Abstract
Metal-catalyzed cross-coupling reactions are extensively employed in both academia and industry for the synthesis of biaryl derivatives for applications to both medicine and material science. Application of these methods to prepare tetra-ortho-substituted biaryls leads to chiral atropisomeric products that introduces the opportunity to use catalyst-control to develop asymmetric cross-coupling procedures to access these important compounds. Asymmetric Pd-catalyzed Suzuki-Miyaura and Negishi cross-coupling reactions to form tetra-ortho-substituted biaryls were studied employing a collection of P-chiral dihydrobenzooxaphosphole (BOP) and dihydrobenzoazaphosphole (BAP) ligands. Enantioselectivities of up to 95:5 and 85:15 er were identified for the Suzuki-Miyaura and Negishi cross-coupling reactions, respectively. Unique ligands for the Suzuki-Miyaura reaction vs the Negishi reaction were identified. A computational study on these Suzuki-Miyaura and Negishi cross-coupling reactions enabled an understanding in the differences between the enantiodiscriminating events between these two cross-coupling reactions. These results support that enantioselectivity in the Negishi reaction results from the reductive elimination step, whereas all steps in the Suzuki-Miyaura catalytic cycle contribute to the overall enantioselection with transmetalation and reductive elimination providing the most contribution to the observed selectivities.
Keywords: cross-coupling, Palladium, catalysis, asymmetric, phosphines
Graphical Abstract

INTRODUCTION
Transition metal catalyzed sp2-sp2 cross-coupling is one of the most powerful and extensively utilized reactions in organic synthesis and materials science to access biaryl compounds.1 Of these methods, the Pd-catalyzed Suzuki-Miyaura2 and Negishi3 cross-coupling have received considerable development over the past 40 years.1,4 Significant advances in the development of highly active Pd-catalysts5 to achieve efficient processes have even allowed for these reactions to be employed on industrial scale.6 In general, bulky electron-rich phosphines7 and N-heterocyclic carbenes8,9 have been the ligands of choice to enable these cross-coupling reactions. Despite these advances, cross-coupling catalysts for the formation of sterically demanding biaryls, such as tetra-ortho-substituted systems, remains limited.9 Additionally, because appropriately substituted tri- and tetra-ortho-substituted biaryls exist as chiral atropisomers about the biaryl axis, methods to forge this bond in an enantioselective fashion are important since many biologically active natural products10 and potential pharmaceuticals11 contain single atropisomeric biaryl groups within their structure (Scheme 1). Therefore, one desirable strategy to access the biaryl functionality in these compounds is to utilize a catalyst-controlled asymmetric Suzuki-Miyaura or Negishi sp2–sp2 cross-coupling.12 While transition metal catalyzed asymmetric Suzuki-Miyaura13 and Negishi14 sp2–sp2 cross-coupling reactions have been developed, significantly fewer methods are available for the asymmetric construction of tetra-ortho-substituted biaryls.13c,d,p–t,14,15
Scheme 1:

Natural Products and Bioactive Atropisomeric Compounds
The first highly selective sp2–sp2 cross-coupling reaction was reported in 1988 by Hayashi and coworkers15 to afford atopisomeric binaphthalenes in up to 95% ee using a Kumada-Corriu16 coupling. It was not until the year 2000 that the first reports of Pd-catalyzed asymmetric Suzuki-Miyaura cross-coupling reactions were disclosed by the groups of Cammidge13c,d and Buchwald.13a,b Cammidge applied Hayashi’s15 chiral ferrocene ligand (PFNMe) to access tri-ortho-substituted biaryl 8a with modest enantioselection (82:18 er, Scheme 2). At the same time, Buchwald discovered that KenPhos could be employed as the chiral ligand affording high enantioselectivities in the cross-coupling if the R-substituent of the aryl halide coupling partner contained a Lewis basic donating group at the ortho-position (8b-d). Mechanistic studies on this class of cross-coupling by Buchwald13b and others11b,13n,x typically consider reductive elimination as the enantiodetermining event in the catalytic cycle and have shown that these Lewis basic ortho-substituents aid in the stabilization of the dominating transition state. However, restricted rotation of the intermediate Pd-complexes as a function of the ortho-substitution pattern of the aryl halide coupling partner can lead to scenarios where oxidative addition13f and transmetalation13d,f may be involved in enantioselection.17 After these initial reports, intense research into the area of catalyst-controlled asymmetric biaryl synthesis by a variety of groups has led to the introduction of multiple chiral ligands that affect this transformation (Scheme 2).11,13 The majority of these studies have focused on the investigation of the formation of tri-ortho-substituted binapthalene derivatives.11,13a,b,e-o,u-bb Of these, asymmetric synthesis of the aryl phosphonate products (8b–8d) by asymmetric cross-coupling have been one of the more common systems studied between different chiral catalysts with variable results.13a,b,j,n,u–y For example, our group11,13n,o introduced the P-chiral dihydrobenzooxaphosphole (BOP) class of ligands (BIDIME, 9, 10) that are effective in asymmetric cross-coupling reactions and are highly tunable to enable ligand modification for enhancement of stereoselectivities. Suzuki-Miyaura cross-coupling to afford 8b employing BOP ligand 9b13n gave improved enantioselectivity to that of KenPhos;13a however, in the coupling to produce 8b, KenPhos resulted in improved selectivity over ligand 9b. These results highlight the fact that asymmetric cross-coupling reactions are typically very substrate dependent. Synthesis of 8b–8d has also been acheived by asymmetric cross-coupling employing the atropisomeric, hemi-labile ligands Cy-MOP,13w 11,13u,x and 12,13v and using a helically-chiral polymeric phosphine ligand (PQXPhos).13i Other ligand classes including chiral dienes (13),13h bis(hydrazones) (14),13f and a resin-supported chiral ligand (15)13p have all been shown to afford high enantioselectivities in certain asymmetric Suzuki-Miyaura cross-coupling reactions. Despite these significant advances in asymmetric cross-coupling methodologies, a general catalyst for a wide-array of substitution patterns is still and unsolved problem. However, these technologies have found utility in natural product13k,o and API synthesis (e.g. 5 via 8).11
Scheme 2:

Asymmetric Suzuki-Miyaura Cross-Coupling Reactions to Prepare Tri-ortho-Substituted Biaryls
While asymmetric Suzuki-Miyaura cross-coupling to generate tri-ortho-substituted biaryls (Scheme 2) has been successful with a variety of chiral ligand types,13a,b,e-o,u-bb the synthesis of tetra-ortho-substituted biaryls by the same means has proven more difficult.13c,d,p–t Tetra-ortho-substituted biaryl product 16 (Scheme 3)13c,d,p–q has typically been the standard cross-coupling reaction analyzed when developing new catalysts for this challenging coupling.12c While the Ni-catalyzed Kumada-Corriu16 coupling to prepare 16 was reported by Hayashi15 with good enantioselectivity, the analogous Suzuki-Miyaura coupling has proven more challenging.13c,d,p–t Cammidge’s initial report13c,d demonstrated that the PFNMe ligand gave good enantioselectivity (93:7 er) in this reaction; however, long reaction time (6 days) was required affording only modest yield (Scheme 3). Espinet13r was able to optimize this result to reduce the reaction time to 3 days at 5 mol % Pd loading by using Pd2(dba)3•CHCl3 as the Pd-source in the presence of excess ligand (20 mol %) to improve the yield to 85%. The catalyst loading could be reduced to employ 3 mol % Pd and 12 mol % ligand by utilizing Pd(MeCN)4(BF4)2 in place of Pd2(dba)3•CHCl3 as the precatalyst with an improvement in enantioselectivity of up to 95:5 er. However, only 55% yield was obtained after 4 day reaction time. While these conditions can be applied to the synthesis of 16, reaction times are still quite long (days), and the scope of the reaction to other substrates is limited or has not been further investigated. Of the catalytic systems reported for the synthesis of 16 by Suzuki-Miyaura cross-coupling, Uozumi’s13p conditions employing resin-supported phosphine ligand 15 are performed in water under heterogeneous conditions that allow for recycling of the catalyst and show the most generality in terms of reaction scope and enantioselectivity; however, only simple substituted naphthalenes have been disclosed in the coupling. Recently there has been increased focus on the development of Suzuki-Miyaura cross-coupling reactions performed using water18,19 as the solvent due to the potential “greenness” of replacing typical organic solvents by water.20,21 However, none of these methods have been extended to enantioselective cross-coupling. Uozumi’s13p report represents the only known catalytic asymmetric cross-coupling carried out in neat water.22 Future studies are needed in this area to determine if improvements in catalytic asymmetric coupling reactions can be achieved using water as solvent.
Scheme 3.

Asymmetric Suzuki-Miyaura Cross-Coupling Reactions to Prepare Tetra-ortho-Substituted Biaryls
As noted above (Scheme 2), previous work in our laboratories identified the P-chiral dihydrobenzooxaphosphole (BOP) ligand class (BIDIME, 9, 10) as a highly tunable scaffold for applications in many different types of asymmetric transformations by modulating the substituents of ligand structure 23 (Scheme 4). By variation of the oxaphosphole C-2 R-substituent and the bis(alkoxy)phenyl aryl-group of ligand 23, catalysts for Suzuki-Miyaura7p–t,9a,19a,23 and asymmetric Suzuki-Miyaura11,13n–o cross-coupling, Buchwald-Hartwig amination,24 borylation,25 asymmetric propargylation of aldehydes26 and imines,27 asymmetric conjugate addition,28 asymmetric addition of boronic acids to imines,29 asymmetric allenylation of imines,30 asymmetric hydrogenation of polar31 and unfunctionalized32 alkenes, asymmetric hydrogenation of heterocycles,33 asymmetric hydroformylation,34 and asymmetric hydrogenation of ketones35 have all been developed. Due to the lack of a general solution for the asymmetric synthesis of tetra-ortho-substituted biaryls using the Suzuki-Miyaura cross-coupling reaction, we sought to investigate application of ligand family 23 as a potential solution to this problem because of the modular nature of 23 and because of its good activity in the asymmetric cross-coupling to furnish tri-ortho-substituted biaryls.11,13n,o In this article, we disclose our studies towards the application of the dihydrobenzooxaphosphole (23, X = O, BOP) and dihydrobenzoazaphosphole (23, X = NR, BAP) ligand scaffold for both the asymmetric Pd-catalyzed Suzuki-Miyaura and Negishi cross-coupling reaction to form tetra-ortho-substituted biaryls (Scheme 4). The differences in the enantiodiscrimination mechanism between these two processes was probed computationally and is also disclosed herein.
Scheme 4.

Tunable P-Chiral Phosphine Ligands for Asymmetric Pd-Catalyzed Cross-Coupling.
RESULTS AND DISCUSSION
Ligand Synthesis.
The ligands employed in the current study were prepared by the general strategy outlined in Scheme 5. Starting from the known P-chiral triflates (26, 27),13n,35b Negishi7t or Suzuki-Miyaura cross-coupling was used to install the lower aryl ring of the ligand biaryl axis affording oxides 29 or 30. Deprotonation of 29 or 30 with LDA followed by electrophilic trapping allowed for substitution of the C-2 position of the oxaphosphole or azaphosphole ring providing 31 or 32, respectively. Oxide reduction then provided the final BOP (X = O) or BAP (X = NR) ligands (33). Additionally, the 3,5-positions of the lower aryl ring of the biaryl group of 34 can be functionalized by bromination11b to afford 35 followed by cross-coupling and alkylation to afford ligand family 36. These procedures have enabled the synthesis of a tunable family of P-chiral phosphine ligands. The ligands prepared and used in this study are shown in Chart 1.
Scheme 5.

Ligand Synthesis.
Chart 1.

