Abstract
We disclose a 1,4,7-triazacyclononane (TACN) ligand featuring an appended boron Lewis acid. Metalation with Cu(I) affords a series of tetrahedral complexes including a boron capped cuprous hydride. We demonstrate distinct reactivity modes as a function of chemical oxidation: hydride transfer to CO2 in the copper(I) state and oxidant-induced H2 evolution as well as alkyne reduction.
Graphical Abstract
INTRODUCTION
The most prominent strategy for tuning the reactivity of metal sites within the synthetic community is primary-sphere ligand tuning;1 less emphasis has been placed on reactivity changes imparted by modifying the secondary coordination sphere.2 Extensive analyses of primary sphere structure/function relationships have provided a blueprint for how metal geometry influences substrate activation. For many substrates, metals in pseudo-tetrahedral geometries provide appropriate metal-substrate orbital overlap to stabilize reactive intermediates.3–5
1,4,7-Triazacyclononane (TACN) ligands are facially binding chelates for metal complexes that are traditionally reported in 5- or 6-coordinate geometries,6 and have been widely employed within the bioinorganic, coordination chemistry, and catalysis fields.7–9 Building upon work by Wieghardt and others,10 TACN ligands containing sterically-encumbering tertiary-alkyl groups have been shown to stabilize transition metal complexes in pseudotetrahedral geometries (Figure 1, bottom).11 This bulky subset of TACN ligands has been used to access complexes that stabilize reactive oxygen species.12–13 For example, Scarborough and coworkers reported the formation of a room temperature stable dicopper(II) μ-η2:η2-peroxo complex that is competent for catalytic aerobic oxidation of 3,5-di-tertbutylcatechol at and above 25 °C.12 More recently, Ray and coworkers reported a highly reactive Fe(IV)-oxo supported by tri-tert-butyl TACN that is an electronic and functional model of taurine dioxygenase.13 An alternative avenue to tune reactivity in TACN manifolds is to incorporate secondary-sphere acidic groups: a modification that has received little attention.14
Secondary-sphere interactions with metal-coordinated substrates are widely used by most metalloenzymes to facilitate small molecule activation/functionalization.15–16 These interactions can enhance the activation of typically-inert molecules by stabilizing otherwise unstable reaction intermediates, thereby providing accessible reaction pathways. Although routinely exploited in biology, the use of secondary sphere interactions has yet to be widely employed in the field of synthetic chemistry. Synthetic ligand architectures containing appended acidic groups may be used to clarify the molecular-level details that inform on mechanism and ultimately enhance function. Hydrogen-bonding interactions are the most common appended groups that have been used to stabilize high-valent intermediates (i.e. metal-oxos):2, 17–20 such intermediates are not generally susceptible to protonation by weak Brønsted acids (hydrogen bond donors). However, these types of secondary sphere donors are less compatible with hydridic metal hydrides because of competitive protonation (H2 elimination).21 Ligands containing appended Lewis acidic groups overcome this incompatibility22–27 and are ideally suited to facilitate small molecule binding/reduction.
Our group has investigated cooperative binding modes and subsequent reactivity enabled by Lewis acidic boron-appended ligands.28–33 A key design principle we uncovered within the context of otherwise unstable metal hydrides was that moderately acidic Lewis acids provide an optimal balance between stability and reactivity.34 Leveling of ligand donor properties through M–H–BR3 interactions provided enhanced stability, without the loss of metal-hydride reactivity (e.g. reductive elimination). We found that there are clear changes to M-H redox properties that are uniquely accessible to Lewis-acid capped metal hydrides. We hypothesized that these principles could be extended to other metals and potentially unveil access to new reactivity modes.
One class of metal hydrides that could benefit from the addition of an appended Lewis acidic group is copper hydrides. Considerable work in the last decade has focused on regio- and enantioselective functionalization of organic substrates including olefins,35–36 alkynes,37 imines,38 ketones,39–40 and α,β-unsaturated carbonyls facilitated by discrete as well as proposed intermediate Cu(I) hydrides.41–42 One precedented pathway through which Cu(I) hydride complexes react with alkyne substrates is through insertion into the Cu-H bond to form a Cu(I) alkenyl complex.42–45 Importantly, we note that this pathway is only precedented at the Cu(I) state. Cu(I) hydride complexes with exogenous borane interactions are also known,46–50 although oxidation studies of these types of complexes have not been explored. In contrast to well-defined reactivity at the Cu(I) state, reactivity pathways accessible to oxidized Cu-H adducts are not clear. The Norton group investigated oxidation of Stryker’s reagent-type clusters and found that electron transfer can be mechanistically relevant for net hydride transfer reactions (Figure 1, top),51 although this concept has not been extended to monomeric or dimeric Cu(I) hydride complexes. A rare example of an oxidized Cu-H-Cu fragment reported by the Warren group demonstrated phenylacetylene reduction to styrene.52 In this case, the authors formulate the oxidized adduct as a Cu1.5-H-Cu1.5. To clarify the requirement of the second Cu atom for oxidized Cu-H reactivity, we hypothesized that a monometallic Cu-H stabilized by a borane Lewis acid might provide similar access to oxidized Cu-H adducts and thus we targeted reactivity modes of both cuprous and oxidized Cu-H adducts.
RESULTS AND DISCUSSION
We targeted a sterically-encumbered heteroalkylated TACN ligand that features two tert-butyl groups in addition to a modifiable auxiliary, by adapting a similar literature protocol originally used for the synthesis of 1,4,7-tri-tert-butyl-1,4,7-triazacyclononane.11 Ring closure of N,N’-(ethane-1,2-diyl)bis(N-(tert-butyl)-2-chloroacetamide) with allylamine followed by LiAlH4 reduction provided 1-allyl-4,7-di-tert-butyl-1,4,7-triazacyclononane (allylTACNtBu) as a colorless oil in 16% yield over 3 steps.
We chose to employ copper(I) for metalation/reactivity studies because it favors low coordination numbers and also participates in single electron redox processes.53 Metalation with [Cu(MeCN)4][PF6] or CuI in THF solvent afforded the cationic or neutral complexes [(allylTACNtBu)Cu(MeCN)][PF6] or (allylTACNtBu)CuI, respectively(1-PF6 and 1-I; Figure 2). Single crystal X-ray diffraction (XRD) experiments were employed to confirm connectivity, and importantly, each molecular structure confirmed that the allylic moiety is not interacting with copper and is available for post-metalation hydroboration.
