Abstract
The reversible activation of dihydrogen with a molecular zinc anilide complex is reported. The mechanism of this reaction has been probed through stoichiometric experiments and density functional theory (DFT) calculations. The combined evidence suggests that H2 activation occurs by addition across the Zn–N bond via a four-membered transition state in which the Zn and N atoms play a dual role of Lewis acid and Lewis base. The zinc hydride complex that results from H2 addition has been shown to be remarkably effective for the hydrozincation of C=C bonds at modest temperatures. The scope of hydrozincation includes alkynes, alkenes, and a 1,3-butadiyne. For alkynes, the hydrozincation step is stereospecific leading exclusively to the syn-isomer. Competition experiments show that the hydrozincation of alkynes is faster than the equivalent alkene substrates. These new discoveries have been used to develop a catalytic system for the semi-hydrogenation of alkynes. The catalytic scope includes both aryl- and alkyl-substituted internal alkynes and proceeds with high alkene: alkane, Z:E ratios, and modest functional group tolerance. This work offers a first example of selective hydrogenation catalysis using zinc complexes.
Introduction
The activation of dihydrogen is a phenomenon that, until recent years, was associated primarily with transition metals.1,2 Significant developments have seen main group complexes reported for the activation of dihydrogen3−6 and applications of these systems in catalytic hydrogenation.7−10 The activation of dihydrogen using zinc compounds has received little attention. Zinc, as a post-transition metal, does not employ its d-electrons, and so its reactivity is likely to mimic that of a main group center. Dihydrogen activation at zinc is an unusual observation. Although there have been previous reports of the use of zinc-based catalysts for hydrogenation, the fundamental understanding of this process has received limited attention. For example, Beller and co-workers reported the use of zinc triflate, Zn(OTf)2, as a pre-catalyst for the hydrogenation of imines.11 Stephan and co-workers also used a zinc-based catalyst for the hydrogenation of imines and ketones.12,13 Milstein and co-workers provided an example of zinc-catalyzed hydrogenation of imines and ketones, using a PNP pincer ligand to facilitate dihydrogen activation through metal-ligand cooperation.14 Very recently, Lacy and co-workers have reported a well-defined Zn catalyst for the hydrogenation of benzophenone and N-benzyl-1-phenylmethanimine.15 Despite these reports, the application of zinc to hydrogenation catalysis remains understudied, with substrates limited to highly polarized species. Furthermore, the issue of selective hydrogenation with these types of catalysts is yet to be addressed.
The semi-hydrogenation of alkynes to selectively form alkenes is an important transformation in the synthesis of vitamins and other natural products,16,17 as well as being industrially relevant in fields such as polymerization catalysis. The current industrial standard with regard to this process is the palladium-based heterogeneous Lindlar catalyst;18 however, in recent years, many homogeneous transition metal systems have also been developed.19−33 Due to the current imperative to replace expensive and toxic transition metal catalyst systems, such as those based on palladium, rhodium, ruthenium, or iridium, we were interested in the application of nontransition metal elements to this reaction.
Frustrated Lewis pair (FLP) systems, featuring boranes in combination with a Lewis basic moiety, are some of the most successful homogeneous nontransition metals capable of the selective semi-hydrogenation of internal alkynes.34−40 During the revision of this work, Milstein and co-workers reported a Mg-based catalyst for alkyne semi-hydrogenation.41 Herein, we report the activation of H2 by the 1,2-addition across a Zn–N bond of a novel zinc anilide complex. Further, we show that this key step can be exploited in a highly chemo- and stereoselective semi-hydrogenation of alkynes to form Z-alkenes. The approach is inspired by pioneering studies that showed copper hydride reagents can be generated by the 1,2-addition of H2 across a Cu–O bond.42
Results
Synthesis of Zinc Anilides
The zinc anilide complex 1 was synthesized in one-pot by the stepwise reaction of ZnCl2, with the lithiated ligand and LiNHPh. The family of terminal, primary zinc–anilide complexes has only a couple of structurally characterized examples,43,44 and the reactivity of these species has not been investigated in any detail. 1 was proposed to react with H2. An NMR scale reaction of 1 with H2 (1 bar) in C6D6 at 100 °C however yielded only <1% of aniline and the known zinc hydride 2.45 The equilibrium is displaced almost entirely toward the starting materials. Repeating this experiment in the presence of a suitable trapping agent for 2 led to unambiguous evidence of dihydrogen activation at zinc. Hence, the reaction of 1, H2, and diphenylacetylene after 18 h at 100 °C formed the zinc vinyl complex 3a in a low but non-negligible yield of ∼10%, again the mass balance is unreacted starting materials (Scheme 1).
Scheme 1. Stoichiometric Reactions of 1 with H2 and Diphenylacetylene.
1H NMR yields.
In separate experiments, 2 reacted cleanly with diphenylacetylene at 100 °C over 28 h to form 3a in near quantitative yield. These results demonstrate slow dihydrogen activation by 1 and the hydrozincation of an alkyne with the resultant zinc hydride 2.