Ligands Used in This Study.
Asymmetric Suzuki-Miyaura cross-coupling.
To investigate the Suzuki-Miyaura cross-coupling to prepare tetra-ortho-substituted biaryls, we initially examined the “benchmark12c” cross-coupling reaction to form tetra-ortho-substituted biaryl compound 18 (Table 1) employing 1-bromo-2-methylnaphthalene (17b) and (2-methylnaphthalen-1-yl)boronic acid (16a) as the coupling partners. Initial experiments probed conditions that previously were shown to be suitable for hindered cross-coupling reactions (Entries 1–3). These consisted of the use of K3PO4 as base and employing BIDIME (L1), AntPhos, or C2-i-Pr-BIDIME (L4) as the ligand. Gratifyingly, under these conditions, product 18 was observed. While almost no enantioinduction was observed using BIDIME (L1) or AntPhos, which lack substitution at the C-2 position of the oxaphosphole ring (entries 1 and 2), C2-i-Pr-BIDIME (L4) afforded modest enantioselectivity and yield in the reaction (entry 3). As has been previously noted for reactions forming tetra-ortho-substituted biaryls,13d,r protodeboronation was a significant side-reaction that prevented the coupling from proceeding to full conversion. To circumvent this issue, alternate bases were examined in an effort to reduce protodeboronation; however, no improvements over the use of K3PO4 were identified (entries 3–8). Additionally, solvents such as toluene and 1,4-dioxane were found to be equally effective in the coupling and did not improve enantioselectivity or yield (entries 9 and 10). Finally, to achieve full reaction conversion, the effect of catalyst loading was examined (entries 11–14). Use of 5 mol % Pd was needed to achieve full conversion and allowed for high yield in the desired reaction (entry 13). Addition of water as a co-solvent to these reactions when employing THF or n-BuOH as solvent afforded < 10% yield of product due to significant amounts of protodeboronation (data not shown) that is not uncommon considering that the addition of excess water is often avoided in sterically demanding and asymmetric cross-coupling reactions.9,13
Table 1.
Asymmetric Suzuki-Miyaura Cross-Coupling Towards Tetra-ortho-Substituted Biaryls: Base and Solvent Survey.

| Entry | Mol % Pd | Ligand (mol%) | Base (3 equiv) | Solvent | %Yielda | erb |
|---|---|---|---|---|---|---|
| 1 | 1 | L1 (2) | K3PO4 | THF | 40 | 55:44 |
| 2 | 1 | (R)-AntPhos (2) | K3PO4 | THF | 30 | 50:50 |
| 3 | 1 | L4 (2) | K3PO4 | THF | 50 | 86:14 |
| 4 | 1 | L4 (2) | CsF | THF | 18 | ND |
| 5 | 1 | L4 (2) | KF | THF | < 5 | ND |
| 6 | 1 | L4 (2) | K2CO3 | THF | 9 | ND |
| 7 | 1 | L4 (2) | Na2CO3 | THF | < 2 | ND |
| 8 | 1 | L4 (2) | NaOtBu | THF | < 3 | ND |
| 9 | 1 | L4 (2) | K3PO4 | Toluene | 48 | 85:15 |
| 10 | 1 | L4 (2) | K3PO4 | 1,4-dioxane | 50 | 85.5: 14.5 |
| 11 | 3 | L4 (6) | K3PO4 | THF | 64 | 86:14 |
| 12 | 3 | L4 (6) | K3PO4 | Toluene | 56 | 85:15 |
| 13 | 5 | L4 (10) | K3PO4 | THF | 90 | 86:14 |
| 14 | 5 | L4 (10) | K3PO4 | Toluene | 85 | 85:15 |
Isolated yield.
Determined by chiral HPLC analysis with Chiralpak OD-H column.
With moderate conditions in hand for the Suzuki-Miyaura cross-coupling reaction (Table 1, entry 13), we next attempted to modify the ligand structure to improve enantioselection (Table 2). The substitution at the C-2 position of the oxaphosphole ring of the ligand was first examined (entries 1–6). This enabled the identification of a ligand that provided enantioselectivities of up to 95:5 er when a dimethylmethoxy (DMM) substituent was at the C-2 position of the oxaphosphole ring (L7, NitinPhos, entry 4). Furthermore, it was critical that the ligand contained a dimethoxyphenyl ring as part of the biaryl axis of the ligand. When this group was replaced with phenyl, anthracene, or methoxy, inferior yields and selectivities were obtained (entry 4 vs entries 9–11). The importance of the lower dimethoxyphenyl ring in enhancing enantioselectivities through an interaction between one methoxy-group and Pd has been postulated previously,11b,13n and is further supported by DFT modeling of this system (vide infra). Replacement of the methoxy-groups on this lower aryl ring with iso-propoxy groups afforded no improvement (entry 1 vs entries 7–8).
Table 2.
Ligand Survey in the Asymmetric Suzuki-Miyaura Cross-Coupling Towards Tetra-ortho-Substituted Biaryls.

| Entry | Ligand | % yielda | erb |
|---|---|---|---|
| 1 | L4 | 92 | 86:14 |
| 2 | L5 | 80 | 90:10 |
| 3 | L6 | 75 | 86:14 |
| 4 | L7 | 97 | 94.5:5.5 |
| 5 | L8 | 45 | 81.5:18.5 |
| 6 | L10 | 91 | 83:17 |
| 7 | L13 | 73 | 84.5:15.5 |
| 8 | L14 | 75 | 83:17 |
| 9 | L16 | 75 | 73:27 |
| 10 | L17 | 35 | 88:12 |
| 11 | L18 | 50 | 64:36 |
| 12 | L20 | 89 | 78:22 |
Isolated yield.
Determined by chiral HPLC analysis with Chiralpak OD-H column.
The effect of Pd-precatalyst and Pd:ligand loading was next examined (Table 3). The use of excess ligand enabled the highest reaction yield (entry 1 vs 2). Additionally, Pd(OAc)2 gave less side-product formation than compared to the use of Pd2(dba)3 or [PdCl(allyl)]2 (entry 2 vs entries 3–4).
Table 3.
Effect of Pd-Precatalyst and Pd:Ligand Loading.

| Entry | Pd Source (mol %) | Mol % L7 | % yielda | Erb |
|---|---|---|---|---|
| 1 | Pd(OAc)2 (5) | 10 | 97% | 94.5:5.5 |
| 2 | Pd(OAc)2 (5) | 5 | 75%c | 94.5:5.5 |
| 3 | Pd2(dba)3 (2.5) | 5 | 64%d | 94:6 |
| 4 | [PdCl(allyl)]2 (5) | 5 | 55%e | 93.7:6.3 |
Isolated yield.
Determined by chiral HPLC analysis with Chiralpak OD-H column.
1–2% side products were observed with 23 % of bromide substrate remaining,
5–7% side products were observed with 29% of bromide substrate,
8–9% side products were observed with 35% of bromide substrate.
In an effort to further reduce the loading of the boron coupling partner, we next examined other boron derivatives in the coupling reaction in an effort to reduce protodeboronation side-reactions (Scheme 6). Use of the glycol ester (16b) that was optimal in Cammidge’s system13d and has been recently shown to give improved transmetalation rates in Suzuki-Miyaura cross-coupling reactions30 gave similar enantioselectivities as that observed with boronic acid 16a but did not afford full conversion in the reaction, even at prolonged reaction times. Use of the pinacol ester 16c allowed for reduction of the boron coupling partner to 2 equiv; however, long reaction times were required and a slight loss in enantiopurity was obtained. The change in enantioselectivity based off of the nature of the boronate employed has been observed previously13d and hints that the transmetalation step in the catalytic cycle for this reaction may be involved in the enantiodeterming event (vide infra). Finally, use of the potassium trifluoroborate salt (16d), which are known for their resistance to protodeboronation,37 was not effective in the cross-coupling reaction.
Scheme 6.

Effect of Boronate Coupling Partner.
With identification of a reactive catalytic system to form tetra-ortho-substituted biaryl product 18 with good enantioselectivities, the scope of this catalytic system was then briefly examined (Scheme 7). Unfortunately, it was quickly determined that the present system did not show good generality, as has been a common problem in asymmetric Pd-catalyzed Suzuki-Miyaura reactions providing tetra-ortho-substituted biaryls. Only subtle modification to the coupling partners led to decreased reaction efficiency due to increased amounts of protodeboronation.
Scheme 7.

Asymmetric Suzuki-Miyaura Cross-Coupling Reaction Scope.
Asymmetric Negishi cross-coupling.
Because of the issues identified with the Suzuki-Miyaura cross-coupling reaction (poor generality, high boronate loading), we decided to investigate the analogous reaction employing Negishi cross-coupling with aryl-zinc reagent 40 prepared by lithium-halogen exchange of 17b with t-BuLi followed by addition of a THF solution of anhydrous ZnBr2. ZnBr2 was our preferred choice over ZnCl2 due to its higher solubility in THF. Towards this end, we surveyed our family of P-chiral oxaphosphole and azaphosphole ligands in the Negishi cross-coupling reaction (Table 4). The highest yield and selectivity was obtained with L13 having iso-propoxy groups on the lower aryl-ring of the ligand biaryl axis (entry 10). Surprisingly, the optimal ligand for the asymmetric Suzuki-Miyaura reaction (NitinPhos, L7, entry 5) gave poor yield and selectivity in the Negishi coupling considering that the presumed enantiodetermining step of reductive elimination should be identical for the Negishi and Suzuki-Miyaura reactions. This difference in selectivity was something we wanted to understand further (vide infra). In general, C2-substituion on the oxaphosphole or azaphosphole ring of the ligand was required for good selectivity (entry 1 vs entries 2–10, and entries 15–16 vs entries 17–21). Further increasing the steric bulk at this C2-position by introducing a second methyl group enabled similar selectivities but led to a reduction in yield (entries 8 & 12). Additionally, a trend was observed where increasing the electron-donating ability of the lower aryl-ring bearing the alkoxy-substituents led to improved enantioselection (compare entries 2, 9, and 10). Lastly, the azaphosphole series of ligands were extremely efficient in terms of reaction yield; however, enantioselectivities were lower with this series (entries 15, 17, 18, 19).
Table 4.
Ligand Survey in the Asymmetric Negishi Cross-Coupling Towards Tetra-ortho-Substituted Biaryls.

| Entry | Ligand | % yielda | erb |
|---|---|---|---|
| 1 | L1 | 69 | 52.5:47.5 |
| 2 | L2 | 49 | 78.5:21.5 |
| 3 | L3 | ND | 81.5:18.5 |
| 4 | L4 | 38 | 82.5:17.5 |
| 5 | L7 | 10 | 80.5:19.5 |
| 6 | L9 | 82 | 76.5:23.5 |
| 7 | L10 | 89 | 83.0:17.0 |
| 8 | L11 | 69 | 85.5:14.5 |
| 9 | L12 | 81 | 71.3:28.7 |
| 10 | L13 | 93 | 85.0:15.0 |
| 11 | L14 | 45 | 82.0:18.0 |
| 12 | L15 | 24 | 83.0:17.0 |
| 13 | L21 | 67 | 77.0:23.0 |
| 14 | L22 | 36 | 65.0:35.0 |
| 15 | L23 | 91 | 60.5:39.5 |
| 16 | L24 | 68 | 61.0:39.0 |
| 17 | L25 | 96 | 82.5:17.5 |
| 18 | L26 | 89 | 79.0:21.0 |
| 19 | L27 | 99 | 53.0:47.0 |
| 20 | L28 | 76 | 60.0:40.0 |
| 21 | L29 | 82 | 78.0:22.0 |
Isolated yield.
Determined by chiral HPLC analysis with Chiralpak OD-H column.
With identification of L13 as the best ligand out of the ligand survey in Table 4, we further optimized parameters for the Negishi coupling reaction with ligand L13 in an effort to improve enantioselection (Table 5). The Pd-precatalyst, metal-ligand ratio, Zn-source, and solvent were examined in the reaction (Table 5). The choice of Pd-precatalyst had little impact on the reaction enantioselectivity, but the yield varied from 70–80 % (entries 1, 4, 5, 7). The metal-ligand ratio also had no impact on selectivity; however, the yield improved if a 1:1.5 Pd:L13 ratio was employed (entries 8–10). Finally, use of other Zn-salt precursors or other solvents did not lead to improvements in reaction efficiency (entries 2, 3, 6, 7).
Table 5.
Survey of Pd-precatalyst, Pd:L13 ratio, Zn-source, and solvent in the Negishi Cross-Coupling Reaction.

| Entry | Pd | Metal:L13 ratio | ZnX2 | Solvent | Yield %a | Erb |
|---|---|---|---|---|---|---|
| 1 | Pd(OAc) 2 | 1:2 | ZnBr2 | THF | 81 | 85.0:15.0 |
| 2 | Pd2(dba)3 | 1:2 | ZnBr2 | Toluene | 80 | 85.0:15.0 |
| 3 | Pd2(dba)3 | 1:2 | ZnBr2 | Dioxane | 90 | 85.0:15.0 |
| 4 | PdCl2(PhCN)2 | 1:2 | ZnBr2 | THF | 71 | 82.0:18.0 |
| 5 | [crotylPdCl]2 | 1:2 | ZnBr2 | THF | 73 | 83.0:17.0 |
| 6 | Pd2(dba)3 | 1:2 | Zn(OAc)2 | THF | 56 | 78.0:22.0 |
| 7c | Pd2(dba)3 | 1:2 | Zn(OPiv)2 | THF | 77 | 84.0:16.0 |
| 8 | Pd2(dba)3 | 1:2 | ZnBr2 | THF | 80 | 85.0:15.0 |
| 9 | Pd2(dba)3 | 1:1.5 | ZnBr2 | THF | 93 | 85.0:15.0 |
| 10 | Pd2(dba)3 | 1:1 | ZnBr2 | THF | 76 | 85.0:15.0 |
Isolated yield.
Determined by chiral HPLC.
The zinc reagent was prepared from the Grignard according to Knochel’s protocol (ref 38a).
To determine if enantioselection in the reaction was variable with reaction progress, the enantioselectivity was monitored over time at both 40 and 60 oC (Scheme 8). Enantioselection was independent of conversion implying that the active catalyst was not changing over time. Reduction of the reaction temperature led to slight improvements in enantioselectivity, but incomplete conversion was obtained at 40 oC.
Scheme 8.

Enantioselectivity vs Time at 40 and 60 °C.
The most efficient catalyst conditions identified in our studies (Table 5, entry 9) were next utilized to examine the scope of the Negishi cross-coupling to afford other tetra-ortho-substitued biaryls (Scheme 9). Notably, the Negishi reaction showed much better scope in terms of reaction yield relative to the Suzuki-Miyaura reaction albeit with lower enantioselection. For the first time, a variety of R-alkyl substituents other than R = Me or OMe have been analyzed in an asymmetric cross-coupling reaction to provide tetra-ortho-substituted products (39d-f). In general, the cross-coupling employing 35 with R = n-alkyl afforded comparable results (18, 39a,c-f). Methoxy-substitution (39b) was not well tolerated and afforded reduced yield due to significant amounts of reduction of the aryl bromide as a byproduct. Adding an additional methyl-substituent on the naphthalene ring also led to decreased reaction efficiency and enantioselectivity (39c); however, product formation was observed whereas the corresponding Suzuki-Miyaura coupling afforded no reaction with this substrate class (Scheme 7). Finally, the present system was specific to coupling naphthalene derivatives. Cross-coupling to afford 43 proceeded in moderate to good yields, but almost no enantioselection was observed despite the nature of the alkoxy-substituent.
Scheme 9.

