Next Article in Journal / Special Issue
Surface Sampling of a Dry Aerosol Deposited Ricin
Previous Article in Journal
New Roles Assigned to the α1–β1 (and α2–β2) Interface of the Human Hemoglobin Molecule from Physiological to Cellular
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Chiral β-Amino Alcohols as Ligands for the Ruthenium-Catalyzed Asymmetric Transfer Hydrogenation of N-Phosphinyl Ketimines

Departamento de Química Orgánica, Facultad de Ciencias and Instituto de Síntesis Orgánica (ISO), Universidad de Alicante, Apdo. 99, 03080 Alicante, Spain
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2012, 2(1), 1-12; https://doi.org/10.3390/app2010001
Submission received: 1 November 2011 / Revised: 22 December 2011 / Accepted: 28 December 2011 / Published: 9 January 2012
(This article belongs to the Special Issue Feature Papers)

Abstract

:
Some chiral β-amino alcohols have been evaluated as potential ligands for the ruthenium-catalyzed asymmetric transfer hydrogenation (ATH) of N-phosphinyl ketimines in isopropyl alcohol. The ruthenium complex prepared from [RuCl2(p-cymene)]2 and (1S,2R)-1-amino-2-indanol has shown to be an efficient catalyst for the ATH of several N-(diphenylphosphinyl)imines, affording the reduction products in very good isolated yields and enantiomeric excesses up to 82%. The inherent rigidity of the indane ring system present in the ligand seems to be very important to achieve good enantioselectivities.

Graphical Abstract

1. Introduction

Amines are very important targets in organic chemistry since they are key substrates in organic synthesis, as well as in pharmaceutical and agrochemical research fields. In addition, chiral amines have been extensively used as resolving agents [1,2,3,4], starting materials for the preparation of therapeutic drugs [5] and chiral auxiliaries in asymmetric synthesis [6,7,8]. Due to this, synthetic chemists have made great efforts in order to implement efficient procedures for the preparation of chiral amines [9,10] among which the reduction of iminic substrates plays a capital role. In recent years, the asymmetric transfer hydrogenation (ATH) has emerged as a powerful, convenient and wide-spread methodology to perform the reduction of C=X bonds (X = O or NR) [11] and it has become the second most used reduction methodology for this kind of functional group, just behind hydrogenations with molecular hydrogen. The ATH protocol presents several advantages: it is operationally simple, proceeds under mild reaction conditions, generally uses low catalyst loadings and avoids the handling of hazardous chemicals such as metallic hydrides or hydrogen gas. The ATH of prochiral ketones affords enantiomerically enriched secondary alcohols with excellent results [12,13,14,15,16,17,18,19]. Nevertheless, the asymmetric reduction of iminic compounds by transfer hydrogenation yielding chiral amines still remains underdeveloped and represents a challenging objective [20,21].
Due to the low electrophilic character of the iminic carbon, substrates in which the nitrogen atom is bonded to an activating group (aryl, alkoxy, amino, sulfinyl, sulfonyl or phosphinyl) are generally used in reduction methodologies leading to chiral amines. The phosphinyl group is very attractive since its removal from the reduction products can be conveniently achieved under mild acidic conditions, leading to the free primary amines [22,23]. N-phosphinyl imines have found a variety of synthetic applications in asymmetric processes [10,24,25]. For instance, the preparation of chiral amines can be achieved by stereoselective reduction of N-phosphinyl ketimines through different synthetic approaches such as hydrogenation [26,27,28], hydrosilylation [29,30,31,32,33], Meerwein-Schmidt-Ponndorf-Verley (MSPV) reaction [34] or the use of boron [35] and aluminium hydrides [36].
Although the transfer hydrogenation protocol has been already applied to the reduction of different iminic substrates bearing alkyl [18], benzyl [17,18], aryl [17,18], sulfonyl [13,18,37] or sulfinyl [38,39,40,41] groups attached to the nitrogen atom, there are only a few examples on the use of N-phosphinyl imines as substrates [18,42]. In the last years, we have been interested in the use of β-amino alcohols as ligands for ruthenium complexes used as catalysts for the ATH of imines, especially N-(tert-butylsulfinyl)imines [38,39,40,41]. Since, to the best of our knowledge, the use of β-amino alcohols as chiral ligands has not been reported so far for the ATH of N-(diphenylphosphinyl)imines, we became interested in trying to apply our methodology to those substrates. Thereby, herein we present our preliminary results on the reduction of phosphinyl imines in isopropyl alcohol with ruthenium complexes bearing chiral β-amino alcohols.