Dihydrogen Activation
To explore the energetic feasibility of the postulated dihydrogen activation process, we performed density functional theory (DFT) calculations using the M06L functional and a def2TZVPP basis-set (see Supporting Information for details). In the solid state, 1 possesses an arrangement in which the core of the β-diketiminate ligand, the Zn–N motif, and the phenyl ring of the anilide all lie within the same plane (Figure 1). This conformation is reproduced by DFT calculations and results in an electronic structure in which the frontier molecular orbitals of 1 are perfectly set up to react with dihydrogen. Hence, the natural localized molecular orbitals (NLMOs) of 1 allow the visualization of the 2p-orbital on N and the 4s-orbital on Zn, which contribute to the HOMO and LUMO + 1, respectively (Figure 2a). The dual Lewis acidic/Lewis basic behavior of the Zn–N bond of 1 was further supported by attempts to prepare an analogue of this compound. Variation of the steric demands of the ligand on zinc (Ar = 2,6-diisopropylphenyl vs 2,6-diethylphenyl) led to the isolation of the zincate complex 4, in which an equiv of Li–Cl is coordinated to the Zn–N bond (Figure 1).
Figure 1.

Crystal structure 1 (a) end-on and (b) side-on, along with (c) 4 (50% probability ellipsoids, selected hydrogen atoms omitted for clarity). Only one molecule (1A) of the two molecules in the unit cell of 1 is shown. Hydrogen atoms omitted for clarity. Selected bond distances (Å) 1: Zn–N1 1.855(2), Zn–N2 1.9140(19), Zn–N3 1.9424(18), N1–Zn–N2 120.38(8), N2–Zn–N3 99.37(8), N1–Zn–N3 139.82(8). 4: Zn–N1 1.980(5), Zn–N2 1.995(3), Zn–N3 1.995(3) Zn–Cl, 2.3873(12).
Figure 2.

(a) Calculated NLMOs showing the key donor and acceptor orbitals of 1. (b) Calculated Gibbs Free Energy profile for H2 activation using 1, at standard conditions using the M06L functional and def2TZVPP basis-set. Energies in kcal mol–1.
Dihydrogen activation by 1 was calculated to occur through a stepwise process involving the formation of a weakly bound encounter complex Int-1, which leads to TS-1. Int-1 is an unstable encounter complex of zinc with dihydrogen. While there is a component of bonding that involves the donation of electron density from the H–H bond to the vacant 4s-orbital of Zn in Int-1, it is not significant enough to assign this as a dihydrogen complex of zinc.
Dihydrogen splitting occurs by translation of the H2 molecule across the zinc center, toward the anilide nitrogen atom, elongating both the H–H and Zn–N distances as it nears the transition state geometry. In TS-1, the population of the σ*-orbital of H2 occurs through the donation of electron density from the N-based LP, breaking the H–H bond and forming Int-2, a weakly bound adduct of 2 and H2NPh. The observations that dihydrogen activation by 1 occurs slowly at 100 °C and that there is a strong preference for the reverse reaction suggest a large energy barrier to H2 activation and an endergonic reaction. These expectations were confirmed by the calculated ΔG298K° and ΔG298K‡ of dihydrogen activation by 1 which at 298 K and 1 bar are +4.4 and +28.2 kcal mol–1, respectively (Figure 2).
Hydrozincation
Direct hydrozincation of unsaturated bonds is well known, with most reports focusing on polar substrates.46−52 However, the direct hydrozincation of alkynes has received little attention. Ingleson and co-workers reported that a cationic two-coordinate zinc hydride species was capable of adding across the triple bond of various alkyne substrates.53 They also went on to develop a catalytic system for di- and triborylation of terminal alkynes and showed that their cationic, two-coordinate complex outperformed complex 2 in this system.54 Direct hydrozincation was also proposed to occur in the zinc-catalyzed hydroboration of alkynes reported by Geetharani and co-workers.55 These reports cover terminal and internal alkynes; however, methods for the uncatalyzed hydrozincation of alkenes are rare.
The reactivity of intermediate zinc hydride 2 toward a range of unsaturated substrates was investigated. Alkynes readily underwent hydrozincation using 2, as did terminal alkenes, while internal alkene substrates were generally more sluggish to react. Reaction of 2 with oct-4-yne cleanly forms 3b. Reaction of 2 with 1-phenyl-1-propyne at 80 °C resulted in the formation of a mixture of two regioisomeric hydrozincation products, 3c and 3d, in an approximately 1:1 ratio, 42% yield. The reaction of 2 with 1,4-diphenybutadiyne led to the selective hydrozincation of one of the alkynyl groups to yield 3e, and in this case, only a single regioisomer was observed in quantitative yield at 80 °C. Reactions of 2 with terminal alkenes 1-allylbenzene and hex-1-ene at 80 °C also formed the corresponding zinc alkyl compounds 3f and 3g, respectively (Scheme 2). Based on the consistent stereochemistry of the zinc alkenyl products 3a–e (Figure 3), it is clear the reaction proceeds via a syn-addition of Zn–H to the alkyne substrate.53 The regiochemistry of hydrozincation is notable. The addition to a phenylprop-1-yne is nonselective, yielding a mixture of Markovnikov and anti-Markovnikov products. 1,4-Diphenylbutadiene proceeds with high selectivity for the Markovnikov product. Terminal alkenes yield anti-Markovnikov products exclusively. These results are predicted based on established trends in hydrometallation chemistry and the expected stabilization of partial charges in the insertion transition state. The hydrozincation step also appeared susceptible to steric effects as internal alkenes react only very slowly with 2. For example, (E)-1,2-diphenylethylene and (Z)-1,2-diphenylethylene give only partial conversion (<15%) to the corresponding hydrozincated products after at least 15 h at 100 °C. 1,3-bis(trimethylsilyl)acetylene did not react with 2 under these conditions.