Asymmetric Negishi Cross-Coupling Reaction Scope.
Mechanistic Modeling by DFT Analysis.
The results of our combined studies of the asymmetric Suzuki-Miyaura and Negishi cross-coupling reactions employing ligand series 33 and 36 prompted us to investigate the mechanistic differences between these two reactions in terms of the enantiodiscrimination steps. Specifically, these ligands gave very different results in the coupling reaction when employed in a Suzuki-Miyaura vs a Negishi type protocol, yet mechanistically, the final step in the catalytic cycle for both processes would presumably proceed through the same reductive elimination intermediate (vide supra). To help understand these differences and to aid in the development of new ligand structures to improve enantioselectivity, we modeled the cross-coupling process using density functional theory (DFT) calculations.39
Negishi Cross-Coupling
Most analyses of the stereoselectivity in cross-couplings regard the final step as stereodetermining.11b,13b,n,x Such an approach assumes equal accessibility to all of the different conformers of the transition state precursors. The percentage of product that forms via each of the possible reductive elimination transitions states can then be determined by Boltzmann analysis of their relative energies.
For the system in question here, a defining feature of the Pd biaryl intermediates preceding reductive elimination is the hindered rotation along the aryl-Pd bonds due to the large size of the aryl groups employed. As such, direct interconversion13f,15a of the different conformations of these intermediates does not occur. However, we hypothesize that all the conformers will be equally accessible due to the reversibility of transmetalation and oxidative addition in the Negishi coupling.40 In such a scenario, a classical analysis of just the reductive elimination transition states pertains. Thus, all four of the possible biaryl palladium isomers need to be considered, each giving rise to two reductive elimination transition states, pro-S and pro-R (Figure 1).
Figure 1.

Stereoselection in Negishi Coupling
Analyzing reductive eliminations for the Me-BOP ligand L2 in such a fashion, we found that there is one dominant transition state for each of pro-S and pro-R sets, controlling the overall selectivity (Figure 2).
Figure 2.

Lowest energy pro-S and pro-R reductive elimination transition states for Me-BOP ligand L2. Total of 8 transition states were located and analyzed. Free energies were computed using M06/6–311+G(d,p)-LANL2DZ(Pd)-SMD-THF//B3LYP/6–31G(d)-LANL2DZ(Pd); values are in kcal/mol.
Coordination of Pd to the methoxy-group of the lower aryl ring of the Me-BOP ligand provides extra stabilization for the pro-S transition state, resulting in highly S-selective reaction. For the pro-R transition state, such an arrangement is not favorable due to the steric interactions with the ligand. A correlation between the electron-donating properties of the lower ring substituent (σp) and the experimental er of the product supports the contribution of this specific substituent to the enantioselection mechanism (Figure 3).
Figure 3.

Efficiency of the chiral Me-BOP ligand vs electron donating property of the lower aryl-ring substituent.
When the Negishi coupling is performed with the alternate DMM-substituted BOP ligand (NitinPhos, L7), enantioselectivity drops. This result can be rationalized using the same approach. Analysis of the selectivity-determining transition states revealed that the additional coordinating group of the DMM-substituent of NitinPhos ligand L7 allows selective stabilization of the higher energy pro-R transition state (Figure 4), which diminishes the pro-R– pro-S energy gap and hence the selectivity of the overall process.
Figure 4.

Lowest energy pro-S and pro-R reductive elimination transition states for the DMM-BOP ligand (NitinPhos, L7). Total of 32 transition states were located and analyzed. Free energies were computed using M06/6–311+G(d,p)-LANL2DZ(Pd)-SMD-THF//B3LYP/6–31G(d)-LANL2DZ(Pd); values are in kcal/mol.
This analysis of the reductive elimination transition states for the Negishi reaction rationalizes the experimentally observed enantioselectivities and highlights the key interactions (Figure 5).
Figure 5.

Comparison of the experimental and calculated efficiencies of the ligands in Negishi coupling.
Suzuki-Miyaura Coupling
The different performance of the DMM-BOP ligand (NitinPhos, L7) in the Suzuki-Miyaura coupling requires a different enantioselection mechanism. Irreversible transmetalation41 in Suzuki-Miyaura coupling would prevent pre-equilibration of pre-reductive elimination Pd(II) species. As such, enantioselectivity of the overall process would be determined not only by the energetics of reductive elimination, but also by the relative population of the pre-reductive elimination Pd(II) intermediates. Thus, it is necessary to consider the relative rates of formation of corresponding species by modeling the transmetalation step. Based on the proposed structures of the pre-transmetalation intermediates,42 the same hindered rotation issues are expected as for the pre-reductive elimination intermediates. Therefore, pre-transmetalation intermediates partition into two groups that react via independent pathways. In turn, the population of the pre-transmetalation intermediates is controlled by the energetics of oxidative addition. Therefore, all three major steps of the cross-coupling mechanism contribute to the overall selectivity of the reaction (Figure 6). Earlier computational studies of the stereoselectivity in Suzuki-Miyaura cross-coupling suggest that oxidative addition is the stereodetermining step.43 We further develop this idea of early enantioinduction by employing the experimental findings of Denmark,42a,b Hartwig,42c and others42d,e that point to the mechanism of the transmetalation step proceeding through a Pd-OH complex.17 The exact mechanism for the transmetalation step in Suzuki-Miyaura cross-coupling is still somewhat controversial39 as there are some reports44 that support the alternate “boronate”-transmetalation pathway.45 The studies reported herein are calculated using the Pd-OH transmetalation pathway;42 however, we would like to stress that the model proposed herein is also compatible with the “boronate”-transmetalation44,45 mechanism: both possibilities ultimately lead to the same transmetalation transition states. The nature of the incoming boron species (boronate vs boronic acid) is not crucial for the selection of the productive oxidative addition step (discussed later) as the same steric interactions are present.
Figure 6.

Proposed stereoselection mechanism in Suzuki coupling. Species separated by dashed lines cannot equilibrate. Rotation around the bonds marked in red is hindered.
Hence, detailed computational analysis of all the steps of the coupling is required. The resultant three-layer selectivity mechanism is shown on Figure 6. This important difference between the Suzuki-Miyaura and Negishi couplings is the result of the irreversible nature of transmetalation in Suzuki-Miyaura coupling. As such, the difference in the energetics of transmetalation significantly complicates the stereoselection process, which explains the well-known challenges in the design of stereoselective Suzuki-Miyaura couplings.
This analysis of the Suzuki-Miyaura coupling focuses two very similar ligands, DMM- and DME-BOP (L7 vs L6, respectively). Despite minor differences in structure, these ligands provide substantially different selectivity levels. The following study rationalizes these selectivity differences within the proposed multi-level selectivity framework. To begin, a detailed analysis of the DME-BOP (L6) ligand is outlined.
Oxidative addition
Oxidative addition to Pd(0) can occur via two pathways, A and B (Figure 7). Exchange of the bromide ultimately leads to several isomers of the PdOH species (Figure 8).
Figure 7.

Oxidative addition pathways in Suzuki coupling. Intermediates marked in red are inert in transmetallation step. Enthalpies and free energies (in brackets) were computed using B3LYP-d3/6–31G(d)-lanl2dz(Pd, Br); values are in kcal/mol. Reactions are balanced using water as the OH source and HBr as the by-product of the exchange.
Figure 8.

Isomeric PdOH intermediates.
In considering the PdOH isomers (Figure 8), only iso1 is ultimately relevant to the next step, transmetalation. For transmetalation to take place, the hydroxyl group and the vacant orbital in PdOH species must be positioned cis on the metal center, which rules out iso2 and iso3. Isomers iso4 and iso5 have severe destabilizing steric interaction between ligand and the naphthyl group, making them unlikely to be formed. Furthermore, iso6 is not active in transmetalation as the vacant orbital is blocked by the ligand, so the approach of incoming boronic acid is problematic. As a consequence, only iso1, which arises from oxidative addition pathway A, needs to be analyzed. Given the activity of iso1 PdOH in transmetalation, we theorize that the other species may eventually convert to iso1.
Pathway A oxidative addition can occur via two conformations, corresponding to different orientations of the aryl group (‘up’ and ‘down’, see Figure 9). The relative energies of these conformations ultimately translate to the ratio of the PdOH species after exchange of the bromide. Based on the energies of oxidative addition transition states, 85% of the formed PdOH arises from OA A1 where the methyl of the naphthyl unit is pointing ‘away’. This selectivity is attributed to an unfavorable methyl-ligand interaction in the transition state OA A2 (red arrow, Figure 9 insert). This differentiation sets up the first layer in the enantioselection mechanism.
Figure 9.

Energetics and selectivity of the productive oxidative addition pathway in Suzuki coupling. Total of 8 oxidative addition transition states were located and analyzed. Enthalpies and free energies (in brackets) were computed using B3LYP-d3/6–31G(d)-LANL2DZ (Pd, Br); values are in kcal/mol. From here on the reactions are balanced using KOH as the OH source and KBr as the by-product of PdOH formation.
Transmetalation
The resultant PdOH intermediates form a pre-transmetalation complex with the boronic acid (preTM, insert in Figure 10). The incoming boronic acid can coordinate such that the methyl substituents on the naphthyl rings are cis or trans. Starting from the major PdOH adduct (PdOH_iso 1_a, Figure 9), the trans adduct is more favorable due to the lack of Me-Me repulsion. This interaction also controls the energies of the concomitant transmetalation transition states (TM TS cis and trans, figure 10). The energy gap between transition states results in 85% of the pre-RE-trans product vs 15% of the pre-RE-cis product. For the minor PdOH adduct (PdOH_iso 1_b, Figure 9), these trends are reversed (see Figure 11) due to the greater spacing between the naphthyl units. As a result, the cis adduct (pre-RE-cis, figure 11) predomominates over the trans (pre-RE-trans, figure 11) adduct (76% vs 24%). During the transmetalation process, an equivalent of boronic acid is released that exists as the borate under the basic reaction conditions.46
Figure 10.

Energetics and selectivities of the transmetalation and reductive eliminations arising from the major PdOH isomer. Total of 8 transmetalation transition states and 16 reductive elimination transition states were located and analyzed. Enthalpies and free energies (in brackets) were computed using B3LYP-d3/6–31G(d)-LANL2DZ (Pd, Br); values are in kcal/mol.
Figure 11.

Energetics and selectivities of the transmetalation and reductive eliminations arising from the minor PdOH isomer. Total of 10 transmetalation transition states and 16 reductive elimination transition states were located and analyzed. Enthalpies and free energies (in brackets) were computed using B3LYP-d3/6–31G(d)-lanl2dz (Pd, Br); values are in kcal/mol.
Reductive Elimination
Both the cis and trans isomers of the diaryl Pd species (preRE intermediates in Figures 10 and 11) can reductively eliminate to form either the S or R binaphthyl. Analyzing the energetics of reductive eliminations resulting from the major PdOH isomer, we found that cis pre-reductive elimination intermediate affords the R-isomer with excellent selectivity, while the trans isomer furnishes exclusively the S-product (reductive eliminations, Figure 10).
The extremely high selectivity of the individual reductive eliminations is due to the compact nature of the transition states. Even minor deviations from the optimal arrangement result in significant energetic penalties. For example, the reductive elimination transition states from the major PdOH isomer depicted in Figure 10, illustrate these phenomena (Figure 12). For the pre-RE-trans isomer, tr S transition state (100%) is much more favorable than tr R, in which interactions of the methyl of the upper naphthyl unit (arising from the aryl bromide) and the back side of the ligand destabilize the structure. Similarly, cis R (100%) is more stable than cis S for the pre-RE-cis isomer.
Figure 12.

Lowest energy pro-S and pro-R reductive elimination transition states from the major trans and cis isomers of diaryl Pd.
As the orientation of the naphthyl fragment is important for the high stereoselectivity of reductive elimination, one might expect lower selectivities for the case of reductive elimination from the minor PdOH isomer (Figure 11), where the upper naphthyl unit is rotated in the opposite direction. Indeed, the pre-RE-cis is only moderately selective towards reductive elimination vs cis R (65%). Reductive elimination from pre-RE-trans remains highly selective (100% tr R) because the methyl group on the lower ring is pointing toward a sterically crowded portion of the transition state.
However, the relative energies of the reductive elimination transition states do not tell the whole story as the intermediates needed for this step form at different rates with both oxidative addition and transmetalation contributing to the partitioning. Figure 13 illustrates a simplified version where different conformational isomers are not shown for each of the branches. Oxidative addition leads to two branches (left = minor, right = major), which again bifurcate at transmetalation and reductive elimination. The pathway highlighted with blue arrows is the only S-selective channel on the diagram. Even though there are more pathways to the R-enantiomer, the S-selective channel is lower in energy and is the main route responsible for formation of the product.
Figure 13.

Simplified energy diagram for Suzuki coupling with DME-BOP (L6) ligand. Selectivities of each step were calculated based on the Boltzmann distribution of the energies of corresponding transition states. (Alternatively, such multi-level processes can be analyzed using a system of rate equations, i.e. microkinetic modeling.47)
Analysis of the DMM-BOP (NitinPhos, L7) ligand generates a similar outcome (Figure 14). The entire transformation is still S-selective and just like in a previous case, there is only one pathway responsible for the production of the major product. However, this pathway is more favorable than it is in the DME-BOP (L6) case. For the DMM ligand (NitinPhos, L7) transmetalation selectivity in the ‘major branch’ (Figure 14) is 90:10, whereas for the DME ligand it is at a slightly smaller 85:15. This difference causes selectivity levels to be higher for DMM (NitinPhos, L7).
Figure 14.