2. Experimental Section

2.1. General

All glassware was dried in an oven at 100 °C and cooled to room temperature under argon before use. All reactions were carried out under an argon atmosphere. Ketone oximes were prepared according to a literature procedure [43]. Commercially available triethylamine, chlorodiphenylphosphine, [RuCl2(p-cymene)]2, ligand L1 and the anhydrous solvents CH2Cl2, Et2O and i-PrOH were used as received. Potassium tert-butoxide was heated in a Kugel-Rohr apparatus at 170–180 °C under vacuum for 4 h before use. Column chromatography was performed with silica gel 60 of 230–400 mesh. When mentioned, purification by column chromatography on deactivated silica gel means that, before adding the reaction crude to the column, the latter was washed with a mixture of 5% triethylamine in hexane until the eluent coming from the column was basic according to pH paper. Thin layer chromatography (TLC) was performed on precoated silica gel plates (Merck 60, F254, 0.25 mm); detection was done by UV254 light and staining with phosphomolybdic acid (solution of 1 g of phosphomolybdic acid in 24 mL of absolute ethanol); Rf values are given under these conditions. When mentioned, an Rf value measured on deactivated silica gel means that the TLC plate was eluted with a mixture of 5% triethylamine in hexane and dried before applying the sample. Melting points (mp) are uncorrected and were measured on a Reichert Thermovar apparatus. NMR spectra were recorded on a Bruker AC-300 spectrometer using, as internal references, tetramethylsilane (TMS) for 1H NMR and CDCl3 for 13C NMR; chemical shifts are given in δ (ppm) and coupling constants (J) in Hz, unless otherwise stated. NMR samples were prepared using CDCl3 or C6D6 as solvent. Infrared (FT-IR) spectra were obtained on a Nicolet 510 P-FT spectrophotometer. Mass spectra (EI) were obtained at 70 eV on a Hewlett Packard HP-5890 GC/MS instrument equipped with a HP-5972 selective mass detector. Optical rotations were measured on a Perkin-Elmer 341 polarimeter. HPLC analyses were performed at 25 °C on a JASCO apparatus, equipped with a PU-2089 Plus pump, a MD-2010 Plus detector and an AS-2059 Plus automatic injector.

2.2. Procedure for the Preparation of N-Phosphinylimines 1

Triethylamine (1.35 mL, 10.5 mmol) was added to a solution of the ketone oxime (10 mmol) in anhydrous CH2Cl2 (25 mL) and Et2O (50 mL) under argon and the solution was cooled to −45 °C. A solution of chlorodiphenylphosphine (1.87 mL, 10 mmol) in CH2Cl2 (3 mL) was added dropwise over a period of 45 min. The reaction mixture was stirred at −45 °C for 1 hour, slowly warmed (without removing the Dewar) to room temperature, and then stirred overnight at room temperature. Insoluble salts (Et3N·HCl) were filtered off and washed with CH2Cl2/Et2O (1:1). Solvents were then evaporated under vacuum from the liquid layer and a yellow solid was obtained, to which CH2Cl2/Et2O (1:1, 40 mL) was added. The insoluble residue was filtered off and the liquid phase was concentrated under vacuum to give a solid that was purified by column chromatography (deactivated silica gel; hexane/acetone 90:10 to 30:70), giving the expected imines in 58% (for 1a), 52% (for 1b) and 56% (for 1c) yields. The corresponding physical and spectroscopic data follow:
Applsci 02 00001 i001
N-(1-Phenylethylidene)diphenylphosphinamide (1a) [29]: pale yellow solid; mp 120–122 °C; IR (KBr) 3064, 1637, 1192 cm−1; 1H NMR (300 MHz, CDCl3) δ 2.97 (3H, d, J = 2.0 Hz), 7.39–7.60 (9H, m), 7.92–8.04 (4H, m), 8.09 (2H, d, J = 7.2 Hz); 13C NMR (75 MHz) δ 26.5 (d, J = 12.7 Hz), 127.9, 128.2 (d, J = 12.9 Hz), 128.6, 131.1 (d, J = 3.0 Hz), 131.3 (d, J = 9.2 Hz), 132.2, 134.4 (d, J = 134.1 Hz), 139.1 (d, J = 24.8 Hz), 181.2 (d, J = 7.9 Hz); m/z 319 (M+, 53%), 318 (23), 277 (12), 242 (14), 202 (11), 201 (100), 125 (28), 77 (50), 51 (21).
Applsci 02 00001 i002
N-(1-Phenylpropylidene)diphenylphosphinamide (1b) [44]: white solid; mp 130–132 °C; IR (KBr) 3058, 1630, 1190 cm−1; 1H NMR (300 MHz, C6D6) δ 1.06 (3H, t, J = 7.5 Hz), 3.35 (2H, q, J = 7.5 Hz), 6.99–7.14 (9H, m), 7.83–7.86 (2H, m), 8.23–8.25 (4H, m); 13C NMR (75 MHz, C6D6) δ 13.5, 29.0 (d, J = 11.9 Hz), 128.6 (d, J = 9.0 Hz), 128.7 (d, J = 4.8 Hz), 131.3 (d, J = 2.9 Hz), 132.1, 132.2 (d, J = 8.7 Hz), 132.4 (d, J = 9.2 Hz), 136.6 (d, J = 131.3 Hz), 138.8 (d, J = 24.3 Hz), 186.4 (d, J = 7.6 Hz); m/z 333 (M+, 46%), 256 (18), 201 (100), 132 (12), 77 (52).
Applsci 02 00001 i003
N-[1-(p-Chlorophenyl)ethylidene]diphenylphosphinamide (1c) [44]: white solid; mp 142–144 °C; IR (KBr) 3058, 1640, 1199 cm−1; 1H NMR (300 MHz, C6D6) δ 2.62 (3H, d, J = 1.8 Hz), 7.02–7.20 (8H, m), 7.50 (2H, d, J = 8.8 Hz), 8.17–8.26 (4H, m); 13C NMR (75 MHz, C6D6) δ 22.0 (d, J = 12.5 Hz), 127.7 (d, J = 11.8 Hz), 127.8, 128.6, 130.5 (d, J = 2.9 Hz), 131.1 (d, J = 8.1 Hz), 135.3 (d, J = 131.5 Hz), 137.4 (d, J = 26.8 Hz), 137.5, 179.0 (d, J = 7.7 Hz); m/z 355 (M++2, 11%), 353 (M+, 39), 278 (21), 276 (59), 201 (100), 154 (12), 152 (38), 77 (43).