Scheme 2. Reactions of 2 with Alkenes and Alkynes to Form Products 3a–g1H NMR Yields Shown (Isolated Yields in Parentheses).
Figure 3.
Crystal structures of the products 3a–f (50% probability ellipsoids, selected hydrogen atoms omitted for clarity).
There is the potential for the hydrozincation step to be reversible with regeneration of the starting materials via a β-hydride elimination mechanism. Mixtures of 3c/3d were used to investigate the reversibility. Monitoring the formation 3c/3d as a function of time revealed no change in the 1:1 product ratio with conversion, suggesting that these products do not interconvert at 80 °C. Furthermore, manual separation of crystals of 3c and 3d allowed isolation of a sample enriched in 3d (2.4:1). Re-exposing the enriched mixture to reaction conditions did not result in an equilibration to a 1:1 mixture of 3c:3d. Similarly, an experiment in which 3g was heated with allylbenzene to 80 °C did not lead to the formation of the cross-over products. These findings strongly suggest that the hydrozincation step is non-reversible and β-hydride elimination is not in operation for either alkyne or alkene substrates up to temperatures of 80 °C.
An Eyring analysis on the reaction of 2 with diphenylacetylene in C6D6 under pseudo-first-order conditions (10 equiv) was conducted. Data sets across the 313–353 K temperature range at 10 K intervals were fitted using initial rates and returned activation parameters of ΔH‡ = 20.6 kcal mol–1, ΔS‡ = −20.5 cal K–1 mol–1, and ΔG298K‡ = 26.7 kcal mol–1. The negative activation entropy is consistent with an intermolecular hydrozincation step, which involves an ordered transition state.
Catalytic Semi-Hydrogenation of Alkynes
We postulated that if aniline could protonate the vinyl fragment of 3a, then the system could be rendered catalytic. The addition of aniline to 3a at 80 °C led to slow production of 1 and (Z)-1,2-diphenylethylene. From these stoichiometric reactions, it is easy to envisage a catalytic cycle very similar to those put forward for copper alkoxide catalysis,42 toward (Z)-selective alkyne semi-hydrogenation. DFT calculations were used to ascertain that each of the steps involved in catalytic turnover had an accessible Gibbs activation energy (Figure 4). The global energy barrier for turnover is Δ298≠ = +33.9 kcal mol–1, consistent with a reaction that requires forcing conditions. This barrier is associated with two steps, a reversible H2 splitting and subsequent diphenylacetylene hydrozincation.
Figure 4.
Alkyne semi-hydrogenation catalyzed by 1. DFT calculated barriers and thermodynamics for R = Ph (kcal mol–1).
We calculated the local barrier to hydrozincation to be ΔG298K‡ = +29.5 kcal mol–1. Similar barriers for the hydrozincation of 1-phenyl-1-propyne with cationic zinc complexes ΔG298K‡ = +28.6 to +31.2 kcal mol–1 have been calculated by Ingleson and co-workers.53 The calculated thermochemistry is in reasonable agreement with the experimental activation parameters determined for the hydrozincation of diphenylacetylene by 2, ΔG298K‡ = 26.7 kcal mol–1.
Hydrogenation reactions were performed at a substrate concentration of 0.17 M in 2–7 mL of solvent. At 145 °C and 23 bar H2 pressure, the optimal yield and selectivity for the semi-hydrogenation of diphenylacetylene with 10 mol % 1 were obtained (Figure 5). This reaction produces (Z)-1,2-diphenylethylene in >80% with >99:1 stereoselectivity and ∼97:3 chemoselectivity with only small amounts of over hydrogenation to the alkane. Further NMR scale experiments show that 1 is a remarkably robust, albeit slow, catalyst. 1 is stable for 18 h at 178 °C in toluene-d8 in a flame-sealed tube. A negative control using no catalyst showed no conversion. 10 mol % 3a is a poor catalyst leading to only a small amount of product (7%) likely from a stoichiometric reaction. Addition of aniline however could be used to re-establish catalysis. Hence, a mixture of 10 mol % 3a + 10 mol % PhNH2 catalytically converts diphenylacetylene to (Z)-1,2-diphenylethylene in >99% with >99:1 stereoselectivity and ∼86:14 chemoselectivity after 4 h at 145 °C and 23 bar H2 pressure. 4 was also catalytically active for the semi-hydrogenation of diphenylacetylene to (Z)-1,2-diphenylethylene in high yield and selectivity (Table S3).
Figure 5.

Scope of semi-hydrogenation using 1.
The proposed mechanism based on DFT calculations predicts that a combination of a reversible hydrogen splitting step and non-reversible alkyne hydrozincation is turnover limiting (pre-equilibrium case). This model predicts that the rate of the reaction should be dependent on both H2 pressure and initial alkyne concentration. Turnover frequencies after 18 h at 140 °C (TOF18h) were measured for a series of batch reactions run at different H2 pressures. TOF18h was found to increase with increasing pressure (5 bar, TOF18h = 0.019 h–1; 10 bar, TOF18h = 0.067 h–1; 15 bar, TOF18h = 0.411 h–1; 20 bar, TOF18h = 0.550 h–1). The effect of initial alkyne concentration proved more complex. Turnover frequencies were measured after 3 h at 160 °C (TOF3h) for a series of initial alkyne concentrations. These experiments suggest that the rate of reaction increases with increasing [alkyne]0 but quickly reaches a plateau and then begins to decrease (see Supporting Information). The data are consistent with catalyst deactivation occurring at higher [alkyne]0, due to the alkyne acting as an inhibitor or promoting off-cycle events.