Simplified energy diagram for Suzuki coupling with DMM-BOP (NitinPhos, L7) ligand.
If one assumes that the entire reaction is controlled only by the selectivity of the major branch, then the difference in the transmetalation energetics would predict the experimentally observed selectivities with improved accuracy (see ‘major branch’ selectivities on each of the diagrams). Thus, our model would predict 85% S selectivity for DME-BOP (L6) ligand vs 86% experimental and 90% S selectivity for DMM-BOP (NitinPhos, L7) ligand vs 94% experimental. Given the complex nature of the selectivity mechanism and the large number of transition states that contribute to the finalized percentages (26 transition states conformers were analyzed and utilized in Boltzmann distributions), this result may indicate that the major branch is indeed the only controlling channel. Such a situation would occur if our estimates of the pre-transmetalation intermediates are not entirely accurate. In our work we tested five computational methods (different combinations of various optimization and single-point calculations) with 36–56 transition state conformations depending on the method. These analyses did not improve the quality of the predictions. This high sensitivity of the selectivity estimates to the type of method used may indicate that our method of choice is not optimal and leads to incorrect estimates of the energetics of major/minor branch splitting. It might also point to the existence of other selectivity-contributing steps (for example, Br exchange) that were not considered.
Since the transmetalation selectivity in the major branch seems to account for the observed difference in the performance of the ligands, we undertook further study of the transmetalation step. As discussed earlier, each of the energy levels shown on the diagrams represents only the lowest energy structure out of a set of calculated conformational isomers. Importantly, molecules traversing the reaction coordinate proceed through a distribution of available conformational transition states rather than through the single lowest energy pathway;48 therefore, the calculated selectivities given in this study were determined using the energies of all the available conformations identified by calculation. This consideration accounts for the computationally observed different transmetalation selectivities of the DMM-BOP (NitinPhos, L7) and DME-BOP (L6) ligands (Figure 15). For the DMM-BOP (NitinPhos, L7) ligand, the transmetalation pathways leading to S-selective reductive elimination are more populated than that for the DME (L6) ligand (90% vs 85%), while the other pathways are less populated (10% vs 15%). This result arises from the greater partitioning to the S-selective transition states for the DMM-BOP ligand which are clustered close together relative to those for the DME-BOP ligand which cover a larger range and are less populated overall relative to the R-selective transition state. This phenomenon accounts for the improved transmetalation selectivity obtained with the DMM-BOP ligand (L7) vs the DME-BOP ligand (L6), and can be explained structurally by the presence of hydrogen bonding interactions between weak C–H hydrogen bond donors of the methylnaphthalene coupling partners with the methoxy-group of the DMM-BOP ligand that is not possible in the DME-BOP ligand case (i.e. 42 vs 45). Hydrogen bond contacts between the methoxy-group of NitinPhos (L7) and C–H bonds of the methylnaphthalene coupling partners are present in transmetalation transition states (42, 43). The bond distances for these O–H interactions are within the distance and angles typical of weak C–H–O hydrogen bonds (2.2–2.8 A; 42: Csp2–H–O = 119o, Csp3–H–O = 157o; 43: Csp3–H–O = 158o, 135o).13b,49 These stabilizing interactions cause fewer accessible low-energy transmetalation transition state conformations to be available in the DMM-BOP ligand case because rotation of the DMM-group results in loss of these interactions which creates a higher energy gap between these conformations such that they cannot contribute to product formation. In contrast, the DME-BOP ligand does not have these interactions, and therefore, the energy splitting between conformers is less severe and leads to several accessible pathways of various energy levels with slight differences in steric interactions due to the orientation of the ethyl-group of the DME-substituent. Finally, it is worth noting that direct coordination of the methoxy-substituent of the DMM-BOP ligand (L7) to the metal center was not found in any of the relevant transition states.
Figure 15.