Complexes 1 were subjected to hydroboration experiments to install an appended Lewis acid. Treating 1-I with one equiv. of 9-borabicyclo[3.3.1]nonane (9-BBN) in THF at room temperature for 16 h afforded the anti-Markovnikov hydroboration product (BBNTACNtBu)CuI (2-I) in 55% yield, assessed by 1H NMR spectroscopy. We attribute the moderate yield to competitive decomposition pathways: in addition to 2-I, we observed an insoluble grey precipitate. To overcome this hurdle, we undertook an alternate synthetic approach (Figure 2). Treating a C6H6 solution of allylTACNtBu with 9-BBN (50 °C, 90 min) resulted in successful in situ anti-Markovnikov hydroboration of the ligand which, when followed by metalation with CuI or Cu2(OTf)2(C6H6) afforded (BBNTACNtBu)Cu(I) (2-I) or (BBNTACNtBu)Cu(OTf) (2-OTf) as a white powders in 67% and 90% yield respectively over the two steps. In contrast to the post-metallation hydroboration route, the pre-metallation hydroboration route provided the desired borane-appended Cu(I) complexes in high yield.54
To ascertain the molecular structures of the Lewis acid appended complexes, we carried out single crystal XRD studies on complexes 2. Data refinement revealed each molecule displays a pseudo-tetrahedral coordination environment (τ4: 2-I = 0.66; 2-OTf = 0.71)55 around copper that is similar to that of other mononuclear Cu(I) TACN complexes.11, 56–57 Each structure contains a planar, three-coordinate borane (ΣBα: 2-I = 359.6(3)°; 2-OTf = 359.4(4)°) that is far removed from any Lewis basic sites (>3.6 Å), consistent with an unquenched Lewis acid in each case.
Voltammetry experiments suggest the X-type ligand in 2-OTf is labile in comparison to 2-I. The voltammogram (0.2 M [Bu4N][PF6], THF, 200 mV/s) of 2-OTf displays an oxidative event at Epa = 0.38 V (vs. Cp2Fe/Cp2Fe+) that shifts to Epa = 0.16 V upon titration with [Bu4N][OTf] (~2, 4 , and 23 equiv); Δ = −0.22 V). This potential, assessed by square wave voltammetry, observed in the presence of [Bu4N][OTf] (0.09 V vs. Cp2Fe/Cp2Fe+) is similar to that measured for 2-I (−0.03 V), consistent with similar electronic environments in the absence of ligand exchange.
The lability of the X-type ligand in 2-OTf is further borne out by reactivity studies. To baseline the electronics of the new chelating ligand, BBNTACNtBu, we treated each complex 2 with CO. Although substitution of the iodide in 2-I did not occur, the trifluoromethanesulfonate ligand in 2-OTf was readily displaced to afford [(BBNTACNtBu)Cu(CO)][OTf] (2-CO) as a vacuum stable white solid.58 The IR spectrum (KBr) of 2-CO displays a sharp ν(CO) stretch at 2069.5 cm−1 that is at comparable energy to related cationic Cu(I)-CO species chelated by triamine-type ligands59–60 suggesting that Lewis acid incorporation in BBNTACNtBu has negligible effects on its ligand field properties and does not interact with CO.
The absence of a Lewis acid/CO interaction in 2-CO may indicate either a geometric or basicity mismatch with the appended Lewis acid. Previous work from our group showed that cooperative binding can be enhanced by modifying the appended Lewis acid tether length as well as the basicity of the substrate.61 To provide a single atom ligand with less directionality requirements that we anticipated would better accommodate a metal-ligand-interaction, we targeted a copper hydride. Treating a freshly thawed toluene solution of 2-OTf with KBHEt3 afforded a new hydride-containing complex, (BBNTACNtBu)CuH (3). The formation of 3 is reversible: treating 3 with [Ph2NH2][OTf] cleanly regenerates 2-OTf with extrusion of H2 (Figure 3). IR spectroscopy of 3 revealed a strong absorption (KBr) at 1831 cm−1 that is sensitive to deuteride incorporation (Δ = 195 cm−1) and assigned as the ν(Cu-H-B) stretch. This absorption is at lower energy than most reported Cu-H-B interactions, notably from scorpionate complexes (~2350–2250 cm−1),46–48 and is more comparable to two-coordinate (IMes)Cu-HBPh3 (ν(Cu-H-B) = 1693 cm−1).49 The 1H NMR spectrum of 3 (C6D6) displays a broad resonance at −1.79 ppm that disappeared in the (BBNTACNtBu)CuD isotopologue. 11B NMR spectroscopy reveals a broad singlet at −6.44 ppm, indicative of a 4-coordinate, tetrahedral boron environment. Together, these data suggest a copper-hydride that interacts with the Lewis acidic trialkylborane.62
To further clarify the binding mode in 3, we analyzed single, X-ray quality crystals that were obtained by layering a C6H6 solution of 3 with hexamethyldisiloxane at room temperature (Figure 3). Data refinement from the XRD experiment revealed a pseudo-tetrahedral (τ4 = 0.67) copper-hydride that is capped by the appended trialkylborane (ΣBα = 322.7(8)°). The position of the hydride atom in 3 was refined and displayed a Cu-H and B-H distance of 1.560(18) and 1.315(19) Å, respectively (Cu-H-B = 125.77°; Cu-B = 2.5617(15) Å). While the Cu-H distance is similar to that observed in (NHC)CuH-BR3 species,49–50 the B-H distance in 3 is considerably longer than for other reported Cu-H-BR3 examples.63–65 The primary coordination sphere of the Cu(I)-TACN in 3 is very similar to 2-OTf and 2-I (τ4 within 0.04).
Two limiting descriptions of complex 3 can be formulated where it is viewed as either a 1) copper-hydride capped by a trialkylborane, or 2) borohydride anion coordinated to the copper center. While the X-ray and spectroscopic data provide a structural description of the Cu–H–BR3 interaction, we undertook a DFT study of 3 (B3LYP-D3/SVP (PCM = THF)) to provide a clearer electronic description. Wiberg bond analyses revealed a Cu-H bond index of 0.13 and a B-H bond index of 0.79. We compared these values to both limiting bonding descriptions: a terminal Cu–H with no borane interaction ((nBuTACNtBu)CuH, 3’; 0.65) and R3B–H (0.94) (see Figure 3B). These data suggest that 3 is best described as a borohydride anion interacting with copper, with some copper hydride resonance contribution. This depiction has also been invoked in related compounds,50, 66 and we note that similar (NHC)CuH-BR3 complexes can be induced to exhibit Cu-hydride-type reactivity (carbonyl hydrosilylation) in select organometallic transformations.49 Importantly, our analysis of 3 suggests that the Cu-H-BR3 moiety contains both Cu-H and B-H character.
