Tetrahedron: Asymmetry xxx (2017) xxx–xxx
Contents lists available at ScienceDirect
Tetrahedron: Asymmetry journal homepage: www.elsevier.com/locate/tetasy
Tartaric acid-derived chiral phosphite-type P,N-ligands: behavioural features in Pd-catalyzed asymmetric transformations Konstantin N. Gavrilov a,⇑, Sergey V. Zheglov a, Ilya V. Chuchelkin a, Marina G. Maksimova a, Ilya D. Firsin a, Andrew N. Fitch b, Vladimir V. Chernyshev c,d, Alexander V. Maximychev e, Alexander M. Perepukhov e a
Department of Chemistry, Ryazan State University, 46 Svoboda Street, 390000 Ryazan, Russian Federation European Synchrotron Radiation Facility, B. P. 220, 38043 Grenoble Cedex, France Department of Chemistry, M. V. Lomonosov Moscow State University, Leninskie Gory 1-3, 119991 Moscow, Russian Federation d A. N. Frumkin Institute of Physical Chemistry and Electrochemistry, Russian Academy of Sciences, Leninskiy Prospect 31/4, 119071 Moscow, Russian Federation e Department of General Physics, Moscow Institute of Physics and Technology, Institutskii per. 9, 141700 Dolgoprudny, Moscow Region, Russian Federation b c
a r t i c l e
i n f o
Article history: Received 13 August 2017 Accepted 12 September 2017 Available online xxxx
a b s t r a c t A practical synthesis of new phosphoramidite, phosphite and diamidophosphite P,N-ligands derived from (R,R)-tartaric acid was carried out. A study of these chiral inducers showed them to provide up to 93% ee in the Pd-catalyzed asymmetric allylations of (E)-1,3-diphenylallyl acetate, up to 84% ee in the Pd-catalyzed asymmetric alkylation of cinnamyl acetate with ethyl 2-oxocyclohexane-1-carboxylate and up to 63% ee in the Pd-catalyzed desymmetrization of N,N0 -ditosyl-meso-cyclopent-4-ene-1,3-diol biscarbamate. The effects of the structural modules, such as the nature of the phosphorus-containing ring or exocyclic substituent as well as of the nature of palladium source on the catalytic activity and enantioselectivity were investigated. Ó 2017 Elsevier Ltd. All rights reserved.
1. Introduction Transition metal-catalyzed asymmetric processes are one of the most powerful and indispensable tools for the synthesis of enantiopure and, in general, enantiomerically enriched compounds. Such products are widely used in the preparation of pharmaceutical agents, vitamins, crop protection chemicals, fragrances, and food additives, as well as ferroelectric liquid crystals and chiral polymers. In turn, the activity and stereoselectivity of metal complex catalysts are governed to a considerable degree by an adequate strategy of the design and synthesis of the corresponding chiral ligands, primarily, phosphorus-containing ones.1 While thousands of chiral trivalent P-ligands have been successfully explored in various transition metal-catalyzed reactions and a large number of such compounds are commercially available, the search for an efficient asymmetric inducer for a given catalytic process often represents a challenging task. A substantial problem is the substrate specificity, i.e., the ability of most of phosphorus stereoselectors (forming a part of corresponding metal complexes) to catalyze with a certain efficiency either a specific reaction or a group of related reactions. To this must be added the high cost of commercial chiral ligands coupled with patent restraints upon ⇑ Corresponding author. Tel.: +7 4912 280580; fax: +7 4912 281435. E-mail address:
[email protected] (K.N. Gavrilov).
the use of some known ligand series. For these reasons, the development of novel inexpensive phosphorus-containing chiral inducers derived preferentially from accessible enantiopure sources remains a topical research goal.1k,l,2 In recent times, TADDOL-based phosphites and phosphoramidites have excelled in a wide range of catalytic asymmetric reactions. Indeed, tetraaryl-1,3-dioxolane-4,5-dimethanols (TADDOLs) are readily obtained from inexpensive naturally occurring tartaric acid, which is available in both enantiomeric forms. As a consequence, TADDOLs are some of the most accessible enantiopure C2-symmetric diols and are extensively used as versatile chiral auxiliaries.3 Phosphite-type ligands are characterized by their oxidative stability, considerable p-acidity, simplicity of preparation by condensation processes (including techniques of parallel and solid-phase syntheses), ease of variation to enable fine-tuning and low cost.1f,g,k,m,2e,4 Among the TADDOL-derived phosphite-type stereoinducers, P, N-ligands are not so numerous but are a very attractive group. Such compounds have been successfully applied to a variety of asymmetric transformations, including the Cu-catalyzed conjugate addition and allylic substitution, the Rh-catalyzed hydrosilylation and hydroboration, and the Ir-mediated hydroboration and hydrogenation.5 Only a few P,N-ligands based on TADDOL (LA–C, Fig. 1) have been used in Pd-catalyzed enantioselective reactions, such as allylic alkylations, Suzuki coupling and Nazarov-type
https://doi.org/10.1016/j.tetasy.2017.09.011 0957-4166/Ó 2017 Elsevier Ltd. All rights reserved.
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
2
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx
R Ph Ph
O
O P
O O
Ph Ph
N
O
O N
O
O
Ph Ph
O P
N
O
Bn
N N R
O Ph
Ph Ph
LA
Ph
O
O P O
O
R
Ph
Ph
LC
LB
Figure 1. Structures of reported TADDOL-based P,N-phosphite-type ligands.
cyclizations.6 It should be noted that in the case of P,N-bidentate ligands, the enantioselectivity of a catalyst is determined not only by the stereochemistry of the ligand, but also by its electron asymmetry due to the presence of donor atoms of different nature.7 Herein we describe the synthesis of phosphoramidite, phosphite, and diamidophosphite P,N-ligands derived from (R,R)-tartaric acid and their use in Pd-catalyzed asymmetric allylic substitutions and desymmetrizations. The enantioselective formation of carbon–carbon and carbon–heteroatom bonds catalyzed by palladium complexes of chiral ligands is a classical, powerful and practically focused approach for the construction of allyl-substituted molecular frameworks.8 On the one hand, the choice of these catalytic processes is defined by the need for a thorough study of the asymmetrizing ability of new chiral inducers.8c,i,9 On the other hand, the products of the alkylation with dimethyl malonate can be readily converted into esters and amides of chiral unsaturated carboxylic acids under mild conditions.8j Allyl amines are extensively used for the preparation of azaheterocycles, a- and b-amino acids, alkaloids, hydrocarbon derivatives, and peptides.10 The Pd-catalyzed enantioselective synthesis of a quaternary carbon center is also a relatively complex problem and, alongside this, provides access to practically useful building blocks.8a,11 Both enantiomers of the desymmetrization product of N,N0 -ditosyl-meso-cyclopent-4-ene-1,3-diol biscarbamate are key precursors for the pharmacologically active natural compounds mannostatin A and ()-swainsonine.8a,e
of boiling PCl3 in the presence of a catalytic amount of Nmethylpyrrolidone. In each case, after azeotropic removal of excess PCl3 with toluene, the crude chlorophosphite was dissolved in toluene in the presence of Et3N and a solution of the corresponding enantiomer of 2-(anilinomethyl)-pyrrolidine or (2S)-2-(ferrocenylideneamino)-3,3-dimethylbutan-1-ol in toluene was added dropwise. After 24 h keeping the temperature at 20 °C, the resulting slurry was heated to 40 °C with stirring for 1 h and then cooled to 20 °C (in the case of 3, the reaction mixture was stirred for 5 min at 110 °C). After work-up, the new ligands 2a–c and 3 were obtained in good yields (77–89%) as white and orange solids, respectively. As shown in Scheme 2, the synthesis of P,N-ligand 7 included the preparation of compounds 5 and 6. Thus, anilide 5 was obtained according to the literature procedure starting from the readily accessible acid 4.13 The reduction of 5 with LiAlH4 led to amino alcohol 6 whose direct phosphorylation gave ligand 7 in 78% yield as a colorless viscous oil. All ligands could be prepared on a gram scale. Advantageously, they were found to be sufficiently stable to allow handling in air and could be stored under dry conditions at room temperature for at least a few months with little to negligible decomposition. They were fully characterized by 31P, 1H, 13C NMR, and 2D NMR (COSY and HSQC) spectroscopy, as well as by elemental analysis. The obtained data were in agreement with the assigned structures (see Table 1 and Experimental section). It should be noted that during the preparation of 2a–c and 7, phosphorylation of the starting 1,2-diamine or amino alcohol occurs exclusively at the pyrrolidine amino group or at the hydroxyl group. The presence of characteristic broad signals for the protons of the peripheral aniline NH group at dH = 4.07–4.32 in the 1H NMR spectra of ligands should be noted. The IR spectra of solutions of 2a–c and 7 in CHCl3 display ~(NAH) symmetric absorption bands of medium intensity m
2. Results and discussion The novel P,N-ligands 2a–c and 3 were synthesized as outlined in Scheme 1. We used the known method12 of preparation of chlorophosphites based on (R,R)-TADDOL 1a and (R,S)-semi-TADDOL 1b. The above-mentioned diols were treated with an excess
Ph Ph O P N O
O O
Ph Ph
O
Ph Ph
NH
i, ii
i, ii
O P N O Ph Ph 2a
i, iii R
Ph Ph
O
NH
2c
2b
O
O P N O
O
R OH
O O
OH Ph
NH
Ph
1a,b R = Ph, a R = H, b
Ph i, iv
Ph O P O O
O O Ph
N
Ph 3
Fe
Scheme 1. Synthesis of P,N-phosphoramidites 2a–c and P,N-phosphite 3 (reagents: (i) PCl3, NMP (cat.); (ii) (S)-2-(anilinomethyl)-pyrrolidine, Et3N; (iii) (R)-2(anilinomethyl)-pyrrolidine, Et3N; (iv) (2S)-2-(ferrocenylideneamino)-3,3-dimethylbutan-1-ol, Et3N).
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
3
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx
O
O
O O
OH
O
OEt
i
O
N H OEt
O
ii
N H OH
O
iii
O
NH
N P N
6
O
O
5
4
O
O
7
Scheme 2. Synthesis of P,N-diamidophosphite 7 (reagents: (i) aniline, DCC, DMAP (cat.); (ii) LiAlH4; (iii) (5S)-2-chloro-3-phenyl-1,3-diaza-2-phosphabicyclo[3.3.0]octane, Et3N.
