Reinvestigation of the Substitutions Reaction of Stereogenic

Oct 18, 2017 - Nucleophilic substitutions at P centers are of high importance in biological processes and asymmetric synthesis. However, detailed stud...
0 downloads 14 Views 1MB Size
Article pubs.acs.org/joc

Reinvestigation of the Substitutions Reaction of Stereogenic Phosphoryl Compounds: Stereochemistry, Mechanism, and Applications Li-Juan Liu,†,§ Wei-Min Wang,†,§ Lan Yao,†,§ Fan-Jie Meng,† Yong-Ming Sun,† Hao Xu,† Zhong-Yuan Xu,† Qiang Li,† Chang-Qiu Zhao,*,† and Li-Biao Han*,‡ †

College of Chemistry and Chemical Engineering, Liaocheng University, Liaocheng, Shandong 252059, China National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba, Ibaraki 305-8565, Japan



S Supporting Information *

ABSTRACT: Nucleophilic substitutions at P centers are of high importance in biological processes and asymmetric synthesis. However, detailed studies on this topic are rare. P-Stereogenic compounds containing P−Cl, P−O, and P−S bonds were diastereoselectively prepared and then used to study the substitution of Cl, O, and S at phosphorus centers with organometallic reagents. It was proposed that with alkynyl metallic reagents an SN2-like mechanism (route A1) and a Berry pseudorotation (BPR) of pentacoordinated phosphorus intermediates (route B1) were involved and afforded P-inverted and P-retained products, respectively. The Pinverted conversion of a P−Cl functionality to a P−C functionality can be controlled by either the temperature or the order of addition of the starting materials. The introduction of a P−Cl bond using an alkyl metallic reagent proceeded through routes A2 and A2′. At higher temperatures, P-inverted products were predominantly afforded via SN2-like route A2. At lower temperatures, bis-substituted products were formed via route A2′ and cleavage of the P−O bond. The P−S bonds were accompanied by the epimerization of the starting materials, triggered by the alkylthio anion, via route C. The epimerization can be suppressed by the use of a poorly soluble magnesium alkylthiolate, and the P-retained compounds will be formed as the major products via route B3 and BPR of the intermediates.



INTRODUCTION Phosphorus is one of the elements that is essential for life. Pcontaining substances such as DNA, RNA, DIPP, phosphopeptide, and adenosine triphosphate (ATP) are fundamental to all living things. P-involved reactions such as phosphorylation and the formation and cleavage of P−O bonds represent vital biological processes.1−4 These reactions are also involved in the metabolism and degradation of agricultural pesticides, herbicides, insecticides, and neurotoxins.5 In addition to their involvement in biological processes, Pcontaining compounds are widely used in synthetic chemistry, not only as phosphine ligands of transition-metal catalysts6 but also as organocatalysts that are solely responsible for promoting organic reactions.7 Many chemists are engaged in elucidating the mechanisms of P-centered reactions and developing new methods for the generation of various P-containing compounds. Because of the unique properties of the stereogenic phosphorus center, chiral phosphines have an important role in asymmetric synthesis. It is expected that compared to their Cstereogenic analogues, P- or P,C-stereogenic phosphines might show better asymmetric induction because the stereogenic center of the catalyst is closer to the active site.8 Several recent © 2017 American Chemical Society

reviews have documented the preparation and applications of chiral phosphines.9 Stereogenic phosphorus compounds can be generated via catalytic asymmetric reactions in high yields. However, the stereoselectivities are often not sufficient.10 Another strategy for forming P-chiral centers is the conversion of P-stereogenic starting materials.11 However, a d-orbital is usually involved in P-centered reactions via the formation of sp3d- or sp2d2hybridized pentacoordinated phosphorus derivatives, which leads to variations in the mechanism and stereochemistry. The stereoselective conversions of P-stereogenic centers are therefore quite challenging. For example, early examples of substitutions of alkylthio groups on phosphorus afforded Pretention and P-inversion products,12 with methyl and methoxy anions as attacking reagents, respectively.13,14 As reported by Imamoto and co-workers, the introduction of a P−Br bond afforded P-inversion and P-retention products with alkynyl and aliphatic alkyl lithium reagents, respectively.15 However, Han Received: May 30, 2017 Published: October 18, 2017 11990

DOI: 10.1021/acs.joc.7b01326 J. Org. Chem. 2017, 82, 11990−12002

Article

The Journal of Organic Chemistry reported that similar introduction of P−Cl bond with alkyl metallic reagents gave P-inversion products.16 Substitutions at P centers are proposed to occur via SN2-like and Berry pseudorotation (BPR) mechanisms to afford the Pinversion and P-retention products, respectively.17 However, detailed and thorough investigations on this topic, including the tuning of the stereoselectivity, are rarely reported. Considering that P-centered substitutions are important for both understanding the biological processes of living things and obtaining optically pure P-stereogenic substances,9 we are actively investigating this topic, as well as the controllable and bidirectional conversion of the same starting material to the RP or SP products. For phosphorus derivatives that contain more than one P− heteroatom bond, the nucleophile can attack the phosphorus from different directions. For example, as seen in Chart 1, the

We found that sulfuryl chloride was a better chlorinating reagent. In ether, the chlorination of 1 was complete within several hours. After removing all of the solvent and volatile substances, 2 was obtained in excellent yield and highly optical purity (Scheme 1). Comparison to the NMR spectrum of 2 Scheme 1. Preparation of 2 and 3 from 1

obtained from the reaction with copper chloride16 confirmed that the product of chlorination with sulfuryl chloride was in the SP configuration and that the stereochemistry of P was retained. The method avoided the usage of the heavy metal copper, and the byproducts were easy to remove. Compound 2 slowly epimerized at room temperature; however, when stored at low temperature, it remained unchanged for several weeks. The chloride of 2 was substituted with methanol to afford Omethyl O-menthyl phenylphosphinate (3). Direct phosphorylation of methanol with 1 in the presence of carbon tetrachloride also afforded 3 via 2 as the putative intermediate (Scheme 1).16 We found that when methanol was added to the solution of 1 and triethylamine in carbon tetrachloride, two stereoisomers, 3 and 3′, were formed in an 80:20 ratio, which was improved to >99:1 when carbon tetrachloride was added to the solution of 1 and triethylamine in methanol. The SPconfiguration of 3 was confirmed by comparison of its spectroscopic data to that of the reported compound, and it was further confirmed through the P-inverted substitution of a chloride with an alcohol.16 S-Alkyl phosphonothioates 4 were stereospecifically prepared by two methods. The reaction of 1 with elemental sulfur followed by S-alkylation with alkyl halides afforded S-aliphatic 4. S-Aromatic 4 was stereospecifically obtained from the direct reaction of 1 with a diaryl disulfide (Scheme 2). Both reactions were confirmed to proceed via P-retained mechanisms by X-ray crystallography.20

Chart 1. Possible Routes for the Substitution of Various P− Heteroatom Bonds with a Nucleophile

nucleophile can attack the phosphorus opposite heteroatom X or Y. In each case, the direct SN2-like substitution (routes A and A′) or indirect substitution after BPR (routes B and B′) could occur and form the retained or inverted products, respectively. The attacking preference and the preference for the BPR mechanism were controlled by the features of the nucleophile, X and Y. In this paper, phosphorus species with P−Cl, P−O, and P−S bonds were stereospecifically prepared from RP(−)-menthyl phenylphosphinate 1. Employing these compounds, substitutions with alkynyl, alkyl, and aryl groups were examined. Routes A and B, including their group selectivity, stereoselectivity, and mechanism, were investigated. The formation of P-stereogenic substances with R or S configurations could be controlled by altering the reaction conditions.