2.3. Procedure for the Asymmetric Transfer Hydrogenation of Imines 1

A mixture of [RuCl2(p-cymene)]2 (15 mg, 0.025 mmol), ligand L (0.05 mmol) and i-PrOH (2 mL) was heated to 90 °C (oil bath temperature) for 20 minutes. During this heating period, the initially orange reaction mixture turned into a dark red colour. The reaction was then cooled to −20 °C, and a solution of the imine 1 (0.5 mmol) in i-PrOH (9 mL) and t-BuOK (2.5 mL of a 0.1 M solution in i-PrOH, 0.25 mmol) were successively added. The solution was stirred for 1 h at the same temperature, slowly warmed (without removing the Dewar) to room temperature (ca. 6 h) and then stirred at room temperature for 14 h. After completion of the reaction (monitored by TLC), the mixture was passed through a small column of silica gel, the column was eluted with ethyl acetate and the combined organic phases were concentrated under vacuum. The obtained residue was purified by column chromatography (silica gel, hexane/acetone 9:1), yielding the expected reduction products 2 in the yields indicated in Table 2. Phosphinamides 2 were analyzed by HPLC on a ChiralCel OD-H column using a 254 nm UV detector, 10% i-PrOH in hexane as eluent and a flow rate of 0.5 mL/min, giving the enantiomeric excesses (ee’s) indicated in Table 2. The retention times were: 16.5 (R) and 23.4 (S) for 2a (OD-H column), 13.9 (R) and 18.9 (S) for 2b (OD-H column), 12.7 (R) and 28.1 (S) for 2c (Chiralpak AS-H column, 20% i-PrOH in hexane as eluent and a flow rate of 1.0 mL/min). For every product 2, the absolute configuration of the major enantiomer was assigned according to the HPLC retention times of the two enantiomers described in the literature under the same conditions [45,46]. The corresponding physical and spectroscopic data follow:
Applsci 02 00001 i004
(R)-N-(1-Phenylethyl)diphenylphosphinamide (2a) [23]: white solid; mp 166–168 °C; [α]20D +32.0 (c 1.1, CH2Cl2, 82% ee) {literature [23] [α]20D +28.5 (c 1.0, CH2Cl2, 75% ee)}; IR (KBr) 3159, 3059, 1179 cm−1; 1H NMR (300 MHz, CDCl3) δ 1.58 (3H, d, J = 6.7 Hz), 3.14–3.21 (1H, m), 4.33–4.48 (1H, m), 7.25–7.50 and 7.78–7.95 (11H and 4H, respectively, 2 m); 13C NMR (75 MHz, CDCl3) δ 26.3 (d, J = 3.0 Hz), 51.3, 126.0, 127.2, 128.3, 128.4, 128.47, 128.56, 128.59, 131.7, 131.82, 131.86, 132.0, 132.1, 132.5, 132.6, 145.1 (d, J = 6.2 Hz); m/z 321 (M+, 5%), 307 (19), 306 (72), 202 (22), 201 (100), 120 (98), 77 (34).
Applsci 02 00001 i005
(R)-N-(1-Phenylpropyl)diphenylphosphinamide (2b) [23]: white solid; mp 153–155 °C; [α]20D +26.0 (c 1.0, MeOH, 82% ee) {literature [23] [α]20D +31.3 (c 1.35, MeOH, 81% ee)}; IR (KBr) 3145, 3060, 1198 cm−1; 1H NMR (300 MHz, CDCl3) δ 0.77 (3H, t, J = 7.5 Hz), 1.75–2.10 (2H, m), 3.12–3.35 (1H, m), 4.02–4.20 (1H, m), 7.10–7.55 and 7.70–7.95 (11H and 4H, respectively, 2 m); 13C NMR (75 MHz, CDCl3) δ 10.5, 32.5 (d, J = 3.1 Hz), 57.1, 126.5, 127.1, 128.2, 128.3, 128.4, 128.5, 131.9, 132.5, 132.6, 143.5 (d, J = 5.4 Hz); m/z 335 (M+, 3%), 307 (21), 306 (100), 201 (53), 134 (16), 77 (22).
Applsci 02 00001 i006
(R)-N-[1-(p-Chlorophenyl)ethyl]diphenylphosphinamide (2c) [42]: white solid; mp 161–163 °C; [α]20D +60.5 (c 1.0, MeOH, 80% ee) {literature for ent-2c [28] [α]20D −73.4 (c 1.14, MeOH, 94% ee)}; IR (KBr) 3152, 3062, 1189 cm−1; 1H NMR (400 MHz, CDCl3) δ 1.54 (3H, d, J = 6.9 Hz), 3.15–3.34 (1H, m), 4.29–4.45 (1H, m), 7.19–7.30, 7.38–7.51, 7.78–7.93 (4H, 6H and 4H, respectively, 3 m); 13C NMR (100 MHz, CDCl3) δ 25.7 (d, J = 3.8 Hz), 50.3, 127.4, 128.2, 128.3, 128.4, 128.6, 131.4, 132.8, 133.5, 143.5 (d, J = 6.1 Hz); m/z 357 (M++2, <1%), 355 (M+, 3), 340 (24), 306 (100), 202 (16), 201 (70), 154 (100), 77 (26).