Relative energies (in bold), and relative populations of the major branch transmetalation transition states for DMM-BOP (NitinPhos, L7) and DME-BOP (L6) ligands. Green lines represent C–H---O hydrogen bonding interactions. Dashed lines indicate additional high-energy conformations that do not significantly contribute to the population distribution.
This DFT study of the stereoselectivity of Suzuki-Miyaura coupling allowed us to propose a multi-step stereoselection framework. This mechanism is likely to operate in the coupling of sterically hindered substrates, where isomeric organometallic intermediates do not readily interconvert.
CONCLUSION
In conclusion, Pd-catalyzed asymmetric Suzuki-Miyaura and Negishi cross-coupling reactions to form tetra-ortho-substituted biaryls were developed. Higher enantioselectivities were obtained for the Suzuki-Miyaura reaction; however, better substrate generality was obtained for the Negishi coupling. For each coupling reaction, a unique ligand was identified to be optimal for the Suzuki-Miyaura vs the Negishi reaction in terms of providing maximum enantioselection. Detailed DFT calculations of these two processes offered an explanation for the differences in the enantiodiscrimination mechanism between the Suzuki-Miyaura and Negishi cross-coupling reactions that likely proceed through the same Pd-intermediate for reductive elimination.
Analysis of only the reductive elimination step in the Negishi cross-coupling reaction by DFT provided good correlation with the experimental enantioselectivity, implying that reductive elimination was enantiodetermining. In contrast, this same analysis for the Suzuki-Miyaura coupling did not accurately predict the observed experimental selectivities. A complete analysis of the oxidative addition, transmetalation, and reductive elimination steps in the Suzuki-Miyaura reaction using DFT revealed that the selectivity in the Suzuki-Miyaura reaction was influenced by all three steps in the catalytic cycle. The delicate interplay of oxidative addition, transmetalation, and reductive elimination calculated for stereocontrol in the Suzuki-Miyaura cross-coupling reaction for the formation of tetra-ortho-substituted biaryls may be the reason why the Suzuki-Miyaura reaction was less tolerant to substrate modifications relative to the analogous Negishi reaction. These findings may aid in the design and development of new catalysts to improve reactivity and selectivity in asymmetric cross-coupling reactions to furnish tetra-ortho-substituted biaryls
Supplementary Material
ACKNOWLEDGEMENTS
We thank the NIH (GM087605 M.C.K.) and Boehringer Ingelheim Pharmaceuticals for financial support. Computational support was provided by XSEDE (TG-CHE120052).
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge on the ACS Publication website at DOI: Experimental procedures, characterization data for new compounds, DFT methods, intermediate and transition state coordinates, and copies of 1H, 13C, and 31P NMR spectra of all new compounds.
The authors declare the following competing financial interest(s): Some of the ligands used in this work are covered by the following patents: US 9096626, US 20130137902, US 20180155377, WO 2011126917.
REFERENCES
- 1.(a).Hassan J; Sevignon M; Gozzi C; Schulz E; Lemaire M Aryl-Aryl Bond Formation One Century after the Discovery of the Ullmann Reaction Chem. Rev 2002, 102, 1359–1470. [DOI] [PubMed] [Google Scholar]; (b) Littke AF; Fu GC Palladium-Catalyzed Coupling Reactions of Aryl Chlorides Angew. Chem. Int. Ed 2002, 41, 4176–4211. [DOI] [PubMed] [Google Scholar]; (c) Tamao K; Miyaura N Cross-Coupling Reactions Top. Curr. Chem 2002, 219, 1–9. [Google Scholar]; (d) Stanforth SP Catalytic Cross-Coupling Reactions in Biaryl Synthesis Tetrahedron 1998, 54, 263–303. [Google Scholar]; (e) Metal-Catalyzed Cross-Coupling Reactions; Diederich F, Stang PJ, Eds.; Wiley-VCH: New York, 1998. [Google Scholar]; (f) Lessene G Advances in the Negishi Coupling Aust. J. Chem 2004, 57, 107. [Google Scholar]; (g) Palladium-Catalyzed Coupling Reactions: Practical Aspects and Future Developments; Molnar A, Ed; Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, 2013. [Google Scholar]
- 2.Miyaura N; Yamada K; Suzuki A A New Stereospecific Cross-Coupling by the Palladium-Catalyzed Reaction of 1-Alkenylboranes with 1-Alkenyl or 1-Alkynyl Halides Tetrahedron Lett 1979, 20, 3437–3440. [Google Scholar]
- 3.Negishi E; King AO; Okukado N Selective Carbon-Carbon Bond Formation via Transition Metal catalysis. 3. A Highly Selective Synthesis of Unsymmetrical Biaryls and Diarylmethanes by the Nickel- or Palladium-Catalyzed Reaction of Aryl- and Benzylzinc Derivatives with Aryl Halides J. Org. Chem 1977, 42, 1821–1823. [Google Scholar]
- 4.Reviews: Suzuki A Recent Advances in the Cross-Coupling Reactions of Organoboron Derivatives with Organic Electrophiles J. Organomet. Chem 1999, 576, 147–168.Miyaura N; Suzuki A Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds Chem. Rev 1995, 95, 2457–2483.Bellina F; Carpita A; Rossi R Palladium Catalysts for the Suzuki Cross-Coupling Reaction: An Overview of Recent Advances Synthesis 2004, 2419–2440.Christmann U; Vilar R Monoligated Palladium Species as Catalysts in Cross-Coupling Reactions Angew. Chem. Int. Ed 2005, 44, 366–374.Alonso F; Beletskaya IP; Yus M Non-Conventional Methodologies for Transition-Metal Catalysed Carbon-Carbon Coupling: a Critical Overview. Part 2: The Suzuki Reaction Tetrahedron 2008, 64, 3047–3101.Miyaura N Organoboron Compounds Top. Curr. Chem 2002, 219, 11–59.Johansson Seechurn CCC; Kitching MO; Colacot TJ; Snieckus V Palladium-Catalyzed Cross-Coupling: A Historical Contextual Perspective to the 2010 Nobel Prize Angew. Chem. Int. Ed 2012, 51, 5062–5085.Lundgren RJ; Stradiotto M Addressing Challenges in Palladium-Catalyzed Cross-Coupling Reactions Through Ligand Design Chem. Eur. J 2012, 18, 9758–9769.
- 5. (a).Miura M Rational Ligand design in Constructing Efficient Catalyst Systems for Suzuki-Miyaura Coupling Angew. Chem. Int. Ed 2004, 43, 2201–2203. [DOI] [PubMed] [Google Scholar]; (b) Martin R; Buchwald SL Palladium-Catalyzed Suzuki-Miyaura Cross-Coupling Reactions Employing Dialkylbiaryl Phosphine Ligands Acc. Chem. Res 2008, 41, 1461–1473. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) DeAngelis AJ; Gildner PG; Chow R; Colacot TJ Generating Active “L-Pd(0)” via Neutral or Cationic p-Allylpalladium Complexes Featuring Biaryl/Bipyrazolylphosphines: Synthetic, Mechanistic, and Structure-Activity Studies in Challenging Cross-Coupling Reactions J. Org. Chem 2015, 80, 6794–6813. [DOI] [PubMed] [Google Scholar]; (d) Kinzel T; Zhang Y; Buchwald SL A New Palladium Precatalyst Allows for the Fast Suzuki-Miyaura Coupling Reactions of Unstable Polyfluorophenyl and 2-Heteroaryl Boronic Acids J. Am. Chem. Soc 2010, 132, 14073–14075. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Yang Y; Oldenhuis NJ; Buchwald SL Mild and General Conditions for Negishi Cross-Coupling Enabled by the Use of Palladacycle Precatalysts Angew. Chem. Int. Ed 2013, 52, 615–619. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Bruno NC; Tudge MT; Buchwald SL Design and Preparation of New Palladium Precatalysts for C-C and C-N Cross-Coupling Reactions Chem. Sci 2013, 4, 916–920. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. (a).Dumrath A; Lubbe C; Beller M “Palladium-Catalyzed Cross-Coupling Reactions–Industrial Applications”, in Palladium-Catalyzed Coupling Reactions: Practical Aspects and Future Developments; Molnar A, Ed.; Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, 2013; pages 445–489. [Google Scholar]; (b) Torborg C; Beller M Recent Applications of Palladium-Catalyzed Coupling Reactions in the Pharmaceutical, Agrochemical, and Fine Chemical Industries Adv. Synth. Catal 2009, 351, 3027–3043. [Google Scholar]; (c) Budarin V; Shuttleworth PS; Clark JH; Luque R Industrial Applications of C-C Coupling Reactions Curr. Org. Synth 2010, 7, 614–627. [Google Scholar]
- 7. (a).Littke AF; Fu GC A Convenient and General Method for Pd-Catalyzed Cross-Couplings of Aryl Chlorides and Arylboronic Acids Angew. Chem. Int. Ed 1998, 37, 3387–3388. [DOI] [PubMed] [Google Scholar]; (b) Littke AF; Dai C; Fu GC Versatile Catalysts for the Suzuki Cross-Coupling of Arylboronic Acids with Aryl and Vinyl Halides and Triflates under Mild Conditions J. Am. Chem. Soc 2000, 122, 4020–4028. [Google Scholar]; (c) Stambuli JP; Kuwano R; Hartwig JF Unparalleled Rates for the Activation of Aryl Chlorides and Bromides: Coupling with Amines and Boronic Acids in Minutes at Room Temperature Angew. Chem. Int. Ed 2002, 41, 4746. [DOI] [PubMed] [Google Scholar]; (d) Fleckenstein CA; Plenio H 9-Fluorenylphosphines for the Pd-Catalyzed Sonogashira, Suzuki, and Buchwald-Hartwig Coupling Reaction in Organic Solvents and Water Chem. Eur. J 2007, 13, 2701–2716. [DOI] [PubMed] [Google Scholar]; (e) Fleckenstein CA; Plenio H Efficient Suzuki-Miyaura Coupling of (Hetero)aryl Chlorides with Thiophene- and Furanboronic Acids in Aqueous n-Butanol J. Org. Chem 2008, 73, 3236–3244. [DOI] [PubMed] [Google Scholar]; (f) Guram AS; King AO; Allen JG; Wang X Schenkel LB; Chan J; Bunel EE; Faul MM; Larsen RD; Martinelli MJ; Reider PJ New Air-Stable Catalysts for General and Efficient Suzuki-Miyaura Cross-Coupling Reactions of Heteroaryl Chlorides Org. Lett 2006, 8, 1787–1789. [DOI] [PubMed] [Google Scholar]; (h) Hoshi T; Nakazawa T; Saitoh I; Mori A; Suzuki T; Sakai J; Hagiwara H Biphenylene-Substituted Ruthenocenylphosphine for Suzuki-Miyaura Coupling of Aryl Chlorides Org. Lett 2008, 10, 2063–2066. [DOI] [PubMed] [Google Scholar]; (i) Hoshi T; Saitoh T; Nakazawa T; Suzuki T; Sakai J; Hagiwara H Biphenylene-Substituted Ruthenocenylphosphine for Suzuki-Miyaura Coupling of Sterically Hindered Aryl Bromides J. Org. Chem 2009, 74, 4013–4016. [DOI] [PubMed] [Google Scholar]; (j) Yin J: Rainka MP; Zhang X–X; Buchwald SL A Highly Active Suzuki Catalyst for the Synthesis of Sterically Hindered Biaryls: Novel Ligand Coordination J. Am. Chem. Soc 2002, 124, 1162–1163. [DOI] [PubMed] [Google Scholar]; (k) Yin J; Buchwald SL A Catalytic Asymmetric Suzuki Coupling for the Synthesis of Axially Chiral Biaryl Compounds J. Am. Chem. Soc 2000, 122, 12051–12052. [Google Scholar]; (l) Walker SD; Barder TE; Martinelli JR; Buchwald SL A Rationally Designed Universal Catalyst for Suzuki-Miyaura Coupling Processes Angew. Chem. Int. Ed 2004, 43, 1871–1876. [DOI] [PubMed] [Google Scholar]; (m) Barder TE; Walker SD; Martinelli JR; Buchwald SL Catalysts for Suzuki-Miyaura Coupling Processes: Scope and Studies of gthe Effect of Ligand Structure J. Am. Chem. Soc 2005, 127, 4685–4696. [DOI] [PubMed] [Google Scholar]; (n) Billingsley KL; Anderson KW; Buchwald SL A Highly Active Catalyst for Suzuki-Miyaura Cross-Coupling Reactions of Heteroaryl Compounds Angew. Chem. Int. Ed 2006, 45, 3484–3488. [DOI] [PubMed] [Google Scholar]; (o) Billingsley K; Buchwald SL A Highly Efficient Monophosphine-Based Catalyst for the Palladium-Catalyzed Suzuki-Miyaura Reaction of Heteroaryl Halides and Heteroaryl Boronic Acids and Esters J. Am. Chem. Soc 2007, 129, 3358–3366. [DOI] [PubMed] [Google Scholar]; (p) Bhayana B; Fors BP; Buchwald SL A Versatile Catalyst System for Suzuki-Miyaura Cross-Coupling Reactions of C(sp2)-Tosylates and Mesylates Org. Lett 2009, 11, 3954–3957. [DOI] [PMC free article] [PubMed] [Google Scholar]; (p) Tang W; Capacci AG; Wei X; Li W; White A; Patel ND; Savoie J; Gao J; Rodriguez S; Qu B; Haddad N; Lu BZ; Krishnamurthy D; Yee NK; Senanayake CH A General and Special Catalyst for Suzuki-Miyaura Coupling Processes Angew. Chem. Int. Ed 2010, 49, 5879–5883. [DOI] [PubMed] [Google Scholar]; (q) Guram AS; King AO; Allen JG; Wang X; Schenkel LB; Chan J Bunel EE; Faul MM; Larsen RD; Martinelli MJ; Reider PJ New Air-Stable Catalysts for General and Efficient Suzuki-Miyaura Cross-Coupling Reactions of Heteroaryl Chlorides Org. Lett 2006, 8, 1787–1789. [DOI] [PubMed] [Google Scholar]; (r) Milne JE; Buchwald SL An Extremely Active Catalyst for the Negishi Cross-Coupling Reaction J. Am. Chem. Soc 2004, 126, 13028–13032. [DOI] [PubMed] [Google Scholar]; (s) Dai C; Fu GC The First General Method for Palladium-Catalyzed Negishi Cross-Coupling of Aryl and Vinyl Chlorides: Use of Commercially Available Pd(P(t-Bu)3)2 as a Catalyst J. Am. Chem. Soc 2001, 123, 2719–2724. [DOI] [PubMed] [Google Scholar]; (t) Sieber JD; Qu B; Rodriguez S; Haddad N; Grinberg N; Lee H; Song JJ; Yee NK; Senanayake CH Synthesis of P-Chiral Dihydrobenzooxaphospholes Through Negishi Cross-Coupling J. Org. Chem 2016, 81, 729–736. [DOI] [PubMed] [Google Scholar]; (u) Luzung MR; Patel JS; Yin J A Mild Negishi Cross-Coupling of 2-Heterocyclic Organozinc Reagents and Aryl Chlorides J. Org. Chem 2010, 75, 8330–8332. [DOI] [PubMed] [Google Scholar]; (v) Zhang Y; Lao KS; Sieber JD; Xu Y; Wu L; Wang X–J; Desrosiers J–N; Lee H; Haddad N; Han S; Yee NK; Song JJ; Howell AR; Senanayake CH Modular Dihydrobenzoazaphosphole Ligands for Suzuki-Miyaura Cross-Coupling Synthesis 2018, DOI: 10.1055/s-0037-1610158 [DOI] [Google Scholar]
