Abstract
In contrast to the rapid growth of synthetic electrochemistry in recent years, enantioselective catalytic methods powered by electricity remain rare. In this work, we report the development of a highly enantioselective method for the electrochemical cyanophosphinoylation of vinylarenes. A new family of serine-derived chiral bisoxazolines with ancillary coordination sites were identified as optimal ligands.
Owing to the prevalence of C=C bonds in feedstock chemicals and synthetic intermediates, the heterodifunctionalization of alkenes provides an efficient strategy for rapidly increasing the complexity of molecules in organic synthesis.1 In this context, enantioselective methods that provide optically active compounds are particularly valuable in synthetic and medicinal applications.2 We3 and others4 have recently demonstrated electrocatalysis as a viable and potentially general approach for the difunctionalization of alkenes. Gifted by the many unique attributes of electrochemistry,5 these reactions are frequently highly efficient, selective, operationally simple, and environmentally friendly. Despite these advances, electro-synthetic methods that enable asymmetric alkene functionalization remain elusive.6 In this report, we describe our development of highly enantioselective cyanophosphinoylation of alkenes through the rational design of ligands and optimization of electrolysis conditions.
In our previous studies, we have demonstrated the use of anodically coupled electrolysis (Scheme 1A) for the chlorotrifluoromethylation7 and chloroalkylation8 of alkenes. This strategy combines two distinct anodic events occurring in parallel to generate two radical intermediates simultaneously. Regulated by a redox-active catalyst, the addition of these radicals across the alkene π-bond takes place chemo- and regioselectively. Given the hypothesized mechanism, it is plausible to render these processes enantioselective when an appropriate chiral ligand is used. We set out to investigate this notion in the context of alkene cyanophosphinoylation by means of electrochemical generation of cyano- and phosphinoyl radical equivalents (Scheme 1B). Organophosphorous groups are prevalent in medicine, agrochemicals, and molecular catalysts;9 nitriles are versatile functional groups in organic synthesis. The envisioned reaction would install two useful functional handles to the alkene in a single step, providing unique structural patterns10 that can be further elaborated to useful products (vide infra).
Scheme 1. Mechanistic Working Hypothesis.
In the envisioned transformation, a secondary phosphine oxide11 is oxidized on the anode to form the corresponding P-centered radical (Scheme 1A, X•).12 This step can be direct or mediated by a catalyst. Meanwhile, CN− is converted to CN• in the form of a metal−cyano complex (catOX−Y). In this context, it has been established that CuII−CN complexes can behave as a latent CN• in radical cyanation reactions.13 We envision, upon the addition of the transient P-centered radical to the alkene, the nascent carbon-centered radical will further react with the persistent metal−cyano complex to furnish the C−CN bond and deliver the desired difunctionalized product.14 We reasoned that judicious choice of chiral ligands could render this C−CN bond formation enantioselective, and this hypothesis is supported in the seminal work by Liu et al. in a series of elegant chemical cyanation reactions.15
A major challenge of implementing this reaction design lies in the reconciliation of multiple oxidation and reduction processes in the same electrochemical system. In particular, the electrochemical compatibility and behavior of copper complexes with common ligand scaffolds that can be readily made chiral [e.g., bisoxazoline (BOX), bisphosphine] have not been systematically studied.16 In addition, for practical reasons, preparative-scale electrochemical reactions are ideally conducted in undivided vessels (e.g., common reaction flasks) using feedstock chemicals as the terminal oxidant or reductant (e.g., H+ as oxidant). This requirement creates a challenge, as Cu ions are substantially easier to reduce than H+ and would plate out on the cathode,17 losing catalytic activity. Therefore, the choice of solvent, proton source, and ligand is critical.
We set out to investigate the feasibility of our reaction design in the difunctionalization of 4-methoxystyrene (1) using TMSCN and diphenylphosphine oxide (2) in the presence of Cu(BOX)-type complexes (Table 1). Using conditions that we previously established for electrochemical alkene difunctionalization in combination with Cu(L4), no desired product 3 was observed (entry 1). The Pt cathode was visibly covered with metallic Cu postelectrolysis. A survey of polar solvents that are ideal for electrochemistry revealed that DMF is optimal, providing 3 in 20% yield with 79% ee (entry 2). However, the Cu catalyst was again irreversibly reduced on Pt and the alkene was fully consumed to form various side products. Replacing HOAc with trifluoroethanol (TFE) as the proton source provided a solution to the product selectivity issue (entry 3). Although the reaction resulted in marginally lower yield and ee, little side product was observed and the remaining styrene was recovered. Finally, screening different electrolytes led to the discovery of the optimal conditions using ligand L4, providing 3 in 51% yield with 84% ee after passing 2 F of charge (entry 4).18
Table 1.