Scheme 2. Preparation of 4 from 1



RESULTS AND DISCUSSION Part I. Preparation of P−Heteroatom Species. Compounds containing phosphorus−heteroatom bonds, such as P− Cl, P−O, P−S, and P−N species, are often used as precursors in the preparation of many substances including drugs and pesticides5 and as intermediates in organic synthesis.16,17c However, the P-stereogenic analogues are rarely studied. An early example of the chlorination of RP-1/SP-1′18 with Nchlorosuccinimide afforded a diastereomeric mixture of Omenthyl phenylphosphonochloridate 2/2′.19 The recently reported chlorination of RP-1 with copper(II) chloride gave optically pure 2.16

Part II. Introduction of a P−Cl Bond Using Alkynyl Metallic Reagents. The introduction of a P−Cl bond in 2 with phenylethynyllithium (5a) was examined first. In most cases, a mixture of two stereoisomers of O-menthyl phenyl(phenylethynyl) phosphinate (6a/6a′) was formed. Similar to what is shown in Chart 1, the pathways to form 6a and 6a′ were the SN2-like route A1 and BPR route B1, respectively (Scheme 3). In route A1, 5a attacked the phosphorus opposite 11991

DOI: 10.1021/acs.joc.7b01326 J. Org. Chem. 2017, 82, 11990−12002

Article

The Journal of Organic Chemistry

the chloride preferentially takes an apical position over an alkoxy group. In addition, groups generally attack and leave from the apical positions because the bonds at these positions are relatively longer and weaker. Akiba reported the selectivity for P−O over P−C bonds in pentacoordinated phosphoranes. The P−O bonds were more reactive toward nucleophilic attack because the reacting LUMO is a low-lying σ*P−O orbital, whereas the corresponding orbital of the P−C bond is a highlying σ*P−C.17c−e,22 Similar results can explain the preference of route A1/B1 in the attack of 2 by 5a. Unlike 5a, when aliphatic 1-hexynyllithium 5b was used in excess at 0 °C, 6b/6b′ was obtained with a poor dr. The ratio of 6b/6b′ could be improved by either conducting the reaction at −20 °C or using alkynyl magnesium as the nucleophile. In these cases, we believe that the competition between route A1 and B1 was lessened. For the reactions with alkynylmagnesium reagents 10, the formation of 7 or 8 was likely influenced by the interaction or coordination with magnesium. Thus, the preference for route A1 or B1 was changed. Meanwhile, the reactive aggregation state of organolithiums is different from that of Grignard reagents; the latter may also react via electron transfer mechanisms or through Lewis acid mediated pathways. When the scope of the reaction was expanded to include the use of 10, various derivatives of 6 were prepared in excellent yields and stereoselectivities (Scheme 5).21

Scheme 3. Proposed Mechanism for the Reaction of 2 with 5a

the chloride via transition state 7, which resulted in the formation of 6a after cleavage of the P−Cl bond. Backside attack opposite the menthoxy formed intermediate 8, which was converted to 9 via a BPR (route B1) and generated 6a′ via a P-retained mechanism upon chloride loss.22 The RP configuration of 6a, as well as the P-inverted SN2-like route A1, was confirmed by X-ray crystallographic analysis. If the reaction is carried out at −45 °C or 2 is added to the solution of 5a at 0 °C, then route A1 will be the dominant pathway. When the reaction proceeded via this route, 6a is formed in >99:1 dr. When 2 and 5a were mixed at −80 °C then rapidly warmed to room temperature or the reaction was carried out at −20 °C, 6a and 6a′ were both formed (Scheme 4).21

Scheme 5. Preparation of 6 via the Reaction of 2 with 5

Part III. Introduction of P−Cl and P−O Bonds with Alkyl Metallic Reagents. Alkyl metallic reagents are stronger nucleophiles and have more steric hindrance than alkynyl metallic reagents. As we will discuss below, these two factors resulted in the obvious differences observed for the substitution with alkyl metallic reagents. Ethyl magnesium bromide, 11b, reacted with a 50:50 mixture of (SP)-2/(RP)-2′ and produced 12b/12b′ in a ratio of approximately 50:50 based on the peaks at 45.19 and 44.20 ppm in the 31P NMR spectrum. When optically pure 2 (>99:1) was used, the major peak at 45.19 ppm (98:2 compared to the peak at 44.20 ppm) was assigned as P-retained SP-12b (vide infra). During the reactions, diethyl phenylphosphine oxide, 13b, was detected in which both the menthoxy and chloride groups had been displaced by ethyl groups (Table 1). At −30 °C, 13b was the major product (entry 1). When 2 and 11b were stirred at −20 °C for 30 min followed by stirring at room temperature for several hours, the ratio of 12b/13b was found to be 82:18. Stirring the reaction longer at −20 °C increased the content of 13b (entries 3−5). At 0 °C for 5 h, 12b/13b were produced in an 84:16 ratio, and 12b was formed in 98:2 dr (12b/12b′). SP-12b was obviously formed from a P-inverted SN2-like pathway, which was assigned as route A2. It appeared that 13b was formed via the further displacement of menthoxy from 12b in a similar manner to what has been reported for the substitution of O-menthoxy with methyllithium or methylmagnesium iodide (Scheme 6).14d,23 However, as shown in Table 1, even at elevated temperatures, excess ethylmagnesium

Scheme 4. Reaction of 2 with 5a under Various Conditions

Route A1 is thought to occur prior to route B1 because chloride is a better leaving group than menthoxy. At low temperatures, such as −80 °C, both routes A1 and B1 are sluggish. At −45 °C, only 6a was slowly generated via route A1. At higher temperatures, such as −20 °C, the reaction can start to proceed through route B1, which can generate 6a′. The stoichiometry of the two starting materials also influenced the reaction mechanism. For example, at 0 °C, excess 5a rapidly consumed 2 via route A1, which prevented the reaction from proceeding through route B. When 2 was in slight excess, both route B1 and route A1 were active. In some cases in which 6a′ was formed significantly, L(−)-menthol was detected by proton NMR spectroscopy. As seen in Scheme 3, other than in the BPR route B1 to form 6a′, 8 could be converted to L-(−)-menthol via direct P−O bond cleavage,21 which was assigned as route A1′. According to Berry’s pseudorotation theory,17 highly electronegative atoms tend to occupy the apical positions of a trigonal bipyramidal structure of a pentacoordinated phosphorus. Thus, 11992

DOI: 10.1021/acs.joc.7b01326 J. Org. Chem. 2017, 82, 11990−12002

Article

The Journal of Organic Chemistry Table 1. Reaction of 2 with Ethylmagnesium Bromide

entry

temp (°C)/time (h) −30/6; 0/5; −20/0.5; −20/0.5; −20/4.5;

1 2 3 4 5

12b/13ba,b

12b/12b′

11:89 84:16 82:18 68:32 54:46

98:2 98:2 98:2c 55:45d

rt/7 rt/6 rt/6 rt/5 rt/17

Scheme 7. Proposed Mechanism for the Reaction of 2 with 11b

a In a typical procedure, 2 equiv of 11b was added to 1 equiv of 2 at low temperature, and then the mixture was stirred at room temperature. bThe ratio was estimated on the basis of the integration of the menthoxy and free menthol protons in the 1H NMR spectrum. c The order of addition was reversed compared to the typical procedure. dEtMgBr (3 equiv) and 2/2′ (50:50) were used.