3. Results and Discussion

In recent years, one of our main research lines has focused on the asymmetric synthesis of chiral amines by enantioselective addition of dialkylzinc reagents to N-(diphenylphosphinyl)imines using β-amino alcohols as ligands [47,48,49]. Since those ligands have shown to give excellent results in the ruthenium-catalyzed asymmetric transfer hydrogenation of ketones [12,13,14,15,16,17,18,19] and N-sulfinylimines [38,39,40,41] in isopropyl alcohol, we decided to investigate if they could also be applied to the reduction of N-phosphinyl ketimines by the same methodology. One of the most noteworthy chiral β-amino alcohols in the ATH of both ketones and N-sulfinyl imines is cis-1-amino-2-indanol [50]. Due to this, we chose (1S,2R)-1-amino-2-indanol L1 as a test ligand for the ruthenium-catalyzed transfer hydrogenation of the N-phosphinyl imine derived from acetophenone 1a (Table 1). The ruthenium catalyst was prepared by refluxing a mixture of the ruthenium dimer [RuCl2(p-cymene)]2 and L1 in isopropyl alcohol.
Table 1. Ruthenium-catalyzed asymmetric transfer hydrogenation of N-(diphenylphosphinyl)- -ketimine 1a in isopropyl alcohol. Optimization of the reaction conditions a. Applsci 02 00001 i007
Table 1. Ruthenium-catalyzed asymmetric transfer hydrogenation of N-(diphenylphosphinyl)- -ketimine 1a in isopropyl alcohol. Optimization of the reaction conditions a. Applsci 02 00001 i007
EntryRu-dimer (mol%)L1(mol%)BaseT (°C)t (h)Product 2a
Basemol%Yield b (%)ee c (%)
1520KOH502555666
2510KOH502537672
355KOH502565564
4510KOH752538466
5510KOH252567866
6510LiOH502556370
7510NaOH502535568
836KOH3025185374
9510KOH505027460
10510KOH500294878
11 d510KOH500 to 25197078
12 d510KOH50−20 to 25228080
13 d510t-BuOK50−20 to 25208782
14 d510t-BuOK50−40 to 25208578
a The solution of the imine 1a (0.5 mmol) in i-PrOH (9 mL) was added to a solution of the ruthenium catalyst [prepared by refluxing a mixture of [RuCl2(p-cymene)]2 and amino alcohol L1 in i-PrOH (2 mL)] at the temperature indicated. Then, the base (as a 0.1 M solution in i-PrOH) was added and the reaction mixture was stirred for the time indicated. b Isolated yield after purification by column chromatography (silica gel, hexane/acetone: 9/1) based on the starting imine 1a. c Determined by HPLC using a ChiralCel OD-H column. The R-enantiomer was the major one in all cases. d The reaction mixture was stirred at the initial temperature for 1 h and then it was allowed to gradually reach room temperature.
Since a proportion Ru-dimer:aminoalcohol:base = 1:4:10 is common for the ATH of ketones, we chose that proportion as the starting point of our study, using 5 mol% of the Ru-dimer. Under those conditions, the Ru complex bearing L1 was able to reduce the imine 1a in isopropyl alcohol in the presence of potassium hydroxide, affording product 2a in a moderate yield and with 66% ee after stirring for 5 h at room temperature (Table 1, entry 1). This promising result encouraged us to try to optimize the reaction conditions. First, the amount of L1 was reduced keeping the loadings of Ru-dimer and base constant and we found that both yield and ee improved when the ratio Ru-dimer: L1 was 1:2 (compare entries 1–3 in Table 1). Hence, we decided to keep that ratio in the following tests. Next, the loading of base was evaluated and a decrease in the enantioselectivity, in comparison with the results of entry 2, was observed when the amount of potassium hydroxide was either increased to 75 mol% (Table 1, entry 4) or decreased to 25 mol% (Table 1, entry 5). Since the optimum ratio Ru-dimer: L1:base seemed to be 1:2:10, it was maintained in further experiments. A different metallic cation in the base did not cause any noticeable effect on the enantioselectivity: lithium and sodium hydroxides gave almost the same ee’s as potassium hydroxide, the yield being higher with the latter (compare entries 2, 6 and 7 in Table 1). We tried to reduce the amount of the Ru-dimer to 3 mol%, but this considerably slowed down the transfer hydrogenation process and resulted in a lower yield of 2a, although the enantioselectivity did not vary (compare entries 2 and 8 in Table 1).