Reaction Optimizationa
![]() | ||||||
|---|---|---|---|---|---|---|
| Entry | Ligand | Electrolyte | [H+] | Solvent | Yield (%) | Ee (%) |
| 1 | L4 | LiClO4 | HOAc | MeCN | <1 | ND |
| 2 | L4 | LiClO4 | HOAc | DMF | 20 | 79 |
| 3 | L4 | LiClO4 | TFE | DMF | 18 | 76 |
| 4 | L4 | TBABF4 | TFE | DMF | 51 | 84 |
| 5 | L5–L14 | TBABF4 | TFE | DMF | <10–63 | 12–79 |

Reaction conditions: 1 (0.2 mmol, 1 equiv), 2 (1.5 equiv), TMSCN (2.0 equiv), Cu(OTf)2 (3 mol %), ligand (6 mol %), electrolyte (2.0 equiv, 0.1 M), [H+] (2.5 equiv), solvent (4.0 mL), C felt anode, Pt cathode, undivided cell, constant current i = 3 mA, corresponding to anodic potential of 165–185 mV vs Fc0/+.
Based on geometrical surface area.
To further improve the reaction enantioselectivity, we surveyed a large number (>20) of chiral ligands (e.g., BOX, PyBOX, BiOX, PyrOX, and BINAP; see Table 1, entry 5 and the SI), including those that are frequently used in Cu-catalyzed enantioselective radical reactions (Scheme 2A).19 However, the optimal ee remained 84% (11:1 er). We reasoned substantial modification to the ligand scaffold would be necessary in order to break the selectivity ceiling that we encountered.
Scheme 2. Development of Serine-Derived BOX Ligands.
In a mechanistically related work by Stahl and Liu (Scheme 2A, eq 1),15d the C−CN bond formation via reductive elimination from a CuIII intermediate was postulated as the enantioselectivity determining step. In their computational model, this key CuIII intermediate adopts a pentacoordinated structure with two CN− bound to the metal in addition to the C-centered radical and BOX ligand. Analogous pentacoordianted CuIII complexes have been observed and proposed to undergo reductive elimination in several other reaction systems.20
Our cyanophosphinoylation reaction differs from the previous reports by Liu et al.15 primarily in three ways. First, the phosphine oxide group in the key intermediate leading to reductive elimination could compete with the ligand or CN− in binding to the Cu center. Second, DMF—a polar solvent that is used to ensure high solution conductivity—could also potentially replace a CN− ligand in the cyanation transition state.21 Finally, the transition state is considerably bulkier owing to the pendant phosphine oxide group.22 All these factors could result in a less organized transition structure and induce unselective reaction pathways.
The above analysis led us to propose that the introduction of an ancillary ligand (e.g., an ester group) to the BOX scaffold would further stabilize the putative pentacoordinated CuIII complex prior to reductive elimination (Scheme 2B). This modification would also increase the rigidity of the transition state and improve the stereochemical fidelity of the cyanation process. Moreover, we hypothesize that this multidentate ligand will further stabilize the Cu catalyst against cathodic demetalation and present a cationic intermediate (I) that is more susceptible to reductive elimination.23
Accordingly, we synthesized a series of bisoxazoline ligands with pendant ester groups derived from serine (sBOX). To the best of our knowledge, these ester-substituted BOX ligands have not been studied in asymmetric catalysis.24 These ligands were then tested in the model reaction under otherwise identical conditions, and the difunctionalization product (3) was isolated in good yield and substantially improved ee (Scheme 3A). In particular, sBOX(iPr) and sBOX(tBu) provided 94% ee (32:1 er) and 95% ee (39:1 er), respectively. We also tested several other substituted styrenes, and sBOX outperformed L4 across the board (Scheme 3B).
Scheme 3. Substrate Scopea.
aReaction conditions: alkene (0.2 mmol, 1 equiv), P source (1.5 equiv), TMSCN (2.0 equiv), Cu(OTf)2 (3 mol %), ligand (6 mol %), TBABF4 (2.0 equiv, 0.1 M), TFE (2.5 equiv), DMF (4.0 mL), C felt anode, Pt cathode, undivided cell, constant current i = 3 mA. bTotal charge = 2 F, reaction time 3.5 h. cTotal charge = 3−3.5 F, reaction time 5−6 h (see SI). d2.0 mmol scale. eEe after ether washes. fUsing p-toluenesulfinic acid or p-Cl-benzenesulfinic acid with pyridine (2 equiv).