Table 1 P NMR chemical shifts (CDCl3) and cone angles h (deg.) of ligands 2a–c, 3 and 7
31
a
Ligand
dP
h
2a 2b (82%)a (18%) 2c 3 7
139.6 145.5 138.4 139.5 130.4 122.4
197 141 199 245 145
Percentage of P*-epimers.
= 3370–3392 cm1 also corresponding to the secondary NH(Ph) amino group.7d,14 We also indicate that, although the starting imino alcohol exists in equilibrium with its oxazolidine tautomer,15 only the imino form takes part in phosphorylation resulting in ligand 3. Compounds 2b and 7 contain an asymmetric phosphorus atom. In the case of phosphoramidite 2b, the ratio of the epimers at the P⁄-stereocenter is 4.6: 1 (see Table 1). On the contrary, diamidophosphite 7 is a stereoindividual compound and has an (R)-configuration at the P⁄-stereocenter, which was confirmed by the presence of a sharp singlet at dP 122.4 in the 31P NMR spectrum of its solution in CDCl3 and a large spin–spin coupling constant in the 13C NMR spectrum, 2JC(NCH2CH2),P = 38.2 Hz (Table 1 and Experimental section). Such a value suggests an anti-orientation of the pseudoequatorial exocyclic substituent at the phosphorus atom and the A(CH2)3A moiety of the pyrrolidine fragment of the phosphabicyclic skeleton and, consequently, a syn-orientation between the lone electron pair on the phosphorus atom and the NCH2CH2 carbon atom.2e,16–18 To estimate the steric demands of novel ligands, we calculated their Tolman cone angles19 by the reported method using semiempirical quantum–mechanical AM1 techniques with full optimization of the geometrical parameters.20 The data obtained (Table 1) show that the steric demands (h) of 2a–c, 3 and 7 vary over a wide range between 141° and 245°. Compounds 2b and 7 are characterized by moderate steric demands (h = 141° and 145°, respectively), while the TADDOL-based ligands 2a and 2c (h = 197–199°) and especially ligand 3 (h = 245°) bearing the ferrocenyl group appear to be significantly more bulky ligands.19,21 The molecular structure of 2c was confirmed by synchrotron X-ray powder diffraction. The crystal data, data collection and refinement parameters are given in the Experimental section; the diffraction profiles after the final bond-restrained Rietveld refinement are shown in Figure 2. Compound 2c crystallizes with two independent molecules in the asymmetric part of the monoclinic unit cell (Fig. 3). The conformations of both molecules correspond to each other, although they have minor differences in the orientations of the terminal rings (Fig. 4). In two independent molecules, the phosphacycle exhibits a boat conformation and the pyrrolidine ring adopts an envelope conformation. A search in the Cambridge Structural Database23 for the crystal structures containing organic ligand with the (3aR,8aR)-2,2-
Figure 2. The final Rietveld plot for 2c. The experimental and difference (experimental minus calculated) diffraction profiles are shown as the black and red lines, respectively. The vertical blue bars correspond to the calculated positions of the Bragg peaks.
dimethyl-4,4,8,8-tetraphenyltetrahydro-[1,3]dioxolo[4,5-e][1,3,2]dioxaphosphepin fragment, where phosphorus atom is not involved in the coordination to any metal center, resulted in only 6 hits with the CCDC refcodes BUXNUE, FAZQAY, KUSHOX, OHEPAU, OHEPEY and XULBEN.24a–e All of the bond lengths and angles in two independent molecules of 2c are normal and comparable with those observed in the aforementioned compounds: the PAO bond lengths lie in the range of 1.642(9)– 1.685(13) Å, the PAN bond lengths are 1.694(12) and 1.725 (11) Å; the NAPAO bond angles lie in the range of 96.2(6)–101.0 (6)°, and the OAPAO bond angles are from 98.0(6) to 98.9(5)°. Having in hand a small series of novel P,N-ligands, we have tested their ability to catalyze the asymmetric allylic alkylation of (E)-1,3-diphenylallyl acetate 8 with dimethyl malonate as a benchmark test (Table 2). As seen from Table 2, compounds 2a, 3 and 7 perform the Pd-catalyzed synthesis of (S)-9a or (R)-9a with quantitative conversion and good enantioselectivity (up to 88%, 93%, and 92% ee, respectively, entries 9, 18 and 31). The moderate enantioselectivity of 28–78% ee (R) was achieved using the semiTADDOL-based phosphoramidite 2b (Table 2, entries 10–13). As opposed to 2a, its diastereomer 2c with the (R)-2-(anilinomethyl)-pyrrolidine fragment was inefficient and catalyzed the reaction with low yield and asymmetric induction (Table 2, entries 14–17). It should be noted that in a standard set of catalytic experiments, we chose [Pd(allyl)Cl]2 as a pre-catalyst and BSA/KOAc as a base. It is clear that CH2Cl2 is the solvent of choice and the best L/ Pd molar ratio is 1 in most cases. Next, for the best ligands 2a, 3 and 7 we tried to further improve asymmetric induction by modifying the reaction conditions (using other starting palladium complexes and/or Cs2CO3 instead of BSA/KOAc). However, under
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
4
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx
Figure 3. Two independent molecules in the asymmetric unit of 2c showing the atomic numbering of non-carbon atoms. Atoms are drawn as spheres with arbitrary radii. H atoms omitted for clarity.
Figure 4. Overlay of two independent molecules (prepared with Mercury22) showing only minor differences in their conformations. H atoms omitted for clarity.
these conditions ligand 2a catalyzed the reaction with comparable or slightly lower enantioselectivity. With the exception of experiments in presence of Cs2CO3, the conversion of starting substrate 8 significantly decreases (Table 2, entries 1 and 5–9). Similar trends were observed in the processes involving phosphite 3 (Table 2, entries 18 and 22–26). At the same time, the allylic alkylation of 8 in the presence of ligand 7 proceeded with quantitative conversion in all cases; the use of Cs2CO3 as the base provided a somewhat higher enantioselectivity (see Table 2, entries 27 and 31, 34 and 35). In the next step, we turned our attention to an asymmetric allylic amination of (E)-1,3-diphenylallyl acetate 8 with pyrrolidine. Catalytic systems based on phosphoramidite 2a showed mediocre activity and enantioselectivity (no more than 48% ee, Table 3, entries 1–6). It is noteworthy that the solvent change resulted in higher enantioselectivity and inversion of the absolute configuration of product 9b (see Table 3, entries 1–4). As Table 3 shows, amination in the presence of the P⁄-chiral ligand 2b proceeded with moderate asymmetric induction (up to 65% ee); the
use of CH2Cl2 as the reaction medium and the 1:1 molar ratio of L/Pd are undoubtedly preferable. As in the case of the allylic alkylation, ligand 2c showed no efficiency (ee is no higher than 10%, Table 3, entries 11–14). Phosphite 3 afforded product (S)-9b with moderate enantiomeric purity (up to 65% ee). In general, both the conversion of substrate 8 and asymmetric induction were poorly sensitive to the L/Pd molar ratio. Alongside this, the solvent effect is quite obvious: the best results were obtained in CH2Cl2 (Table 3, entries 15–20). The amination using P,N-ligand 7 resulted in quantitative conversion and with enantioselectivity up to 90% ee. The best result was obtained in using [Pd2(dba)3]CHCl3 as the pre-catalyst, while the catalysis by [Pd(allyl)Cl]2 and [Pd(allyl)(cod)]BF4 afforded the target product (R)-9b with a somewhat lower enantioselectivity (81% and 78% ee, respectively, Table 3, entries 21 and 26). In general, there are no obvious trends in the effects of the nature of the solvent and the L/Pd molar ratio on the asymmetric induction for this stereoselector (Table 3, entries 21–24). The obtained data for the Pd-catalyzed asymmetric alkylation of cinnamyl acetate 10 with ethyl 2-oxocyclohexane-1-carboxylate 11 in a toluene solution are summarized in Table 4. In this challenging process, a quaternary C⁄-stereocenter is generated on a carbon atom belonging to the nucleophile. Unfortunately, sterically demanding compounds 2a, 2c, and 3 gave no conversion. At the same time, when less bulky ligands 2b and 7 were involved in the process, the quaternary-substituted b-keto ester (S)-12 was produced with moderate conversion and enantioselectivity up to 84% and 74% ee, respectively (Table 4, entries 1–4). The optimal L/Pd molar ratio was found to be 2, especially for stereoinducer 2b. Finally, we also screened novel P,N-ligands in the Pd-catalyzed desymmetrization of N,N0 -ditosyl-meso-cyclopent-4-ene-1,3-diol biscarbamate 14 (Table 5). This transformation can be described as an intramolecular allylic substitution of a meso-substrate with two enantiotopic leaving groups. The catalytic systems were generated in situ from [Pd2(dba)3]CHCl3, the appropriate ligand and using Et3N as a base in the THF medium, which were established as the standard conditions.28 The highest enantiomeric excesses (60–63% ee) and chemical yields (95–96%) were obtained in the presence of phosphoramidite 2a regardless of the L/Pd molar ratio (Table 5, entries 1 and 2). Both ligands 2b and 2c provided good chemical yields of oxazolone (S,R)-15, but mediocre levels of asymmetric induction (26% and 39% ee, respectively, Table 5, entries 3– 6). It is noteworthy that in the case of stereoselector 3, the increase in the L/Pd molar ratio gave rise to product 15 with lower enantiomeric purity and opposite absolute configuration (Table 5,
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
5
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx Table 2 Pd-catalyzed allylic alkylation of (E)-1,3-diphenylallyl acetate 8 with dimethyl malonatea MeO2 C
OAc Ph
Ph 8
a b
c d e f g
CH 2(CO 2Me)2, cat solvents BSA, KOAc
Ph
CO 2Me * Ph
9a
Entry
Ligand
L/Pd
Solvent
Conversion (%)
ee (%)b,c
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35
2a 2a 2a 2a 2a 2a 2a 2a 2a 2b 2b 2b 2b 2c 2c 2c 2c 3 3 3 3 3 3 3 3 3 7 7 7 7 7 7 7 7 7
1 2 1 2 1 1 1 1 1 1 2 1 2 1 2 1 2 1 2 1 2 1 1 1 1 1 1 2 1 2 1 1 1 1 1
CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2
100 100 15 13 100 30 15 64 100 100 100 31 15 27 15 9 7 100 100 93 76 100 96 46 98 100 100 100 100 100 100 100 100 100 100
87 (S) 79 (S) 76 (S) 76 (S) 80 (S)d 82 (S)e 78 (S)f 85 (S)g 88 (S)d,g 78 (R) 34 (R) 46 (R) 28 (R) 2 (S) 12 (S) 10 (S) 6 (S) 93 (R) 91 (R) 82 (R) 84 (R) 86 (R)d 84 (R)e 82 (R)f 84 (R)g 90 (R)d,g 89 (S) 90 (S) 88 (S) 82 (S) 92 (S)d 90 (S)e 90 (S)f 80 (S)g 89 (S)d,g
Reactions were carried out with 2 mol % of [Pd(allyl)Cl]2 at 20 °C for 48 h (BSA, KOAc). The conversion of substrate 8 and enantiomeric excess of 9a were determined by HPLC (Kromasil 5-CelluCoat, C6H14/i-PrOH = 99/1, 0.6 mL/min, 254 nm, t(R) = 18.6 min, t (S) = 20.7 min). The absolute configuration was assigned by comparison of the HPLC retention times reported in the literature.25,26 With Cs2CO3 as a base. With complex [Pd2(dba)3]CHCl3 as a pre-catalyst. With complex [Pd(cod)Cl2] as a pre-catalyst. With complex [Pd(allyl)(cod)]BF4 as a pre-catalyst.