Because 11b is a stronger base than the sp-hybridized alkynyl anion,15 it should more favor attack opposite the oxygen instead of opposite the chloride, and it should lead to facile cleavage of either the P−O or P−Cl bond. The competition between route A2 and route A2′, which occurs via transition states 15 and 16, respectively, resulted in the formation of a significant amount of 13b. Meanwhile, the facile cleavage of the P−O bond suppressed the conversion of 16 to 17 followed by the formation of 12b′ via BPR, which resulted in an excellent dr for 12b/12b′. As seen in Table 1, higher temperatures (0 °C) favored route A2, and lower temperatures (−30 °C) favored route A2′. The temperature dependence could be attributed to the steric hindrance associated with the ethyl group and better leaving group ability of chloride over menthoxy. As seen in Scheme 7, 11b can approach the pentacoordinated phosphorus from the two apical positions to form 15 and 16. In 15, the ethyl group was perpendicular to the bulky menthoxy. The gauche interactions between the groups increased the energy of this configuration such that the conversion of 15 to 12b (route A2) represented a large drop in energy (Figure 1). Thus, 15 was

Scheme 6. Comparison of the Substitution of Menthoxy Groups with Various Organometallic Reagents

bromide cannot convert (SP)-12b or (RP)-12b′ to 13b. In a separate experiment, no menthoxy substitution was observed when O-menthyl cyclohexylphenylphosphinate (12e, Table 2) Table 2. Substitution of 2 with Organometallic Reagents To Afford 12 run

RM

Conditions

yielda (%)

dra

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

MeMgI EtMgBr n-BuMgBr i-PrMgBr i-PrMgBr c-HexMgBr s-BuMgBr s-BuLi t-BuLi BnMgCl allylMgBr p-BrC6H4MgBr p-BrC6H4MgBr p-ClC6H4MgBr o-MeC6H4Li

°C to rt, 9 h °C to rt, 9 h °C to rt, 6 h °C to rt, 5 h rt, 5h 0 °C to rt, 10 h rt, 8h 0 °C, 0.3 h 0 °C, 0.5 h 0 °C to rt, 4 h 0 °C to rt, 8 h rt, 8h 0 °C to rt, 10 h 0 °C to rt, 10 h −80 °C to rt, 10 h

12a, 84 12b, 84 12c, 77 12d, 18 12d, 54 12e, 59 12f, 14 12f, 55 12g, 95 12h, 99 12i, 99 12j, 74 12j, 75 12k, 94 12l, 56

90/10 98/2 98/2 98/2 >99/1 98.5/1.5b 99/1 88.5/11.5 96/4b >99/1 97/3 28/72 11/89b 10/90b 1.5/98.5b

0 0 0 0

Figure 1. Kinetic conversion of 16 to 14 and thermodynamic conversion of 15 to 12b.

a

The yields and dr (SP-12/RP-12′) were determined by 31P{1H} NMR spectroscopy. bAfter isolation, all compounds showed >99:1 or 99:1 dr at room temperature. The reaction of cyclohexylmagnesium bromide gave 12e in 59% yield and 98.5:1.5 dr, and 30% of unreacted 2 was recovered. The reaction with sec-butyl Grignard reagent gave 12f in low yield. When s-BuLi was used, the yield was improved to 55% (entries 7 and 8). The reaction of t-BuLi with 2 at 0 °C for 0.5 h afforded 12g in 95% yield and 96:4 dr. The reactions of allyl and benzyl Grignard reagents afforded 12h and 12i, respectively, in good yields and dr (entries 10 and 11). The organolithium reagents were more reactive in the substitution, perhaps because they tend to aggregate less than Grignard reagents. At room temperature, p-bromophenyl Grignard reagent afforded 12j in 28:72 dr, which was improved to 11:89 at 0 °C (entries 12 and 13). Compound 12k was similarly prepared in 90% yield and 10:90 dr from p-chlorophenyl Grignard reagent. For 12j and 12k, the two stereoisomers were difficult to distinguish by 31P NMR. The peaks of 12j/12j′ and 12k/ 12k′ were shifted by 0.038 and 0.045 ppm, respectively, by comparison to the peaks of pure 12/12′ prepared from the mixture of 2/2′. This allowed us to confirm the structures of 12j and 12k. For o-tolyl Grignard, conducting the reaction at −80 °C improved the dr of 12l/12l′ to 1.5/98.5. After isolation, optically pure 12j, 12k, and 12l were obtained. P-inverted route A2 was confirmed by different analyses. At first, our previously reported free-radical addition of 1 to cyclohexene afforded the P-retained RP stereoisomer 12e′,14b,24 which had a peak at 43.87 ppm in its 31P NMR spectrum instead of the peak observed at 44.96 ppm for the current product. Thus, 12e was determined to be in the S P configuration. Additionally, the X-ray diffraction results of 12h (R = Bn) confirmed that its stereocenter is in the SP configuration (Figure 2). On the basis of the P-inverted mechanism, 12j to 12l were assigned to be in the RP configuration according the Cahn−Ingold−Prelog priority rules. The above procedure was successfully expanded to allow the introduction of two stereogenic phosphorus atoms into a molecule. For example, the dilithium reagent, which was prepared from diaryl dibromide and n-BuLi, was reacted with 2 at −80 °C to stereoselectively afford bisphosphinate 12m in 58% yield and excellent dr (Scheme 8). These compounds are useful precursors for the preparation of P-stereogenic chelating P−N−P ligands.25 The substitution of an oxygen-containing moiety with a metallic reagent was examined using 3 as the substrate. When

Figure 2. Structure of 12h from X-ray crystallographic analysis (hydrogen atoms have been removed for clarity).

Scheme 8. Preparation of Bisphosphinate 12m

heated with 11b in toluene at 50−60 °C for 40 h, 60% of 3 was consumed. The new single peak at 44.20 ppm in the 31P NMR spectrum was different than what was observed in the spectrum of 12b (45.19 ppm). Compared to the aforementioned NMR spectrum of 12b/12b′ (45.19 and 44.20 ppm), the new peak was confidently attributed to 12b′. Although peaks for 12b′ and unconsumed 3 were present, peaks from 13b and 12b were not observed. Methyl- or n-butylmagnesium halide reacted with 3 to afford 12a′ and 12c′, respectively, in good dr (Scheme 9). The Scheme 9. Substitution of 3 with Grignard Reagents To Afford 12′

displacement of menthoxy was not detected. Similar substitutions with secondary alkyl or aryl Grignard reagents did not occur even after prolonged heating at 70−80 °C. When nBuLi was used, the reaction mixture became complicated. The differences in the outcomes of the reactions with organolithiums versus the reactions with Grignard reagents were probably because magnesium is more prone to aggregation than lithium. Additionally, as we will discuss in part IV, the poor solubility of the magnesium salts probably helped increase the stereoselectivity. The formation of RP-12′ indicated the substitution of the methoxy group on 3 proceeded via a P-inverted mechanism.14d,23,26 Compared to 2, 3 was less reactive but showed better group selectivity. We hypothesized that the alkyl group attacked the phosphorus position opposite the better leaving group (methoxy) in an SN2-like cleavage of the P−OMe bond to form 12′. The double inversion of the configuration of 2 to 3 11994