Owing to the ease of synthesis,25 sBOX(iPr) was chosen instead of sBOX(tBu) to study the substrate scope. Styrenes with a variety of substituents proved to be suitable substrates. In particular, functionalities that are potentially sensitive to chemical oxidation, such as aldehyde (21) and sulfide (24), were preserved under the electrolysis conditions. These functional groups have not been shown to be compatible with previously reported cyanation conditions using chemical oxidants, such as Mn(OAc)310 (see SI) and N-fluorobenzene-sulfonimide.15 Benzyl chloride (27), which could undergo nucleophilic substitution by CN−, was also tolerated. Notably, a variety of electron-rich and electron-deficient N-heterocycles (28, 29, 30, 31) were converted to the corresponding products in high yield with excellent enantioselectivity. Substituted diaryl phosphine oxides took part in the difunctionalization smoothly, providing 32 and 33 with high ee. Dibutyl phosphine oxide also reacted (34) at elevated temperature (22 °C) to achieve synthetically useful yield and ee. Introduction of the phosphine oxide group significantly increases the crystallinity of the product, whose enantiomeric excess can be upgraded readily via ether washes with very little mass loss (e.g., 26, 28). The current reaction conditions are not applicable to substrates that are intrinsically challenging to Cu-catalyzed cyanofunctionalization reactions,10,15 such as dialkylphosphites (no desired product) and aliphatic alkenes (low enantioselectivity; see SI). Finally, we demonstrated the synthesis of 3 using ElectraSyn 2.0 instead of our custom-made electrochemical cell and obtained similar results (see SI).
We also extended this electrochemical protocol to the cyanosulfinylation of vinylarenes using sulfinic acids as nucleophiles to furnish product 35−38 in high efficiency and selectivity. To balance the pH of the medium and ensure the facile generation of sulfinyl radical, pyridine was added to the reaction.
The difunctionalized products contain two newly installed groups, granting access to various structurally diverse molecules upon further synthetic elaborations. For instance, thiourea-derived phosphine (39),26 Schiff-based-derived phosphine oxide (40),27 and aminophosphine (41)28 constitute bifunctional scaffolds that resemble catalysts currently used in asymmetric synthesis (Scheme 4).
Scheme 4. Product Derivatizationa.
aReaction conditions: (a) CoCl2 (cat.), NaBH4, MeOH, 0 °C, 83% yield; (b) CeCl3, LiAlH4, THF, 50 °C, 96% yield; (c) 3,5-Bistrifluoromethylphenyl isothiocyanate, DCM, 42% yield; (d) 3,5-Di-tert-butylsalicylaldehyde, EtOH, reflux, 55% yield. (e) NiCl2 (cat.), NaBH4, Boc2O, MeOH, 66% yield.
Finally, we propose an electrocatalytic cycle based on anodically coupled electrolysis (Figure 1A). Cyclic voltammetry (CV) data showed that the oxidation of CuI(sBOX)(OTf) catalyst to the corresponding CuII complex results in a feature at around 0.35 V (vs Fc+/0; Figure 1B, black line). Upon the addition of TMSCN, however, this redox event takes place at a less positive potential of 0.15 V with an enhanced anodic current (red line). This result shows that ligand exchange leads to a complex—presumably CuI(sBOX)(CN)—that is both thermo-dynamically and kinetically more favorable to undergo oxidation. Importantly, the addition of phosphine oxide 2 further augmented the anodic current at 0.15 V (blue line), and this current improvement is more pronounced with a higher concentration of 2 (Figure S3). The observation of such catalytic current is indicative of CuII(sBOX)(CN) being capable of oxidizing 2 in a catalytic fashion,29 likely resulting in a free P-centered radical while returning to the CuI state before anodic reoxidation. The direct oxidation of diphenylphosphine oxide on a carbon anode proved to be difficult (>0.3 V; see SI). Finally, we conducted controlled potential electrolysis with a constant anodic potential of 0.19 V and observed the formation of product 3 in 94% ee. These data together are consistent with an electrochemical difunctionalization reaction mediated by a [CuI/II(sBOX)(CN)]0/+ redox couple.
Figure 1.
(A) Proposed electrocatalytic cycle. (B) Cyclic voltammetry data.