entries 7 and 8). The data in Table 5 showed that ligands 3 and 7 provided, in general, comparable chemical yields and ee values (entries 7, 8 and 9, 10). 3. Conclusion The following conclusions can be drawn: (i) novel phosphoramidite, phosphite, and diamidophosphite P,N-ligands have been successfully designed and synthesized from readily accessible chiral sources. An efficient procedure for the preparation of chlorophosphites from (R,R)-TADDOL and (R,S)-semi-TADDOL as necessary phosphorylating agents has been exploited; (ii) the key advantage of these stereoinducers is that their modular nature enables the structure of the phosphacyclic core and the exocyclic sidearm to be straightforwardly and systematically varied; (iii) the new ligands were applied in the Pd- mediated allylic substitution and desymmetrization. Overall, the results strongly depend on
the ligand nature, reaction conditions (including palladium precatalysts), and the type of catalytic reaction/nucleophile. Ligands 3 and 7 are the best stereoselectors in the allylic alkylation of (E)-1,3-diphenylallyl acetate (compound 7 also showed efficiency in its amination), whereas ligands 2a and 2b are superior in the desymmetrization of N,N0 -ditosyl-meso-cyclopent-4-ene-1,3-diol biscarbamate and alkylation of cinnamyl acetate, respectively. The enantioselectivities achieved with ligand 2a were higher than those with its epimeric counterpart 2c in all cases; (iv) taking the Pd-catalyzed allylic alkylation of (E)-1,3-diphenylallyl acetate with dimethyl malonate as a test reaction, it has been demonstrated that the new ligands worthily compete with the related P,N-phosphite-type chiral inductors LA and LB. Indeed, these compounds gave maximum ee values of 56% and 88%, respectively.6a–c These values are comparable or perceptible lower than those obtained with 2a, 3 and 7 (88–93% ee). Additional studies highlighting the potential of these tartaric acid-derived ligands in other transition
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
6
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx
Table 3 Pd-catalyzed allylic amination of (E)-1,3-diphenylallyl acetate 8 with pyrrolidinea
OAc Ph
Ph
a
c d e
* Ph
Ph
solvents
8
b
N
HN(CH2 )4 , cat 9b
Entry
Ligand
L/Pd
Solvent
Conversion (%)
ee (%)b,c
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26
2a 2a 2a 2a 2a 2a 2b 2b 2b 2b 2c 2c 2c 2c 3 3 3 3 3 3 7 7 7 7 7 7
1 2 1 2 1 1 1 2 1 2 1 2 1 2 1 2 1 2 1 1 1 2 1 2 1 1
CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2 CH2Cl2 CH2Cl2 THF THF CH2Cl2 CH2Cl2
28 45 23 34 0 13 100 100 39 66 12 13 0 0 89 100 13 15 0 100 100 100 100 100 97 100
28 (R) 30 (R) 44 (S) 38 (S) –d 48 (R)e 65 (R) 24 (R) 64 (R) 16 (R) 10 (S) 10 (S) — — 62 (S) 63 (S) 32 (S) 28 (S) —d 65 (S)e 81 (R) 71 (R) 75 (R) 76 (R) 90 (R)d 78 (R)e
Reactions were carried out with 2 mol % of [Pd(allyl)Cl]2 at 20 °C for 48 h. The conversion of substrate 8 and enantiomeric excess of 9b were determined by HPLC (Kromasil 5-AmiCoat, C6H14/i-PrOH = 200/1, 1.0 mL/min, 254 nm, t(R) = 5.5 min, t (S) = 6.0 min). The absolute configuration was assigned based on the literature data.27 With complex [Pd2(dba)3]CHCl3 as the precatalyst. With complex [Pd(allyl)(cod)]BF4 as the precatalyst.
Table 4 Pd-catalyzed allylic alkylation of cinnamyl acetate 10 with ethyl 2-oxocyclohexane-1-carboxylate 11a O OAc
Ph 10
a b
c
EtO
O OEt
+
cat PhCH 3
Ph O
11
12
O *
Entry
Ligand
L/Pd
Conversion (%)
eeb,c (%)
1 2 3 4
2b 2b 7 7
1 2 1 2
24 67 45 49
42 84 73 74
(S) (S) (S) (S)
Reactions were carried out with 2 mol % of [Pd(allyl)Cl]2 in toluene at 20 °C for 48 h (BSA, Zn(OAc)2). The conversion of substrate 10 and enantiomeric excess of 12 were determined by HPLC (Kromasil 5-CelluCoat, C6H14/i-PrOH = 95/5, 0.4 mL/min, 254 nm, t(R) = 14.3 min, t(S) = 16.4 min). The absolute configuration was assigned by comparison of the HPLC retention times reported in the literature.11a,b
metal-catalyzed asymmetric transformations are currently in progress in our laboratories. 4. Experimental 4.1. General 31
P, 13C and 1H NMR spectra were recorded on a Bruker AMX 400 (162.0 MHz for 31P, 100.6 MHz for 13C and 400.13 MHz for
1
H), Varian Inova 500 (202.33 MHz for 31P, 125.69 MHz for 13C and 499.8 MHz for 1H) and a Bruker Avance III 600 (242.9 MHz for 31P, 150.9 MHz for 13C and 600.13 MHz for 1H) instruments in CDCl3 medium. Complete assignment of all the resonances in the 1 H and 13C NMR spectra was achieved by the use of APT and DEPT techniques (together with 1H–1H COSY and 1H–13C HSQC techniques) and published data.2e,7d,29 Chemical shifts (parts per million, ppm) were given relative to Me4Si (1H and 13C) and 85% H3PO4 (31P NMR). Data are represented as follows: chemical shift,
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
7
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx Table 5 Pd-catalyzed desymmetrization of N,N0 -ditosyl-meso-cyclopent-4-ene-1,3-diol biscarbamate 14a Ts
Ts HO
OH
O
TsNCO
O
THF
13
HN
NH
O O
14 Ts cat
N O
THF, Et3 N - CO 2, - TsNH 2
O
a b c
Ligand
L/Pd
1 2 3 4 5 6 7 8 9 10
2a 2a 2b 2b 2c 2c 3 3 7 7
1 2 1 2 1 2 1 2 1 2
N +
H
(S,R)-15
Entry
Ts
H
H
O O
H
(R,S)-15
Yield (%) 96 95 90 92 95 93 60 45 51 48
eeb,c (%) 63 60 10 26 37 39 51 15 43 50
(S,R) (S,R) (S,R) (S,R) (S,R) (S,R) (R,S) (S,R) (S,R) (S,R)
Reactions were carried out with 5 mol % of [Pd2(dba)3]CHCl3 in THF 15 °C for 24 h. The enantiomeric excess of 15 was determined by HPLC (Knauer AG 250, C6H14/i-PrOH = 1/1, 1.0 mL/min, 245 nm, t(R,S) = 16.0 min, t(S,R) = 23.0 min). The absolute configuration was assigned by comparison of the HPLC retention times reported in the literature.28b