The effect of the temperature was also studied. The ee decreased to 60% when the reaction was performed at 50 °C (Table 1, entry 9). A decrease of the temperature to 0 °C drastically lengthened the reaction time and, furthermore, the yield decreased to 48% (Table 1, entry 10). However, the enantioselection of the process slightly improved to 78% ee. In order to try to improve the yield, the reaction was repeated and an upward temperature gradient was applied: after stirring for 1 h at 0 °C, the reaction was allowed to gradually reach room temperature. We were delighted to see that, under these conditions, a 70% yield was obtained with no detriment to the enantioselectivity (Table 1, entry 11). A slight improvement of both yield and enantioselectivity were observed when the initial temperature was −20 °C instead of 0 °C (Table 1, entry 12). Since in our previous studies concerning the ATH of N-(tert-butylsulfinyl)imines [38,39,40,41] we found that potassium tert-butoxide was a more convenient base than potassium hydroxide, we decided to evaluate the former in the ATH of phosphinylimine 1a and we obtained a further improvement, giving 87% yield and 82% ee (Table 1, entry 13). A decrease of the initial temperature to −40 °C slightly reduced both the yield and the enantiomeric excess (Table 1, entry 14). After having performed this screening, we chose the conditions shown in entry 13 as the optimal ones.
The influence of the arene ligand in the ruthenium catalyst was also studied and we found that the use of [RuCl2(benzene)]2 instead of [RuCl2(p-cymene)]2 almost completely prevented any asymmetric induction, the ee of the product 2a being only 8%.
Once we had determined the optimum reaction conditions, we decided to test some other chiralβ-amino alcohols with the aim of evaluating the influence of the ligand structure on the success of the ATH process. Amino alcohols (1R,2S)-norephedrine L2, (1R,2S)-2-amino-1,2-diphenylethanol L3 and prolinols L4 and L5 were tested as potential ligands, and the results are summarized in Table 2. The ruthenium complex bearing norephedrine L2 was able to reduce imine 1a, though the results concerning both yield and enantioselectivity were rather poor (Table 2, entry 2). When the ATH reaction was carried out using ligand L3, the reduction product was obtained in only 39% yield and with a moderate ee. Therefore, ligands L2 and L3 have shown to be clearly inferior to L1 (compare entries 1 to 3 in Table 2). This drop in enantioselectivity could be due to the fact that (1S,2R)-1-amino-2-indanol L1 has a more rigid structure than ligands L2 or L3. Therefore, it seems that the rigidity of the β-amino alcohol seems to be very convenient for achieving good enantioselectivities in this ATH process. Finally, in agreement with our observations in the ATH of sulfinylimines [38,39,40,41], a ligand with a free primary amino group led to higher enantioselectivities than other ligands bearing secondary amino groups: when the ligands with a prolinol backbone L4 and L5 were tested, either no reduction product or only traces of it could be detected after working-up the reaction (Table 2, entries 4 and 5).
Table 2. Test of different amino alcohols as chiral ligands for the ruthenium-catalyzed transfer hydrogenation of N-phosphinylimines 1 a. Applsci 02 00001 i008
Table 2. Test of different amino alcohols as chiral ligands for the ruthenium-catalyzed transfer hydrogenation of N-phosphinylimines 1 a. Applsci 02 00001 i008
Entry Imine Ligandt (h)Product 2
No.R1R2No.Yield b (%)ee c (%)
11a HMeL1 202a 8782
21a HMeL2 202a 4948
31a HMeL3 202a 3953
41a HMeL4 222a d
51a HMeL5 252a e
61b HEtL1 202b 9082
71c ClMeL1 202c 8580
a The solution of the imine 1 (0.5 mmol) in i-PrOH (9 mL) was added to a solution of the ruthenium catalyst [prepared by refluxing a mixture of [RuCl2(p-cymene)]2 (0.025 mmol) and amino alcohol L (0.05 mmol) in i-PrOH (2 mL)] at −20 °C. Then, t-BuOK (0.25 mmol, as a 0.1 M solution in i-PrOH) was added and the reaction mixture was stirred at the same temperature for 1 h and then it was allowed to gradually reach room temperature. b Isolated yield after purification by column chromatography (silica gel, hexane/acetone: 9/1) based on the starting imine 1. c Determined by HPLC using a ChiralCel OD-H column or a Chiralpak AS-H column. The R-enantiomer was the major one in all cases. d Only traces of the product were detected in the crude reaction mixture. e No reduction product was detected in the crude reaction mixture.
After having established that L1 was the ligand of choice, we tested two more imines as substrates. The asymmetric transfer hydrogenation of imines 1b and 1c, derived from propiophenone and 4-chloroacetophenone respectively, under the conditions of entry 13 in Table 1 gave the expected phosphinamides 2b and 2c in high yields and with ee’s of 82 and 80%, respectively (Table 2, entries 6 and 7).