Upon addition of the nascent P-centered radical to the alkene to form the C−P bond, the incipient carbon-centered radical undergoes single-electron oxidative addition to [CuII(sBOX)(CN)]+. This event was followed by reductive elimination to furnish the chiral product. The resultant CuI picks up another molecule of CN−—which becomes easier to oxidize according to CV data—before being turned over on the anode to CuII. The combination of appropriate reaction medium, proton donor, and chelating ligand allows the merger of Cu redox catalysis and electrochemistry without the necessity for an additional electron mediator.
Mechanism-informed ligand design enabled optimization of the enantioselectivity of the electrocatalytic difunctionalization reaction. In our aforementioned working hypothesis, an ester group in sBOX serves as an ancillary ligand to rigidify the key reduction elimination transition state toward high levels of enantioinduction. It is unlikely that these second-sphere groups provide merely steric repulsions to dictate the reaction enantioselectivity, as the ester groups are virtually planar, less bulky, and more flexible than the iPr group in ligand L4. However, it is also possible that the ester groups exert electrostatic stabilization of the major reaction transition state by functioning as a Lewis base. The mechanism of enantioinduction by Cu(sBOX) complexes is a current topic of investigation.
From the perspective of green chemistry, the removal of a conventional chemical oxidant constitutes another attractive feature of our protocol. We anticipate that our electrocatalytic strategy and new bisoxazoline ligands will pave the way for the discovery of other synthetically useful transformations.
Supplementary Material
ACKNOWLEDGMENTS
Financial support was provided by Cornell University and NIGMS (R01GM130928). This study made use of the NMR facility supported by the NSF (CHE-1531632). We thank IKA for the donation of ElectraSyn 2.0 and Dr. Scott McCann and Gregory Sauer for constructive discussion.
Footnotes
The authors declare no competing financial interest.
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/jacs.9b03296.
Procedures and characterization data (PDF)
REFERENCES
- (1).For representative reviews, see: Chemler SR; Fuller PH Heterocycle Synthesis by Copper Facilitated Addition of Heteroatoms to Alkenes, Alkynes and Arenes. Chem. Soc. Rev 2007, 36, 1153–1160.McDonald RI; Liu G; Stahl SS Palladium(II)-Catalyzed Alkene Functionalization via Nucleopalladation: Stereochemical Pathways and Enantioselective Catalytic Applications. Chem. Rev 2011, 111, 2981–3019.Courant T; Masson G Recent Progress in Visible-Light Photoredox-Catalyzed Intermolecular 1,2-Difunctionalization of Double Bonds via an ATRA-Type Mechanism. J. Org. Chem 2016, 81, 6945–6952.Romero RM; Wöste TH; Muñiz K Vicinal Difunctionalization of Alkenes with Iodine(III) Reagents and Catalysts. Chem. - Asian J 2014, 9, 972–983.Williamson KS; Michaelis DJ; Yoon TP Advances in the Chemistry of Oxaziridines. Chem. Rev 2014, 114, 8016–8036.Derosa J; Tran VT; van der Puyl VA; Engle KM Carbon−Carbon π-Bonds as Conjunctive Reagents in Cross-Coupling. Aldrichimica Acta 2018, 51, 21–32.
- (2).For representative recent examples, see: Liu Z; Li X; Zeng T; Engle KM Directed, Palladium(II)-Catalyzed Enantioselective anti-Carboboration of Alkenyl Carbonyl Compounds. ACS Catal. 2019, 9, 3260–3265.Wang H; Bai Z; Jiao T; Deng Z; Tong H; He G; Peng Q; Chen G Palladium-Catalyzed Amide-Directed Enantioselective Hydrocarbofunctionalization of Unactivated Alkenes Using a Chiral Monodentate Oxazoline Ligand. J. Am. Chem. Soc 2018, 140, 3542–3546.Anthony D; Lin Q; Baudet J; Diao T Nickel-Catalyzed Asymmetric Reductive Diarylation of Vinylarenes. Angew. Chem., Int. Ed 2019, 58, 3198–3202.Bovino MT; Chemler SR Catalytic Enantioselective Alkene Aminohalogenation/Cyclization Involving Atom Transfer. Angew. Chem., Int. Ed 2012, 51, 3923–3927.Fang L; Yan L; Haeffner F; Morken JP Carbohydrate-Catalyzed Enantioselective Alkene Diboration: Enhanced Reactivity of 1,2-Bonded Diboron Complexes. J. Am. Chem. Soc 2016, 138, 2508–2511.You W; Brown MK Catalytic Enantioselective Diarylation of Alkenes. J. Am. Chem. Soc 2015, 137, 14578–14581.Chen B; Fang C; Liu P; Ready JM Rhodium-Catalyzed Enantioselective Radical Addition of CX4 Reagents to Olefins. Angew. Chem., Int. Ed 2017, 56, 8780–8784.