multiplicity (br = broad, s = singlet, d = doublet, t = triplet, m = multiplet); coupling constants nJ in Hertz (Hz), ‘‘n” values are reported in the case of their unambiguous determination. IR spectra were recorded on a Specord M80 instrument. Optical rotations were measured on an Atago AP-300 polarimeter. HPLC analyses were performed on a Stayer instrument using KromasilÒ and KnauerÒ columns. Elemental analyses were performed on a CHN-microanalyzer Carlo Erba EA1108 CHNS-O. All reactions were carried out under a dry argon atmosphere in flame-dried glassware and in freshly dried and distilled solvents. For example, toluene and tetrahydrofuran were freshly distilled from sodium benzophenone ketyl before use; dichloromethane was distilled from NaH. Triethylamine and pyrrolidine were distilled over KOH and then over a small amount of LiAlH4 before use. Thin-layer chromatography was performed on E. Merck precoated silica gel 60 F254 and Macherey-Nagel Alugram Alox N/ UV254 plates. Column chromatography was performed using silica gel MN Kieselgel 60 (230–400 mesh) and MN-Aluminum oxide, basic, Brockmann Activity 1. Compounds 1a,b were prepared analogously to known procedures.29 (S)- and (R)-2-(anilinomethyl)pyrrolidine and (2S)-2-(ferrocenylideneamino)-3,3-dimethylbutan-1-ol were obtained as published.17,15 Phosphorylating reagent (5S)-2-chloro-3-phenyl-1,3-diaza-2-phosphabicyclo[3.3.0]octane and ligands 2a–c, 3 and 7 were obtained analogously to known procedures.16,12,30,7d,15,2e The (E)-1,3-diphenylallyl acetate 6, [Pd (allyl)Cl]2, [Pd(allyl)(cod)]BF4, [Pd2(dba)3]CHCl3 and [Pd(Cod)Cl2] were prepared as published.31 Pd-catalyzed reactions: allylic alkylation of (E)-1,3-diphenylallyl acetate 8 with dimethyl malonate and amination with pyrrolidine, allylic alkylation of cinnamyl acetate 10 with ethyl 2-oxocyclohexane-1-carboxylate 11 and desymmetrization of N,N0 -ditosyl-meso-cyclopent-4-ene-1,3-diol biscarbamate 14 were performed according to the appropriate procedures.16,32,11a,b,28,2e N-Methylpyrrolidone (NMP), 1,3-dicyclohexylcarbodiimide (DCC), 4-dimethylamino-pyridine (DMAP), dimethyl malonate, N, O-bis(trimethylsilyl)acetamide (BSA), cinnamyl acetate 10, ethyl
2-oxocyclohexane-1-carboxylate 11, meso-cyclopent-4-ene-1,3diol 13 and tosyl isocyanate were purchased from Aldrich and Acros Organics and used without further purification. 4.2. General procedure for the preparation of ligands 2a–c At first, NMP (0.01 g, 0.1 mmol) was added to a vigorously stirred suspension of compound 1a or 1b (2.0 mmol) in PCl3 (2.6 mL, 30.0 mmol) and the mixture was refluxed for 5 min. The bulk of the excess PCl3 was removed under reduced pressure (40 Torr), and the final traces were removed by azeotropic distillation with toluene (10 mL) in vacuum (40 Torr). The obtained appropriate phosphorylating reagent and Et3N (0.7 mL, 5.0 mmol) were dissolved in toluene (7 mL), and a solution of the (S)- or (R)-2-(anilinomethyl)-pyrrolidine (0.35 g, 2.0 mmol) in toluene (3 mL) was added dropwise with vigorous stirring. The mixture obtained was stirred for 24 h at 20 °C, then heated to 40 °C, stirred at this temperature for 1 h, and cooled to 20 °C. The resulting suspension was filtered through a short plug of Al2O3/SiO2, the column was washed twice with toluene (10 mL), and the solvent was evaporated under reduced pressure (40 Torr). Products 2a,c and 2b were additionally purified by column chromatography (SiO2, toluene/ heptane, 1:2) and by flash chromatography (SiO2, hexane), respectively. 4.2.1. (3aR,8aR)-2,2-Dimethyl-6-[(S)-20 -[(N-phenylamino) methyl]pyrrolidin-10 -yl]-4,4,8,8-tetraphenyltetrahydro[1,3]dioxolo[4,5-e][1,3,2]dioxaphosphepin 2a White powder (1.10 g, yield 82%). [a]25 D = +37.7 (c 0.5, CH2Cl2). 1 H NMR (499.8 MHz, 26 °C): d = 0.33 (s, 3H, CH3), 1.36 (s, 3H, CH3), 1.72–1.80 (m, 1H, CH2), 1.87–1.99 (m, 2H, CH2), 2.05–2.13 (m, 1H, CH2), 3.08–3.16 (m, 2H, N(Ph)CH2), 3.21–3.28 (m, 1H, NCH2), 3.72–3.79 (m, 1H, NCH2), 4.13–4.20 (m, 1H, NCH), 4.32 (br t, 3J = 5.6, 1H, NH), 4.87 (d, 3J = 8.6, 1H, OCH), 5.27 (dd, 3 J = 8.6, 4J = 3.6, 1H, OCH), 6.44 (d, 3J = 7.6, 2H, CHNPh), 6.66 (t, 3 J = 7.6, 1H, CHNPh), 7.09 (t, 3J = 7.6, 2H, CHNPh), 7.18–7.28 (m, 5H,
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
8
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx
CHPh), 7.29–7.38 (m, 7H, CHPh), 7.47–7.52 (m, 4H, CHPh), 7.66 (d, 3 J = 7.4, 2H, CHPh), 7.79 (d, 3J = 6.7, 2H, CHPh). 13C NMR (125.69 MHz, 27 °C): d = 25.3 (s, CH3), 25.4 (s, CH2), 27.7 (s, CH3), 30.4 (s, CH2), 45.2 (d, 2J = 15.0, NCH2), 48.6 (d, 3J = 4.6, N(Ph)CH2), 56.9 (d, 2J = 13.9, NCH), 81.9 (d, 2J = 9.3, C(Ph2)), 82.1 (s, C(Ph2)), 82.3 (d, 3J = 20.8, OCH), 82.6 (d, 3J = 3.8, OCH), 111.6 (s, C(CH3)2), 112.8 (s, CHNPh), 116.9 (s, CHNPh), 127.0 (s, CHPh), 127.1 (s, CHPh), 127.2 (s, CHPh), 127.3 (s, CHPh), 127.4 (s, CHPh), 127.5 (s, CHPh), 127.7 (s, CHPh), 128.2 (s, CHPh), 128.3 (s, CHPh), 128.7 (s, CHPh), 128.8 (s, CHPh), 129.0 (s, CHNPh), 129.1 (s, CHPh), 141.5 (s, CPh), 142.3 (s, CPh), 146.6 (s, CPh), 146.9 (s, CPh), 148.6 (s, CNPh). IR ~ = 3392 (NAH) cm1. Anal. Calcd for C42H43N2O4P: C, (CHCl3): m 75.20; H, 6.46; N, 4.18. Found: C, 75.40; H, 6.54; N, 4.29. 4.2.2. (3aR,8aS)-2,2-dimethyl-6-[(S)-20 -[(N-phenylamino) methyl]pyrrolidin-10 -yl]-4,4-diphenyltetrahydro-[1,3]dioxolo [4,5-e][1,3,2]dioxaphosphepin 2b White powder (0.83 g, yield 80%). [a]26 D = 123.0 (c 0.5, CH2Cl2). 1 H NMR (600.13 MHz, 26 °C): d = 0.66 (s, 3H, CH3), 1.48 (s, 3H, CH3), 1.73–1.79 (m, 1H, CH2), 1.87–1.98 (m, 2H, CH2), 2.04–2.11 (m, 1H, CH2), 3.12 (t, 3J = 6.0, 2H, N(Ph)CH2), 3.23–3.29 (m, 1H, NCH2), 3.71–3.77 (m, 1H, NCH2), 3.90–3.95 (m, 1H, OCH2), 3.98– 4.04 (m, 1H, NCH), 4.21 (br t, 3J = 5.8, 1H, NH), 4.22–4.27 (m, 1H, OCH), 4.32–4.39 (m, 1H, OCH2), 4.98 (dd, 3J = 8.7, 4J = 3.7, 1H, OCH), 6.48 (d, 3J = 7.7, 2H, CHNPh), 6.67 (t, 3J = 7.7, 1H, CHNPh), 7.12 (t, 3J = 7.7, 2H, CHNPh), 7.26–7.31 (m, 3H, CHPh), 7.32–7.36 (m, 2H, CHPh), 7.38–7.43 (m, 3H, CHPh), 7.67 (d, 3J = 7.3, 2H, CHPh). 