4. Conclusions

In conclusion, we have developed a new efficient procedure for the reduction of prochiral N-(diphenylphosphinyl)ketimines through a ruthenium-catalyzed asymmetric transfer hydrogenation process in isopropyl alcohol using a β-amino alcohol as a chiral ligand. The rigid structure present in (1S,2R)-1-amino-2-indanol seems to be very convenient in order to achieve good enantiomeric excesses. The arene ligand on the ruthenium complex also has an influence on the stereochemical outcome of the reaction. Since the phosphinyl group can easily be removed under mild acidic conditions, this procedure represents an efficient method to synthesize optically enriched chiral primary amines. Further efforts to improve the enantioselectivity and to extend the substrate scope of this new ATH process are currently underway in our laboratory.

Acknowledgements

This work was generously supported by the Spanish Ministerio de Ciencia e Innovación (MICINN; Grant No. CONSOLIDER INGENIO 2010, CSD2007-00006, CTQ2007-65218, and CTQ11-24151) and the Generalitat Valenciana (PROMETEO/2009/039 and FEDER). Óscar Pablo thanks the Spanish Ministerio de Educación for a predoctoral fellowship (Grant No. AP2008-00989). We also thank MEDALCHEMY S.L. for a gift of chemicals.

References and Notes

  1. Jacques, J.; Collet, A.; Wilen, S.H. Enantiomers, Racemates and Resolution; John Wiley & Sons: New York, NY, USA, 1981. [Google Scholar]
  2. Juaristi, E.; Escalante, J.; León-Romo, J.L.; Reyes, A. Recent applications of α-phenylethylamine (α-PEA) in the preparation of enantiopure compounds. Part 1. Incorporation in chiral catalysts. Part 2. α-PEA and derivatives as resolving agents. Tetrahedron: Asymmetry 1998, 9, 715–740. [Google Scholar] [CrossRef]
  3. Fogassy, E.; Nógrádi, M.; Kozma, D.; Egri, G.; Pálovics, E.; Kiss, V. Optical resolution methods. Org. Biomol. Chem. 2006, 4, 3011–3030. [Google Scholar] [CrossRef]
  4. Faigl, F.; Fogassy, E.; Nógrádi, M.; Pálovics, E.; Schindler, J. Strategies in optical resolution: A practical guide. Tetrahedron: Asymmetry 2008, 19, 519–536. [Google Scholar] [CrossRef]
  5. Nugent, T.C. Chiral Amine Synthesis. Methods, Developments and Applications; Wiley-VCH: Weinheim, Germany, 2010. [Google Scholar]
  6. Whitesell, J.K. C2 Symmetry and asymmetric induction. Chem. Rev. 1989, 89, 1581–1590. [Google Scholar]
  7. Lucet, D.; Le Gall, T.; Mioskowski, C. The chemistry of vicinal diamines. Angew. Chem. Int. Ed. 1998, 37, 2580–2627. [Google Scholar] [CrossRef]
  8. Waldmann, H. C2-symmetric amines as chiral auxiliaries. In Organic Synthesis Highlights II; Waldmann, H., Ed.; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2008. [Google Scholar]
  9. Braese, S.; Baumann, T.; Dahmen, S.; Vogt, H. Enantioselective catalytic syntheses of α-branched chiral amines. Chem. Commun. 2007, 1881–1890. [Google Scholar]
  10. Nugent, T.C.; El-Shazly, M. Chiral amine synthesis. Recent developments and trends for enamide reduction, reductive amination, and imine reduction. Adv. Synth. Catal. 2010, 352, 753–819. [Google Scholar]