The electronic effect that the appended trialkylborane imparts on the Cu–H–BR3 unit of 3 was probed by electrochemical analysis. Although Stryker’s reagent-type clusters, [(Ar3P)CuH]6, are commonly employed as reducing agents (−1.00 – −1.20 V),51, 67 Cu-H motifs have rarely been subjected to electrochemical analysis, perhaps due to limited stability. In contrast, isolated samples of 3 are indefinitely stable in the solid state at −35 °C and have solution (C6D6 or THF) half-lives of >2 weeks at room temperature.68 Voltammetry experiments of 3 were performed in THF (0.2 M [Bu4N][PF6]) and revealed an irreversible oxidation event at −0.13 V vs. Cp2Fe/Cp2Fe+. The difference in the measured potential of 2-I vs 3 (Δ = 100 mV) is small. Redox potentials between hydride and halide ligands are typically quite different; for instance, in isostructural RhIII-X (X = Cl, H) complexes, the two redox potentials vary by >800 mV.69 We attribute the similar redox potentials for 2-I and 3 to the appended trialkylborane, which we propose syphons electron density from the hydrido-ligand and thus levels the overall redox-potential of the system (redox leveling).
To ascertain the requirement of the boron Lewis acid to provide access to TACN-ligated cuprous hydrides, we attempted to synthesize a Lewis acid-free variant of 3. We designed a new ligand, 1-n-butyl-4,7-di-tert-butyl-1,4,7-triazacyclononane (nBuTACNtBu), where the −(CH2)3BBN substituent was replaced by an inert n-butyl group. Metalation with Cu2(OTf)2(C6H6) proceeded smoothly, affording (nBuTACNtBu)Cu(OTf) (2’-OTf; see SI for full characterization). Unfortunately, attempts to form an analogous copper hydride complex (3’) were unsuccessful. Reactions with 1 equiv. KHBEt3 and 2’-OTf under the same conditions used to prepare 3 yielded free-ligand and an insoluble black precipitate, and neither (nBuTACNtBu)CuH nor (nBuTACNtBu)Cu(HBEt3) were observed, which highlights the importance of the appended Lewis acid in 3 to stabilize the hydrido ligand.
We used DFT analyses to provide further insight into the electronic structure of 2-I vs. 3 as well as their Lewis acid-free counterparts, 2’-I and 3’.70 The HOMOs of 2-I and 2’-I were calculated to be very similar, separated by 0.10 eV (−4.53 and −4.43 eV respectively). In comparison to complexes 2, the corresponding hydride 3 is stabilized by 0.37 eV (−4.90 eV) while 3’ is destabilized by 0.65 eV (−3.78 eV). These results are consistent with the ability of the appended Lewis acidic borane to provide substantial stabilization to the otherwise unstable Cu-H unit (Table S18 and Figure S83).
Related boron-capped Cu-H systems have been shown to reduce C=O bonds,49, 71–73 through hydride transfer. To evaluate the ability of 3 to similarly deliver a hydride, we introduced CO2 to 3 (Figure 4A). Treating a C6D6 solution of 3 with CO2 (20 psig) at room temperature resulted in the precipitation of a new copper-containing species (4) as a white solid. 1H NMR spectroscopy (THF) is consistent with reduction of CO2 to CO2H−, indicated by a diagnostic resonance at 8.21 ppm assigned as the formate C-H bond.74 An XRD experiment revealed that 4 is a dimeric species, [(BBNTACNtBu)Cu(O2CH)]2, containing two pseudotetrahedral copper centers (τ4= 0.70, 0.67), where the pyramidalized boron Lewis acid (ΣBα = 322.14(18)°) from one molecule interacts (B-O = 1.609(4) Å) with the copper-bound formate (Cu-O = 1.9537(17) Å) of an adjacent molecule. We propose that the dimeric form is favored for this substrate due to a geometrical mismatch for intramolecular binding of this substrate through the Cu and borane Lewis acid, which we note is distinct from the hydride and consistent with recently described criteria for cooperative binding.33
To further investigate the preference of 4 to form a dimer, rather than a monomer, we examined the energetic landscape for monomer/dimer equilibra by DFT. The dimeric (Cu-OCHO-BR3)2 complex is calculated to be more stable than its monomeric counterpart (ΔG=−11.49 kcal/mol; Figure 4A, cutout). This result clarifies the thermodynamic driving force to form the dimer, and may indicate that a different geometry or appended Lewis acid tether length may be required for the stabilization of a monomeric formate complex. Analysis of the bonding within formate was consistent with XRD data, with calculated Wiberg bond indices indicating slightly more double bond character in the copper-bound C-O (1.44) than in the boron-bound C-O (1.30). This bonding depiction is comparable to that observed in systems employing strong exogenous Lewis acids.75–77 Importantly, compound 4 demonstrates that the same electronic structure can be attained by using intramolecular Lewis acids that are considerably weaker.78
The availability of single electron redox chemistry to 3 provides a strategy to modify the reactivity of a copper hydride. Further analysis of the electrochemistry of 3 indicated that additional chemical reactivity pathways occur following oxidation. After scanning more positive than −0.13 V, the voltammogram of 3 featured a fully reversible oxidation event at +0.39 V (vs Cp2Fe/Cp2Fe+), identified as [(BBNTACNtBu)Cu]+. (Note: this redox event was observed in our analysis of 2-OTf) This result suggests that oxidation of 3 results in loss of an H-atom. To investigate the products of oxidation, we treated 3 with a chemical oxidant in a sealed vessel. When 1 equiv. ferrocenium trifluoromethanesulfonate ([Cp2Fe][OTf]) was added to a THF solution of 3 in a sealed J-Young NMR tube, we observed 0.42 (±0.03) equiv of H2 (Figure 4B-1), in addition to 2-OTf. To clarify the pathway by which H2 forms, we repeated this reaction using the deuteride analogue of 3 in protio THF solvent. Using these reagents, only D2 was formed (no HD or H2). The stoichiometry of H2 formation and the deuterium labeling results are consistent with a bimolecular process where H2 is derived solely from the copper-(boro)hydride.