- 8. (a).Gstottmayr CWK; Bohm VPW; Herdtweck E; Grosche M; Herrmann WA A Defined N-Heterocyclic Carbene Coplex for the Palladium-Catalyzed Suzuki Cross-Coupling of Aryl Chlorides at Ambient Temperatures Angew. Chem. Int. Ed 2002, 41, 1363–1365. [DOI] [PubMed] [Google Scholar]; (b) Navarro O; Kelly RA; Stevens ED; Nolan SP A General Method for the Suzuki-Miyaura Cross-Coupling of Sterically Hindered Aryl Chlorides: Synthesis of Di- and Tri-ortho-substituted Biaryls in 2-Propanol at Room Temperature J. Am. Chem. Soc 2003, 125, 16194–16195. [DOI] [PubMed] [Google Scholar]; (c) Altenhoff G; Goddard R; Lehmann CW; Glorius F An N-Heterocyclic Carbene Ligand with Flexible Steric Bulk Allows Suzuki Cross-Coupling of Sterically Hindered Aryl Chlorides at Room Temperature Angew. Chem. Int. Ed 2003, 42, 3690–3693. [DOI] [PubMed] [Google Scholar]; (d) Altenhoff G; Goddard GR; Lehmann CW; Glorius F Sterically Demanding, Bioxazoline-Derived N-Heterocyclic Carbene Ligands with Restricted Flexibility for Catalysis J. Am. Chem. Soc 2004, 126, 15195–15201. [DOI] [PubMed] [Google Scholar]; (e) Song C; Ma Y; Chai Q; Ma C; Jiang W; Andrus MB Palladium Catalyzed Suzuki-Miyaura Coupling with Aryl Chlorides using a Bulky Phenanthryl N-Heterocyclic Carbene Ligand Tetrahedron 2005, 61, 7438–7446. [Google Scholar]; (f) Marion N; Navarro O; Mei J; Stevens ED; Scott NM; Nolan SP Modified (NHC)Pd(allyl)Cl (NHC = N-Heterocyclic Carbene) Complexes for Room-Temperature Suzuki-Miyaura and Buchwald-Hartwig Reactions J. Am. Chem. Soc 2006, 128, 4101–4111. [DOI] [PubMed] [Google Scholar]; (g) O’Brien CJ; Kantchev EAB; Valente C; Hadei N; Chass GA; Lough A; Hopkinson AC; Organ MG Easily Prepared Air- and Moisture-Stable Pd-NHC (NHC = N-Heterocyclic Carbene) Complexes: A Reliable, User-Friendly, Highly Active Palladium Precatalyst for the Suzuki-Miyaura Reaction Chem. Eur. J 2006, 12, 4743–4748. [DOI] [PubMed] [Google Scholar]; (h) Organ MG; Calimsiz S; Sayah M; Hoi KH; Lough AJ Pd-PEPPSI-IPent: an Active, Sterically Demanding Cross-Coupling Catalyst and Its Application in the Synthesis of tetra-ortho-substituted Biaryls Angew. Chem. Int. Ed 2009, 48, 2383–2387. [DOI] [PubMed] [Google Scholar]; (i) Valente C; Belowich BE; Hadei N; Organ MG Pd-PEPPSI Complexes and the Negishi Reaction Eur. J. Org. Chem 2010, 4343–4354. [Google Scholar]
- 9. (a).Zhao Q; Li C; Senanayake CH; Tang W An Efficient Method for Sterically Demanding Suzuki-Miyaura Coupling Reactions Chem. Eur. J 2013, 19, 2261–2265. [DOI] [PubMed] [Google Scholar]; (b) Demchuk OM; Yoruk B; Blackburn T; Snieckus V A Mixed Naphthyl-Phenyl Phosphine Ligand Motif for Suzuki, Heck, and Hydrodehalogenation Reactions Synlett 2006, 2908–2913. [Google Scholar]; (c) Lu D–D; He X–X; Liu F–S Bulky Yet Flexible Pd-PEPPSI-IPentAn for the Synthesis of Sterically Hindered Biaryls in Air J. Org. Chem 2017, 82, 10898–10911. [DOI] [PubMed] [Google Scholar]; (d) Lesieur M; Slawin AMZ; Cazin CSJ [Pd(μ-Cl)Cl(IPr*)]2: a Highly Hindered Pre-Catalyst for the Synthesis of Tetra-ortho-substituted Biaryls via Grignard Reagent Cross-Coupling Org. Biomol. Chem 2014, 12, 5586–5589. [DOI] [PubMed] [Google Scholar]; (e) Bastug G; Nolan SP [Pd(IPr*OMe)(cin)Cl] (cin = Cinnamyl): A Versatile Catalyst for C-N and C-C Bond Formation Organometallics 2014, 33, 1253–1258. [Google Scholar]; (f) Tu T; Sun Z; Fang W; Xu M; Zhou Y Robust Acenaphthoimidazolylidene Palladium Complexes: Highly Efficient Catalysts for Suzuki-Miyaura Couplings with Sterically Hindered Substrates Org. Lett 2012, 14, 4250–4253. [DOI] [PubMed] [Google Scholar]; (g) Chartoire A; Lesieur M; Falivene L; Slawin AMZ; Cavallo L; Cazin CSJ; Nolan SP [Pd(IPr*)(cinnamyl)Cl]: An Efficient Pre-Catalyst for the Preparation of Tetra-ortho-substituted Biaryls by Suzuki-Miyaura Cross-Coupling Chem. Eur. J 2012, 18, 4517–4521. [DOI] [PubMed] [Google Scholar]; (h) Li G–Q; Yamamoto Y; Miyaura N Synthesis of Tetra-ortho-Substituted Biaryls Using Aryltriolborates Synlett 2011, 12, 1769–1773. [Google Scholar]; (i) Hoshi T; Nakazawa T; Saitoh I; Mori A; Suzuki T; Sakai J–I; Hagiwara H Biphenylene-Substituted Ruthenocenylphosphine for Suzuki-Miyaura Coupling of Aryl Chlorides Org. Lett 2008, 10, 2063–2066. [DOI] [PubMed] [Google Scholar]; (j) Schaarschmidt D; Grumbt M; Hildebrandt A; Lang H A Planar-Chiral Phosphino(alkenyl)ferrocene for Suzuki-Miyaura C0C Coupling Reactions Eur. J. Org. Chem 2014, 6676–6685. [Google Scholar]; (k) Giannerini M; Hornillos V; Vila C; Fananas-Mastral M; Feringa BL Hindered Aryllithium Reagents as Partners in Palladium-Catalyzed Cross-Coupling: Synthesis of Tri- and Tetra-ortho-Substituted Biaryls under Ambient Conditions Angew. Chem. Int. Ed 2013, 52, 13329–13333. [DOI] [PubMed] [Google Scholar]; (l) Wu L; Drinkel E; Gaggia F; Capolicchio S; Linden A; Falivene L; Cavallo L; Dorta R Room-Temperature Synthesis of Tetra-ortho-Substituted Biaryls by NHC-Catalyzed Suzuki-Miyaura Couplings Chem. Eur. J 2011, 17, 12886–12890. [DOI] [PubMed] [Google Scholar]; (m) Ackermann L; Potukuchi HK; Althammer A; Born R; Mayer P Tetra-orhto-Substituted Biaryls through Palladium-Catalyzed Suzuki-Miyaura Couplings with Diaminochlorophosphine Ligand Org. Lett 2010, 12, 1004–1007. [DOI] [PubMed] [Google Scholar]
- 10. (a).Bringmann G; Gulder T; Gulder TAM; Breuning M Atropselective Total Synthesis of Axially Chiral Biaryl Natural Products Chem. Rev 2011, 111, 563–639. [DOI] [PubMed] [Google Scholar]; (b) Kozlowski MC; Morgan BJ; Linton EC Total Synthesis of Chiral Biaryl Natural Products by Asymmetric Biaryl Coupling Chem. Soc. Rev 2009, 38, 3193–3207. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Zask A; Murphy J; Ellestad GA Biological Stereoselectivity of Atropisomeric Natural Products and Drugs Chirality 2013, 25, 265–274. [DOI] [PubMed] [Google Scholar]; (d) Smyth JE; Butler NM; Keller PA A Twist of Nature–The Significance of Atropisomers in Biological Systems Nat. Prod. Rep 2015, 32, 1562–1583. [DOI] [PubMed] [Google Scholar]; (e) Bringmann G; Price-Mortimer AJ; Keller PA; Gresser MJ; Garmer J; Breuning M Atropselective Synthsis of Axially Chiral Biaryl Compounds Angew. Chem. Int. Ed 2005, 44, 5384–5427. [DOI] [PubMed] [Google Scholar]; (f) Wencel-Delord J; Panossian A; Leroux FR; Colobert F Recent Advances and New Concepts for the Synthesis of Axially Stereoenriched Biaryls Chem. Soc. Rev 2015, 44, 3418–3430. [DOI] [PubMed] [Google Scholar]
- 11. (a).Fandrick KR; Li W; Zhang Y; Tang W; Gao J Rodriguez S; Patel ND; Reeves DC; Wu J–P; Sanyal S; Gonnella N; Qu B; Haddad N; Lorenz JC; Sidhu K; Wang J; Ma S; Grinberg N; Lee H; Tsantrizos Y; Poupart M–A; Busacca CA; Yee NK; Lu BZ; Senanayake CH Concise and Practical Asymmetric Synthesis of a Challenging Atropisomeric HIV Integrase Inhibitor Angew. Chem. Int. Ed 2015, 54, 7144–7148. [DOI] [PubMed] [Google Scholar]; (b) Haddad N; Mangunuru HPR; Fandrick KR; Qu B; Sieber JD; Rodriguez S; Desrosiers J–N; Patel ND; Lee H; Kurouski D; Grinberg N; Yee NK; Song JJ; Senanayake CH Reengineered BI-DIME Ligand Core Based on Computer Modeling to Increase Selectivity in Asymmetric Suzuki-Miyaura Coupling for the Challenging Axially Chiral HIV Integrase Inhibitor Adv. Synth. Catal 2016, 358, 3522–3527. [Google Scholar]
- 12.Reviews on catalytic asymmetric cross-coupling: Baudoin O The Asymmetric Suzuki Coupling Route to Axially chiral Biaryls Eur. J. Org. Chem 2005, 4223–4229.Yang H; Yang X; Tang W Transition-Metal Catalyzed Asymmetric Carbon-Carbon Cross-Coupling with Chiral Ligands Tetrahedron 2016, 72, 6143–6174.Demchuk OM; Kaplon K; Kacka A; Pietrusiewicz KM The Utilization of Chrial Phosphorus Ligands in Atropselective Cross-Coupling Reactions Phosphorous, Sufur, and Silicon 2016, 191, 180–200.Cherney AH; Kadunce NT; Reisman SE Enantioselective and Enantiospecific Transition-Metal-Catalyzed Cross-Coupling Reactions of Organometallic Reagents to Construct C-C Bonds Chem. Rev 2015, 115, 9587–9652.
- 13. (a).Yin JJ; Buchwald SL A Catalytic Asymmetric Suzuki Coupling for the Synthesis of Axiaqlly Chiral Biaryl Compounds J. Am. Chem. Soc 2000, 122, 12051–12052. [Google Scholar]; (b) Shen X; Jones GO; Watson DA; Bhayana B; buchwald SL Enantioselective Synthesis of Axially Chiral Biaryls by the Pd-Catalyzed Suzuki-Miyaura Reaction: Substrate Scope and Quantum Mechanical Investigations J. Am. Chem. Soc 2010, 132, 11278–11287. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Cammidge AN; Crepy KVL The First Asymmetric Suzuki Cross-Coupling Reaction Chem. Commun 2000, 1723–1724. [Google Scholar]; (d) Cammidge AN; Crepy KVL Synthesis of Chiral Binaphthalenes using the Asymmetric Suzuki Reaction Tetrahedron 2004, 60, 4377–4386. [Google Scholar]; (e) Bermejo A; Ros A; Fernandez R; Lassaletta JM C2-Symmetric Bis-Hydrazones as Ligands in the Asymmetric Suzuki-Miyaura Cross-Coupling J. Am. Chem. Soc 2008, 130, 15798–15799. [DOI] [PubMed] [Google Scholar]; (f) Ros A; Estepa B; Bermejo A; Alvarez E; Fernandez R; Lassaletta JM Phosphino Hydrazones as Suitable Ligands in the Asymmetric Suzuki-Miyaura Cross-Coupling J. Org. Chem 2012, 77, 4740–4750. [DOI] [PubMed] [Google Scholar]; (g) Sawai K; Tatumi R; Nakahodo T; Fujihara H Asymmetric Suzuki-Miyaura Coupling Reactions Catalyzed by Chiral Palladium Nanoparticles at Room Temperature Angew. Chem. Int. Ed 2008, 47, 6917–6919. [DOI] [PubMed] [Google Scholar]; (h) Zhang S–S; Wang ZQ; Xu M–H; Lin G–Q Chiral Diene as the Ligand for the Synthesis of Axially chiral Compounds via the Palladium-Catalyzed Suzuki-Miyaura Coupling Reaction Org. Lett 2010, 12, 5546–5549. [DOI] [PubMed] [Google Scholar]; (i) Yamamoto T; Akai Y; Nagata Y; Suginome M Highly Enantioselective synthesis of Axially Chiral biarylphosphonates: Asymmetric Suzuki-Miyaura Coupling Using High-Molecular-Weight, Helically Chiral Polyquinoxaline-Based Phosphines Angew. Chem. Int. Ed 2011, 50, 8844–8847. [DOI] [PubMed] [Google Scholar]; (j) Castanet A–S; Colobert F; Broutin P–E; Obringer M Asymmetric Suzuki Cross-Coupling Reaction: Chirality Reversal Depending on the Palladium-Chiral Phosphine Ratio Tetrahedron: Asymmetry 2002, 13, 659–665. [Google Scholar]; (k) Herrbach A; Marinetti A; Baudoin O; Guenard D; Gueritte F Asymmetric Synthesis of an Axially Chiral Antimitotic biaryl via an Atrop-Enantioselective Suzuki Cross-Coupling J. Org. Chem 2003, 68, 4897–4905. [DOI] [PubMed] [Google Scholar]; (l) Mikami K; Miyamoto T; Hatano M A Highly Efficient Asymmetric Suzuki-Miyaura Coupling Reaction Catalyzed by Cationic Chiral Palladium(II) Complexes Chem. Commun 2004, 2082–2083. [DOI] [PubMed] [Google Scholar]; (m) Sun L; Dai W–M Determination of Absolute Configuration of 2-Methyl-1-(o-tolyl)naphthalene and the Related Axially chiral Biaryls Tetrahedron 2011, 67, 9072–9079. [Google Scholar]; (n) Tang W; Patel ND; Xu G; Xu X; Savoie J; Ma S; Hao M–H; Keshipeddy S; Capacci AG; Wei X; Zhang Y; Gao J; Li W; Rodriguez S; Lu BZ; Yee NK Senanayake CH Efficient Chiral Monophosphorous Ligands for Asymmetric Suzuki-Miyaura Coupling Reactions Org. Lett 2012, 14, 2258–2261. [DOI] [PubMed] [Google Scholar]; (o) Xu G; Fu W; Liu G; Senanayake CH; Tang W Efficient Syntheses of Korupensamines A, B and Michellamine B by Asymmetric Suzuki-Miyaura Coupling Reactions J. Am. Chem. Soc 2014, 136, 570–573. [DOI] [PubMed] [Google Scholar]; (p) Uozumi Y; Matsuura T; Arakawa T; Yamada YM A. Asymmetric Suzuki-Miyaura Coupling in Water with a Chiral Palladium Catalyst Supported on an Amphiphilic Resin Angew. Chem. Int. Ed 2009, 48, 2708–2710. [DOI] [PubMed] [Google Scholar]; (q) Jensen JF; Johannsen M New Air-Stable Planar Chiral Ferrocenyl Monophosphine Ligands: Suzuki Cross-Coupling of Aryl Chlorides and Bromides Org. Lett 2003, 5, 3025–3028. [DOI] [PubMed] [Google Scholar]; (r) Genov M; Almorin A; Espinet P Efficient Synthesis of Chiral 1,1’-Binaphthalenes by the Asymmetric Suzuki-Miyaura Reacdtion: Dramatic Synthetic Improvement by Simple Purification of Naphthylboronic Acids Chem.