13 C NMR (150.9 MHz, 27 °C): d = 25.3 (d, 3J = 2.6, CH2), 25.5 (s, CH3), 27.5 (s, CH3), 30.3 (d, 3J = 3.5, CH2), 44.5 (d, 2J = 10.9, NCH2), 48.4 (d, 3 J = 4.2, N(Ph)CH2), 56.6 (d, 2J = 16.4, NCH), 65.8 (d, 2J = 9.8, OCH2), 75.3 (d, 3J = 4.0, OCH), 81.4 (d, 2J = 6.4, C(pH2)), 86.3 (d, 3J = 19.3, OCH), 110.9 (s, C(CH3)2), 112.7 (s, CHNPh), 117.0 (s, CHNPh), 126.9 (s, CHPh), 127.0 (s, CHPh), 127.1 (s, CHPh), 127.6 (s, CHPh), 128.2 (s, CHPh), 128.6 (s, CHPh), 129.1 (s, CHNPh), 141.0 (s, CPh), 146.6 (s, CPh), 148.3 (s, CNPh) (major epimer) and 25.2 (d, 3J = 2.9, CH2), 25.6 (s, CH3), 26.8 (s, CH3), 29.9 (d, 3J = 2.9, CH2), 45.1 (d, 2 J = 15.9, NCH2), 48.6 (d, 3J = 3.9, N(Ph)CH2), 57.0 (d, 2J = 13.8, NCH), 65.4 (d, 2J = 8.2, OCH2), 75.2 (s, OCH), 81.1 (d, 2J = 5.8, C (pH2)), 86.0 (d, 3J = 19.6, OCH), 111.0 (s, C(CH3)2), 112.9 (s, CHNPh), 117.1 (s, CHNPh), 127.2 (s, CHPh), 127.3 (s, CHPh), 127.4 (s, CHPh), 127.8 (s, CHPh), 128.3 (s, CHPh), 128.5 (s, CHPh), 129.2 (s, CHNPh), 141.3 (s, CPh), 146.8 (s, CPh), 148.4 (s, CNPh) (minor epimer). IR ~ = 3381 (NAH) cm1. Anal. Calcd for C30H35N2O4P: C, (CHCl3): m 69.48; H, 6.80; N, 5.40. Found: C, 69.66; H, 6.88; N, 5.55. 4.2.3. (3aR,8aR)-2,2-Dimethyl-6-[(R)-20 -[(N-phenylamino) methyl]pyrrolidin-10 -yl]-4,4,8,8-tetraphenyltetrahydro[1,3]dioxolo[4,5-e][1,3,2]dioxaphosphepin 2c White powder (1.03 g, yield 77%). [a]25 D = 53.0 (c 0.5, CH2Cl2). 1 H NMR (600.13 MHz, 26 °C): d = 0.30 (s, 3H, CH3), 1.39 (s, 3H, CH3), 1.74–1.80 (m, 1H, CH2), 1.81–1.86 (m, 1H, CH2), 1.90–1.97 (m, 1H, CH2), 2.13–2.18 (m, 1H, CH2), 3.02–3.13 (m, 2H, N(Ph) CH2), 3.26–3.32 (m, 1H, NCH2), 3.44–3.51 (m, 1H, NCH2), 4.23 (br s, 1H, NH), 4.36–4.41 (m, 1H, NCH), 4.89 (d, 3J = 8.5, 1H, OCH), 5.26 (dd, 3J = 8.5, 4J = 3.5, 1H, OCH), 6.26 (d, 3J = 7.3, 2H, CHNPh), 6.66 (t, 3J = 7.3, 1H, CHNPh), 7.08 (t, 3J = 7.3, 2H, CHNPh), 7.20–7.29 (m, 5H, CHPh), 7.30–7.38 (m, 7H, CHPh), 7.47 (t, 3J = 8.6, 4H, CHPh), 7.62 (d, 3J = 7.3, 2H, CHPh), 7.84 (d, 3J = 7.3, 2H, CHPh). 13C NMR (150.9 MHz, 27 °C): d = 28.1 (s, CH3), 28.2 (s, CH2), 30.5 (s, CH3), 33.4 (s, CH2), 48.1 (d, 2J = 18.2, NCH2), 51.4 (d, 3J = 2.8, N(Ph)CH2), 59.4 (d, 2J = 10.5, NCH), 84.4 (d, 2J = 8.0, C(Ph2)), 85.0 (s, C(Ph2)), 85.1 (d, 3J = 21.6, OCH), 85.7 (d, 3J = 3.3, OCH), 114.4 (s, C(CH3)2), 115.6 (s, CHNPh), 119.7 (s, CHNPh), 130.0 (s, CHPh), 130.1 (s, CHPh), 130.2 (s, CHPh), 130.3 (s, CHPh), 130.4 (s, CHPh), 130.6 (s, CHPh), 130.8 (s, CHPh), 131.1 (s, CHPh), 131.7 (s, CHPh), 131.8 (s, CHPh), 131.9 (s, CHPh), 132.0 (s, CHNPh), 144.7 (s, CPh), 148.9 (s, CPh),
~ = 3389 149.0 (s, CPh), 150.0 (s, CPh), 151.3 (s, CNPh). IR (CHCl3): m (NAH) cm1. Anal. Calcd for C42H43N2O4P: C, 75.20; H, 6.46; N, 4.18. Found: C, 75.50; H, 6.51; N, 4.06. 4.3. Procedure for the preparation of ligands 3 At first, NMP (0.01 g, 0.1 mmol) was added to a vigorously stirred suspension of compound 1a (0.93 g, 2.0 mmol) in PCl3 (2.6 mL, 30.0 mmol) and the mixture was refluxed for 5 min. The bulk of the excess PCl3 was removed under reduced pressure (40 Torr), and the final traces were removed by azeotropic distillation with toluene (10 mL) in vacuum (40 Torr). The obtained appropriate phosphorylating reagent and Et3N (0.7 mL, 5.0 mmol) were dissolved in toluene (7 mL), and a solution of the (2S)-2-(ferrocenylideneamino)-3,3-dimethylbutan-1-ol (0.63 g, 2.0 mmol) in toluene (3 mL) was added dropwise with vigorous stirring. The mixture obtained was heated to the boiling point, stirred for 5 min, and cooled to 20 °C. The resulting suspension was filtered through a short plug of Al2O3/SiO2, the column was washed twice with toluene (10 mL), and the solvent was evaporated under reduced pressure (40 Torr). Product 3 was additionally purified by crystallization from heptane. 4.3.1. (3aR,8aR)-2,2-dimethyl-6-[(S)-20 -(ferrocenylideneamino)30 ,30 -dimethylbutoxy]-4,4,8,8-tetraphenyltetrahydro[1,3]dioxolo[4,5-e][1,3,2]dioxaphosphepin 3 Orange powder (1.44 g, yield 89%). [a]27 D = 25.6 (c 0.5, CH2Cl2). 1 H NMR (600.13 MHz, 26 °C): d = 0.52 (s, 3H, CH3), 0.87 (s, 9H, C (CH3)3), 0.92 (s, 3H, CH3), 2.71–2.76 (m, 1H, NCH), 3.94–4.0 (m, 2H, OCH2), 4.17 (s, 5H, CHCp), 4.26 (s, 1H, CHFc), 4.31 (s, 1H, CHFc), 4.59 (s, 1H, CHFc), 4.64 (s, 1H, CHFc), 5.02 (dd, 3J = 8.2, 4J = 1.6, 1H, OCH), 5.16 (d, 3J = 8.2, 1H, OCH), 7.20–7.26 (m, 6H, CHPh), 7.27– 7.32 (m, 4H, CHPh), 7.34 (t, 3J = 7.5, 2H, CHPh), 7.45–7.50 (m, 4H, CHPh), 7.52–7.56 (m, 4H, CHPh), 7.97 (s, 1H, N = CH). 13C NMR (150.9 MHz MHz, 25 °C): d = 26.1 (s, CH3), 27.0 (s, CH3), 27.1 (s, C (CH3)3), 33.1 (s, C(CH3)3), 63.0 (s, OCH2), 67.9 (s, CHFc), 68.9 (s, CHCp), 69.1 (s, CHFc), 69.9 (s, CHFc), 70.0 (s, CHFc), 80.1 (d, 3J = 2.3, NCH), 80.9 (d, 3J = 4.6, OCH), 81.4 (s, CFc-ipso), 82.1 (d, 3J = 16.1, OCH), 82.2 (s, C(Ph2)), 85.1 (d, 2J = 10.3, C(Ph2)), 112.8 (s, C (CH3)2), 127.0 (s, CHPh), 127.1 (s, CHPh), 127.2 (s, CHPh), 127.3 (s, CHPh), 127.4 (s, CHPh), 127.5 (s, CHPh), 127.6 (s, CHPh), 128.0 (s, CHPh), 128.1 (s, CHPh), 128.9 (s, CHPh), 129.0 (s, CHPh), 129.1 (s, CHPh), 141.3 (s, CPh), 141.7 (s, CPh), 146.0 (s, CPh), 146.1 (s, CPh), 160.6 (s, N = CH). Anal. Calcd for C48H50FeNO5P: C, 71.37; H, 6.24; N, 1.73. Found: C, 71.60; H, 6.20; N, 1.65. 4.4. Procedure for the preparation of compound 5 To a vigorously stirred solution of (4R,5R)-5-(ethoxycarbonyl)2,2-dimethyl-1,3-dioxolane-4-carboxylic acid 4 (3.3 g, 15.1 mmol), aniline (1.93 mL, 21.1 mmol) and DMAP (0.09 g, 0.76 mmol) in CH2Cl2 (25 mL) was slowly added DCC (3.43 g, 16.6 mmol). After the resultant mixture was stirred for 48 h at 20 °C, the by-product dicyclohexylurea was removed by filtration. The solvent was evaporated under reduced pressure (40 Torr) and the residue was purified by flash chromatography (SiO2, EtOAc/hexane, 4:1) and bulbto-bulb distillation under vacuum (1 Torr) to give product 5. 4.4.1. Ethyl (4R,5R)-2,2-dimethyl-5-(phenylcarbamoyl)-1,3dioxolane-4-carboxylate 5 Colorless viscous oil that solidified on standing (3.01 g, yield 1 68%). [a]27 D = 44.4 (c 0.5, CH2Cl2). H NMR (499.8 MHz, 27 °C): d = 1.35 (t, 3J = 7.1, 3H, CH3), 1.55 (s, 3H, CH3), 1.58 (s, 3H, CH3), 4.32 (q, 3J = 7.1, 2H, CH2), 4.85 (d, 3J = 5.3, 1H, OCH), 4.91 (d, 3 J = 5.3, 1H, OCH), 7.15 (t, 3J = 7.6, 1H, CHPh), 7.35 (t, 3J = 7.6, 2H, CHPh), 7.57 (d, 3J = 7.6, 2H, CHPh), 8.26 (br s, 1H, NH). 13C NMR
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx
(125.69 MHz, 25 °C): d = 14.0 (s, CH3), 26.3 (s, CH3), 26.7 (s, CH3), 61.9 (s, OCH2), 77.7 (s, OCH), 78.1 (s, OCH), 113.6 (s, C(CH3)2), 119.7 (s, CHPh), 124.8 (s, CHPh), 129.1 (s, CHPh), 136.8 (s, CPh), 167.5 (s, OC), 169.9 (s, OC). Anal. Calcd for C15H19NO5: C, 61.42; H, 6.53; N, 4.78. Found: C, 61.67; H, 6.57; N, 4.59.