The reaction conditions were tested with 22 alkyne substrates (Figure 5). Diarylethynes were ideal substrates, affording high conversion and high selectivities within relatively short reaction times, with electron-rich and -deficient systems giving similar results. Mixed alkyl/aryl and dialkyl internal alkynes also gave good conversion and selectivities, although the longer reaction times required. Halogen, ether, thienyl, pyridyl, nitrile, trifluoromethyl, and cyclopropyl functional groups were all tolerated. While in some cases yields were low, Z:E stereoselectivity and alkene:alkane chemoselectivity were very high. The hydrogenation of dimethoxybut-2-ene leads to a useful building block for the potential synthesis of ring systems such as pyridoxine (relevant to vitamin B6).56
Carbonyl and nitro functional groups were less well tolerated under the reaction conditions. Both ketones and nitro groups underwent competitive reduction with the alkyne, leading to mixtures containing either alkenes/alcohols or alkenes/amines. Nevertheless, in the case of 1-nitro-4-(phenylethynyl)benzene, the Z-alkene remains the major product of the reaction. An ester-substituted substrate did not participate in hydrogenation catalysis, suggesting this group may act as an inhibitor for turnover. Bis(trimethylsilyl)ethyne and 1-phenyl-2-trimethylsilylethyne did not undergo hydrogenation, consistent with the observations regarding the hydrozincation of these substrates using 2 (vide supra). The hydrogenation of terminal alkynes (hex-1-yne, phenylacetylene) was also investigated in this system, and low conversions were achieved for these substrates (Figure 5 and Table S3). Terminal alkynes can potentially deactivate the catalyst. For example, when 1 was combined with hex-1-yne, facile protonolysis of the anilide ligand of 1 led to the formation of the corresponding catalytically inactive zinc acetylide species.54 1,4-Diphenylbutadiyne could also be selectively hydrogenated to either form the Z-enyne product in low yield, while dodeca-5,7-diyne formed a mixture including the Z,Z-diene as the major product.
For comparison, copper(I) or magnesium(II) semi-hydrogenation catalysts show limited scope for nitrile and nitro functional groups. There are only a handful of examples of sp2C–CN substituted substrates that undergo selective hydrogenation with homogeneous copper catalysts, and those that are reported tend not to be diarylalkynes.57−62 Similarly, there is only one example of a copper(1) semi-hydrogenation system that reports the reaction with a substrate bearing a sp2C–NO2 group, and this does not use H2 a reductant. Examples that do use H2 either lead to no conversion or complete and unselective over-reduction of the sp2C–NH2 amine.57,59,61,63−65 Furthermore, selective diyne reduction with copper catalysis has been reported to form Z,Z-dienes for diaryl-substituted substrates and E,E-dienes for dialkyl-substituted substrates.66,67 The zinc-base catalyst reported herein, while proceeding with low conversion appears to give access to complementary selectivities.
Origin of Selectivity
The origin of the remarkable selectivity achieved by 1 in semi-hydrogenation catalysis is likely determined by the reactivity of the intermediate 2 toward unsaturated substrates. The hydrozincation of alkynes with 2 occurs by a stereospecific reaction to form the syn-isomer, and this leads exclusively to the Z-alkene on protonation, resulting in high Z:E ratios. The high alkene:alkane ratios can be understood by considering the relative rates of hydrozincation of alkynes and alkenes with 2. Competition reactions in which a 1:1 mixture of diphenylacetylene: (Z)-1,2-diphenylethene (or diphenylacetylene: (E)-1,2-diphenylethene) were reacted with 2 at 100 °C for 15 h resulted in exclusive and quantitative formation of 3a, and no zinc-alkyl species was observed (Scheme 3).
Scheme 3. Competition Reactions of 2 with Internal Alkenes and Alkynes.
Conclusions
In summary, a novel system for H2 activation using a zinc anilide complex has been demonstrated, where dihydrogen is split across a Zn–N bond to yield a reactive zinc hydride species. This observation has been supported by a detailed mechanistic investigation using a combined experimental and computational approach. Molecular orbital analysis allows us to visualize the synergic donor-acceptor interactions that leads to this activation event. Additionally, we have investigated the scope of direct hydrozincation using a terminal zinc hydride species and probed the selectivity of this process. Through this mechanistic understanding, we were able to rationally design a system capable of semi-hydrogenation of alkynes. This system was shown to be effective for the highly Z-selective semi-hydrogenation of a range of alkynes, at moderate temperature and pressure. This system represents a rare post-transition metal alternative for semi-hydrogenation catalysis. This work introduces a novel solution to the challenge of dihydrogen activation and significantly broadens the remit of zinc and post-transition metal catalysis.
Acknowledgments
Nikolaus Gorgas, Maria Batuecas, Daniel Sheldon, Feriel Rekhroukh, Andreas Phanopoulos, and Joseph Parr are thanked for conducting experiments during the peer-review process. David Baker, Margaret Baker, and Ellen Nelson are thanked for their support.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c02301.
Synthetic procedures, kinetic experiments, NMR spectra of all compounds, crystallographic data, and computational methods (PDF)
The authors are grateful to the EPSRC for the provision of a Ph.D. Scholarship to G.J.B. and to Johnson-Matthey for in-kind support.
The authors declare no competing financial interest.
Author Status
The authors dedicate this paper to our friend and colleague Greg Baker, he was adored by all and is missed every day.