DOI: 10.1021/acs.joc.7b01326 J. Org. Chem. 2017, 82, 11990−12002

Article

The Journal of Organic Chemistry

and poor dr; meanwhile a substantial proportion of 4c was epimerized (part A of Scheme 10). In a separate experiment, 4c (>99:1) was fully racemized to 4c/4c′ (50:50) when heated with 0.02 equiv of lithium phenylthiolate at 60 °C for 16 h. For comparison, no change was detected during the heating of a solution 4c with no additives (part B of Scheme 10). Other bases such as menthoxylithium similarly resulted in racemization. When MeMgI was used, the 2:98 dr of 12a/12a′ indicated the formation of 12a via direct attack opposite sulfur, which occurs via transition state 18, was insignificant (route A3 of Scheme 11). Compound 12a′ was obviously generated through a pathway that retains the stereochemistry, assigned as route B3, in which the methyl attacked opposite the oxygen to form 19. The BPR of 19 and P−S bond cleavage of 20 afforded 12a′. In 19, loss of the menthoxy group and subsequent displacement of the methylthio moiety afforded 13a (route A3′). The methylthio anion, an extra nucleophile generated as a byproduct, can reversibly attack 4a (route C), leading to its epimerization. Mislow reported the P-inverted cleavage of the P−S bond by an alkyloxy anion12,16 via the SN2-like attack of RO opposite sulfur. The alkylthiolate anion was thought to behave in a manner similar to the alkyloxy anion, attacking 4a via an SN2-like and P-inverted mechanism. As seen in Scheme 10A, the excellent dr obtained for the substitution with MeMgI can be attributed to the poor solubility of magnesium alkylthiolate; it precipitated from the reaction mixture and was therefore unable to cause epimerization. In contrast, the corresponding lithium salt was soluble, leading the reaction to proceed through route C. In fact, a white cloudy mixture was formed by the addition of MeMgI to 4c, whereas a clear solution was observed for the reaction with methyllithium. When the above procedure was utilized for the preparation of various 12′ derivatives, derivatives of 4 with smaller alkylthio groups should be chosen because of the poorer solubility of the generated alkylthiolate salt. However, a significant amount of 13 was formed in the reaction of 4a (via route A3′). Therefore, the addition of 4b (R1 = Et) to excess Grignard reagents 11 was adopted as the preferred procedure. The presence of excess 11 competed and suppressed route C, ensuring substitution occurred prior to epimerization. The substitutions of 4b with various metallic reagents are summarized in Table 3. For primary alkyl, allyl, and benzyl

and then 3 to 12′ resulted in total retention of the original phosphorus configuration. Part IV. Substitution of P−Sulfur Functional Groups with Alkyl Metallic Reagents. S-Alkyl phosphonothioates 4 contain both oxygen and sulfur. The main difference in their electronegativities and leaving group abilities greatly influenced the substitution of 4. As we discuss below, when alkyl groups were used as nucleophiles, the sulfur was displaced, and the main products had retained the stereochemistry of the starting material; however, the reaction would be influenced by epimerization of 4. Mislow and co-workers briefly reported the substitution of 4a/4a′ (R1 = Me, 35:65) with methyl magnesium bromide via a P-retained mechanism, which formed a mixture of 12a′/12a in the same ratio.27 We found the reaction of 4b/4b′ (R1 = Et, 50:50) with 11b afforded 12b/12b′ in a manner similar to that of the reaction of 2/2′. When optically pure 4b (>99:1) was used, the RP stereoisomer 12b′ was predominantly formed in 2:98 dr. Similarly, 4c (R1 = Ph) reacted with methylmagnesium iodide to afford 12a′ in 99 12c′, 96 12d′, 57 12f′, 88 12g′, 91 12h′, 97 12i′, 83 12j′, 54 12l′, 50 12n′, 62

99:1. Mp: 154−155 °C. 1H NMR (400 MHz, CDCl3) δ (ppm): 7.70−7.58 (m, 2H), 7.53− 7.44 (m, 1H), 7.43−7.32 (m, 2H), 7.26−7.08 (m, 5H), 4.01 (ddd, J = 14.8 Hz, 11.0 Hz, 4.5 Hz, 1H), 3.31 (d, J = 17.7 Hz, 2H), 2.18 (d, J = 12.2 Hz, 1H), 1.99−1.79 (m, 1H), 1.59 (d, J = 11.9 Hz, 2H), 1.44− 1.23 (m, 2H), 1.15 (dd, J = 23.5 Hz, 11.8 Hz, 2H), 0.86 (d, J = 6.4 Hz, 3H), 0.82 (d, J = 7.0 Hz, 3H), 0.36 (d, J = 6.9 Hz, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 38.41. 13C NMR (101 MHz, CDCl3) δ (ppm): 131.8 (d, J = 2.7 Hz), 131.7 (d, J = 9.4 Hz), 131.6, 130.8 (d, J = 126.1 Hz), 129.9 (d, J = 5.8 Hz), 128.1 (d, J = 2.8 Hz), 127.9 (d, J = 12.6 Hz), 126.4 (d, J = 3.3 Hz), 48.7 (d, J = 6.5 Hz), 43.5, 39.1, 38.2, 33.9, 31.4, 25.2, 22.5, 21.8, 20.9, 15.1. IR (KBr) ν/cm−1: 3054, 2869, 1638, 1456, 1214, 1016. [α]D30 = −18 (c 0.054, CH2Cl2). (SP)-O-Menthyl Allylphenylphosphinate (12i). The product was prepared from the reaction of 2 with allylmagnesium bromide (1.2 equiv, commercially available 1.0 mol/L solution in diethyl ether) at 0 °C for 2 h then at room temperature for 6 h, and purified by preparative TLC (chloroform/AcOEt (v/v) = 20/1 as eluent), as a colorless oil, yield 90% (95 mg), dr 98:2. 1H NMR (400 MHz, CDCl3) δ (ppm): 7.81−7.68 (m, 2H), 7.56−7.33 (m, 3H), 5.82−5.62 (m, 1H), 5.15−4.95 (m, 2H), 4.29−4.32 (m, 0.01H), 4.15−3.86 (m, 0.99H), 2.97−2.60 (m, 2H), 2.32 (d, J = 12.0 Hz, 1H), 2.22 (s, 1H), 2.01−1.77 (m, 1H), 1.69−1.44 (m, 2H), 1.43−1.08 (m, 2H), 0.87 (d, J = 6.5 Hz, 3H), 0.85−0.80 (m, 2H), 0.79 (d, J = 7.0 Hz, 3H), 0.33 (d, J = 6.9 Hz, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 39.02 (s, 98%), 38.47 (s, 2%). 13C NMR (101 MHz, CDCl3) δ (ppm): 132.2 (d, J = 2.6 Hz), 131.8 (d, J = 9.5 Hz), 128.4 (d, J = 12.5 Hz), 127.8 (d, J = 9.0 Hz), 120.3 (d, J = 13.0 Hz), 48.9 (d, J = 6.4 Hz), 44.1, 37.3, 36.3, 34.3, 31.8, 25.6, 22.9, 22.2, 21.2, 15.4. HRMS (ESI+): calcd for C19H30O2P [M + H]+ 321.1983, found 321.1984. IR (KBr) ν/cm−1: 3056, 1637, 1455, 1219, 1014. [α]D30 = −64 (c 0.019, CH2Cl2). (RP)-O-Menthyl p-Bromophenylphenylphosphinate (12j). The product was prepared from the reaction of 2 to pbromophenylmagnesium bromide (1.2 equiv) at 0 °C for 2 h then at room temperature for 8 h and purified by preparative TLC (chloroform/AcOEt (v/v) = 20/1 as eluent), as a colorless oil, yield 70% (100 mg), dr < 1:99. 1H NMR (400 MHz, CDCl3) δ (ppm): 7.79−7.59 (m, 4H), 7.59−7.50 (m, 2H), 7.50−7.32 (m, 3H), 4.23 (ddd, J = 14.9 Hz, 10.9, 4.5, 1H), 2.24 (s, 1H), 2.18−1.96 (m, 2H), 1.70−1.51 (m, 2H), 1.40 (dd, J = 27.5 Hz, 16.0 Hz, 2H), 1.17 (dd, J = 23.3 Hz, 12.0 Hz, 1H), 1.02−0.89 (m, 1H), 0.84 (d, J = 7.1 Hz, 3H), 0.82 (d, J = 6.5 Hz, 3H), 0.51 (d, J = 6.9 Hz, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 28.96. 13C NMR (101 MHz, CDCl3) δ (ppm): 133.3 (d, J = 10.8 Hz), 133.0 (d, J = 28.6 Hz), 132.2 (d, J = 2.7 Hz), 131.8 (d, J = 13.2 Hz), 131.7, 131.5, 128.6 (d, J = 13.2 Hz), 127.2 (d, J = 3.6 Hz), 49.1 (d, J = 6.3 Hz), 43.8, 34.3, 31.8, 25.8, 22.9, 22.1, 21.3, 15.5. HRMS (ESI+): calcd for C22H28BrO2PNa [M + Na]+ 457.0908, found 457.0897. IR (KBr) ν/cm−1: 2955, 1644, 1456, 1229, 1127, 1008. [α]D30 = −40 (c 0.479, CH2Cl2). (RP)-O-Menthyl p-Chlorophenylphenylphosphinate (12k). The product was prepared from the reaction of 2 to pchlorophenylmagnesium bromide (1.2 equiv) at 0 °C for 2 h and then at room temperature for 8 h, and purified by preparative TLC (chloroform/AcOEt (v/v) = 20/1 as eluent), as colorless oil, yield 88% (113 mg), dr < 1:99. 1H NMR (400 MHz, CDCl3) δ (ppm): 7.88−7.63 (m, 4H), 7.56−7.30 (m, 5H), 4.32−4.15 (m, 1H), 2.19− 11998