–Eur. J 2006, 12, 9346–9352. [DOI] [PubMed] [Google Scholar]; (s) Meskova M; Putala M Highly Sterically Hindered Binaphthalene-Based Monophosphane Ligands: Synthesis and Application in Stereoselective Suzuki-Miyaura Reactions Tetrahedron Asymmetry 2013, 24, 894–902. [Google Scholar]; (t) Castillo AB; Perandones BF; Zangrando E; Gladiali S; Godard C; Claver C Pd-Catalysed Asymmetric Suzuki-Miyaura Reactions using Chiral Mono- and Bidentate Phosphorous Ligands J. Organomet. Chem 2013, 743, 31–36. [Google Scholar]; (u) Wang S; Li J; Miao T; Wu W; Li Q; Zhuang Y; Zhou Z; Qiu L Highly Efficient Synthesis of a Class of Novel Chiral-Bridged Atropisomeric Monophosphine Ligands via Simple Desymmetrization and Their Applications in Asymmetric Suzuki-Miyaura Coupling Reaction Org. Lett 2012, 14, 1966–1969. [DOI] [PubMed] [Google Scholar]; (v) Wu W; Wang S; Zhou Y; He Y; Zhuang Y; Li L; Wan P Wang L; Zhou Z; Qiu L Highly Diastereoselective Synthesis of Atropisomeric Bridged P,N-Ligands and Their Applications in Asymmetric Suzuki-Miyaura Coupling Reaction Adv. Synth. Catal 2012, 354, 2395–2402. [Google Scholar]; (w) Zhou Y; Wang S; Wu W; Li Q; He Y; Zhuang Y; Li L; Pang J; Zhou Z; Qiu L Enantioselective Synthesis of Axially Chiral Multifunctionalized Biaryls via Asymmetric Suzuki-Miyaura Coupling Org. Lett 2013, 15, 5508–5511. [DOI] [PubMed] [Google Scholar]; (x) Zhou Y; Zhang X; Liang H; Cao Z; Zhao X; He Y; Wang S; Pang J; Zhou Z Ke Z; Qiu L Enantioselective Synthesis of Axially Chiral Biaryl Monophosphine Oxides via Direct Asymmetric Suzuki Coupling and DFT Investigations of the Enantioselectivity ACS Catalysis 2014, 4, 1309–1397. [Google Scholar]; (y) Xia W; Li Y; Zhou Z; Chen H; Liang H; Yu S; He X; Zhang Y; Pang J; Zhou Z; Qiu L Synthesis of Chiral-Bridged Atropisomeric Monophosphine Ligands with Tunable Dihedral Angles and their Applications in Asymmetric Suzuki-Miyaura Coupling Reactions Adv. Synth Catal 2017, 359, 1656–1662. [Google Scholar]; (z) Chatterjee A; Mallin H; Klehr J; Vallapurackal J; Finke AD; Vera L; Marsh M; Ward TR An Enantioselective Artificial Suzukiase Based on the Biotin-Streptavidin Technology Chem. Sci 2016, 7, 673–677. [DOI] [PMC free article] [PubMed] [Google Scholar]; (aa) Benhamou L; Besnard C; Kundig EP Chrial PEPPSI Complexes: Synthesis, Characterization, and Application in Asymmetric Suzuki-Miyaura Coupling Reactions Organometallics 2014, 33, 260–266. [Google Scholar]; (bb) Li Y; Tang J; Gu J; Wang Q; Sun P; Zhang D Chiral 1,2-Cyclohexane-Bridged Bis-NHC Palladium Catalysts for Asymmetric Suzuki-Miyaura Coupling: Synthesis, Characterization, and Steric Effects on Enantiocontrol Organometallics 2014, 33, 876–884. [Google Scholar]; (cc) Yang X; Xu G; Tang W Efficient Synthesis of Chiral Biaryls via Asymmetric Suzuki-Miyaura Cross-Coupling of ortho-Bromo Aryl Triflates Tetrahedron 2016, 72, 5178–5183. [Google Scholar]
- 14. (a).Genov M; Fuentes B; Espinet P; Pelaz B Asymmetric Negishi Reaction for Sterically Hindered Couplings: Synthesis of Chiral Binaphthalenes Tetrahedron: Asymmetry 2006, 17, 2593–2595. [Google Scholar]; (b) Genov M; Almorin A; Espinet P Microwave Assisted Asymmetric Suzuki-Miyaura Cross-Coupling Reactions: Synthesis of chiral Binaphthalenes Tetrahedron Asymmetry 2007, 18, 625–627. [Google Scholar]
- 15. (a).Hayashi T; Hayashizaki K; Kiyoi T; Ito Y Asymmetric Synthesis Catalyzed by Chiral Ferrocenylphosphine-Transition-Metal Complexes. 6. Practical Asymmetric Synthesis of 1,1’-Binaphthyls via Asymmetric Cross-Coupling with a Chiral [(Alkoxyalkyl)ferrocenyl]monophosphine/Nickel Catalyst J. Am. Chem. Soc 1988, 110, 8153–8156. [Google Scholar]; (b) Hayahsi T; Niizuma S; Kamikawa T; Suzuki N; Uozumi Y Catalytic Asymmetric Synthesis of Axially Chiral Biaryls by Palladium-Catalyzed Enantioposition-Selective Cross-Coupling J. Am. Chem. Soc 1995, 117, 9101–9102. [Google Scholar]
- 16. (a).Tamao K; Sumitani K; Kumada M Selective Carbon-Carbon Bond Formation by Cross-Coupling of Grignard Reagents with Organic Halides. Catalysis by Nickel-Phosphine Complexes. J. Am. Chem. Soc 1972, 94, 4374–4376. [Google Scholar]; (b) Corriu RJP; Masse JP Activation of Grignard Reagents by Transition-metal Complexes. A New and Simple Synthesis of trans-Stilbenes and Polyphenyls. Chem. Commun 1972, 144a–144a. [Google Scholar]; (c) Smith RS; Kochi JK Mechanistic Studies of Iron Catalysis in the Cross Coupling of Alkenyl Halides and Grignard Reagents. J. Org. Chem 1976, 41, 502–509. [Google Scholar]; (d) Adrio J; Carretero JC Functionalized Grignard Reagents in Kumada Cross-Coupling Reactions. ChemCatChem. 2010, 2, 1384–1386. [Google Scholar]; (e) Kambe N; Iwasaki T; Terao J Pd-Catalyzed Cross-Coupling Reactions of Alkyl Halides. Chem. Soc. Rev 2011, 40, 4937–4947. [DOI] [PubMed] [Google Scholar]; (f) Heravi MM; Hajiabbasi P Recent Advances in Kumada-Tamao-Corriu Cross Coupling Reaction Catalyzed by Different Ligands. Montash Chem 2012, 143, 1575–1592. [Google Scholar]
- 17.For an overview of the selectivity determining step in Pd-catalyzed cross-coupling reactions, see: For an overview of the selectivity determining step in Pd-catalyzed cross-coupling reactions, see: Denmark SE; Chang W–TT; Houk KN; Liu P Development of Chiral Bis-Hydrazone Ligands for the Enantioselective Cross-Coupling Reactions of Aryldimethylsilanolates J. Org. Chem 2015, 80, 313–366.
- 18.Reviews: Lipshutz BH; Ghorai S; Cortes-Clerget M The Hydrophobic Effect Applied to Organic Synthesis: Recent Synthetic Chemistry “in Water” Chem. Eur. J 2018, 24, 6672–6695.Demchuk OM; Jasinski R Organophosphorous Ligands: Recent Developments in Design, Synthesis and Application in Environmentally Benign Catalysis Phosphorous, Sulfur, and Silicon 2016, 191, 245–253.Dallinger D; Kappe CO Microwave-Assisted Synthesis in Water as Solvent. Chem. Rev 2007, 107, 2563–2591.Lamlin M; Nassar-Hardy L; Hierso JC; Fouquet E; Felpin FX Recyclable Heterogeneous Palladium Catalysts in Pure Water: Sustainable Developments in Suzuki, Heck, Sonogashira and Tsuji-Trost Reactions. Adv. Synth. Catal 2010, 352, 33–79.Lipshutz BH; Ghorai S “Designer”-Surfactant-Enabled Cross-Couplings in Water at Room Temperature. Aldrichimica Acta 2012, 45, 3–16.
- 19.Selected examples: Patel ND; Rivalti D; Buono FG; Chatterjee A; Qu B; Desrosiers J–N; Rodriguez S; Sieber JD; Haddad N; Fandrick KR; Lee H; Yee NK; Busacca CA; Senanayake CH Effective BI-DIME Ligand for Suzuki-miyaura Cross-Coupling Reactions in Water with 500 ppnm Palladium Loading and Triton X Asian J. Org. Chem 2017, 6, 1285–1291.Leadbeater NE; Marco M Rapid and Amenable Suzuki Coupling Reaction in Water Using Microwave and Conventional Heating J. Org. Chem 2003, 68, 888–892.Amini M; Tarassoli A; Yousefi S; Delsouz-Hafshejani S; Bigdeli M; Salehifar M Suzuki-Miyaura Cross-Coupling Reactions in Water using in situ Generated Palladium(II)-Phosphazane Complexes Chinese Chem. Lett 2014, 25, 166–168.Demchuk OM; Kaplon K; Mazur L; Strzelecka D; Pietrusiewicz KM Readily Available Catalysts for Demanding Suzuki-Miyaura Couplings Under Mild Conditions Tetrahedron 2016, 72, 6668–6677.Li Z; Gelbaum C; Heaner WL IV; Fisk J; Jaganathan A; Holden B; Pollet P; Liotta CL Palladium-Catalyzed Suzuki Reactions in Water with No Added Ligand: Effects of Reaction Scale, Temperature, pH of Aqueous Phase, and Substrate Structure Org. Process Res. Dev 2016, 20, 1489–1499.Bumagin NA; Bykov VV; Beletskaya IP Synthesis of Diaryls from Phenylboric Acin and Aryl Iodides in an Aqueous Medium. Bull. Acad. Sci. USSR, Div. Chem. Sci 1989, 38, 2206–2206.Bumagin NA; Bykov VV Ligandless Palladium Catalyzed Reactions of Arylboronic Acids and Sodium Tetraphenylborate with Aryl Halides in Aqueous Media. Tetrahedron 1997, 53, 14437–14450.Liu C; Zhang YX; Liu N; Qiu JS A Simple and Efficient Approach for the Palladium-Catalyzed Ligand-Free Suzuki Reaction in Water. Green Chem 2012, 14, 2999–3003.Hoffmann I; Blumenroder B; Thumann SON; Dommer S; Schatz J Suzuki Cross-Coupling in Aqueous Media. Green Chem 2015, 17, 3844–3857.Zhou CS; Wang JY; Li LY; Wang RH; Hong MC A Palladium Chelatin Complex of Ionic Water-Soluble Nitrogen-Containing Ligand: the Efficient Precatalyst for Suzuki-Miyaura Reaction in Water. Green Chem. 2011, 13, 2100–2106.Handa S; Wang Y; Gallou F; Lipshutz BH Sustainable Fe-ppm Pd Nanoparticle Catalysis of Suzuki-Miyaura Cross-Couplings in Water. Science 2015, 349, 1087–1091.Botella L; Najera CA Convenient Oxime-Carbapalladacycle-Catalyzed Suzuki Cross-Coulpling of Aryl Chlorides in Water. Angew. Chem. Int. Ed 2002, 41, 179–181.Rohlich C; Wirth AS; Kohler K Suzuki Coupling Reaction in Neat Water as the Solvent: Where in the Biphasic Reaction Mixture Do the Catalytic Reaction Steps Occur? Chem. Eur. J 2012, 18, 15485–15494.Lipshutz BH; Abela AR Micellar Catalysis of Suzuki-Miyaura Cross-Couplings with Heteroaromatics in Water. Org. Lett 2008, 10, 5329–5332.Handa S; Andersson MP; Gallou F; Reilly J; Lipshutz BH HandaPhos: A General Ligand Enabling Sustainable ppm Levels of Palladium-Catalyzed Cross-Couplings in Water at Room Temperature. Angew. Chem. Int. Ed 2016, 55, 4914–4918.
- 20. (a).Breslow R The Principles of and Reasons for Using Water as a Solvent for Green Chemistry In Handbook of Green Chemistry; Li C–J Ed.; Wiley-VCH: Weinheim, Germany, 2010; Vol. 5, pp 1–29. [Google Scholar]; (b) Hailes HC Reaction Solvent Selection: The Potential of Water as a Solvent for Organic Transformations. Org. Process Res. Dev 2007, 11, 114–120. [Google Scholar]; (c) Welton T Solvents an Sustainable Chemistry. Proc. R. Soc. A 2015, 471, 20150502. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Jessop PG Searching for Green Solvents. Green Chem 2011, 12, 1391–1398. [Google Scholar]; (e) Clarke CJ; Tu W–C; Levers O; Brohl A; Hallett JP Green and Sustainable Solvents in Chemical Processes. Chem. Rev 2018, 118, 747–800. [DOI] [PubMed] [Google Scholar]; (f) Simon M–O; Li C–J Green Chemistry Oriented Organic Synthesis in Water. Chem. Soc. Rev 2012, 41, 1415–1427. [DOI] [PubMed] [Google Scholar]; (g) Li C–J; Chen L Organic Chemistry in Water. Chem. Soc. Rev 2006, 35, 68–82. [DOI] [PubMed] [Google Scholar]; (h) Akiya N; Savage PE Roles of Water for Chemical Reaction in High Temperature Water. Chem. Rev 2002, 102, 2725–2750. [DOI] [PubMed] [Google Scholar]; (i) Lindstrom UM Stereoselective Organic Reactions in Water. Chem. Rev 2002, 102, 2751–2772. [DOI] [PubMed] [Google Scholar]
- 21.Despite the recognized potential “green” properties of water for use as the reaction solvent (ref. 19), reactions run in water may or may not actually be “green.” A truly green process employing water as the solvent must be designed correctly to address other important factors related to reaction concentration, waste management, and contaminated water purification/recycling, see: Blackmond DG; Armstrong A; Coombe V; Wells A Water in Organocatalytic Processes: Debunking the Myths. Angew. Chem. Int. Ed 2007, 46, 3798–3800.Ni Y; Holtmann D; Hollmann F How Green is Biocatalysis? To Calculate is to Know ChemCatChem 2014, 6, 930–943.Dominguez de Maria P; Hollmann F On the (Un)greenness of Biocatalysis: Some Challenging Figures and Some Promising Options Front. Microbiol 2015, 6, 1257–1261.Clark JH; Tavener SJ Alternative Solvents: Shades of Green Org. Proc. Res. & Dev 2007, 11, 149–155.
- 22.Two other reports of asymmetric Suzuki-Miyaura cross-coupling in water have been reported (ref. 13z, 13aa); however, these both are performed in the presence of an organic co-solvent.
- 23.Qu B; Haddad N; Rodriguez S; Sieber JD; Desrosiers J–N; Patel ND; Zhang Y; Grinberg N; Lee H; Ma S; Ries UJ; Yee NK; Senanayake CH Ligand-Accelerated Stereoretentive Suzuki-Miyaura Coupling of Unprotected 3,3’-Dibromo-BINOL J. Org. Chem 2016, 81, 745–750. [DOI] [PubMed] [Google Scholar]