The requirement of the appended Lewis acid to stabilize the Cu-H in 3 suggested that the Lewis acidic group might also enable unique reactivity. We undertook a series of control reactions to examine the extent to which H2 generation could be accessed from different synthetic routes, and influenced by the appended Lewis acid. Mayer and coworkers have tabulated the H-atom BDFE values of mixtures of acids and reductants that can be used to deliver H• equivalents.79 We selected a combination of pyridinium tetrakis[3,5-bis(trifluoromethyl)phenyl]borate ([pyr-H][BArF24]) with decamethylferrocene (Cp*2Fe), which has a BDFE of 58.0 kcal/mol in acetonitrile and is thus a weak H-atom donor.79 While the combination of [pyr-H][BArF24] and Cp*2Fe alone does not generate H2, addition of [pyr-H][BArF24] and Cp*2Fe to a C6D6 solution of 2-OTf resulted in the formation of H2 along with the reformation of 2-OTf (Figure 4B-2).80 The observation of H2 formation suggests that this reaction may share a common intermediate with that of chemical oxidation of 3: a transient Cu(II)-H. To evaluate the role(s) of the appended Lewis acid, we examined reactivity employing 2’-OTf, which does not contain a Lewis acid, and found that H2 does not form under analogous conditions (Figure 4B-3).
Finally, we sought to circumvent the H2 formation pathway by intercepting the product of chemical oxidation of 3 with a sterically-accessible alkyne substrate. We found that there is no reaction between 3 and 10 equiv. phenylacetylene (Figure 4C-1). In contrast, when the reaction was repeated in the presence of 1.2 equiv. [Cp2Fe][OTf] oxidant, the reduction product (styrene) formed, albeit in low yield (Figure 4C-2).81
We examined two sets of control reactions to ascertain the requirements of each component of 3 to form styrene: the Lewis acid and copper ion. To evaluate the role(s) of the appended Lewis acid, we examined analogous reactivity using mixtures of 2’-OTf and KHBEt3 in place of 3, providing analogue of 3 with an exogenous borane. Solutions of 2’-OTf, KHBEt3, phenylacetylene, and [Cp2Fe][OTf] were sequentially layered and frozen in a J-Young NMR tube to prevent unwanted side reactions between individual reagents. Subsequent thawing and mixing of the reagents revealed only H2, and no styrene (Figure 4C-3). Similarly, to evaluate the role(s) of the Cu ion, when a solution of KHBEt3 was treated under analogous conditions, only H2 formed.82–83 Combined, these results confirm that in 3, both the Cu and the appended borane are crucial cooperative components for phenylacetylene reduction and we propose this is due to unique stabilization imparted to an otherwise unstable copper hydride.
Although an oxidant was added to 3 in the above reactions, we did not observe solution colors that are characteristic of Cu(II) at the end of reactions. We further investigated this observation using EPR and NMR spectroscopy to assess the oxidation state of Cu. Addition of 0.9 equiv. [Cp2Fe][OTf] to a mixture of 3 and phenylacetylene at −78 °C generated a transient EPR signal assigned as Cu(II) which decayed over time, and no paramagnetic signals were observed at the end of the reaction (2h at room temperature; see SI:Figure S75–S76). These results are consistent with 1H NMR spectra obtained after completion of the reaction, which indicated 2-OTf formed in high yield. The results of both experiments provide key insight into the reactivity of 3: addition of an oxidant to the Cu(I)-H oxidizes at the Cu center, which induces reduction of phenylacetylene and reforms Cu(I). This reactivity is distinct from previous Cu(I)-H complexes that react directly with alkynes and represents an unusual pathway because addition of an oxidant is required for alkyne reduction with 3.
CONCLUSION
In conclusion, we have presented a tetrahedral-enforcing tridentate ligand scaffold containing an appended borane Lewis acid and metalation with Cu(I). The appended Lewis acid enables the isolation of an otherwise unstable Cu(I) hydride that exhibits distinct reactivity modes as a function of chemical oxidation: hydride transfer to CO2 in the copper(I) state and oxidant-induced H2 evolution as well as alkyne reduction. Access to these divergent reactivity modes is attributed to an intramolecular appended Lewis acid. Ongoing work is focused on clarifying the mechanistic details as well as extending these concepts to promote selective reduction of more inert small molecules.
Supplementary Material
ACKNOWLEDGMENT
We thank Michela Maiola for preliminary synthetic efforts, Prof. Charles McCrory for insightful discussions regarding electrochemistry, and Prof. Neal Mankad for thoughtful discussions regarding copper hydride reactivity.
Funding Sources
This work was supported by the NIGMS of the NIH under Award 1R35GM136360–01. N.K.S.is a Camille Dreyfus Teacher-Scholar. The X-ray diffractometers were funded by the NSF (CHE 1625543).
Footnotes
The authors declare no competing financial interest.
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website.
Experimental procedures, spectroscopic characterization of all species, and cartesian coordinates (PDF).