9
CHNPh), 117.6 (s, CHNHPh), 119.2 (s, CHNPh), 129.2 (s, CHNHPh), 129.3 (s, CHNPh), 145.4 (d, 2JC,P = 15.7, CNPh), 148.1 (s, CNHPh). IR ~ = 3370 (NAH) cm1. Anal. Calcd for C24H32N3O3P: C, (CHCl3): m 65.29; H, 7.31; N, 9.52. Found: C, 65.54; H, 7.22; N, 9.48. 4.7. Catalytic reactions
4.5. Procedure for the preparation of compound 6 Compound 5 (1.72 g, 5.86 mmol) was added portionwise to a vigorously stirred suspension of LiAlH4 (0.56 g, 14.66 mmol) in THF (25 mL) cooled in an ice-water bath. The mixture was allowed to warm up to 20 °C, refluxed for 8 h and hydrolyzed with 15% aqueous KOH solution (1.3 mL) at 0 °C. The reaction mixture was then shortly heated up to boiling point, cooled down to 20 °C and filtered. The filter cake was washed with THF (25 mL) and CH2Cl2 (2 15 mL) and the combined filtrates were evaporated under reduced pressure (40 Torr). All volatiles were removed under vacuum and the crude product was purified by bulb-to-bulb distillation under vacuum (1 Torr) to give the amino alcohol 6. 4.5.1. (4S,5S)-2,2-Dimethyl-5-(phenylamino)methyl)-1,3dioxolan-4-yl)methanol 6 Slightly yellow viscous oil (1.22 g, yield 88%). [a]27 D = 32.4 (c 0.5, CH2Cl2). 1H NMR (400.13 MHz, 26 °C): d = 1.46 (s, 3H, CH3), 1.48 (s, 3H, CH3), 3.29–3.42 (m, 2H, NCH2), 3.70–3.85 (m, 2H, OCH2), 3.97–4.01 (m, 1H, OCH), 4.03 (br s, 1H), 4.16–4.20 (m, 1H, OCH), 6.66 (d, 3J = 7.7, 2H, CHPh), 6.76 (t, 3J = 7.7, 1H, CHPh), 7.21 (t, 3J = 7.7, 2H, CHPh). 13C NMR (150.9 MHz, 25 °C): d = 30.0 (s, CH3), 30.2 (s, CH3), 48.8 (s, NCH2), 65.1 (s, OCH2), 79.4 (s, OCH), 82.2 (s, OCH), 112.3 (s, C(CH3)2), 116.2 (s, CHPh), 121.0 (s, CHPh), 132.3 (s, CHPh), 150.8 (s, CPh). Anal. Calcd for C13H19NO3: C, 65.80; H, 8.07; N, 5.90. Found: C, 65.92; H, 8.14; N, 6.09. 4.6. Procedure for the preparation of ligand 7 To a vigorously stirred solution of the phosphorylating reagent (5S)-2-chloro-3-phenyl-1,3-diaza-2-phosphabicyclo[3.3.0]octane (0.48 g, 2 mmol) and Et3N (0.56 mL, 4 mmol) in toluene (10 mL) was added amino alcohol 6 (0.48 g, 2 mmol) in one portion. The mixture obtained was stirred for 24 h at 20 °C, then heated to 40 °C, stirred at this temperature for 1 h, and cooled to 20 °C. The resulting suspension was filtered through a short plug of Al2O3/ SiO2, the column was washed twice with toluene (10 mL), and the solvent was evaporated under reduced pressure (40 Torr). Product 7 was additionally purified by flash chromatography (SiO2, EtOAc/hexane, 1:5). 4.6.1. (2R,5S)-2-[(4S,5S)-2,2-Dimethyl-5-(phenylamino)methyl)1,3-dioxolan-4-yl)methoxy]-3-phenyl-1,3-diaza-2phosphabicyclo[3.3.0]octane 7 Colorless viscous oil (0.69 g, yield 78%). [a]25 D = 114.4 (c 0.5, CH2Cl2). 1H NMR (600.13 MHz, 26 °C): d = 1.42 (s, 3H, CH3), 1.43 (s, 3H, CH3), 1.62–1.67 (m, 1H, NCHCH2), 1.76–1.82 (m, 1H, NCH2CH2), 1.84–1.90 (m, 1H, NCH2CH2), 2.02–2.09 (m, 1H, NCHCH2), 3.16–3.26 (m, 1H, NH(Ph)CH2; 1H, NCH2; 1H, NCH2CH2), 3.31– 3.35 (m, 1H, NH(Ph)CH2), 3.57–3.64 (m, 1H, NCH2CH2), 3.75–3.83 (m, 1H, NCH2; 2H, OCH2), 3.88–3.91 (m, 1H, OCH), 4.07 (br t, 3 J = 6.2, 1H, NH), 4.11–4.14 (m, 1H, OCH), 4.15–4.19 (m, 1H, NCH), 6.63 (d, 3J = 7.6, 2H, CHNHPh), 6.75 (t, 3J = 7.6, 1H, CHNHPh), 6.88 (t, 3J = 7.7, 1H, CHNPh), 7.05 (d, 3J = 7.7, 2H, CHNPh), 7.20 (t, 3 J = 7.6, 2H, CHNHPh), 7.26 (t, 3J = 7.7, 2H, CHNPh). 13C NMR (150.9 MHz, 26 °C): d = 26.3 (d, 3JC,P=3.8, NCH2CH2), 27.0 (s, CH3), 27.2 (s, CH3), 32.2 (s, NCHCH2), 46.1 (s, NH(Ph)CH2), 48.7 (d, 2JC, 2 2 P = 38.2, NCH2CH2), 54.9 (d, JC,P = 7.3, NCH2), 62.2 (d, JC,P = 4.4, OCH2), 63.4 (d, 2JC,P = 8.7, NCH), 77.9 (d, 3JC,P = 2.7, OCH), 78.6 (s, OCH), 109.2 (s, C(CH3)2), 113.1 (s, CHNHPh), 114.9 (d, 3JC,P = 12.0,
4.7.1. Pd-catalyzed allylic alkylation of (E)-1,3-diphenylallyl acetate 8 with dimethyl malonate A solution of [Pd(allyl)Cl]2 (0.0019 g, 0.005 mmol) (or [Pd2(dba)3]CHCl3 (0.005 g, 0.005 mmol), [Pd(cod)Cl2] (0.0029 g, 0.01 mmol), [Pd(allyl)(cod)]BF4 (0.0034 g, 0.01 mmol)) and the appropriate ligand (0.01 mmol or 0.02 mmol) in the appropriate solvent (1.5 mL) was stirred for 40 min. (E)-1,3-Diphenylallyl acetate (0.05 mL, 0.25 mmol) was added and the solution stirred for 15 min. Dimethyl malonate (0.05 mL, 0.44 mmol), BSA (0.11 mL, 0.44 mmol), and potassium acetate (0.002 g) or dimethyl malonate (0.05 mL, 0.44 mmol) and cesium carbonate (0.163 g, 0.5 mmol) were added. The reaction mixture was stirred for 48 h, diluted with CH2Cl2 or THF (2 mL), and filtered through a thin layer of SiO2. The filtrate was evaporated at reduced pressure (40 Torr) and dried in vacuum (10 Torr, 12 h) affording a residue containing (E)-dimethyl 2-(1,3-diphenylallyl)malonate 9a.33 In order to evaluate the ee and conversion, the obtained residue was dissolved in an appropriate eluant mixture (8 mL) and a sample was taken for HPLC analysis. 4.7.2. Pd-catalyzed allylic amination of (E)-1,3-diphenylallyl acetate 8 with pyrrolidine A solution of [Pd(allyl)Cl]2 (0.0019 g, 0.005 mmol) (or [Pd2(dba)3]CHCl3 (0.005 g, 0.005 mmol), [Pd(cod)Cl2] (0.0029 g, 0.01 mmol), [Pd(allyl)(cod)]BF4 (0.0034 g, 0.01 mmol)) and the appropriate ligand (0.01 mmol or 0.02 mmol) in the appropriate solvent (1.5 mL) was stirred for 40 min. (E)-1,3-Diphenylallyl acetate (0.05 mL, 0.25 mmol) was added and the solution stirred for 15 min, then freshly distilled pyrrolidine (0.06 mL, 0.75 mmol) was added. The reaction mixture was stirred for 48 h, diluted with CH2Cl2 or THF (2 mL), and filtered through a thin layer of SiO2. The filtrate was evaporated at reduced pressure (40 Torr) and dried in vacuum (10 Torr, 12 h) affording a residue containing (E)-1-(1,3diphenylallyl)pyrrolidine 9b.27,34 In order to evaluate the ee and conversion, the obtained residue was dissolved in an appropriate eluant mixture (8 mL) and a sample was taken for HPLC analysis. 4.7.3. Pd-catalyzed allylic alkylation of cinnamyl acetate 10 with ethyl 2-oxocyclohexane-1-carboxylate 11 A solution of [Pd(allyl)Cl]2 (0.0019 g, 0.005 mmol) and the appropriate ligand (0.01 mmol or 0.02 mmol) in toluene (1.5 mL) was stirred for 40 min. Cinnamyl acetate (0.04 mL, 0.25 mmol) was added and the solution stirred for 15 min. Ethyl 2-oxocyclohexane-1-carboxylate (0.06 mL, 0.375 mmol), BSA (0.25 mL, 1 mmol) and Zn(OAc)2 (0.005 g) were added. The reaction mixture was stirred for 48 h, diluted with toluene (2 mL) and filtered through a thin layer of SiO2. The filtrate was evaporated at reduced pressure (40 Torr) and dried in vacuum (10 Torr, 12 h) affording a residue containing ethyl 1-cinnamyl-2-oxocyclohexanecarboxylate 12.11a,b In order to evaluate the ee and conversion, the obtained residue was dissolved in an appropriate eluent mixture (8 mL) and a sample was taken for HPLC analysis. 4.7.4. Pd-catalyzed desymmetrization of N,N0 -ditosyl-mesocyclopent-4-ene-1,3-diol biscarbamate 14 A solution of [Pd2(dba)3]CHCl3 (0.005 g, 0.005 mmol) and the appropriate ligand (0.01 mmol or 0.02 mmol) in THF (1 mL) was stirred for 40 min. A solution of N,N0 -ditosyl-meso-cyclopent-4ene-1,3-diol biscarbamate 14 and Et3N (14 mL, 0.099 mmol) in THF (0.5 mL) was added dropwise with vigorous stirring at
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
10
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx
15 °C (compound 14 was prepared in situ as follows: to a solution of the meso-cyclopent-4-ene-1,3-diol 13 (0.01 g, 0.099 mmol) in THF (0.5 mL), tosyl isocyanate (35 mL, 0.232 mmol) was added; the mixture was stirred at 20 °C for 15 min, heated to 55 °C for 1 h, and cooled down to 20 °C). The reaction mixture was stirred for 24 h at 15 °C and then warmed to room temperature. The solvent was removed under reduced pressure (40 Torr) and the residue was purified by flash chromatography on a short pad of SiO2 (toluene). The solvent was evaporated at reduced pressure (40 Torr) and the residue was dried in vacuum (1 Torr, 2 h) gave the desired 3-tosyl-3,3a,6,6a-tetrahydro-2H-cyclopenta[d]oxazol2-one 15 as a slightly yellow oil that solidified on standing.28c,35 4.8. X-ray analysis 4.8.1. General information High-resolution X-ray powder diffraction measurements were carried out at room temperature at beam line ID22 of the European Synchrotron Radiation Facility (ESRF, Grenoble, France). The instrument was equipped with a cryogenically-cooled, doublecrystal Si (1 1 1) monochromator and Si (1 1 1) analyzers.36 The powder was loaded into a 1-mm-diameter thin-walled borosilicate capillary, which was rotated during measurements at a rate of 1200 rpm to improve the powder averaging. Calibration of the instrument and refinement of the X-ray wavelength (0.399962 (4) Å) were performed using NIST silicon standard 640c. The monoclinic unit-cell dimensions were determined with the indexing program ITO.37 Based on systematic extinctions the space group P21 was assigned and further tested using a Pawley fit.38 The crystal structure has been solved with the use of simulated annealing technique39 realized in MRIA40 and parallel tempering41 realized in FOX42 following the known procedure described by us earlier.7d,43 The solution found was fitted with the program MRIA in the bond-restrained Rietveld refinement taking into account anisotropic line broadening.44 The March–Dollase45 formalism was used for correction of preferred orientation in the [100] direction. One common isotropic displacement parameter Uiso was refined for all non-H atoms. H atoms were positioned geometrically with CAH 0.93–0.98 Å and NAH 0.90 Å, and not refined. Crystallographic data for 2c has been deposited with the Cambridge Crystallographic Data Centre as supplementary publication no. CCDC 1565227. Copies of data can be obtained free of charge on application to CCDC, 12 Union Road, Cambridge CB21EZ, UK (fax: (44)1223-336-033; e-mail:
[email protected]). 4.8.2. Crystal data for 2c Empirical formula C42H43N2O4P, Mr = 670.75, white powder, crystal system monoclinic, space group P21, diffractometer ID22 (ESRF), wavelength 0.399962(4) Å, a = 9.4230(9) Å, b = 46.029 (2) Å, c = 9.3432(8) Å, b = 119.072(12)°, volume 3541.9(6) Å3, Z = 4, Dx = 1.258 Mg m3, l = 0.026 mm1, 2hmin 2hmax, D2h = 0.900–17.000°, 0.002°, no. params/restraints 327/402, Rp, Rwp, Rexp 0.0377, 0.0490, 0.0148 (Rp, Rwp and Rexp are defined according to Young & Wiles46). Acknowledgements The reported study was supported by Russian Foundation for Basic Research, (research Project No. 16-33-00237-mol-a, ligands 2a–c) and Ministry of Education and Science of Russian Federation (research Project No. 4.9515.2017, ligands 3 and 7). The X-ray part of this research was supported by the Ministry of Education and Science of Russian Federation (grant No. RFMEFI61616X0069). We also thank ESRF for the access to ID22 station, experiment MA-3313.