- (3).For a perspective, see: Sauer GS; Lin S An Electrocatalytic Approach to the Radical Difunctionalization of Alkenes. ACS Catal. 2018, 8, 5175–5187. For examples, see:Siu JC; Parry JB; Lin S Aminoxyl-Catalyzed Electrochemical Diazidation of Alkenes Mediated by a Metastable Charge-Transfer Complex. J. Am. Chem. Soc 2019, 141, 2825–2831.Siu JC; Sauer GS; Saha A; Macey RL; Fu N; Chauvire T; Lancaster KL; Lin S Electrochemical Azidooxygenation of Alkenes Mediated by a TEMPO−N3 Charge-Transfer Complex. J. Am. Chem. Soc 2018, 140, 12511–12520.Fu N; Sauer GS; Lin S Electrocatalytic Radical Dichlorination of Alkenes with Nucleophilic Chlorine Sources. J. Am. Chem. Soc 2017, 139, 15548–15553.Fu N; Sauer GS; Saha A; Loo A; Lin S Metal-Catalyzed Electrochemical Diazidation of Alkenes. Science 2017, 357, 575–579.
- (4).For a review, see: Martins GM; Shirinfar B; Hardwick T; Ahmed N A Green Approach: Vicinal Oxidative Electrochemical Alkene Difunctionalization. ChemElectroChem 2019, 6, 1300–1315. For representative recent examples, see:Ashikari Y; Shimizu A; Nokami T; Yoshida J -i. Halogen and Chalcogen Cation Pools Stabilized by DMSO. Versatile Reagents for Alkene Difunctionalization. J. Am. Chem. Soc. 2013, 135, 16070–16073.Xu H-C; Campbell JM; Moeller KD Cyclization Reactions of Anode-Generated Amidyl Radicals. J. Org. Chem 2014, 79, 379–391.Chen J; Yan W-Q; Lam CM; Zeng C-C; Hu L-M; Little RR Electrocatalytic Aziridination of Alkenes Mediated by n-Bu4NI: A Radical Pathway. Org. Lett 2015, 17, 986–989.Yuan Y; Cao Y; Lin Y; Li Y; Huang Z; Lei A Electrochemical Oxidative Alkoxysulfonylation of Alkenes Using Sulfonyl Hydrazines and Alcohols with Hydrogen Evolution. ACS Catal. 2018, 8, 10871–10875.Xiong P; Long H; Song J; Wang Y; Li J-F; Xu H-C Electrochemically Enabled Carbohydroxylation of Alkenes with H2O and Organotrifluoroborates. J. Am. Chem. Soc 2018, 140, 16387–16391.Li J; Huang W; Chen J; He L; Cheng X; Li G Electrochemical Aziridination by Alkene Activation Using a Sulfamate as the Nitrogen Source. Angew. Chem., Int. Ed 2018, 57, 5695–5698.