Cartesian coordinates (XYZ)
REFERENCES
- 1.Lundgren RJ; Stradiotto M, Key Concepts in Ligand Design. In Ligand Design in Metal Chemistry, 2016; pp 1–14. [Google Scholar]
- 2.Drover MW, A guide to secondary coordination sphere editing. Chem. Soc. Rev 2022, 51 (6), 1861–1880. [DOI] [PubMed] [Google Scholar]
- 3.Moret M-E; Peters JC, Terminal Iron Dinitrogen and Iron Imide Complexes Supported by a Tris(phosphino)borane Ligand. Angew. Chem. Int. Ed 2011, 50 (9), 2063–2067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Speelman AL; Čorić I; Van Stappen C; DeBeer S; Mercado BQ; Holland PL, Nitrogenase-Relevant Reactivity of a Synthetic Iron–Sulfur–Carbon Site. J. Am. Chem. Soc 2019, 141 (33), 13148–13157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Saouma CT; Peters JC, M≡E and M=E Complexes of Iron and Cobalt that Emphasize Three-fold Symmetry (E = O, N, NR). Coord. Chem. Rev 2011, 255 (7–8), 920–937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Wainwright KP, Synthetic and structural aspects of the chemistry of saturated polyaza macrocyclic ligands bearing pendant coordinating groups attached to nitrogen. Coord. Chem. Rev 1997, 166, 35–90. [Google Scholar]
- 7.Haines RI, Pendant-Arm Tri- and Tetraazamacrocycles: Synthesis, Structure and Properties of their First-Row Transition Metal Complexes. Rev. Inorg. Chem 2001, 21 (3–4), 165–206. [Google Scholar]
- 8.Hage R; Iburg JE; Kerschner J; Koek JH; Lempers ELM; Martens RJ; Racherla US; Russell SW; Swarthoff T; van Vliet MRP; Warnaar JB; Wolf L.v. d. ; Krijnen B, Efficient manganese catalysts for low-temperature bleaching. Nature 1994, 369 (6482), 637–639. [Google Scholar]
- 9.Sugiarto; Kawamoto K; Hayashi Y, Artificial bioinorganic clusters of dinuclear 3d-transition metal ions coordinated by an inorganic coordination ligand. J. Inorg. Biochem 2019, 201, 110821. [DOI] [PubMed] [Google Scholar]
- 10.Wieghardt K, 1,4,7-Triazacyclononane and N,N’N”-trimethyl-1,4,7-triazacyclononane - two versatile macrocycles for the synthesis of monomeric and oligomeric metal complexes. Pure. Appl. Chem 1988, 60 (4), 509–516. [Google Scholar]
- 11.Thangavel A; Wieliczko M; Bacsa J; Scarborough CC, 1,4,7-Triazacyclononane Ligands Bearing Tertiary Alkyl Nitrogen Substituents. Inorg. Chem 2013, 52 (22), 13282–13287. [DOI] [PubMed] [Google Scholar]
- 12.Karahalis GJ; Thangavel A; Chica B; Bacsa J; Dyer RB; Scarborough CC, Synthesis and Catalytic Reactivity of a Dicopper(II) μ-η2:η2-Peroxo Species Supported by 1,4,7-Tri-tert-butyl-1,4,7-triazacyclononane. Inorg. Chem 2016, 55 (3), 1102–1107. [DOI] [PubMed] [Google Scholar]
- 13.Warm K; Paskin A; Kuhlmann U; Bill E; Swart M; Haumann M; Dau H; Hildebrandt P; Ray K, A Pseudotetrahedral Terminal Oxoiron(IV) Complex: Mechanistic Promiscuity in C−H Bond Oxidation Reactions. Angew. Chem. Int. Ed 2021, 60 (12), 6752–6756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. We note that there are reports of TACN ligands that incorporate H-bond donor groups. For this work, we draw the disctinction of referring to non-protic Lewis acidic groups. (See ref.(a)Macedi, E.; Bencini, A.; Caltagirone, C.; Lippolis, V., The design of TACN-based molecular systems for different supramolecular functions. Coord. Chem. Rev. 407 (2020) 213151;(b)Joshi T.; Kubeil M.; Nsubuga A.; Singh G.; Gasser G.; Stephan H., Harnessing the Coordination Chemistry of 1,4,7-Triazacyclononane for Biomimicry and Radiopharmaceutical Applications. ChemPlusChem, 2018, 83, 554 –564 and references therein.) [DOI] [PubMed]
- 15.Liu J; Chakraborty S; Hosseinzadeh P; Yu Y; Tian S; Petrik I; Bhagi A; Lu Y, Metalloproteins Containing Cytochrome, Iron–Sulfur, or Copper Redox Centers. Chemical Reviews 2014, 114 (8), 4366–4469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Sippel D; Rohde M; Netzer J; Trncik C; Gies J; Grunau K; Djurdjevic I; Decamps L; Andrade Susana LA; Einsle O, A bound reaction intermediate sheds light on the mechanism of nitrogenase. Science 2018, 359 (6383), 1484–1489. [DOI] [PubMed] [Google Scholar]
- 17.MacBeth CE; Golombek AP; Young VG; Yang C; Kuczera K; Hendrich MP; Borovik AS, O2 Activation by Nonheme Iron Complexes: A Monomeric Fe(III)-Oxo Complex Derived From O2. Science 2000, 289 (5481), 938. [DOI] [PubMed] [Google Scholar]
- 18.Matson EM; Park YJ; Fout AR, Facile Nitrite Reduction in a Non-heme Iron System: Formation of an Iron(III)-Oxo. J. Am. Chem. Soc 2014, 136 (50), 17398–17401. [DOI] [PubMed] [Google Scholar]
- 19.Wallen CM; Bacsa J; Scarborough CC, Hydrogen Peroxide Complex of Zinc. J. Am. Chem. Soc 2015, 137 (46), 14606–14609. [DOI] [PubMed] [Google Scholar]
- 20.Widger LR; Davies CG; Yang T; Siegler MA; Troeppner O; Jameson GNL; Ivanović-Burmazović I; Goldberg DP, Dramatically Accelerated Selective Oxygen-Atom Transfer by a Nonheme Iron(IV)-Oxo Complex: Tuning of the First and Second Coordination Spheres. J. Am. Chem. Soc 2014, 136 (7), 2699–2702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Reduction of typically inert small molecule substrates often requires highly reducing metal hydrides (See ref. 70)
- 22.Miller AJ; Labinger JA; Bercaw JE, Reductive coupling of carbon monoxide in a rhenium carbonyl complex with pendant Lewis acids. J. Am. Chem. Soc 2008, 130 (36), 11874–5. [DOI] [PubMed] [Google Scholar]
- 23.Miller AJM; Labinger JA; Bercaw JE, Homogeneous CO Hydrogenation: Ligand Effects on the Lewis Acid-Assisted Reductive Coupling of Carbon Monoxide. Organometallics 2010, 29 (20), 4499–4516. [Google Scholar]
- 24.Podiyanachari SK; Fröhlich R; Daniliuc CG; Petersen JL; Mück-Lichtenfeld C; Kehr G; Erker G, Hydrogen Activation by an Intramolecular Boron Lewis Acid/Zirconocene Pair. Angew. Chem. Int. Ed 2012, 51 (35), 8830–8833. [DOI] [PubMed] [Google Scholar]
- 25.Ostapowicz TG; Merkens C; Hölscher M; Klankermayer J; Leitner W, Bifunctional Ruthenium(II) Hydride Complexes with Pendant Strong Lewis Acid Moieties: Structure, Dynamics, and Cooperativity. J. Am. Chem. Soc 2013, 135 (6), 2104–2107. [DOI] [PubMed] [Google Scholar]
- 26.Barnett BR; Moore CE; Rheingold AL; Figueroa JS, Cooperative Transition Metal/Lewis Acid Bond-Activation Reactions by a Bidentate (Boryl)iminomethane Complex: A Significant Metal–Borane Interaction Promoted by a Small Bite-Angle LZ Chelate. J. Am. Chem. Soc 2014, 136 (29), 10262–10265. [DOI] [PubMed] [Google Scholar]