A. Supplementary data Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.tetasy.2017.09. 011. References 1. (a) Brown, J. M. In Comprehensive Asymmetric Catalysis; Jacobsen, E. N., Pfaltz, A., Yamamoto, H., Eds.; ; Springer: Berlin, 1999; Vol. 1, pp 121–182; (b) Ohkuma, T.; Kitamura, M.; Noyori, R. In Catalytic Asymmetric Synthesis; Ojima, I., Ed.; Wiley-VCH: New York, 2000; pp 1–110; (c)Asymmetric Catalysis on Industrial Scale; Blaser, H. U., Federsel, H.-J., Eds., 2nd ed.; Wiley-VCH: New York, 2010; pp 1–110; (d)Application of Transition Metal Catalysis in Drug Discovery and Development; Crawley, M. L., Trost, B. M., Eds.; Wiley-VCH: Hoboken, 2012; (e) Phipps, R. J. Synlett 2016, 27, 1024–1026; (f) van Leeuwen, P. W. N. M.; Kamer, P. C. J.; Claver, C.; Pamies, O.; Dieguez, M. Chem. Rev. 2011, 111, 2077–2118; (g) Claver, C.; Pamies, O.; Dieguez, M. In Phosphorus Ligands in Asymmetric Catalysis; Börner, A., Ed.; ; Wiley-VCH: Weinheim, 2008; Vol. II, pp 507–528; (h) Falciola, C. A.; Alexakis, A. Eur. J. Org. Chem. 2008, 3765–3780; (i) Hargaden, G. C.; Guiry, P. J. Chem. Rev. 2009, 109, 2505–2550; (j) Bini, L.; Muller, C.; Vogt, D. Chem. Commun. 2010, 8325–8334; (k) Teichert, J. F.; Feringa, B. L. Angew. Chem. Int. Ed. 2010, 49, 2486–2528; (l) Luhr, S.; Holz, J.; Börner, A. ChemCatChem 2011, 3, 1708–1730; (m)Phosphorus(III) Ligands in Homogeneous Catalysis: Design and Synthesis; Kamer, P. C. J., van Leeuwen, P. W. N. M., Eds.; -VCH: Chichester, 2012; (n) Liu, Q.-L.; Chen, W.; Jiang, Q.-Y.; Bai, X.-F.; Li, Z.; Xu, Z.; Xu, L.-W. ChemCatChem 2016, 8, 1495–1499. 2. (a) Wassenaar, J.; Reek, J. N. H. Org. Biomol. Chem. 2011, 9, 1704–1713; (b) Tolstikov, A. G.; Khlebnikova, T. B.; Tolstikova, O. V.; Tolstikov, G. A. Russ. Chem. Rev. 2003, 72, 803–822; (c) Dieguez, M.; Ruiz, A.; Claver, C. Dalton Trans. 2003, 2957–2963; (d) Zhou, Q.-L. In Privileged Chiral Ligands and Catalysts; Zhou, Q.-L., Ed.; Wiley-VCH: Weinheim, 2011; (e) Gavrilov, K. N.; Zheglov, S. V.; Novikov, I. M.; Lugovsky, V. V.; Zimarev, V. S.; Mikhel, I. Tetrahedron: Asymmetry 2016, 27, 1260–1268. 3. (a) Pellissier, H. Tetrahedron 2008, 64, 10279–10317; (b) Lam, H. W. Synthesis 2011, 2011–2043; (c) Seebach, D.; Beck, A. K.; Heckel, A. Angew. Chem. Int. Ed. 2001, 40, 92–138; (d) Bhowmick, K. C.; Joshi, N. Tetrahedron: Asymmetry 2006, 17, 1901–1929. 4. (a) Ansell, J.; Wills, M. Chem. Soc. Rev. 2002, 31, 259–268; (b) Alexakis, A.; Benhaim, C. Eur. J. Org. Chem. 2002, 3221–3236; (c) Molt, O.; Schrader, T. Synthesis 2002, 2633–2670; (d) Reetz, M. T.; Mehler, G.; Meiswinkel, A.; Sell, T. Tetrahedron Lett. 2002, 43, 7941–7943; (e) Dieguez, M.; Pamies, O. Tetrahedron: Asymmetry 2004, 15, 2113–2122; (f) Gavrilov, K. N.; Bondarev, O. G.; Polosukhin, A. I. Russ. Chem. Rev. 2004, 73, 671–699; (g) Jerphagnon, T.; Renaud, J.-L. Tetrahedron: Asymmetry 2004, 15, 2101–2111; (h) Reetz, M. T.; Mehler, G.; Bondarev, O. Chem. Commun. 2006, 2292–2294; (i) de Vries, J. G.; Lefort, L. Chem. Eur. J. 2006, 12, 4722–4734; (j) Jagt, R. B. C.; Toullec, P. Y.; Geerdink, D.; de Vries, J. G.; Feringa, B. L.; Minnaard, A. J. Angew. Chem. Int. Ed. 2006, 45, 2789–2791; (k) Mata, Y.; Pamies, O.; Dieguez, M. Chem. Eur. J. 2007, 13, 3296–3304; (l) Swennenhuis, B. H. G.; Chen, R.; van Leeuwen, P. W. N. M.; de Vries, J. G.; Kamer, P. C. J. Org. Lett. 2008, 10, 989–992; (m) Qiao, X.-C.; Zhu, S.-F.; Zhou, Q.-L. Tetrahedron: Asymmetry 2009, 20, 1254–1261; (n) Gavrilov, K. N.; Zheglov, S. V.; Gavrilova, M. N.; Chuchelkin, I. V.; Groshkin, N. N.; Rastorguev, E. A.; Davankov, V. A. Tetrahedron Lett. 2011, 52, 5706–5710. 5. (a) Alexakis, A.; Burton, J.; Vastra, J.; Benhaim, C.; Fournioux, X.; van den Heuvel, A.; Leveque, J.-M.; Maze, F.; Rosset, S. Eur. J. Org. Chem. 2000, 4011– 24027; (b) Alexakis, A. Tetrahedron: Asymmetry 2001, 12, 1151–1157; (c) Vuagnoux-d’Augustin, M.; Alexakis, A. Eur. J. Org. Chem. 2007, 5852–5860; (d) Alexakis, A.; Malan, C.; Lea, L.; Benhaim, C.; Fournioux, X. Synlett 2001, 927– 930; (e) Heldmann, D. K.; Seebach, D. Helv. Chim. Acta 1999, 82, 1096–1100; (f) Yao, S.; Meng, J.-C.; Siuzdak, G.; Finn, M. G. J. Org. Chem. 2003, 68, 2540–2546; (g) Blume, F.; Zemolka, S.; Fey, T.; Kranich, R.; Schmalz, H.-G. Adv. Synth. Catal. 2002, 344, 868–883; (h) Alexakis, A.; Polet, D.; Bournaud, C.; Bonin, M. Tetrahedron: Asymmetry 2005, 16, 3672–3675; (i) Pfaltz, A.; Blankenstein, J.; Hilgraf, R.; Hörmann, E.; McIntyre, S.; Menges, F.; Schönleber, M.; Smidt, S. P.; Wustenberg, B.; Zimmermann, N. Adv. Synth. Catal. 2003, 345, 33–43. 6. (a) Hilgraf, R.; Pfaltz, A. Synlett 1999, 1814–1816; (b) Hilgraf, R.; Pfaltz, A. Adv. Synth. Catal. 2005, 347, 61–77; (c) Bronger, R. P. J.; Guiry, P. J. Tetrahedron: Asymmetry 2007, 18, 1094–1102; (d) Zhou, Z.; Tius, M. A. Angew. Chem. Int. Ed. 2015, 54, 6037–6040; (e) Kitamura, K.; Shimada, N.; Stewart, C.; Atesin, A. C.; Atesin, T. A.; Tius, M. A. Angew. Chem. Int. Ed. 2015, 54, 6288–6291. 7. (a) Espinet, P.; Soulantica, K. Coord. Chem. Rev. 1999, 193–195, 499–556; (b) Chelucci, G.; Orru, G.; Pinna, G. A. Tetrahedron 2003, 59, 9471–9515; (c) Guiry, P. J.; Saunders, C. P. Adv. Synth. Catal. 2004, 346, 497–537; (d) Gavrilov, K. N.; Mikhel, I. S.; Chuchelkin, I. V.; Zheglov, S. V.; Gavrilov, V. K.; Birin, K. P.; Tafeenko, V. A.; Chernyshev, V. V.; Goulioukina, N. S.; Beletskaya, I. P. ChemistrySelect 2016, 1, 4173–4186. 8. (a) Trost, B. M.; Crawley, M. L. Chem. Rev. 2003, 103, 2921–2943; (b) Crepy, K. V. L.; Imamoto, T. Adv. Synth. Catal. 2003, 345, 79–101; (c) Fernandez-Perez, H.; Etayo, P.; Panossian, A.; Vidal-Ferran, A. Chem. Rev. 2011, 111, 2119–2176; (d) McCarthy, M.; Guiry, P. J. Tetrahedron 2001, 57, 3809–3844; (e) Graening, T.; Schmalz, H.-G. Angew. Chem. Int. Ed. 2003, 42, 2580–2584; (f) Chapsal, B. D.;
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011
K. N. Gavrilov et al. / Tetrahedron: Asymmetry xxx (2017) xxx–xxx
9. 10. 11.
12. 13.
14. 15.
16.
17.
18. 19. 20. 21. 22.
23. 24.