Supplementary Material
References
- Jessop P. G.; Morris R. H. Reactions of Transition Metal Dihydrogen Complexes. Coord. Chem. Rev. 1992, 121, 155–284. 10.1016/0010-8545(92)80067-2. [DOI] [Google Scholar]
- Heinekey D. M.; Oldham W. J. Coordination Chemistry of Dihydrogen. Chem. Rev. 1993, 93, 913–926. 10.1021/cr00019a004. [DOI] [Google Scholar]
- Spikes G. H.; Fettinger J. C.; Power P. P. Facile Activation of Dihydrogen by an Unsaturated Heavier Main Group Compound. J. Am. Chem. Soc. 2005, 127, 12232–12233. 10.1021/ja053247a. [DOI] [PubMed] [Google Scholar]
- Welch G. C.; Juan R. R. S.; Masuda J. D.; Stephan D. W. Reversible, Metal-Free Hydrogen Activation. Science 2006, 314, 1124–1126. 10.1126/science.1134230. [DOI] [PubMed] [Google Scholar]
- Power P. P. Main-Group Elements as Transition Metals. Nature 2010, 463, 171–177. 10.1038/nature08634. [DOI] [PubMed] [Google Scholar]
- Stephan D. W. Diverse Uses of the Reaction of Frustrated Lewis Pair (FLP) with Hydrogen. J. Am. Chem. Soc. 2021, 143, 20002–20014. 10.1021/jacs.1c10845. [DOI] [PubMed] [Google Scholar]
- Spielmann J.; Buch F.; Harder S. Early Main-Group Metal Catalysts for the Hydrogenation of Alkenes with H2. Angew. Chem., Int. Ed. 2008, 47, 9434–9438. 10.1002/anie.200804657. [DOI] [PubMed] [Google Scholar]
- Feng X.; Du H. Metal-Free Asymmetric Hydrogenation and Hydrosilylation Catalyzed by Frustrated Lewis Pairs. Tetrahedron Lett. 2014, 55, 6959–6964. 10.1016/j.tetlet.2014.10.138. [DOI] [Google Scholar]
- Scott D. J.; Fuchter M. J.; Ashley A. E. Designing Effective ‘Frustrated Lewis Pair’ Hydrogenation Catalysts. Chem. Soc. Rev. 2017, 46, 5689–5700. 10.1039/C7CS00154A. [DOI] [PubMed] [Google Scholar]
- Lam J.; Szkop K. M.; Mosaferi E.; Stephan D. W. FLP Catalysis: Main Group Hydrogenations of Organic Unsaturated Substrates. Chem. Soc. Rev. 2018, 48, 3592–3612. 10.1039/C8CS00277K. [DOI] [PubMed] [Google Scholar]
- Werkmeister S.; Fleischer S.; Zhou S.; Junge K.; Beller M. Development of New Hydrogenations of Imines and Benign Reductive Hydroaminations: Zinc Triflate as a Catalyst. ChemSusChem 2012, 5, 777–782. 10.1002/cssc.201100633. [DOI] [PubMed] [Google Scholar]
- Jochmann P.; Stephan D. W. H2 Cleavage, Hydride Formation, and Catalytic Hydrogenation of Imines with Zinc Complexes of C5Me5 and N-Heterocyclic Carbenes. Angew. Chem., Int. Ed. 2013, 52, 9831–9835. 10.1002/anie.201303968. [DOI] [PubMed] [Google Scholar]
- Jochmann P.; Stephan D. W. Zincocene and Dizincocene N-Heterocyclic Carbene Complexes and Catalytic Hydrogenation of Imines and Ketones. Chem. - Eur. J. 2014, 20, 8370–8378. 10.1002/chem.201402875. [DOI] [PubMed] [Google Scholar]
- Rauch M.; Kar S.; Kumar A.; Avram L.; Shimon L. J. W.; Milstein D. Metal–Ligand Cooperation Facilitates Bond Activation and Catalytic Hydrogenation with Zinc Pincer Complexes. J. Am. Chem. Soc. 2020, 142, 14513–14521. 10.1021/jacs.0c05500. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paul S.; Morgante P.; MacMillan S. N.; Autschbach J.; Lacy D. C. Hydrogenative Catalysis with Three-Coordinate Zinc Complexes Supported with PN Ligands Is Enhanced Compared to PNP Analogs. Chem. Eur. J. 2022, e202201042 10.1002/chem.202201042. [DOI] [PubMed] [Google Scholar]
- Cirla A.; Mann J. Combretastatins: From Natural Products to Drug Discovery. Nat. Prod. Rep. 2003, 20, 558–564. 10.1039/b306797c. [DOI] [PubMed] [Google Scholar]
- Baur J. A.; Sinclair D. A. Therapeutic Potential of Resveratrol: The in Vivo Evidence. Nat. Rev. Drug Discovery 2006, 5, 493–506. 10.1038/nrd2060. [DOI] [PubMed] [Google Scholar]
- Lindlar H. Ein Neuer Katalysator Für Selektive Hydrierungen. Helv. Chim. Acta 1952, 35, 446–450. 10.1002/hlca.19520350205. [DOI] [Google Scholar]
- Schrock R. R.; Osborn J. A. Catalytic Hydrogenation Using Cationic Rhodium Complexes. II. The Selective Hydrogenation of Alkynes to Cis Olefins. J. Am. Chem. Soc. 1976, 98, 2143–2147. 10.1021/ja00424a021. [DOI] [Google Scholar]
