc. 2017, 139 (6), 2484–2503. 10.1021/jacs.6b13155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smaligo A. J.; Swain M.; Quintana J. C.; Tan M. F.; Kim D. A.; Kwon O. Hydrodealkenylative C(sp3)-C(sp2) Bond Fragmentation. Science 2019, 364 (6441), 681. 10.1126/science.aaw4212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smaligo A. J.; Wu J.; Burton N. R.; Hacker A. S.; Shaikh A. C.; Quintana J. C.; Wang R.; Xie C.; Kwon O. Oxodealkenylative Cleavage of Alkene C(sp3)-C(sp2) Bonds: A Practical Method for Introducing Carbonyls into Chiral Pool Materials. Angew. Chem., Int. Ed. 2020, 59 (3), 1211–1215. 10.1002/anie.201913201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smaligo A. J.; Kwon O. Dealkenylative Thiylation of C(sp3)-C(sp2) Bonds. Org. Lett. 2019, 21 (21), 8592–8597. 10.1021/acs.orglett.9b03186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- For additional opportunities, see the following recent publication:; Chen S.-C.; Zhu Q.; Cao Y.; Li C.; Guo Y.; Kong L.; Che J.; Guo Z.; Chen H.; Zhang N.; Fang X.; Lu J.-T.; Luo T. Dealkenylative Ni-Catalyzed Cross-Coupling Enabled by Tetrazine and Photoexcitation. J. Am. Chem. Soc. 2021, 143 (35), 14046–14052. 10.1021/jacs.1c05092. [DOI] [PubMed] [Google Scholar]
- For previous syntheses of this and related building blocks, see refs (72) and (73) and the following:; Li D. B.; Rogers-Evans M.; Carreira E. M. Construction of Multifunctional Modules for Drug Discovery: Synthesis of Novel Thia/Oxa-Azaspiro[3.4]octanes. Org. Lett. 2013, 15 (18), 4766–4769. 10.1021/ol402127b. [DOI] [PubMed] [Google Scholar]
- Painter T. O.; Bunn J. R.; Schoenen F. J.; Douglas J. T.; Day V. W.; Santini C. Skeletal Diversification via Heteroatom Linkage Control: Preparation of Bicyclic and Spirocyclic Scaffolds from N-Substituted Homopropargyl Alcohols. J. Org. Chem. 2013, 78 (8), 3720–3730. 10.1021/jo400077m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kumar S.; Thornton P. D.; Painter T. O.; Jain P.; Downard J.; Douglas J. T.; Santini C. Synthesis of a Family of Spirocyclic Scaffolds: Building Blocks for the Exploration of Chemical Space. J. Org. Chem. 2013, 78 (13), 6529–6539. 10.1021/jo400738b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mitsudo T.-A. The Chemistry of η3-Vinylcarbene Complexes. Bull. Chem. Soc. Jpn. 1998, 71 (7), 1525–1538. 10.1246/bcsj.71.1525. [DOI] [Google Scholar]
- The structures of several ordinary ruthenium carbenes formed by gem-hydrogenation show this peripheral interaction; see refs (2), (3), and (9). Moreover, it is manifested in gentle deshielding of the quarternary C-atom carrying the O-substituent as well as the −OMe group. For details, see the Supporting Information.
- Gómez-Gallego M.; Sierra M. A. Kinetic Isotope Effects in the Study of Organometallic Reaction Mechanisms. Chem. Rev. 2011, 111 (8), 4857–4963. 10.1021/cr100436k. [DOI] [PubMed] [Google Scholar]
- Simmons E. M.; Hartwig J. F. On the Interpretation of Deuterium Kinetic Isotope Effects in C-S-H Bond Functionalizations by Transition-Metal Complexes. Angew. Chem., Int. Ed. 2012, 51 (13), 3066–3072. 10.1002/anie.201107334. [DOI] [PubMed] [Google Scholar]
- Kwart H. Temperature Dependence of the Primary Kinetic Hydrogen Isotope Effect as a Mechanistic Criterion. Acc. Chem. Res. 1982, 15 (12), 401–408. 10.1021/ar00084a004. [DOI] [Google Scholar]
- Haibach M. C.; Seidel D. C-H-H Bond Functionalization through Intramolecular Hydride Transfer. Angew. Chem., Int. Ed. 2014, 53 (20), 5010–5036. 10.1002/anie.201306489. [DOI] [PubMed] [Google Scholar]
- Wang L.; Xiao J. Advancement in Cascade [1, n]-Hydrogen Transfer/Cyclization: A Method for Direct Functionalization of Inactive C(sp3)-H Bonds. Adv. Syn. Catal. 2014, 356 (6), 1137–1171. 10.1002/adsc.201301153. [DOI] [Google Scholar]
- See the Supporting Information accompanying ref (8).
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.