DOI: 10.1021/acs.joc.7b01326 J. Org. Chem. 2017, 82, 11990−12002

Article

The Journal of Organic Chemistry

iodide at 70−80 °C for 4 h and purified by preparative TLC (hexane/ AcOEt (v/v) = 5/1 as eluent), as a white solid, yield 66% (124 mg), dr < 1:99. Mp: 80−81 °C. 1H NMR (400 MHz, CDCl3) δ (ppm): 7.80 (dd, J = 12.1 Hz, 7.0 Hz, 2H), 7.62−7.36 (m, 3H), 4.34−4.17 (m, 1H), 2.28−2.06 (m, 1H), 1.81 (d, J = 11.8 Hz, 1H), 1.66 (s, 1H), 1.62 (s, 4H), 1.35 (d, J = 10.3 Hz, 2H), 1.26 (s, 2H), 1.02 (d, J = 11.9 Hz, 1H), 0.96 (d, J = 7.0 Hz, 3H), 0.89 (d, J = 6.9 Hz, 3H), 0.77 (d, J = 6.5 Hz, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 40.41. 13C NMR (101 MHz, CDCl3) δ (ppm): 132.3, 131.3 (d, J = 10.2 Hz), 128.8 (d, J = 12.7 Hz), 49.3 (d, J = 6.0 Hz), 43.7, 34.6, 31.9, 31.3, 30.2, 26.4, 23.5, 21.9 (d, J = 81.3 Hz), 17.5, 16.4 (d, J = 15.4 Hz). (RP)-O-Menthyl Ethylphenylphosphinate (12b′). The product was prepared from the reaction of 3 with ethylmagnesium bromide at 50−60 °C for 40 h and purified by preparative TLC (chloroform/ AcOEt (v/v) = 5/1 as eluent), as a colorless oil, yield 55% (108 mg), dr < 1:99. 1H NMR (400 MHz, CDCl3) δ (ppm): 7.77 (dd, J = 10.9 Hz, 7.7, 2H), 7.59−7.36 (m, 3H), 4.28 (ddd, J = 18.3 Hz, 11.0 Hz, 4.5 Hz, 1H), 2.35−2.13 (m, 1H), 2.04−1.69 (m, 3H), 1.68−1.51 (m, 2H), 1.45−1.17 (m, 2H), 1.14−0.97 (m, 5H), 0.94 (d, J = 7.2 Hz, 3H), 0.87 (d, J = 6.9 Hz, 3H), 0.84−0.76 (m, 1H), 0.72 (d, J = 6.5 Hz, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 44.20. 13C NMR (101 MHz, CDCl3) δ (ppm): 133.3 (d, J = 123.1 Hz), 132.2 (d, J = 2.7 Hz), 131.8 (d, J = 9.6 Hz), 128.7 (d, J = 12.3 Hz), 49.3 (d, J = 6.0 Hz), 43.6, 34.6, 31.9, 26.2, 24.4, 23.4, 22.3, 21.5, 16.2, 6.3 (d, J = 4.5 Hz). HRMS (ESI +): calcd for C18H30O2P [M + H]+ 309.1973, found 309.1978. [α]D30 = −54 (c 0.341, CH2Cl2). (RP)-O-Menthyl n-Butylphenylphosphinate (12c′). The product was prepared from the reaction of 3 with n-butylmagnesium bromide at 70−80 °C for 5 h and purified by preparative TLC (hexane/AcOEt (v/v) = 2.5/1 as eluent), as a white solid, yield 60% (129 mg), dr 2:98. Mp: 42−43 °C. 1H NMR (400 MHz, CDCl3) δ (ppm): 7.86−7.65 (m, 2H), 7.44 (ddd, J = 11.2 Hz, 10.6 Hz, 6.1 Hz, 3H), 4.24 (ddd, J = 15.0 Hz, 11.1 Hz, 4.5 Hz, 1H), 2.31−2.11 (m, 2H), 1.96−1.65 (m, 3H), 1.66−1.42 (m, 3H), 1.43−1.12 (m, 6H), 1.03−0.94 (m, 1H), 0.92 (d, J = 7.0 Hz, 3H), 0.87−0.73 (m, 6H), 0.70 (d, J = 6.5 Hz, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 42.99. 13C NMR (101 MHz, CDCl3) δ (ppm): 133.5 (d, J = 123.1 Hz), 132.1 (d, J = 2.7 Hz), 131.6 (d, J = 9.8 Hz), 128.6 (d, J = 12.3 Hz), 49.2 (d, J = 6.0 Hz), 43.5, 34.4, 31.7, 30.9, 29.9, 25.9, 24.3, 24.1, 24.1, 23.2, 22.2, 21.4, 16.0, 13.8. HRMS (ESI+): calcd for C20H34O2P [M + H]+ 337.2277, found 337.2291. [α]D30 = −63 (c 0.119, CH2Cl2). Reaction of 4a with MeMgI or with MeLi. To the solution of 4c (39 mg, 0.1 mmol) in THF (2 mL) was added a solution of methylmagnesium iodide (prepared from methyl iodide and magnesium in ether, as 1 mmol/mL solution, 0.1 mL) dropwise at room temperature. After the clear solution was stirred for 10 min, cloudiness developed. The mixture was stirred at the same temperature for 3 h, quenched with saturated aqueous ammonium chloride, and extracted with ethyl ether. The organic layer was washed with water and dried over magnesium sulfate. After the solvents were removed in vacuo, the residue was analyzed with NMR spectroscopy. The two single peaks at 40.05 (major) and 40.74 (trace, less than 1%) ppm were observed, which were assigned as 12a′ and 12a, respectively. The reaction of methyllithium with 4c was carried out following a similar procedure. Commercially available MeLi (1.6 M solution in ether, 0.063 mL) was used. When the solution was stirred at room temperature for 10 min, no cloudy was developed. After quenched, except for the peaks of 12a and 12a′ (as above), the peaks at 39.77 and 39.57 ppm were observed in the ratio of 50:50, which was assigned as 4c and 4c′.20 The epimerization of 4c with base. To the solution of 4c (39 mg, 0.1 mmol) in THF (2 mL), the solution of lithium phenylthiolate (0.1 M in THF, 0.02 mL), which was prepared from benzenethiol (110 mg, 1 mmol) and n-butyllithium (1.57 mol/L solution in hexand, 0.64 mL) in THF (9.5 mL). The clear solution was heated at 60 °C for 16 h. After being cooled to room temperature, the saturated aqueous ammonium chloride solution was added, and the mixture was extracted with ether (3 × 20 mL). The combined organic layer was dried over anhydrous magnesium sulfate. After the solvents were removed in vacuo, the residue was analyzed by NMR spectroscopy.