- 24.Rodriguez S; Qu B; Haddad N; Reeves D; Tang W; Krishnamurthy D; Senanayake CH Oxaphosphole-Based Monophosphorus Ligands for Palladium-Catalyzed Amination Reactions Adv. Synth. Catal 2011, 353 533–537. [Google Scholar]
- 25.Tang W; Keshipeddy S; Zhang Y; Wei X; Savoie J; Patel ND; Yee NK; Senanayake CH Efficient Monophosphorus Ligands for Palladium-catalyzed Miyaura Borylation Org. Lett 2011, 13, 1366–1369. [DOI] [PubMed] [Google Scholar]
- 26.Fandrick DR; Fandrick KR; Reeves JT; Tan Z; Tang W; Capacci AG; Rodriguez S; Song JJ; Lee H; Yee NK; Senanayake CH Copper Catalyzed Asymmetric Propargylation of Aldehydes J. Am. Chem. Soc 2010, 132, 7600–7601. [DOI] [PubMed] [Google Scholar]
- 27.Fandrick DR; Hart CA; Okafor IS; Mercadante MA; Sanyal S; Masters JT; Sarvestani M; Fandrick KR; Stockdill JL; Grinberg N; Gonnella N; Lee H; Senanayake CH Copper-Catalyzed Asymmetric Propargylation of Cyclic Aldimines. Org. Lett 2016, 18, 6192–6195. [DOI] [PubMed] [Google Scholar]
- 28.Sieber JD; Rivalti DR; Herbage MA; Masters JT; Fandrick KR; Fandrick DR; Haddad N; Lee H; Yee NK; Gupton BF; Senanayake CH Rh-Catalysed Asymmetric Conjugate Addition of Boronic Acids to Nitroalkenes Employing a P-Chiral P,π-Hybrid Ligand Org. Chem. Front 2016, 3, 1149–1153. [Google Scholar]
- 29.Sieber JD; Chennamadhavuni D; Fandrick KR; Qu B; Han ZS; Savoie J; Ma S; Samankumara LP; Grinberg N; Lee H; Song JJ; Senanayake CH Development of New P-Chiral P,π-Dihydrobenzooxaphosphole Hybrid Ligands for Asymmetric Catalysis Org. Lett 2014, 16, 5494–5497. [DOI] [PubMed] [Google Scholar]
- 30.Sieber JD; Angeles-Dunham VV; Chennamadhavuni D; Fandrick DR; Haddad N; Grinberg N; Kurouski D; Lee H; Song JJ; Yee NK; Mattson AE; Senanayake CH Rhodium-Catalyzed Asymmetric Allenylation of Sulfonylimines and Application to the Stereospecific Allylic Allenylation Adv. Synth. Catal 2016, 358, 3062–3068. [Google Scholar]
- 31. (a).Tang W; Qu B; Capacci AG; Rodriguez S; Wei X; Haddad N; Narayanan B; Ma S; Grinberg N Yee NK; Krishnamurthy D; Senanayake CH Novel, Tunable, and Efficient Chiral Bisdihydrobenzooxaphosphole Ligands for Asymmetric Hydogenation Org. Lett 2010, 12, 176–179. [DOI] [PubMed] [Google Scholar]; (b) Tang W; Capacci AG; White A; Ma S; Rodriguez S; Qu B; Savoie J; Patel ND; Wei X; Haddad N; Grinberg N; Yee NK; Krishnamurthy D; Senanayake CH Novel and Efficient Chiral Bisphosphorus Ligands for Rhodium-Catalyzed asymmetric Hydrogenation Org. Lett 2010, 12, 1104–1108. [DOI] [PubMed] [Google Scholar]; (c) Li G; Zatolochnaya OV; Wang X–J; Rodriguez S; Qu B; Sieber JD; Desrosiers J–N; Mangunuru HPR; Biswas S; Rivalti D; Karyakarte S; Grinberg N; Wu L; Lee H; Haddad N; Fandrick DR; Yee NK; Song JJ; Senanayake CH BABIPhos Family of Biaryl Dihydrobenzooxaphosphole Ligands for Asymmetric Hydrogenation Org. Lett 2018, 20, 1725–1729. [DOI] [PubMed] [Google Scholar]
- 32. (a).Qu B; Samankumara LP; Savoie J; Fandrick DR; Haddad N; Wei X; Ma S; Lee H; Rodriguez S; Busacca CA; Yee NK; Song JJ; Senanayake CH Synthesis of Pyridyl-dihydrobenzooxaphosphole Ligands and Their Application in Asymmetric Hydrogenation of Unfunctionalized Alkenes J. Org. Chem 2014, 79, 993–1000. [DOI] [PubMed] [Google Scholar]; b) Qu B; Samankumara LP; Ma S; Fandrick KR; Desrosiers J-N; Rodriguez S; Li Z; Haddad N; Han ZS; McKellop K; Pennino S; Grinberg N; Gonnella NC; Song JJ; Senanayake CH A Mild Dihydrobenzooxaphosphole Oxazoline/Iridium Catalytic System for Asymmetric Hydrogenation of Unfunctionalized Dialins Angew. Chem., Int. Ed 2014, 53, 14428–14432. [DOI] [PubMed] [Google Scholar]
- 33. (a).Wei X; Qu B; Zeng X; Savoie J; Fandrick KR; Desrosiers J–N; Tcyrulnikov S; Marsini MA; Buono FG; Li Z; Yang B–S; Tang W; Haddad N; Gutierrez O; Wang J; Lee H; Ma S; Campbell S; Lorenz JC; Eckhardt M; Himmelsbach F; Peters S; Patel ND; Tan Z; Yee NK; Song JJ; Roshcanger F; Kozlowski MC; Senanayake CH Sequential C-H Arylation and Enantioselective Hydrogenation Enables Ideal Asymmetric Entry to the Indenopiperidine Core of an 11β-HSD-1 Inhibitor J. Am. Chem. Soc 2016, 138, 15473–15481. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Qu B; Mangunuru HPR; Wei X; Fandrick KR; Desrosiers J–N; Sieber JD; Kurouski D; Haddad N; Samankumara LP; Lee H; Savoie J; Ma S; Grinberg N; Sarvestani M; Yee NK; Song JJ; Senanayake CH Synthesis of Enantioenriched 2-Alkyl Piperidine Derivatives through Asymmetric Reduction of Pyridinium Salts Org. Lett 2016, 18, 4920–4923. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Qu B; Mangunuru HPR; Tcyrulnikov S; Rivalti D; Zatolochnaya OV; Kurouski D; Radomkit S; Biswas S; Karyakarte S; Fandrick KR; Sieber JD; Rodriguez S; Desrosiers J–N; Haddad N; McKellop K; Pennino S; Lee H; Yee NK; Song JJ; Kozlowlsi MC; Senanayake CH Enantioselective synthesis of α-(Hetero)aryl Piperidines through Asymmetric Hydrogenation of Pyridinium Salts and Its Mechanistic Insights Org. Lett 2018, 20, 1333–1337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Tan R; Zheng X; Qu B; Sader A; Fandrick KR; Senanayake CH; Zhang X Tunable P-Chiral Bisdihydrobenzooxaphosphole Ligands for Enantioselective Hydroformylation Org. Lett 2016, 18, 3346–3349. [DOI] [PubMed] [Google Scholar]
- 35. (a).Rodriguez S; Qu B; Fandrick KR; Buono F; Haddad N; Xu Y; Herbage MA; Zeng X; Ma S; Grinberg N; Lee H; Han ZS; Yee NK; Senanayake CH Amine-Tunable Ruthenium Catalysts for Asymmetric Reduction of Ketones Adv. Synth. Catal 2014, 356, 301–307. [Google Scholar]; (b) Zatolochnaya OV; Rodriguez S; Zhang Y; Lao KS; Tcyrulnikov S; Li G; Wang X–J; Qu B; Biswas S; Mangunuru HPR; Rivalti D; Sieber JD; Desrosier J–N; Leung JC; Grinber N; Lee H; Haddad N; Yee NK; Song JJ; Kozlowski MC; Senanayake CH Copper-Catalyzed Asymmetric Hydrogenation of 2-Substituted Ketones via Dynamic Kinetic Resolution Chem. Sci 2018, 9, 4505–4508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Thomas AA; Zahrt AF; Delaney CP; Denmark SE Elucidating the Role of the Boronic Esters in the Suzuki-Miyaura Reaction: Structural, Kinetic, and Computational Investigations J. Am. Chem. Soc 2018, 140, 4401–4416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Molander GA; Ellis N Organotrifluoroborates: Protected Boronic Acids That Expand the Versatility of the Suzuki Coupling Reaction Acc. Chem. Res 2007, 40, 275–286. [DOI] [PubMed] [Google Scholar]
- 38.Manolikakes SM; Ellwart M; Stathakis CI; Knochel P Air-Stable Solid Aryl and Heteroaryl Organozinc Pivalates: syntheses and Applications in Organic Synthesis Chem. Eur. J 2014, 20, 12289–12297. [DOI] [PubMed] [Google Scholar]
- 39.Reviews: García-Melchor M; Braga AAC, Lledós A, Ujaque G, Maseras F Computational Perspective on Pd-Catalyzed C-C Cross-Coupling Reaction Mechanisms. Acc. Chem. Res, 2013, 46, 2626–2634.Sperger T, Sanhueza IA, Kalvet I, Schoenebeck F Computational Studies of Synthetically Relevant Homogeneous Organometallic Catalysis Involving Ni, Pd, Ir, and Rh: An Overview of Commonly Employed DFT Methods and Mechanistic Insights. Chem. Rev, 2015, 115, 9532–9586.Busch M; Wodrich MD; Corminboeuf C A Generalized Picture of C-C Cross-Coupling. ACS Catalysis 2017, 7, 5643–5653
- 40. (a).Casares JA; Espinet P; Fuentes B; Salas G Insights into the Mechanism of the Negishi Reaction: ZnRX versus ZnR2 Reagents J. Am. Chem. Soc 2007, 129, 3508–3509. [DOI] [PubMed] [Google Scholar]; (b) Fuentes B; García-Melchor M; Lledós A; Maseras F; Casares JA; Ujaque G; Espinet P Palladium Round Trip in the Negishi Couploing of trans-[PdMeCl(PMePh2)2] with ZnMeCl: An Experimental and DFT Study of the Transmetalation Step Chem. Eur. J 2010, 16, 8596–8599. [DOI] [PubMed] [Google Scholar]; (c) Liu Q; Lan Y; Liu J; Li G; Wu Y –D.; Lei, A. Revelaing a Second Transetalation Step in the Negishi Coupling and Its Competition with Reductive Elimination: Improvement in the Interpretation of the Mechanism of Biaryl Syntheses J. Am. Chem. Soc 2009, 131, 10201–10210. [DOI] [PubMed] [Google Scholar]
- 41.Mlynarski SN; Schuster CH; Morken JP Asymmetric Synthesis from Terminal Alkenes by Cascades of Diboration and Cross-Coupling Nature, 2014, 505, 386–390. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. (a).Thomas AA; Denmark SE Pre-transmetalation Intermediates in the Suzuki-Miyaura Reaction Revealed: The Missing Link Science, 2016, 352, 329–332. [DOI] [PubMed] [Google Scholar]; (b) Thomas AA; Wang H; Zahrt AF; Denmark SE Structural, Kinetic, and Computational Characterization of the Elusive Arylpalladium(II)boronate Complexes in the Suzuki-Miyaura Reaction J. Am. Chem. Soc 2017, 139, 3805–3821. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Carrow BP; Hartwig JF Distinguishing Between Pathways for Transmetalation in Suzuki-Miyaura Reactions J. Am. Chem. Soc 2011, 133, 2116–2119. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Lennox AJJ; Lloyd-Jones GC Transmetalation in the Suzuki-Miyaura Coupling: The Fork in the Trail Angew. Chem., Int. Ed 2013, 52, 7362. [DOI] [PubMed] [Google Scholar]; (e) Amatore C; Jutand A; Le Duc G Kinetic Data for the Transmetalation/Reductive Elimination in Palladium-Catalyzed Suzuki-Miyaura Reactions: Unexpected Triple Role of Hydroxide Ions Used as Base Chem. Eur. J 2011, 17, 2492–2503. [DOI] [PubMed] [Google Scholar]; (f) Jover J; Fey N; Purdie M; Lloyd-Jones GC; Harvey JN A Computational Study of Phosphine Ligands Effects in Suzuki-Miyaura Coupling. J. Mol. Catal. A 2010, 324, 39–47. [Google Scholar]
- 43.Jasinki R; Demchuk OM; Babyuk D A Quantum-Chemical DFT Approach to Elucidation of the Chirality Transfer Mechanism of the Enantioselective Suzuki-Miyaura Cross-Coupling Reaction J. Chem 2017, Article ID 3617527, 12 pages, 10.1155/2017/3617527 (Accessed Jun 6, 2018). [DOI] [Google Scholar]
- 44. (a).Ortuno MA; Lledos A; Maseras F; Ujaque G The Transmetalation Process in Suzuki-Miyaura Reactions: Calculations Indicate Lower Barrier via Boronate Intermediate. ChemCatChem. 2014, 6, 3132–3138. [Google Scholar]; (b) Sicre C, Braga AAC, Maseras F, Cid MM Mechanistic Insights into the Transmetalation Step of a Suzuki-Miyaura Reaction of 2(4)-Bromopyridines: Characterization of an Intermediate. Tetrahedron, 2008, 64, 7437–7443. [Google Scholar]; (c) Braga AAC; Morgon NH; Ujaque G; Maseras F Computational Characterization of the Role of the Base in the Suzuki-Miyaura Cross-Coupling Reaction. J. Am. Chem. Soc 2005, 127, 9298–9307. [DOI] [PubMed] [Google Scholar]; (d) Braga AAC; Morgon NH; Ujaque G; Lledos A; Maseras FJ Organomet. Chem 2006, 691, 4459–4466. [Google Scholar]; (e) Braga AA; Ujaque G; Maseras F A DFT Study of the Full Catalytic Cycle of the Suzuki-Miyaura Cross-Coupling on a Model System. Organometallics 2006, 25, 3647–3658. [Google Scholar]
- 45.Miyaura N Cross-Coupling Reaction of Organoboron Compounds via Base-Assisted Transmetalation to Palladium(II) Complexes. J. Organomet. Chem 2002, 653, 54–57. [Google Scholar]
- 46.Zeebe RE; Sanyal A; Ortiz JD; Wolf-Gladrow DA A Theoretical Study of the Kinetics of the Boric Acid-Borate Equilibrium in Seawater Marine Chemistry, 2001, 73, 113–124. [Google Scholar]
- 47.Davis ME, Davis RJ (2003) Fundamentals of chemical reaction engineering. McGraw-Hill Higher Education, New York, NY. [Google Scholar]
- 48. (a).Guan Y, Wheeler SE Automated Quantum Mechanical Predictions of Enantioselectivity in Rhodium-Catalyzed Asymmetric Hydrogenation Angew. Chem. Int. Ed 2017, 56, 9101–9105. [DOI] [PubMed] [Google Scholar]; (b) Santoro S, Kalek M, Huang G, Himo F Elucidation of Mechanisms and Selectivities of Metal-Catalyzed Reactions using quantum Chemical Metholdology Acc. Chem. Res 2016, 49, 1006–1018. [DOI] [PubMed] [Google Scholar]
- 49. (a).Desiraju G R.S. T The Weak Hydrogen Bond In Structural Chemistry and Biology; Oxford University Press: New York, NY, 1999. [Google Scholar]; (b) Steiner T; Desiraju GR Chem. Commun 1998, 891–892. [Google Scholar]; (c) Jeffrey GA An Introduction to Hydrogen Bonding; Oxford University Press: New York, NY, 1997. [Google Scholar]; (d) Taylor R; Kennard O Crystallographic evidence for the Existence of C-H…O, C-H…N, and C-H…Cl Hydrogen bonds J. Am. Chem. Soc 1982, 104, 5063–5070. [Google Scholar]; (e) Jeffrey GA; Maluszynska H A Survey of Hydrogen Bond Geometries in the Crystal Structures of Amino Acids Int. J. Biol. Macromol 1982, 4, 173–185. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