- Laren M. W.; van Elsevier C. J. Selective Homogeneous Palladium(0)-Catalyzed Hydrogenation of Alkynes to (Z)-Alkenes. Angew. Chem., Int. Ed. 1999, 38, 3715–3717. . [DOI] [PubMed] [Google Scholar]
- Hauwert P.; Boerleider R.; Warsink S.; Weigand J. J.; Elsevier C. J. Mechanism of Pd(NHC)-Catalyzed Transfer Hydrogenation of Alkynes. J. Am. Chem. Soc. 2010, 132, 16900–16910. 10.1021/ja1062407. [DOI] [PubMed] [Google Scholar]
- Richmond E.; Moran J. Ligand Control of E/Z Selectivity in Nickel-Catalyzed Transfer Hydrogenative Alkyne Semireduction. J. Org. Chem. 2015, 80, 6922–6929. 10.1021/acs.joc.5b01047. [DOI] [PubMed] [Google Scholar]
- Karunananda M. K.; Mankad N. P. E-Selective Semi-Hydrogenation of Alkynes by Heterobimetallic Catalysis. J. Am. Chem. Soc. 2015, 137, 14598–14601. 10.1021/jacs.5b10357. [DOI] [PubMed] [Google Scholar]
- Semba K.; Kameyama R.; Nakao Y. Copper-Catalyzed Semihydrogenation of Alkynes to Z-Alkenes. Synlett 2015, 26, 318–322. 10.1055/s-0034-1379896. [DOI] [Google Scholar]
- Pape F.; Thiel N. O.; Teichert J. F. Z-Selective Copper(I)-Catalyzed Alkyne Semihydrogenation with Tethered Cu–Alkoxide Complexes. Chem. - Eur. J. 2015, 21, 15934–15938. 10.1002/chem.201501739. [DOI] [PubMed] [Google Scholar]
- Tseng K.-N. T.; Kampf J. W.; Szymczak N. K. Modular Attachment of Appended Boron Lewis Acids to a Ruthenium Pincer Catalyst: Metal–Ligand Cooperativity Enables Selective Alkyne Hydrogenation. J. Am. Chem. Soc. 2016, 138, 10378–10381. 10.1021/jacs.6b03972. [DOI] [PubMed] [Google Scholar]
- Higashida K.; Mashima K. E -Selective Semi-Hydrogenation of Alkynes with Dinuclear Iridium Complexes under Atmospheric Pressure of Hydrogen. Chem. Lett. 2016, 45, 866–868. 10.1246/cl.160410. [DOI] [Google Scholar]
- Wakamatsu T.; Nagao K.; Ohmiya H.; Sawamura M. Copper-Catalyzed Semihydrogenation of Internal Alkynes with Molecular Hydrogen. Organometallics 2016, 35, 1354–1357. 10.1021/acs.organomet.6b00126. [DOI] [Google Scholar]
- Monfredini A.; Santacroce V.; Marchiò L.; Maggi R.; Bigi F.; Maestri G.; Malacria M. Semi-Reduction of Internal Alkynes with Prototypical Subnanometric Metal Surfaces: Bridging Homogeneous and Heterogeneous Catalysis with Trinuclear All-Metal Aromatics. ACS Sustainable Chem. Eng. 2017, 5, 8205–8212. 10.1021/acssuschemeng.7b01847. [DOI] [Google Scholar]
- Landge V. G.; Pitchaimani J.; Midya S. P.; Subaramanian M.; Madhu V.; Balaraman E. Phosphine-Free Cobalt Pincer Complex Catalyzed Z -Selective Semi-Hydrogenation of Unbiased Alkynes. Catal. Sci. Technol. 2018, 8, 428–433. 10.1039/C7CY01994G. [DOI] [Google Scholar]
- Johnson C.; Albrecht M. Z-Selective Alkyne Semi-Hydrogenation Catalysed by Piano-Stool N -Heterocyclic Carbene Iron Complexes. Catal. Sci. Technol. 2018, 8, 2779–2783. 10.1039/C8CY00681D. [DOI] [Google Scholar]
- Gorgas N.; Brünig J.; Stöger B.; Vanicek S.; Tilset M.; Veiros L. F.; Kirchner K. Efficient Z-Selective Semihydrogenation of Internal Alkynes Catalyzed by Cationic Iron(II) Hydride Complexes. J. Am. Chem. Soc. 2019, 141, 17452–17458. 10.1021/jacs.9b09907. [DOI] [PubMed] [Google Scholar]
- Ramirez B. L.; Lu C. C. Rare-Earth Supported Nickel Catalysts for Alkyne Semihydrogenation: Chemo- and Regioselectivity Impacted by the Lewis Acidity and Size of the Support. J. Am. Chem. Soc. 2020, 142, 5396–5407. 10.1021/jacs.0c00905. [DOI] [PubMed] [Google Scholar]
- Xu B.; Kehr G.; Fröhlich R.; Wibbeling B.; Schirmer B.; Grimme S.; Erker G. Reaction of Frustrated Lewis Pairs with Conjugated Ynones-Selective Hydrogenation of the Carbon–Carbon Triple Bond. Angew. Chem., Int. Ed. 2011, 50, 7183–7186. 10.1002/anie.201101051. [DOI] [PubMed] [Google Scholar]
- Chernichenko K.; Madarász Á.; Pápai I.; Nieger M.; Leskelä M.; Repo T. A Frustrated-Lewis-Pair Approach to Catalytic Reduction of Alkynes to Cis-Alkenes. Nat. Chem. 2013, 5, 718–723. 10.1038/nchem.1693. [DOI] [PubMed] [Google Scholar]