- (5).(a) For representative recent reviews, see: Wiebe A; Gieshoff T; Mohle S; Rodrigo E; Zirbes M; Waldvogel SR Electrifying Organic Synthesis. Angew. Chem., Int. Ed 2018, 57, 5594–5619. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Tang S; Liu Y; Lei A Electrochemical Oxidative Cross-coupling with Hydrogen Evolution: A Green and Sustainable Way for Bond Formation. Chem. 2018, 4, 27–45. [Google Scholar]; (c) Yan M; Kawamata Y; Baran PS Synthetic Organic Electrochemical Methods Since 2000: On the Verge of a Renaissance. Chem. Rev 2017, 117, 13230–13319. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Feng R; Smith JA; Moeller KD Anodic Cyclization Reactions and the Mechanistic Strategies That Enable Optimization. Acc. Chem. Res 2017, 50, 2346–2352. [DOI] [PubMed] [Google Scholar]
- (6) (a).Torii S; Liu P; Bhuvaneswari N; Amatore C; Jutand A Chemical and Electrochemical Asymmetric Dihydroxylation of Olefins in I2−K2CO3−K2OsO2(OH)4 and I2−K3PO4/K2HPO4−K2OsO2(OH)4 Systems with Sharpless’ Ligand. J. Org. Chem 1996, 61, 3055–3060. [DOI] [PubMed] [Google Scholar]; (b) Nguyen BH; Redden A; Moeller KD Sunlight, electrochemistry, and sustainable oxidation reactions. Green Chem. 2014, 16, 69–72. [Google Scholar]; (c) Tanaka H; Kuroboshi M; Takeda H; Kanda H; Torii S Electrochemical asymmetric epoxidation of olefins by using an optically active Mn−salen complex. J. Electroanal. Chem 2001, 507, 75–81. [Google Scholar]
- (7) (a).Ye K-Y; Pombar G; Fu N; Sauer GS; Keresztes I; Lin S Anodically Coupled Electrolysis for the Heterodifunctionalization of Alkenes. J. Am. Chem. Soc 2018, 140, 2438–2441. [DOI] [PubMed] [Google Scholar]; (b) Ye K-Y; Song Z; Sauer GS; Harenberg JH; Fu N; Lin S Synthesis of Chlorotrifluoromethylated Pyrrolidines by Electrocatalytic Radical Ene-Yne Cyclization. Chem. - Eur. J 2018, 24, 12274–12279. [DOI] [PubMed] [Google Scholar]
- (8).Fu N; Shen Y; Allen AR; Song L; Ozaki A; Lin S Mn-Catalyzed Electrochemical Chloroalkylation of Alkenes. ACS Catal. 2019, 9, 746–754. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).For representative reviews on chiral phosphine/phosphine oxide in catalysis, see: Guo H; Fan YC; Sun Z; Wu Y; Kwon O Phosphine Organocatalysis. Chem. Rev 2018, 118, 10049–10293.Hayashi T Chiral Monodentate Phosphine Ligand MOP for Transition-Metal-Catalyzed Asymmetric Reactions. Acc. Chem. Res 2000, 33, 354–362.
- (10).For a racemic example of alkene cyanophosphinoylation, see: Zhang P-Z; Zhang L; Li J-A; Shoberu A; Zou J-P; Zhang W Phosphinoyl Radical Initiated Vicinal Cyanophosphinoylation of Alkenes. Org. Lett 2017, 19, 5537–5540. [DOI] [PubMed] [Google Scholar]
- (11).Secondary phosphine oxides are commercially available or readily prepared from diethyl phosphite: Hays HR Reaction of diethyl phosphonate with methyl and ethyl Grignard reagents. J. Org. Chem 1968, 33, 3690–3694. For examples of phosphinoyl radical addition to alkenes, see:Wei W; Ji J-X Catalytic and Direct Oxyphosphorylation of Alkenes with Dioxygen and H-Phosphonates Leading to β-Ketophosphonates. Angew. Chem., Int. Ed 2011, 50, 9097–9099.Zhang C; Li Z; Zhu L; Yu L; Wang Z; Li C Silver-Catalyzed Radical Phosphonofluorination of Unactivated Alkenes. J. Am. Chem. Soc 2013, 135, 14082–14085.Li J-A; Zhang P-Z; Liu K; Shoberu A; Zou J-P; Zhang W Phosphinoyl Radical-Initiated α,β-Aminophosphinoylation of Alkenes. Org. Lett 2017, 19, 4704–4706.Yang B; Hou S-M; Ding S-Y; Zhao X-N; Gao Y; Wang X; Yang S-D Cerium(IV)-Promoted Phosphinoylation-Nitratation of Alkenes. Adv. Synth. Catal 2018, 360, 4470–4474.
- (12).Li Q-Y; Swaroop TR; Hou C; Wang Z-Q; Pan Y-M; Tang H-T Electrochemical Dehydrogenative Coupling of Alcohols with Hydrogen Phosphoryl Compounds: A Green Protocol for P−O Bond Formation. Adv. Synth. Catal 2019, 361, 1761–1765. [Google Scholar]
- (13).Wang F; Chen P; Liu G Copper-Catalyzed Radical Relay for Asymmetric Radical Transformations. Acc. Chem. Res 2018, 51, 2036–2046. [DOI] [PubMed] [Google Scholar]
- (14).The regioselectivity of the addition steps are governed by the persistent radical effect: Studer, A. The Persistent Radical Effect in Organic Synthesis. Chem. - Eur. J 2001, 7, 1159–1164. Also see ref 7a. [DOI] [PubMed] [Google Scholar]
- (15).For examples, see: Wang D; Zhu N; Chen P; Lin Z; Liu G Enantioselective Decarboxylative Cyanation Employing Cooperative Photoredox Catalysis and Copper Catalysis. J. Am. Chem. Soc 2017, 139, 15632–15635.Wang D; Wang F; Chen P; Lin Z; Liu G Enantioselective Copper−Catalyzed Intermolecular Amino- and Azidocyanation of Alkenes in a Radical Process. Angew. Chem., Int. Ed 2017, 56, 2054–2058.Wang F; Wang D; Wan X; Wu L; Chen P; Liu G Enantioselective Copper-Catalyzed Intermolecular Cyanotrifluoromethylation of Alkenes via Radical Process. J. Am. Chem. Soc 2016, 138, 15547–15550.Zhang W; Wang F; McCann SD; Wang D; Chen P; Stahl SS; Liu G Science 2016, 353, 1014–1018.Yang S; Wang L; Zhang H; Liu C; Zhang L; Wang X; Zhang G; Li Y; Zhang Q Copper-Catalyzed Asymmetric Aminocyanation of Arylcyclopropanes for Synthesis of γ-Amino Nitriles. ACS Catal. 2019, 9, 716–721.