- 27.Ostapowicz TG; Merkens C; Holscher M; Klankermayer J; Leitner W, Bifunctional ruthenium(II) hydride complexes with pendant strong Lewis acid moieties: structure, dynamics, and cooperativity. J. Am. Chem. Soc 2013, 135 (6), 2104–7. [DOI] [PubMed] [Google Scholar]
- 28.Tutusaus O; Ni C; Szymczak NK, A Transition Metal Lewis Acid/Base Triad System for Cooperative Substrate Binding. J. Am. Chem. Soc 2013, 135 (9), 3403–3406. [DOI] [PubMed] [Google Scholar]
- 29.Tseng K-NT; Kampf JW; Szymczak NK, 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 (33), 10378–10381. [DOI] [PubMed] [Google Scholar]
- 30.Kiernicki JJ; Zeller M; Szymczak NK, Hydrazine Capture and N–N Bond Cleavage at Iron Enabled by Flexible Appended Lewis Acids. J. Am. Chem. Soc 2017, 139 (50), 18194–18197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Kiernicki JJ; Zeller M; Szymczak NK, Requirements for Lewis Acid-Mediated Capture and N–N Bond Cleavage of Hydrazine at Iron. Inorg. Chem 2019, 58 (2), 1147–1154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Kiernicki JJ; Norwine EE; Zeller M; Szymczak NK, Tetrahedral iron featuring an appended Lewis acid: distinct pathways for the reduction of hydroxylamine and hydrazine. Chem. Commun 2019, 55 (79), 11896–11899. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Kiernicki JJ; Norwine EE; Lovasz MA; Zeller M; Szymczak NK, Mobility of Lewis acids within the secondary coordination sphere: toward a model for cooperative substrate binding. Chem. Commun 2020, 56 (86), 13105–13108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Kiernicki JJ; Shanahan JP; Zeller M; Szymczak NK, Tuning ligand field strength with pendent Lewis acids: access to high spin iron hydrides. Chem. Sci 2019, 10 (21), 5539–5545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Pirnot MT; Wang Y-M; Buchwald SL, Copper Hydride Catalyzed Hydroamination of Alkenes and Alkynes. Angew. Chem. Int. Ed 2016, 55 (1), 48–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Rendler S; Oestreich M, Polishing a Diamond in the Rough: “Cu-H” Catalysis with Silanes. Angew. Chem. Int. Ed 2007, 46 (4), 498–504. [DOI] [PubMed] [Google Scholar]
- 37.Uehling MR; Rucker RP; Lalic G, Catalytic Anti-Markovnikov Hydrobromination of Alkynes. J. Am. Chem. Soc 2014, 136 (24), 8799–8803. [DOI] [PubMed] [Google Scholar]
- 38.Lipshutz BH; Shimizu H, Copper(I)-Catalyzed Asymmetric Hydrosilylations of Imines at Ambient Temperatures. Angew. Chem. Int. Ed 2004, 43 (17), 2228–2230. [DOI] [PubMed] [Google Scholar]
- 39.Díez-González S; Nolan SP, Copper, Silver, and Gold Complexes in Hydrosilylation Reactions. Acc. Chem. Res 2008, 41 (2), 349–358. [DOI] [PubMed] [Google Scholar]
- 40.Tran BL; Neisen BD; Speelman AL; Gunasekara T; Wiedner ES; Bullock RM, Mechanistic Studies on the Insertion of Carbonyl Substrates into Cu-H: Different Rate-Limiting Steps as a Function of Electrophilicity. Angew. Chem. Int. Ed 2020, 59 (22), 8645–8653. [DOI] [PubMed] [Google Scholar]
- 41.Mahoney WS; Stryker JM, Hydride-mediated homogeneous catalysis. Catalytic reduction of α,β-unsaturated ketones using [(Ph3P)CuH]6 and H2. J. Am. Chem. Soc 1989, 111 (24), 8818–8823. [Google Scholar]
- 42.Jordan AJ; Lalic G; Sadighi JP, Coinage Metal Hydrides: Synthesis, Characterization, and Reactivity. Chem. Rev 2016, 116 (15), 8318–8372. [DOI] [PubMed] [Google Scholar]
- 43.Suess AM; Uehling MR; Kaminsky W; Lalic G, Mechanism of Copper-Catalyzed Hydroalkylation of Alkynes: An Unexpected Role of Dinuclear Copper Complexes. J. Am. Chem. Soc 2015, 137 (24), 7747–7753. [DOI] [PubMed] [Google Scholar]
- 44.Hazra A; Kephart JA; Velian A; Lalic G, Hydroalkylation of Alkynes: Functionalization of the Alkenyl Copper Intermediate through Single Electron Transfer Chemistry. J. Am. Chem. Soc 2021, 143 (21), 7903–7908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Speelman AL; Tran BL; Erickson JD; Vasiliu M; Dixon DA; Bullock RM, Accelerating the insertion reactions of (NHC)Cu–H via remote ligand functionalization. Chem. Sci 2021, 12 (34), 11495–11505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Dyson G; Hamilton A; Mitchell B; Owen GR, A new family of flexible scorpionate ligands based on 2-mercaptopyridine. Dalton Trans. 2009, (31), 6120–6126. [DOI] [PubMed] [Google Scholar]
- 47.Neshat A; Varestan S; Halvagar MR, Cu(I) complexes of dihydrobis(2-mercapto-benzimidazolyl)borate and dihydrobis(2-mercapto-benzothiazolyl)borate ligands: structural, photophysical and computational studies. New J. Chem 2018, 42 (3), 2036–2046. [Google Scholar]
- 48.Owen GR; Gould PH; Moore A; Dyson G; Haddow MF; Hamilton A, Copper and silver complexes bearing flexible hybrid scorpionate ligand mpBm. Dalton Trans. 2013, 42 (31), 11074–11081. [DOI] [PubMed] [Google Scholar]
- 49.Ritter F; Mukherjee D; Spaniol TP; Hoffmann A; Okuda J, A Masked Cuprous Hydride as a Catalyst for Carbonyl Hydrosilylation in Aqueous Solutions. Angew. Chem. Int. Ed 2019, 58 (6), 1818–1822. [DOI] [PubMed] [Google Scholar]
- 50.Collins LR; Rajabi NA; Macgregor SA; Mahon MF; Whittlesey MK, Experimental and Computational Studies of the Copper Borate Complexes [(NHC)Cu(HBEt3)] and [(NHC)Cu(HB(C6F5)3)]. Angew. Chem. Int. Ed 2016, 55 (50), 15539–15543. [DOI] [PubMed] [Google Scholar]
- 51.Eberhart MS; Norton JR; Zuzek A; Sattler W; Ruccolo S, Electron Transfer from Hexameric Copper Hydrides. J. Am. Chem. Soc 2013, 135 (46), 17262–17265. [DOI] [PubMed] [Google Scholar]
- 52.Zhang S; Fallah H; Gardner EJ; Kundu S; Bertke JA; Cundari TR; Warren TH, A Dinitrogen Dicopper(I) Complex via a Mixed-Valence Dicopper Hydride. Angew. Chem. Int. Ed 2016, 55 (34), 9927–9931. [DOI] [PubMed] [Google Scholar]
- 53.Rubino JT; Franz KJ, Coordination chemistry of copper proteins: How nature handles a toxic cargo for essential function. J. Inorg. Biochem 2012, 107 (1), 129–143. [DOI] [PubMed] [Google Scholar]
- 54. Treating 1-PF6 with one equiv. of 9-BBN resulted in gradual precipitation of an unidentified solid without 1H NMR evidence of allylic hydroboration. Recently, we identified ancillary ligand lability as a key consideration for late-stage hydroboration (see ref. Kiernicki, J. J.; Zeller, M.; Szymczak, N. K., Requirements for Late-Stage Hydroboration of Pyridine N-Heterocyclic Carbene Iron(0) Complexes: The Role of Ancillary Ligands. Organometallics 2021, 40 (15), 2658–2665). We propose the enhanced lability of acetonitrile vs. iodide results in deleterious reactivity of the hydroborating agent.