- Liu Y.; Hu L.; Chen H.; Du H. An Alkene-Promoted Borane-Catalyzed Highly Stereoselective Hydrogenation of Alkynes to Give Z- and E-Alkenes. Chem. - Eur. J. 2015, 21, 3495–3501. 10.1002/chem.201405388. [DOI] [PubMed] [Google Scholar]
- Szeto K. C.; Sahyoun W.; Merle N.; Castelbou J. L.; Popoff N.; Lefebvre F.; Raynaud J.; Godard C.; Claver C.; Delevoye L.; Gauvin R. M.; Taoufik M. Development of Silica-Supported Frustrated Lewis Pairs: Highly Active Transition Metal-Free Catalysts for the Z -Selective Reduction of Alkynes. Catal. Sci. Technol. 2016, 6, 882–889. 10.1039/C5CY01372K. [DOI] [Google Scholar]
- Chernichenko K.; Kótai B.; Nieger M.; Heikkinen S.; Pápai I.; Repo T. Replacing C6F5 Groups with Cl and H Atoms in Frustrated Lewis Pairs: H 2 Additions and Catalytic Hydrogenations. Dalton Trans. 2017, 46, 2263–2269. 10.1039/C6DT04649E. [DOI] [PubMed] [Google Scholar]
- Wech F.; Hasenbeck M.; Gellrich U. Semihydrogenation of Alkynes Catalyzed by a Pyridone Borane Complex: Frustrated Lewis Pair Reactivity and Boron–Ligand Cooperation in Concert. Chem. - Eur. J. 2020, 26, 13445–13450. 10.1002/chem.202001276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wech F.; Gellrich U. In Situ Formation of an Efficient Catalyst for the Semihydrogenation of Alkynes from Imidazolone and BH3. ACS Catal. 2022, 12, 5388–5396. 10.1021/acscatal.2c00722. [DOI] [Google Scholar]
- Liang Y.; Das U. K.; Luo J.; Diskin-Posner Y.; Avram L.; Milstein D. Magnesium Pincer Complexes and Their Applications in Catalytic Semihydrogenation of Alkynes and Hydrogenation of Alkenes: Evidence for Metal–Ligand Cooperation. J. Am. Chem. Soc. 2022, 144, 19115–19126. 10.1021/jacs.2c08491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mahoney W. S.; Stryker J. M. Hydride-Mediated Homogeneous Catalysis. Catalytic Reduction of Alpha,Beta-Unsaturated Ketones Using [(Ph3P)CuH]6 and H2. J. Am. Chem. Soc. 1989, 111, 8818–8823. 10.1021/ja00206a008. [DOI] [Google Scholar]
- Sánchez-Barba L. F.; Alonso-Moreno C.; Garcés A.; Fajardo M.; Fernández-Baeza J.; Otero A.; Lara-Sánchez A.; Rodríguez A. M.; López-Solera I. Synthesis, Structures and Ring-Opening Polymerization Studies of New Zinc Chloride and Amide Complexes Supported by Amidinate Heteroscorpionate Ligands. Dalton Trans. 2009, 8054–8062. 10.1039/b910048b. [DOI] [PubMed] [Google Scholar]
- Luo B.; Gladfelter W. L. On the Way to Monodispersed ZnO Nanocrystals: Structure of a Zinc Dimer Bearing Primary Amido Ligands. J. Coord. Chem. 2011, 64, 82–92. [Google Scholar]
- Spielmann J.; Piesik D.; Wittkamp B.; Jansen G.; Harder S. Convenient Synthesis and Crystal Structure of a Monomeric Zinc Hydride Complex with a Three-Coordinate Metal Center. Chem. Commun. 2009, 3455–3456. 10.1039/b906319f. [DOI] [PubMed] [Google Scholar]
- Roy M. M. D.; Omaña A. A.; Wilson A. S. S.; Hill M. S.; Aldridge S.; Rivard E. Molecular Main Group Metal Hydrides. Chem. Rev. 2021, 121, 12784–12965. 10.1021/acs.chemrev.1c00278. [DOI] [PubMed] [Google Scholar]
- Sattler W.; Parkin G. Synthesis, Structure, and Reactivity of a Mononuclear Organozinc Hydride Complex: Facile Insertion of CO2 into a Zn–H Bond and CO2-Promoted Displacement of Siloxide Ligands. J. Am. Chem. Soc. 2011, 133, 9708–9711. 10.1021/ja2035706. [DOI] [PubMed] [Google Scholar]
- Boone C.; Korobkov I.; Nikonov G. I. Unexpected Role of Zinc Hydride in Catalytic Hydrosilylation of Ketones and Nitriles. ACS Catal. 2013, 3, 2336–2340. 10.1021/cs400581w. [DOI] [Google Scholar]
- Dawkins M. J. C.; Middleton E.; Kefalidis C. E.; Dange D.; Juckel M. M.; Maron L.; Jones C. Two-Coordinate Terminal Zinc Hydride Complexes: Synthesis, Structure and Preliminary Reactivity Studies. Chem. Commun. 2016, 52, 10490–10492. 10.1039/C6CC05445E. [DOI] [PubMed] [Google Scholar]
- Lortie J. L.; Dudding T.; Gabidullin B. M.; Nikonov G. I. Zinc-Catalyzed Hydrosilylation and Hydroboration of N-Heterocycles. ACS Catal. 2017, 7, 8454–8459. 10.1021/acscatal.7b02811. [DOI] [Google Scholar]