- (16).For examples of Cu-catalyzed electrochemical reactions under ligand-free conditions, see: Yang Q-L; Wang X-Y; Lu J-Y; Zhang L-P; Fang P; Mei T-S Copper-Catalyzed Electrochemical C−H Amination of Arenes with Secondary Amines. J. Am. Chem. Soc 2018, 140, 11487–11494.Kathiravan S; Suriyanarayanan S; Nicholls IA Electrooxidative Amination of sp2 C−H Bonds: Coupling of Amines with Aryl Amides via Copper Catalysis. Org. Lett 2019, 21, 1968–1972.Yi X; Hu X Formal Aza−Wacker Cyclization by Tandem Electrochemical Oxidation and Copper Catalysis. Angew. Chem., Int. Ed 2019, 58, 4700–4704.
- (17).Merchant RR; Oberg KM; Lin Y; Novak AJE; Felding J; Baran PS Divergent Synthesis of Pyrone Diterpenes via Radical Cross Coupling. J. Am. Chem. Soc 2018, 140, 7462–7465. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (18).The electrolyte may affect the catalyst structure and the electrokinetics. At this point, we do not have an explanation for this ee-dependence on the electrolyte.
- (19).For examples, see: Zhu R; Buchwald SL Versatile Enantioselective Synthesis of Functionalized Lactones via Copper-Catalyzed Radical Oxyfunctionalization of Alkenes. J. Am. Chem. Soc 2015, 137, 8069–8077.Zhu R; Buchwald SL Enantioselective Functionalization of Radical Intermediates in Redox Catalysis: Copper-Catalyzed Asymmetric Oxytrifluoromethylation of Alkenes. Angew. Chem., Int. Ed 2013, 52, 12655–12658.Lin J-S; Dong X-Y; Li T-T; Jiang N-C; Tan B; Liu X-Y A Dual-Catalytic Strategy to Direct Asymmetric Radical Aminotrifluoromethylation of Alkenes. J. Am. Chem. Soc 2016, 138, 9357–9360.
- (20).For reports that propose pentacoordianted Cu intermediates, see: Addison AW; Burke PJ; Henrick K; Rao TN; Sinn E Pentacoordinate copper complexes of nitrogen-sulfur donors: structural chemistry of two complexes of bis(2-(2-benzimidazolyl)ethyl) sulfide with the sulfur alternatively in equatorial and axial coordination modes. Inorg. Chem 1983, 22, 3645–3653.Casitas A; King AE; Parella T; Costas M; Stahl SS; Ribas X Direct observation of CuI/CuIII redox steps relevant to Ullmann-type coupling reactions. Chem. Sci 2010, 1, 326–330.Zhang S-L; Liu L; Fu Y; Guo Q-X Theoretical Study on Copper(I)-Catalyzed Cross-Coupling between Aryl Halides and Amides. Organometallics 2007, 26, 4546–4554.Fier PS; Luo J; Hartwig JF Copper-Mediated Fluorination of Arylboronate Esters. Identification of a Copper(III) Fluoride Complex. J. Am. Chem. Soc 2013, 135, 2552–2559.
- (21).DMF has been shown to provide lower ee in Cu radical catalysis with BOX-type ligands compared with nonpolar solvents. See: Sha W; Deng L; Ni S; Mei H; Han J; Pan Y Merging Photoredox and Copper Catalysis: Enantioselective Radical Cyanoalkylation of Styrenes. ACS Catal. 2018, 8, 7489–7494. Also see ref 15a. [Google Scholar]
- (22).Indeed, large tert-leucine-derived BOX ligands consistently perform poorly in comparison to valine derivatives (e.g., 84% ee with 4 vs 50% ee with 5, 66% ee with 11 vs 12% ee with 12). See SI.