- 55.Yang L; Powell DR; Houser RP, Structural variation in copper(i) complexes with pyridylmethylamide ligands: structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, (9), 955–964. [DOI] [PubMed] [Google Scholar]
- 56.Pal N; Majumdar A, Controlling the Reactivity of Bifunctional Ligands: Carboxylate-Bridged Nonheme Diiron(II) Complexes Bearing Free Thiol Groups. Inorg. Chem 2016, 55 (6), 3181–3191. [DOI] [PubMed] [Google Scholar]
- 57.Lam BMT; Halfen JA; Young VG; Hagadorn JR; Holland PL; Lledós A; Cucurull-Sánchez L; Novoa JJ; Alvarez S; Tolman WB, Ligand Macrocycle Structural Effects on Copper−Dioxygen Reactivity. Inorg. Chem 2000, 39 (18), 4059–4072. [DOI] [PubMed] [Google Scholar]
- 58. Elevated CO pressures result in demetalation of the TACN ligand in 2-I, see SI.
- 59.Matsumoto J; Kajita Y; Masuda H, Synthesis and Characterization of a (μ-η2:η2-Peroxido)dicopper(II) Complex with N,N′,N″-Triisopropyl-cis,cis-1,3,5-triaminocyclohexane (R3TACH, R = iPr): Selective Preparation of (μ-η2:η2-Peroxido)dicopper(II) and Bis(μ-oxido)dicopper(III) Species Regulated by Substituent Groups. Eur. J. Inorg. Chem 2012, (26), 4149–4158. [Google Scholar]
- 60.Kujime M; Kurahashi T; Tomura M; Fujii H, 63Cu NMR Spectroscopy of Copper(I) Complexes with Various Tridentate Ligands:– CO as a Useful 63Cu NMR Probe for Sharpening 63Cu NMR Signals and Analyzing the Electronic Donor Effect of a Ligand. Inorg. Chem 2007, 46 (2), 541–551. [DOI] [PubMed] [Google Scholar]
- 61.Kiernicki JJ; Norwine EE; Zeller M; Szymczak NK, Substrate Specific Metal–Ligand Cooperative Binding: Considerations for Weak Intramolecular Lewis Acid/Base Pairs. Inorg. Chem 2021, 60 (18), 13806–13810. [DOI] [PubMed] [Google Scholar]
- 62. Note that the 1H NMR chemical shift of the hydride resonance in two-coordinate complexes of the type (NHC)CuH-BR3 are highly dependent upon the identity of the Lewis acid (BEt3 = −2.60 ppm; B(C6F5)3 = 2.08 ppm; BPh3 = 3.42 ppm) and therefore the extent of metal-hydride (or borohydride) character in complex 3 cannot be inferred from NMR alone (see ref 49 and 50).
- 63.van Dijkman TF; de Bruijn HM; Siegler MA; Bouwman E, Bright Cyan Phosphorescence of a (Phosphane)copper(I) Complex of the Tri-hydrido-pyrazolylborate Ligand H3B(3,5-Ph2Pz)–. Eur. J. Inorg. Chem 2015, 2015 (32), 5387–5394. [Google Scholar]
- 64.Ghilardi CA; Midollini S; Orlandini A, A unidentate attachment of the tetrahydroborate group. Crystal and molecular structure of (tetrahydroborato)[1,1,1-tris((diphenylphosphino)methyl)ethane-P,P’,P”]copper(I). Inorg. Chem 1982, 21 (11), 4096–4098. [Google Scholar]
- 65.Takusagawa F; Fumagalli A; Koetzle TF; Shore SG; Schmitkons T; Fratini AV; Morse KW; Wei C-Y; Bau R, Neutron and x-ray diffraction studies of tris(methyldiphenylphosphine)[tetrahydroborato(1-)]copper, Cu[P(C6H5)2CH3]3(BH4). The first accurate characterization of an unsupported metal-hydrogen-boron bridge bond. J. Am. Chem. Soc 1981, 103 (17), 5165–5171. [Google Scholar]
- 66.Aloisi A; Crochet É; Nicolas E; Berthet J-C; Lescot C; Thuéry P; Cantat T, Copper–Ligand Cooperativity in H2 Activation Enables the Synthesis of Copper Hydride Complexes. Organometallics 2021, 40 (13), 2064–2069. [Google Scholar]
- 67.Liu S; Eberhart MS; Norton JR; Yin X; Neary MC; Paley DW, Cationic Copper Hydride Clusters Arising from Oxidation of (Ph3P)6Cu6H6. J. Am. Chem. Soc 2017, 139 (23), 7685–7688. [DOI] [PubMed] [Google Scholar]
- 68. A black precipitate forms with evidence of H2 formation.