Ojima, I. Org. Lett. 2006, 8, 1395–1398; (g) Lu, Z.; Ma, S. Angew. Chem. Int. Ed. 2008, 47, 258–297; (h) Dieguez, M.; Pamies, O. Acc. Chem. Res. 2010, 43, 312– 322; (i) Lam, F. L.; Kwong, F. Y.; Chan, A. S. C. Chem. Commun. 2010, 4649–4667; (j) Lafrance, D.; Bowles, P.; Leeman, K.; Rafka, R. Org. Lett. 2011, 13, 2322–2325; (i) Trost, B. M. Tetrahedron 2015, 71, 5708–5733; (k) Xu, J.-X.; Ye, F.; Bai, X.-F.; Zhang, J.; Xu, Z.; Zheng, Z.-J.; Xu, L.-W. RSC Adv. 2016, 45495–45502. Nemoto, T.; Hamada, Y. Tetrahedron 2011, 67, 667–687. (a) Nag, S.; Batra, S. Tetrahedron 2011, 67, 8959–9061; (b) Chavan, S. P.; Khairnar, L. B.; Chavan, P. N. Tetrahedron Lett. 2014, 55, 5905–5907. (a) Nemoto, T.; Matsumoto, T.; Masuda, T.; Hitomi, T.; Hatano, K.; Hamada, Y. J. Am. Chem. Soc. 2004, 126, 3690–3691; (b) Nemoto, T.; Masuda, T.; Matsumoto, T.; Hamada, Y. J. Org. Chem. 2005, 70, 7172–7178; (c) Nemoto, T.; Fukuda, T.; Matsumoto, T.; Hitomi, T.; Hamada, Y. Adv. Synth. Catal. 2005, 347, 1504–1506; (d) Hawner, C.; Alexakis, A. Chem. Commun. 2010, 7295–7306; (e) Punirun, T.; Peewasan, K.; Kuhakarn, C.; Soorukram, D.; Tuchinda, P.; Reutrakul, V.; Kongsaeree, P.; Prabpai, S.; Pohmakotr, M. Org. Lett. 2012, 14, 1820–1823; (f) Liu, Y.; Han, S.-J.; Liu, W. B. M. Acc. Chem. Res. 2015, 48, 740–751. Gavrilov, K. N.; Zheglov, S. V.; Novikov, I. M.; Chuchelkin, I. V.; Gavrilov, V. K.; Lugovsky, V. V.; Zamilatskov, I. A. Russ. Chem. Bull. Int. Ed. 2015, 64, 1595–1601. (a) Seo, J. S.; Whang, D.; Lee, H.; Jun, S. I.; Oh, J.; Jeon, Y. J.; Kim, K. Nature 2000, 404, 982–986; (b) Yeager, A. R.; Finney, N. S. Bioorg. Med. Chem. 2004, 12, 6451– 6460. Sutherland, I. O. In Comprehensive Organic Chemistry; Barton, D., Ollis, W. D., Eds.; Nitrogen Compounds; Pergamon Press: Oxford, 1979; Vol. 2, p 917. Tsarev, V. N.; Lyubimov, S. E.; Bondarev, O. G.; Korlyukov, A. A.; Antipin, M. Yu.; Petrovskii, P. V.; Davankov, V. A.; Shiryaev, A. A.; Benetsky, E. B.; Vologzhanin, P. V.; Gavrilov, K. N. Eur. J. Org. Chem. 2005, 2097–2105. Tsarev, V. N.; Lyubimov, S. E.; Shiryaev, A. A.; Zheglov, S. V.; Bondarev, O. G.; Davankov, V. A.; Kabro, A. A.; Moiseev, S. K.; Kalinin, V. N.; Gavrilov, K. N. Eur. J. Org. Chem. 2004, 2214–2222. (a) Brunel, J. M.; Constantieux, T.; Buono, G. J. Org. Chem. 1999, 64, 8940–8942; (b) Barta, K.; Hölscher, M.; Franciò, G.; Leitner, W. Eur. J. Org. Chem. 2009, 4102– 4116. Kimura, M.; Uozumi, Y. J. Org. Chem. 2007, 72, 707–714. Tolman, C. A. Chem. Rev. 1977, 77, 313–348. Polosukhin, A. I.; Kovalevskii, A. I.; Gavrilov, K. N. Russ. J. Coord. Chem. 1999, 25, 758–761. Bunten, K. A.; Chen, L.; Fernandez, A. L.; Poe, A. J. Coord. Chem. Rev. 2002, 233– 234, 41–51. Macrae, C. F.; Bruno, I. J.; Chisholm, J. A.; Edgington, P. R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P. A. J. Appl. Crystallogr. 2008, 41, 466–470. Groom, C. R.; Allen, F. H. Angew. Chem. Int. Ed. 2014, 53, 662–671. (a) Mewald, M.; Weickgenannt, A. Tetrahedron: Asymmetry 2010, 21, 1232– 1237; (b) Keller, E.; Maurer, J.; Naasz, R.; Schader, T.; Meetsma, A.; Feringa, B. Tetrahedron: Asymmetry 1998, 9, 2409–2415; (c) Bichler, B.; Holzhacker, C.; Glatz, M.; Stoger, B.; Kirchner, K. Acta Crystallogr., Sect. B 2015, B71, 524–534; (d) Holzhacker, C.; Stoger, B.; Carvalho, M. D.; Ferreira, L. P.; Pittenauer, E.;
25.
26.
27. 28.
29.
30.
31.
32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46.
11
Allmaier, G.; Veiros, L. F.; Realista, S.; Gil, A.; Calhorda, M. J.; Muller, D.; Kirchner, K. Dalton Trans. 2015, 44, 13071–13079; (e) Spek, A. L.; Siegler, M. A.; Patureau, F. W.; Reek, J. N. H. CSD Commun. 2015. (a) Gavrilov, K. N.; Shiryaev, A. A.; Zheglov, S. V.; Gavrilov, V. K.; Groshkin, N. N.; Maksimova, M. G.; Volov, A. N.; Zamilatskov, I. A. Tetrahedron 2014, 70, 616–624; (b) Benetskiy, E. B. Tetrahedron: Asymmetry 2011, 22, 373–378. (a) Thiesen, K. E.; Maitra, K.; Olmstead, M. M.; Attar, S. Organometallics 2010, 29, 6334–6342; (b) Ramillien, M.; Vanthuyne, N.; Jean, M.; Gheraseb, D.; Giorgic, M.; Naubronc, J.-V.; Piras, P.; Roussel, C. J. Chromatogr. A 2012, 1269, 82–93. Smyth, D.; Tye, H.; Eldred, C.; Alcock, N. W.; Wills, M. J. Chem. Soc., Perkin Trans. 2001, 1, 2840–2849. (a) Trost, B. M.; Patterson, D. E. J. Org. Chem. 1998, 63, 1339–1341; (b) Buschmann, N.; Ruckert, A.; Blechert, S. J. Org. Chem. 2002, 67, 4325–4329; (c) Benessere, V.; De Roma, A.; Del Litto, R.; Lega, M.; Ruffo, F. Eur. J. Org. Chem. 2011, 5779–5782; (d) Lega, M.; Margalef, J.; Ruffo, F.; Pàmies, O. Tetrahedron: Asymmetry 2013, 24, 995–1000. (a) Seebach, D.; Beck, A. K.; Imwinkelried, R.; Roggo, S.; Wonnacott, A. Helv. Chim. Acta 1987, 70, 954–974; (b) Seebach, D.; Devaquet, E.; Ernst, A.; Hayakawa, M.; Kuhnle, F. N. M.; Schweizer, W. B.; Weber, B. Helv. Chim. Acta 1995, 78, 1636–1650. Gavrilov, K. N.; Maksimova, M. G.; Chuchelkin, I. V.; Lugovsky, V. V.; Zheglov, S. V.; Gavrilov, V. K.; Perepukhov, A. M. Russ. Chem. Bull. Int. Ed. 2017, 66 (Izv. Akad. Nauk, Ser. Khim. 2017, 1265–1268). (a) Auburn, P. R.; Mackenzie, P. B.; Bosnich, B. J. Am. Chem. Soc. 1985, 107, 2033–2046; (b) White, D. A. Inorg. Synth. 1972, 13, 55–65; (c) Ukai, T.; Kawazura, H.; Ishii, Y.; Bonnet, J. J.; Ibers, J. A. J. Organomet. Chem. 1974, 65, 253–266; (d) Drew, D.; Doyle, J. R. Inorg Synth. 1972, 13, 47–55. Gavrilov, K. N.; Lyubimov, S. E.; Zheglov, S. V.; Benetsky, E. B.; Davankov, V. A. J. Mol. Catal. A: Chemical 2005, 231, 255–260. (a) Breeden, S.; Wills, M. J. Org. Chem. 1999, 64, 9735–9738; (b) Mei, L.-Y.; Yuan, Z.-L.; Shi, M. Organometallics 2011, 30, 6466–6475. Chen, J.; Lang, F.; Li, D.; Cun, L.; Zhu, J.; Deng, J. Tetrahedron: Asymmetry 2009, 20, 1953–1956. Zhao, D.; Wang, Z.; Ding, K. Synlett 2005, 2067–2071. Fitch, A. N. J. Res. Natl. Inst. Stand. Technol. 2004, 109, 133–142. Visser, J. W. J. Appl. Cryst. 1969, 2, 89–95. Pawley, G. S. J. Appl. Cryst. 1981, 14, 357–361. Zhukov, S. G.; Chernyshev, V. V.; Babaev, E. V.; Sonneveld, E. J.; Schenk, H. Z. Kristallogr. 2001, 216, 5–9. Zlokazov, V. B.; Chernyshev, V. V. J. Appl. Crystallogr. 1992, 25, 447–451. Favre-Nicolin, V.; Cerny, R. J. Appl. Crystallogr. 2002, 35, 734–743. Favre-Nicolin, V.; Cerny, R. Z. Kristallogr. 2004, 219, 847–856. Gavrilov, K. N.; Zheglov, S. V.; Gavrilov, V. K.; Maksimova, M. G.; Tafeenko, V. A.; Chernyshev, V. V.; Birin, K. P.; Mikhel, I. S. Tetrahedron 2017, 73, 461–471. Popa, N. C. J. Appl. Crystallogr. 1998, 31, 176–180. Dollase, W. A. J. Appl. Crystallogr. 1986, 19, 267–272. Young, R. A.; Wiles, D. B. J. Appl. Crystallogr. 1982, 15, 430–438.
Please cite this article in press as: Gavrilov, K. N.; et al. Tetrahedron: Asymmetry (2017), https://doi.org/10.1016/j.tetasy.2017.09.011