1.99 (m, 2H), 1.71−1.51 (m, 2H), 1.49−1.25 (m, 2H), 1.27−1.07 (m, 2H), 0.99−0.89 (m, 1H), 0.85 (d, J = 7.1 Hz, 3H), 0.82 (d, J = 6.5 Hz, 3H), 0.51 (d, J = 6.9 Hz, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 28.84. 13C NMR (101 MHz, CDCl3) δ (ppm): 138.1, 132.8 (d, J = 10.8 Hz), 132.4 (d, J = 50.3 Hz), 131.7 (d, J = 2.7 Hz), 131.3 (d, J = 10.2 Hz), 131.3 (d, J = 10.2 Hz), 128.4 (d, J = 13.7 Hz), 128.1 (d, J = 13.2 Hz), 48.6 (d, J = 6.3 Hz), 43.3, 33.8, 31.3, 25.3, 22.5, 21.7, 20.8, 15.1. HRMS (ESI+): calcd for C22H28ClO2PNa [M + Na]+ 413.1413, found 413.1408. [α]D30 = −76 (c 0.519, CH2Cl2). (RP)-O-Menthyl o-Methylphenylphenylphosphinate (12l). The product was prepared from the reaction of 2 with omethylphenyllithium, which was prepared from o-methylphenyl bromide (0.16 mL, 1.36 mmol) and n-BuLi (2.2 M solution in hexane, 0.62 mL, 1.36 mmol) at −30 °C, at −80 °C to room temperature for 10 h, as a colorless oil, yield 50% (61 mg), dr 1:99. The crude product was purified by preparative TLC (chloroform/ AcOEt (v/v) = 20:1 as eluent). 1H NMR (400 MHz, CDCl3) δ (ppm): 8.03 (dd, J = 12.4 Hz, 7.7 Hz, 1H), 7.68 (dd, J = 12.1 Hz, 7.6 Hz, 2H), 7.54−7.35 (m, 3H), 7.34−7.10 (m, 3H), 4.34−4.20 (m, 1H), 2.33 (s, 2H), 2.23 (d, J = 12.1 Hz, 1H), 2.10−1.93 (m, 1H), 1.62 (d, J = 11.0 Hz, 2H), 1.53−1.30 (m, 2H), 1.32−1.14 (m, 3H), 1.01−0.88 (m, 1H), 0.85 (d, J = 4.2 Hz, 3H), 0.83 (d, J = 3.6 Hz, 3H), 0.44 (d, J = 6.9 Hz, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 25.67. 13C NMR (101 MHz, CDCl3) δ (ppm): 142.3 (d, J = 12.1 Hz), 133.4 (d, J = 7.8 Hz), 132.6 (d, J = 2.6 Hz), 132.1 (d, J = 2.8 Hz), 131.9, 131.9 (d, J = 3.0 Hz), 131.8, 128.7 (d, J = 13.1 Hz), 125.8 (d, J = 12.1 Hz), 49.5 (d, J = 6.6 Hz), 44.0, 34.6, 32.1, 30.2, 26.1, 23.3, 22.4, 21.8, 21.7, 21.6, 15.7. HRMS (ESI+): calcd for C23H32O2P [M + H]+ 371.2140, found 371.2141. [α]D30 = −156 (c 0.050, CH2Cl2). Preparation of Bidentate Phenylphosphinate (RP,RP)-12m from 2 and Lithium Reagent. To the solution of di(obromophenyl)-tert-butylamine (126.4 mg, 0.288 mmol) in diethyl ether (3 mL) was added n-BuLi (2.2 M solution in hexane, 0.26 mL, 0.576 mmol) at −30 °C with stirring. After completion of the addition, the solution was allowed to stir at room temperature for 2 h. Then this solution was added to the solution of 2 (199.3 mg, 0.63 mmol, 2.2 equiv) in ether (2 mL) at −80 °C. The mixture was stirred while being warmed to room temperature. After being quenched with aq saturated ammonium chloride, the mixture was extracted with dichloromethane (3 × 20 mL) and the combined organic layer was dried over magnesium sulfate. After removal of the solvents, the residue was purified with prepared TLC (hexane/AcOEt = 5:1) to afford a colorless oil as product in 60% yield (140 mg) and 99:1. 1H NMR (400 MHz, CDCl3) δ (ppm): 8.05 (m, 1H), 7.74 (m, 2H), 7.48−7.37 (m, 4H), 7.29−7.20 (m, 2H), 4.40−4.31 (m, 1H), 2.82- 2.65 (m, 2H), 2.13 (m, 1H), 1.89 (m, 1H), 1.66 (m, 2H), 1.43 (m, 1H), 1.29 (m, 1H), 1.13 (m, 1H), 1.08 (m, nH), 0.96 (m, 1H), 0.86 (d, 3H), 0.79 (m, 6H), 0.64 (d, 3H). 31P NMR (162 MHz, CDCl3) δ (ppm): 29.02. 13C NMR (101 MHz, CDCl3) δ (ppm): 141.9 (d, J = 12.1 Hz), 135.1 (s), 133.7 (s), 133.1 (d, J = 8.0 Hz), 132.7−132.1 (m), 131.8 (d, J = 2.8 Hz), 131.7 (s), 131.6 (s), 131.5 (s), 131.4 (s), 130.0 (s), 128.3 (dd, J = 24.0, 13.3 Hz), 128.3 (dd, J = 24.0, 13.3 Hz), 49.1 (d, J = 5.3 Hz), 43.6, 34.3, 31.7, 31.2, 25.6, 23.1, 22.2, 21.4, 15.7. HRMS (ESI+): calcd for

The two singlet peaks located at 39.77 and 39.57 ppm in a ratio of ca.50:50. For the optically pure 4c, only the peak at 39.77 was observed.20 Preparation of (RP)-O-Menthyl Alkylphenylphosphinates 12′ from 4. Typical Procedure: Preparation of (RP)-O-Menthyl Ethylphenylphosphinate (12b′). To a solution of ethylmagnesium bromide (1 mol/L solution in ether, 0.35 mL, 0.35 mmol) was added a solution of 4b (51 mg, 0.15 mmol) in diethyl ether (1 mL) under an atmosphere of nitrogen. The mixture was stirred at room temperature for 24 h. After the reaction was completed, as monitored by TLC, the mixture was quenched with aqueous ammonium chloride and extracted with ether. The combined organic phase was washed with sodium carbonate, followed with water. After drying over anhydrous magnesium sulfate and evaporation of solvent in vacuo, the residue was purified by preparative TLC (silica gel, chloroform/ethyl acetate 6:1 as eluent) to afford colorless oil in 91% yield (42 mg) and 99:1. 1H NMR (400 MHz, CDCl3) δ = 8.03 (m, 1H), 7.74 (m, 2H), 7.50−7.38 (m, 4H), 7.28 (m, 1H), 7.16 (m, 1H), 4.36 (m, 1H), 2.32 (s, 3H), 2.13 (m, 1H), 1.89 (bd, 1H), 1.65 (m, 2H), 1.45 (m, 1H), 1.35 (m, 1H), 1.14 (m, 1H), 0.98 (m, 1H), 0.87 (d, 3H), 0.83 (m, 1H), 0.65 (d, 2.94H), 0.40 (d, 0.06H). 31P NMR (162 MHz, CDCl3) δ 29.30. 13C NMR (101 MHz, CDCl3) δ = 142.0 (d, J = 125.9), 135.1 (d, J = 2.7), 133.8, 131.6, 132.3 (d, J = 5.8), 131.8, 131.6, 130.0, 128.5, 123.0, 124.5, 49.14, 43.6, 34.3, 31.8, 31.1, 25.8, 23.0, 22.2, 21.5, 21.5, 21.4, 15.5. Anal. Calcd for C23H31O2P: C, 74.57; H, 8.43. Found: C, 74.48; H, 8.50. [α]D30 = −43 (c 0.210, CH2Cl2).