- Ritter F.; Spaniol T. P.; Douair I.; Maron L.; Okuda J. Molecular Zinc Hydride Cations [ZnH]+: Synthesis, Structure, and CO2 Hydrosilylation Catalysis. Angew. Chem., Int. Ed. 2020, 59, 23335–23342. 10.1002/anie.202011480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X.; Xu X. Hydroboration of Nitriles and Imines by Highly Active Zinc Dihydride Catalysts. RSC Adv. 2021, 11, 1128–1133. 10.1039/D0RA09648B. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Procter R. J.; Uzelac M.; Cid J.; Rushworth P. J.; Ingleson M. J. Low-Coordinate NHC–Zinc Hydride Complexes Catalyze Alkyne C–H Borylation and Hydroboration Using Pinacolborane. ACS Catal. 2019, 9, 5760–5771. 10.1021/acscatal.9b01370. [DOI] [Google Scholar]
- Uzelac M.; Yuan K.; Ingleson M. J. A Comparison of Two Zinc Hydride Catalysts for Terminal Alkyne C–H Borylation/Hydroboration and the Formation of 1,1,1-Triborylalkanes by Tandem Catalysis Using Zn–H and B–H Compounds. Organometallics 2020, 39, 1332–1338. 10.1021/acs.organomet.0c00086. [DOI] [Google Scholar]
- Mandal S.; Mandal S.; Geetharani K. Zinc-Catalysed Hydroboration of Terminal and Internal Alkynes. Chem. - Asian J. 2019, 14, 4553–4556. 10.1002/asia.201900839. [DOI] [PubMed] [Google Scholar]
- Bonrath W.; Medlock J.; Schütz J.; Wüstenberg B.; Netscher T.. Hydrogenation in the Vitamins and Fine Chemicals Industry—an Overview, Karamé I., Eds.; InTech, 2012. [Google Scholar]
- Whittaker A. M.; Lalic G. Monophasic Catalytic System for the Selective Semireduction of Alkynes. Org. Lett. 2013, 15, 1112–1115. 10.1021/ol4001679. [DOI] [PubMed] [Google Scholar]
- Pape F.; Thiel N. O.; Teichert J. F. Z-Selective Copper(I)-Catalyzed Alkyne Semihydrogenation with Tethered Cu–Alkoxide Complexes. Chem. - Eur. J. 2015, 21, 15934–15938. 10.1002/chem.201501739. [DOI] [PubMed] [Google Scholar]
- Wakamatsu T.; Nagao K.; Ohmiya H.; Sawamura M. Copper-Catalyzed Semihydrogenation of Internal Alkynes with Molecular Hydrogen. Organometallics 2016, 35, 1354–1357. 10.1021/acs.organomet.6b00126. [DOI] [Google Scholar]
- Korytiaková E.; Thiel N. O.; Pape F.; Teichert J. F. Copper(I)-Catalysed Transfer Hydrogenations with Ammonia Borane. Chem. Commun. 2017, 53, 732–735. 10.1039/C6CC09067B. [DOI] [PubMed] [Google Scholar]
- Bao H.; Zhou B.; Jin H.; Liu Y. Diboron-Assisted Copper-Catalyzed Z-Selective Semihydrogenation of Alkynes Using Ethanol as a Hydrogen Donor. J. Org. Chem. 2019, 84, 3579–3589. 10.1021/acs.joc.9b00321. [DOI] [PubMed] [Google Scholar]
- Huang J.; Li X.; Wen H.; Ouyang L.; Luo N.; Liao J.; Luo R. Substrate-Controlled Cu(OAc)2-Catalyzed Stereoselective Semi-Reduction of Alkynes with MeOH as the Hydrogen Source. ACS Omega 2021, 6, 11740–11749. 10.1021/acsomega.1c01083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fedorov A.; Liu H.-J.; Lo H.-K.; Copéret C. Silica-Supported Cu Nanoparticle Catalysts for Alkyne Semihydrogenation: Effect of Ligands on Rates and Selectivity. J. Am. Chem. Soc. 2016, 138, 16502–16507. 10.1021/jacs.6b10817. [DOI] [PubMed] [Google Scholar]
- Park B. Y.; Lim T.; Han M. S. A Simple and Efficient in Situ Generated Copper Nanocatalyst for Stereoselective Semihydrogenation of Alkynes. Chem. Commun. 2021, 57, 6891–6894. 10.1039/D1CC02685B. [DOI] [PubMed] [Google Scholar]
- Moran M. J.; Martina K.; Bieliunas V.; Baricco F.; Tagliapietra S.; Berlier G.; Borggraeve W. M. D.; Cravotto G. Copper(0) Nanoparticle Catalyzed Z-Selective Transfer Semihydrogenation of Internal Alkynes. Adv. Synth. Catal. 2021, 363, 2850–2860. 10.1002/adsc.202100126. [DOI] [Google Scholar]
- Semba K.; Kameyama R.; Nakao Y. Copper-Catalyzed Semihydrogenation of Alkynes to Z-Alkenes. Synlett 2015, 26, 318–322. 10.1055/s-0034-1379896. [DOI] [Google Scholar]
- Thiel N. O.; Kemper S.; Teichert J. F. Copper(I)-Catalyzed Stereoselective Hydrogenation of 1,3-Diynes and Enynes. Tetrahedron 2017, 73, 5023–5028. 10.1016/j.tet.2017.05.029. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.