- (23).It has been shown that reductive elimination is favored from cationic organometallic complexes involving metals other than Cu. For examples, see: Furuya T; Benitez D; Tkatchouk E; Strom AE; Tang P; Goddard WA III; Ritter T Mechanism of C−F Reductive Elimination from Palladium (IV) Fluorides. J. Am. Chem. Soc 2010, 132, 3793–3807.Procelewska J; Zahl A; Liehr G; van Eldik R; Smythe NA; Williams BS; Goldberg KI Mechanistic Information on the Reductive Elimination from Cationic Trimethylplatinum(IV) Complexes to Form Carbon−Carbon Bonds. Inorg. Chem 2005, 44, 7732–7742.
- (24).Although serine-derived BOX ligands have not been used in enantioselective catalysis, they have been used as intermediates for the synthesis of alcohol-substituted BOX ligands. For an example, see: Perrotta D; Wang M-M; Waser J Lewis Acid Catalyzed Enantioselective Desymmetrization of Donor−Acceptor meso-Diaminocyclopropanes. Angew. Chem., Int. Ed 2018, 57, 5120–5123. [DOI] [PubMed] [Google Scholar]
- (25).The synthesis of sBOX(iPr) requires three steps from commercially available reactants whereas that of sBOX(tBu) requires five steps. See SI.
- (26).For use of chiral thiourea-derived phosphines, see: Fang Y-Q; Jacobsen EN Cooperative, Highly Enantioselective Phosphinothiourea Catalysis of Imine−Allene [3 + 2] Cycloadditions. J. Am. Chem. Soc 2008, 130, 5660–5661.Fang Y-Q; Tadross PM; Jacobsen EN Highly Enantioselective, Intermolecular Hydroamination of Allenyl Esters Catalyzed by Bifunctional Phosphinothioureas. J. Am. Chem. Soc 2014, 136, 17966–17968.Zhong F; Han X; Wang Y; Lu Y Highly Enantioselective [3 + 2] Annulation of Morita−Baylis−Hillman Adducts Mediated by L-Threonine-Derived Phosphines: Synthesis of 3-Spirocyclopentene-2-oxindoles Having Two Contiguous Quaternary Centers. Angew. Chem., Int. Ed 2011, 50, 7837–7841.
- (27).For use of chiral Schiff base-derived phosphines and phosphine oxides, see: Dai H; Hu X; Chen H; Bai C; Zheng Z New efficient P,N,O-tridentate ligands for Ru-catalyzed asymmetric transfer hydrogenation. Tetrahedron: Asymmetry 2003, 14, 1467–1472.Kim T-J; Lee H-Y; Ryu E-S; Park D-K; Cho CS; Shim S-C; Jeong JH Asymmetric addition of diethylzinc to aromatic aldehydes by chiral ferrocene-based catalysts. J. Organomet. Chem 2002, 649, 258–267.Ye F; Zheng Z-J; Deng W-H; Zheng LS; Deng Y; Xia C-G; Xu L-W Modulation of Multifunctional N,O,P Ligands for Enantioselective Copper−Catalyzed Conjugate Addition of Diethylzinc and Trapping of the Zinc Enolate. Chem. -Asian J 2013, 8, 2242–2253.Aleksanyan DV; Nelyubina YV; Dmitrienko AO; Bushmarinov IS; Klemenkova ZS; Kozlov VA Polyhedron 2015, 85, 295–301.
- (28).For a review, see: Li W; Zhang J Recent developments in the synthesis and utilization of chiral b-aminophosphine derivatives as catalysts or ligands. Chem. Soc. Rev 2016, 45, 1657–1677. This precursor could also be readily elaborated to peptide-derived phophines for asymmetric catalysis:Wang T; Han X; Zhong F; Yao W; Lu Y Amino Acid-Derived Bifunctional Phosphines for Enantioselective Transformations. Acc. Chem. Res 2016, 49, 1369–1378.
- (29).Yi H; Yang D; Luo Y; Pao C-W; Lee J-F; Lei A Direct Observation of Reduction of Cu(II) to Cu(I) by P−H Compounds using XAS and EPR Spectroscopy. Organometallics 2016, 35, 1426–1429. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.