(5) (a) Thompson, C. M.; Berkman, C. E.; Ryu, S.; Jackson, J. A.; Quinn, D. A.; Larsen, A. Rev. Pestic. Toxicol. 1993, 2, 133−148. (b) Yang, Y. C.; Baker, J. A.; Ward, J. R. Chem. Rev. 1992, 92, 1729− 1743. (c) Battershill, J. M.; Edwards, P. M.; Johnson, M. K. Food Chem. Toxicol. 2004, 42 (8), 1279−1285. (6) (a) Quin, L. D. A Guide to Organophosphorus Chemistry; WileyInterscience: New York, 2000. (b) Zhou, Q.-L. Privileged Chiral Ligands and Catalysts; Wiley-VCH: Weinheim, 2011; p 6. (c) Busacca, C. A.; Senanayake, C. H. The use of New Phosphines as Powerf ul Tools in Asymmetric Synthesis of Biologically Active Compounds. In Comprehensive Chirality; Carreira, E. M., Yamamoto, H., Eds.; Elsevier: Amsterdam, 2012; pp 167−216. (7) (a) Darcel, C.; Uziel, J.; Jugé, S. In Phosphorous Ligands in Asymmetric Catalysis; Börner, A., Ed.; Wiley-VCH: Weinheim, 2008; Vol. 3 pp 1211−1233. (b) Kamer, P. C. J.; Leeuwen, P. N. M. C. V. In Phosphours (III) Ligands in Homogeneous Catalysis: Design and Synthesis; Wiley & Sons, Ltd., 2012. (8) (a) Cockburn, N.; Karimi, E.; Tam, W. J. Org. Chem. 2009, 74, 5762−5765. (b) Kalek, M.; Johansson, T.; Jezowska, M.; Stawinski, J. Org. Lett. 2010, 12, 4702−4704. (c) Zhang, X.; Huang, K.; Hou, G.; Cao, B.; Zhang, X. Angew. Chem., Int. Ed. 2010, 49, 6421−6424. (d) Burk, M. J. Acc. Chem. Res. 2000, 33, 363−372. (e) Glueck, D. S. Chem. - Eur. J. 2008, 14, 7108−7117. (9) Recent reviews for preparation of P or C stereogenic phosphines: (a) Wauters, I.; Debrouwer, W.; Stevens, C. V. Beilstein J. Org. Chem. 2014, 10, 1064−1096. (b) Chew, R. J.; Leung, P.-H. Chem. Rec. 2016, 16, 141−158. (c) Koshti, V.; Gaikwad, S.; Chikkali, S. H. Coord. Chem. Rev. 2014, 265, 52−73. (d) Pullarkat, S. A. Synthesis 2016, 48, 493− 503. (e) Glueck, D. S. Synlett 2007, 2007, 2627−2634. (f) Glueck, D. S. Coord. Chem. Rev. 2008, 252, 2171−2179. (10) (a) Grabulosa, A.; Muller, J. G. Coord. Chem. Rev. 2007, 251, 25−90. (b) Harvey, J. S.; Gouverneur, V. Chem. Commun. 2010, 46, 7477−7485. (c) Kolodiazhnyi, O. Tetrahedron: Asymmetry 2012, 23, 1−46. (d) Knowles, W. S. Angew. Chem., Int. Ed. 2002, 41, 1998− 2007. (e) Huang, Y.; Pullarkat, S. A.; Li, Y.-X.; Leung, P.-H. Inorg. Chem. 2012, 51, 2533−2540. (f) Carlone, A.; Bartoli, G.; Bosco, M.; Sambri, L.; Melchiorre, P. Angew. Chem., Int. Ed. 2007, 46, 4504−4506. (g) Butti, P.; Rochat, R.; Sadow, A.; Togni, A. Angew. Chem., Int. Ed. 2008, 47, 4878−4881. (h) Chan, V. S.; Chiu, M.; Bergman, R. G.; Toste, F. D. J. Am. Chem. Soc. 2009, 131, 6021−6032. (i) Anderson, B. J.; Guino-o, M. A.; Glueck, D. S.; Golen, J. A.; DiPasquale, A. G.; Liable-Sands, L. M.; Rheingold, A. L. Org. Lett. 2008, 10, 4425−4428. (j) Julienne, D.; Delacroix, O.; Lohier, J.-F.; Oliveira-Santos, J. S. d.; Gaumont, A.-C. Eur. J. Inorg. Chem. 2011, 2011, 2489−2498. (11) (a) Imamoto, T.; Yashio, K.; Crépy, K. V. L; Katagiri, K.; Takahashi, H.; Kouchi, M.; Gridnev, I. D. Organometallics 2006, 25, 908−914. (b) Kumaraswamy, G.; Venkata Rao, G.; RamaKrishna, G. Synlett 2006, 2006, 1122−1124. (c) Imamoto, T.; Saitoh, Y.; Koide, A.; Ogura, T.; Yoshida, K. Angew. Chem., Int. Ed. 2007, 46, 8636− 8639. (d) Cieslikiewicz, M.; Bouet, A.; Jugé, S.; Toffano, M.; Bayardon, J.; West, C.; Lewinski, K.; Gillaizeau, I. Eur. J. Org. Chem. 2012, 2012, 1101−1106. (e) Oshiki, T.; Imamoto, T. J. Am. Chem. Soc. 1992, 114, 3975−3977. (f) Al-Masum, M.; Livinghouse, T. Tetrahedron Lett. 1999, 40, 7731−7734. (12) Mislow, K.; Donohue, J.; Mandel, N.; Farnham, W. B.; Murray, R. K.; Benschop, H. P. J. Am. Chem. Soc. 1971, 93, 3792−3793. (13) (a) Hall, C. R.; Inch, T. D.; Peacock, G.; Pottage, C.; Williams, N. E. J. Chem. Soc., Perkin Trans. 1 1984, 669−674. (b) Aaron, H. S.; Uyeda, R. T.; Frack, H. F.; Miller, J. I. J. Am. Chem. Soc. 1962, 84, 617−621. (c) Michalski, J.; Okruszek, A.; Stec, W. J. Chem. Soc. D 1970, 22, 1495−1497. (d) Farnham, W. B.; Mislow, K.; Mandel, N.; Donohue, J. J. Chem. Soc., Chem. Commun. 1972, 120−121. (e) Hall, C. R.; Inch, T. D.; Pottage, C.; Williams, N. E. Tetrahedron 1985, 41, 4909−4917. (14) (a) Benschop, H. P.; Van den Berg, G. R.; Boter, H. L. Recl. Trav. Chim. Pays-Bas 1968, 87, 387−395. (b) Benschop, H. P.; Platenburg, D. H. J. M. J. Chem. Soc. D 1970, 1098−1099. (c) Van den Berg, G. R.; Platenburg, D. H. J. M.; Benschop, H. P. Recl. Trav. Chim. Pays-Bas 1972, 91, 929−934. (d) Farnham, W. B., Jr.; Murray, R. K.;

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.joc.7b01326. 1 H, 31P, and 13C NMR spectra (PDF) X-ray crystallographic data for compound 12h (CIF) X-ray crystallographic data for compound 12h′ (CIF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Chang-Qiu Zhao: 0000-0002-9016-8151 Li-Biao Han: 0000-0001-5566-9017 Author Contributions §

L.-J.L., W.-M.W., and L.Y. contributed equally to this work.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We acknowledge the financial support of the Natural Science Foundation of China (Grant No. 20772055) and the Natural Science Foundation of Shandong Province (Grant No. ZR2016BM18).