  11. Gladiali, S.; Alberico, E. Asymmetric transfer hydrogenation: Chiral ligands and applications. Chem. Soc. Rev. 2006, 35, 226–236. [Google Scholar] [CrossRef]
  12. Noyori, R.; Hashiguchi, S. Asymmetric transfer hydrogenation catalyzed by chiral ruthenium complexes. Acc. Chem. Res. 1997, 30, 97–102. [Google Scholar] [CrossRef]
  13. Palmer, M.J.; Wills, M. Asymmetric transfer hydrogenation of C=O and C=N bonds. Tetrahedron 1999, 10, 2045–2061. [Google Scholar] [CrossRef]
  14. Wills, M.; Palmer, M.; Smith, A.; Kenny, J.; Walsgrove, T. Recent developments in the area of asymmetric transfer hydrogenation. Molecules 2000, 5, 4–18. [Google Scholar]
  15. Clapham, S.E.; Hadzovic, A.; Morris, R.H. Mechanisms of the H2-hydrogenation and transfer hydrogenation of polar bonds catalyzed by ruthenium hydride complexes. Coord. Chem. Rev. 2004, 248, 2201–2237. [Google Scholar]
  16. Ikariya, T.; Blacker, A.J. Asymmetric transfer hydrogenation of ketones with bifunctional transition metal-based molecular catalysts. Acc. Chem. Res. 2007, 40, 1300–1308. [Google Scholar] [CrossRef]
  17. Wang, C.; Wu, X.; Xiao, J. Broader, greener, and more efficient: Recent advances in asymmetric transfer hydrogenation. Chem. Asian J. 2008, 3, 1750–1770. [Google Scholar] [CrossRef]
  18. Wills, M. Imino reductions by transfer hydrogenation. In Modern Reduction Methods; Andersson, P.G., Munslow, I.J., Eds.; Wiley-VCH: Weinheim, Germany, 2008; pp. 271–296. [Google Scholar]
  19. Malacea, R.; Poli, R.; Manoury, E. Asymmetric hydrosilylation, transfer hydrogenation, and hydrogenation of ketones catalyzed by iridium complexe. Coord. Chem. Rev. 2010, 254, 729–752. [Google Scholar]
  20. Spindler, F.; Blaser, H.-U. Enantioselective hydrogenation of C=N functions and enamines. Handb. Homog. Hydrog. 2007, 3, 1193–1214. [Google Scholar]
  21. Fabrello, A.; Bachelier, A.; Urrutigoïty, M.; Kalck, P. Mechanistic analysis of the transition metal-catalyzed hydrogenation of imines and functionalized enamines. Coord. Chem. Rev. 2010, 254, 273–287. [Google Scholar]
  22. Krzyzanowska, B.; Stec, W.J. A new approach to the synthesis of primary amines, isothiocyanates, and 1-aminoalkanephosphonates via N-phosphinyl aldoximes and ketoxime. Synthesis 1978, 521–524. [Google Scholar] [CrossRef]
  23. Andersson, P.G.; Guijarro, D.; Tanner, D. Preparation and use of aziridino alcohols as promoters for the enantioselective addition of dialkylzinc reagents to N-(diphenylphosphinoyl) imines. J. Org. Chem. 1997, 62, 7364–7375. [Google Scholar] [CrossRef]
  24. For a review on the use of N-phosphinoylimines in stereoselective synthesis, see: Weinreb, S.M.; Orr, R.K. N-Phosphinoylimines: An emerging class of reactive intermediates for stereoselective organic synthesis. Synthesis 2005, 1205–1227. [Google Scholar] [CrossRef]
  25. Vilaivan, T.; Bhanthumnavin, W.; Sritana-Anant, Y. Recent advances in catalytic asymmetric addition to imines and related C=N systems. Curr. Org. Chem. 2005, 9, 1315–1392. [Google Scholar] [CrossRef]