- 69.Hu Y; Li L; Shaw AP; Norton JR; Sattler W; Rong Y, Synthesis, Electrochemistry, and Reactivity of New Iridium(III) and Rhodium(III) Hydrides. Organometallics 2012, 31 (14), 5058–5064. [Google Scholar]
- 70.Wiedner ES; Chambers MB; Pitman CL; Bullock RM; Miller AJM; Appel AM, Thermodynamic Hydricity of Transition Metal Hydrides. Chem. Rev 2016, 116 (15), 8655–8692. [DOI] [PubMed] [Google Scholar]
- 71.Romero EA; Zhao T; Nakano R; Hu X; Wu Y; Jazzar R; Bertrand G, Tandem copper hydride–Lewis pair catalysed reduction of carbon dioxide into formate with dihydrogen. Nat. Catal 2018, 1 (10), 743–747. [Google Scholar]
- 72.Zhang L; Cheng J; Hou Z, Highly efficient catalytic hydrosilylation of carbon dioxide by an N-heterocyclic carbene copper catalyst. Chem. Commun 2013, 49 (42), 4782–4784. [DOI] [PubMed] [Google Scholar]
- 73.Jordan AJ; Wyss CM; Bacsa J; Sadighi JP, Synthesis and Reactivity of New Copper(I) Hydride Dimers. Organometallics 2016, 35 (5), 613–616. [Google Scholar]
- 74.The C=O reduction chemistry observed by the in-tramolecular M-H-BR3 unit in 3 is reminicent of that employed in bis(pyrazolyl)hydroborato zinc complexes as a method to alter the third chelating functionality of scorpionate complexes. See ref.; a) Ghosh P; Parkin G, Mimicking the binding of glutamate to zinc in thermolysin and carboxypeptidase: the synthesis of [η3-(HCO2)BpBu,Pr]ZnCl by insertion of CO2 into a B–H bond of the bis(pyrazolyl)hydroborato zinc complex [BpBu,Pr]ZnCl. J. Chem. Soc., Dalton Trans 1998, (14), 2281–2284. [Google Scholar]; b) Ghosh P; Parkin G, Modeling the active sites of bacteriophage T7 lysozyme, bovine 5-aminolevulinate dehydratase, and peptide deformylase: synthesis and structural characterization of a bis(pyrazolyl)(thioalkoxy)hydroborato zinc complex, [(Ph2CHS)BpBut,Pri]ZnI. Chem. Commun 1998, (3), 413–414. [Google Scholar]; c) Dowling C; Parkin G, Elaboration of the bis(pyrazolyl)hydroborato ligand [BpBut,Pri] into the NNO donor ligand, [(MeO)BpBut,Pri]: Structural characterization of a complex in which the [(MeO)BpBut,Pri] ligand models the binding of zinc to the peptide backbone in thermolysin. Polyhedron 1996, 15 (14), 2463–2465. [Google Scholar]
- 75.Ríos P; Curado N; López-Serrano J; Rodríguez A, Selective reduction of carbon dioxide to bis(silyl)acetal catalyzed by a PBP-supported nickel complex. Chem. Commun 2016, 52 (10), 2114–2117. [DOI] [PubMed] [Google Scholar]
- 76.Jiang Y; Blacque O; Fox T; Berke H, Catalytic CO2 Activation Assisted by Rhenium Hydride/B(C6F5)3 Frustrated Lewis Pairs—Metal Hydrides Functioning as FLP Bases. J. Am. Chem. Soc 2013, 135 (20), 7751–7760. [DOI] [PubMed] [Google Scholar]
- 77.Berkefeld A; Piers WE; Parvez M; Castro L; Maron L; Eisenstein O, Decamethylscandocinium-hydrido-(perfluorophenyl)borate: fixation and tandem tris(perfluorophenyl)borane catalysed deoxygenative hydrosilation of carbon dioxide. Chem. Sci 2013, 4 (5), 2152–2162. [Google Scholar]
- 78.Comparable exogenous Lewis acids were used in two W-OCHO-BEt3 structures and display equivalent C-O bond lengths. See; a) Chen Z; Schmalle HW; Berke H CCDC 647509: Experimental Crystal Structure Determination, 2016, DOI: 10.5517/ccdc.csd.ccpqsdk, and [DOI] [Google Scholar]; b) Chen Z; Schmalle HW; Berke H CCDC 647510: Experimental Crystal Structure Determination, 2016, DOI: 10.5517/ccdc.csd.ccpqsfl. [DOI] [Google Scholar]
- 79.Agarwal RG; Coste SC; Groff BD; Heuer AM; Noh H; Parada GA; Wise CF; Nichols EM; Warren JJ; Mayer JM, Free Energies of Proton-Coupled Electron Transfer Reagents and Their Applications. Chem. Rev 2022, 122 (1), 1–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80. In the absence of the decamethylferrocene, we note that H2 is still observed, albeit in lower yield and we attribute this result to Cu(I) serving as a sacrificial reductant (instead of decamethylferrocene) to generatate a similarly weak source of H•.
- 81. In this reaction, we note that H2 is also observed, and we attribute this result to a competitive bimolecular H2 formation pathway.
- 82.Chemical oxidation of [H-BR3]- adducts are established to follow an ECC process to produce H2 through a Bronsted acid/base pathway. (see ref.; (a) Lawrence EJ; Oganesyan VS; Hughes DL; Ashley; Wildgoose GG, An Electrochemical Study of Frustrated Lewis Pairs: A Metal-Free Route to Hydrogen Oxidation. J. Am. Chem. Soc 2014, 136, 6031–6036; [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Blagg RJ; Wildgoose GG, H2 activation using the first 1 : 1 : 1 hetero-tri(aryl)borane RSC Adv. 2016, 6, 42421–42427; [Google Scholar]; (c) Lawrence EJ; Blagg; Hughes DL; Ashley; Wildgoose GG, A Combined “Electrochemical–Frustrated Lewis Pair” Approach to Hydrogen Activation: Surface Catalytic Effects at Platinum Electrodes Chem. Eur. J 2015, 21, 900–906; [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Lawrence EJ; Herrington TJ; Ashley AE; Wildgoose GG, Metal-Free Dihydrogen Oxidation by a Borenium Cation: A Combined Electrochemical/Frustrated Lewis Pair Approach Angew. Chem. Int. Ed 2014, 53, 9922–9925; [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Blagg RJ; Lawrence EJ; Resner K; Oganesyan VS; Herrington TJ; Ashley AE; Wildgoose GG, Exploring structural and electronic effects in three isomers of tris{bis(trifluoromethyl)phenyl}borane: towards the combined electrochemical-frustrated Lewis pair activation of H2 Dalton trans. 2016, 45, 6023–6031.) [DOI] [PubMed] [Google Scholar]
- 83.We also prepared the analogous Zn(II)-H-B complex, characterized by 1H NMR, 11B NMR, and IR spectroscopy. When the Zn-H complex was employed under the standard conditions for phenylacetylene reduction (10 equiv. phenylacetylene, 1.2 equiv. [Cp2Fe][OTf]), no styrene was observed.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.