REFERENCES

(1) (a) Todd, L. Where there is life, there is phosphorus. Science and Scientists; Japan Sci Press: Tokyo, 1981. (b) Newton, A. C. Chem. Rev. 2001, 101, 2353−2364. (c) Johnson, L. N.; Lewis, R. J. Chem. Rev. 2001, 101, 2209−2242. (d) Saito, H. Chem. Rev. 2001, 101, 2497− 2509. (e) Qvit, N.; Disatnik, M.-H.; Sho, E.; Mochly-Rosen, D. J. Am. Chem. Soc. 2016, 138, 7626−7635. (2) Haskell, M. D.; Slack, J. K.; Parsons, J. T.; Parsons, S. J. Parsons. Chem. Rev. 2001, 101, 2425−2440. (3) Camden, A. J.; Walsh, S. M.; Suk; Silverman, S. K. Biochemistry 2016, 55, 2671−2676. (4) (a) Sasaki, M. In Chirality in Agrochemicals; Kurihara, N., Miyamoto, J., Eds.; Wiley & Sons: Chichester, 1998; pp 85−139. (b) Kuo, L. Y.; Baker, D. C.; Dortignacq, A. K.; Dill, K. M. Organometallics 2013, 32, 4759−4765. (c) Dhar, B. B.; Edwards, D. R.; Brown, R. S. Inorg. Chem. 2011, 50, 3071−3077. (d) Bromberg, L.; Pomerantz, N.; Schreuder-Gibson, H.; Hatton, T. A. Ind. Eng. Chem. Res. 2014, 53, 18761−18774. (e) Onyido, I.; Swierczek, K.; Purcell, J.; Hengge, A. C. J. Am. Chem. Soc. 2005, 127, 7703−7711. 12001

DOI: 10.1021/acs.joc.7b01326 J. Org. Chem. 2017, 82, 11990−12002

Article

The Journal of Organic Chemistry Mislow, K. J. Am. Chem. Soc. 1970, 92, 5809−5810. (e) Mislow, K.; Donohue, J.; Mandel, N.; Farnham, W. B.; Murray, R. K.; Benschop, H. P. J. Am. Chem. Soc. 1971, 93, 3792−3793. (f) DeBruin, K. F.; Johnson, D. M. J. Chem. Soc., Chem. Commun. 1975, 753−754. (g) Moriyama, M.; Bentrude, W. G. J. Am. Chem. Soc. 1983, 105, 4727−4733. (h) Daniel, K. A.; Kopff, L. A.; Patterson, E. V. J. Phys. Org. Chem. 2008, 21, 321−328. (15) Imamoto obtained excellent diastereoselectivity by carrying out the substitution at low temperature. Imamoto, T.; Saitoh, Y.; Koide, A.; Ogura, T.; Yoshida, K. Angew. Chem., Int. Ed. 2007, 46, 8636− 8639. (16) (a) Zhou, Y.; Wang, G.; Saga, Y.; Shen, R.; Goto, M.; Zhao, Y.; Han, L.-B. J. Org. Chem. 2010, 75, 7924−7927. (b) Wang, G.; Shen, R.; Xu, Q.; Goto, M.; Zhao, Y.; Han, L.-B. J. Org. Chem. 2010, 75, 3890−3892. (17) (a) Berry, R. S. J. Chem. Phys. 1960, 32, 933−938. (b) Kuo, L. Y.; Glazier, S. K. Inorg. Chem. 2012, 51, 328−335. (c) Debruin, K. E.; Tang, C.-l. W.; Johnson, D. M.; Wilde, R. L. J. Am. Chem. Soc. 1989, 111, 5871−5879. (d) Seckute, J.; Menke, J. L.; Emnett, R. J.; Patterson, E. V.; Cramer, C. J. J. Org. Chem. 2005, 70, 8649−8660. (e) Menke, J. L.; Patterson, E. V. J. Mol. Struct.: THEOCHEM 2007, 811, 281−291. (18) (a) Wang, W.-M.; Liu, L.-J.; Zhao, C.-Q.; Han, L.-B. Eur. J. Org. Chem. 2015, 2015, 2342−2345. (b) Berger, O.; Montchamp, J.-L. Angew. Chem., Int. Ed. 2013, 52, 11377−11380. (19) For an early preparation of diastereomerically mixture of 2/2′ from chlorination of an RP/SP mixture of 1/1′ with N-chlorosuccinimide, see: Corriu, R. J. P.; Lanneau, G. F.; Leclercq, D. J. Chem. Soc., Chem. Commun. 1978, 104−105. (20) Wang, W.-M.; Liu, L.-J.; Yao, L.; Meng, F.-J.; Sun, Y.-M.; Zhao, C.-Q.; Xu, Q.; Han, L.-B. J. Org. Chem. 2016, 81, 6843−6847. (21) Yao, L.; Liu, L.-J.; Xu, Z.-Y.; Nie, S.-Z.; Xiao, X.-Q.; Zhao, C.-Q. Org. Biomol. Chem. 2016, 14, 1702−1706. (22) Matsukawa, S.; Kojima, S.; Kajiyama, K.; Yamamoto, Y.; Akiba, K.-Y.; Re, S.; Nagase, S. J. Am. Chem. Soc. 2002, 124, 13154−13170. (23) Korpiun, O.; Lewis, R. A.; Chickos, J.; Mislow, K. J. Am. Chem. Soc. 1968, 90, 4842−4846. (24) Han, L.-B.; Zhao, C.-Q. J. Org. Chem. 2005, 70, 10121−10123. (25) P−N−P ligand: (a) Ashburn, B. O.; Carter, R. G.; Zakharov, L. N. J. Am. Chem. Soc. 2007, 129, 9109. (b) Heller, B.; Gutnov, A.; Fischer, C.; Drexler, H.-J; Spannenberg, A.; Redkin, D.; Sundermann, C.; Sundermann, B. Chem. - Eur. J. 2007, 13, 1117. (c) Nishida, G.; Noguchi, K.; Hirano, M.; Tanaka, K. Angew. Chem., Int. Ed. 2007, 46, 3951. (d) Ashburn, B. O.; Carter, R. G. Angew. Chem., Int. Ed. 2006, 45, 6737. (e) Doherty, S.; Knight, J. G.; Smyth, C. H.; Jorgenson, G. A. Adv. Synth. Catal. 2008, 350, 1801. (26) Xu, Q.; Zhao, C.-Q.; Han, L.-B. J. Am. Chem. Soc. 2008, 130, 12648−12655. (27) Farnham, W. B.; Mislow, K. J. Chem. Soc. D 1971, 605. (28) Emmick, T. L.; Letsinger, R. L. J. Am. Chem. Soc. 1968, 90, 3459−3465.

12002

DOI: 10.1021/acs.joc.7b01326 J. Org. Chem. 2017, 82, 11990−12002