  26. Spindler, F.; Blaser, H.-U. The highly enantioselective hydrogenation of N-diphenylphosphinylketimines with cationic Rh ferrocenyldiphosphine catalysts. Adv. Synth. Catal. 2001, 343, 68–70. [Google Scholar] [CrossRef]
  27. Wang, Y.-Q.; Zhou, Y.-G. Highly enantioselective Pd-catalyzed asymmetric hydrogenation of N-diphenylphosphinyl ketimines. Synlett 2006, 1189–1192. [Google Scholar]
  28. Wang, Y.-Q.; Lu, S.-M.; Zhou, Y.-G. Highly enantioselective Pd-catalyzed asymmetric hydrogenation of activated imines. J. Org. Chem. 2007, 72, 3729–3734. [Google Scholar] [CrossRef]
  29. Lipshutz, B.H.; Shimizu, H. Copper(I)-catalyzed asymmetric hydrosilylations of imines at ambient temperatures. Angew. Chem. Int. Ed. 2004, 43, 2228–2230. [Google Scholar] [CrossRef]
  30. Nolin, K.A.; Ahn, R.W.; Toste, F.D. Enantioselective reduction of imines catalyzed by a rhenium(V)-oxo complex. J. Am. Chem. Soc. 2005, 127, 12462–12463. [Google Scholar]
  31. Park, B.-M.; Mun, S.; Yun, J. Zinc-catalyzed enantioselective hydrosilylation of imines. Adv. Synth. Catal. 2006, 348, 1029–1032. [Google Scholar] [CrossRef]
  32. Bandini, M.; Melucci, M.; Piccinelli, F.; Sinisi, R.; Tommasi, S.; Umani-Ronchi, A. New chiral diamino-bis(tert-thiophene): An effective ligand for Pd- and Zn-catalyzed asymmetric transformations. Chem. Commun. 2007, 4519–4521. [Google Scholar]
  33. Nolin, K.A.; Ahn, R.W.; Kobayashi, Y.; Kennedy-Smith, J.J.; Toste, F.D. Enantioselective reduction of ketones and imines catalyzed by (CN-Box)ReV-oxo complexes. Chem. Eur. J. 2010, 16, 9555–9562. [Google Scholar]
  34. Graves, C.R.; Scheidt, K.A.; Nguyen, S.T. Enantioselective MSPV reduction of ketimines using 2-propanol and (BINOL)Al(III). Org. Lett. 2006, 8, 1229–1232. [Google Scholar]
  35. Yamada, T.; Nagata, T.; Sugi, K.D.; Yorozu, K.; Ikeno, T.; Ohtsuka, Y.; Miyazaki, D.; Mukaiyama, T. Enantioselective borohydride reduction catalyzed by optically active cobalt complexes. Chem. Eur. J. 2003, 9, 4485–4509. [Google Scholar]
  36. Dai-Ho, G.; Mariano, P.S. Novel photochemical-diradical cyclization methods for protoberberine alkaloid synthesis. Preparation of (±)-xylopinine and (±)-stylopine. J. Org. Chem. 1987, 52, 704–706. [Google Scholar] [CrossRef]
  37. Kwak, S.H.; Lee, S.A.; Lee, K.-I. Highly enantioselective Rh-catalyzed transfer hydrogenation of N-sulfonyl ketimines. Tetrahedron 2010, 21, 800–804. [Google Scholar] [CrossRef]
  38. Guijarro, D.; Pablo, O.; Yus, M. Ruthenium-catalysed asymmetric transfer hydrogenation of N-(tert-butanesulfinyl)imines. Tetrahedron Lett. 2009, 50, 5386–5388. [Google Scholar] [CrossRef]
  39. Guijarro, D.; Pablo, O.; Yus, M. Asymmetric synthesis of chiral primary amines by transfer hydrogenation of N-(tert-butanesulfinyl)ketimines. J. Org. Chem. 2010, 75, 5265–5270. [Google Scholar] [CrossRef]
  40. Guijarro, D.; Pablo, O.; Yus, M. Achiral β-amino alcohols as efficient ligands for the ruthenium-catalyzed asymmetric transfer hydrogenation of sulfinylimines. Tetrahedron Lett. 2011, 52, 789–791. [Google Scholar] [CrossRef]
  41. Pablo, O.; Guijarro, D.; Kovács, G.; Lledós, A.; Ujaque, G.; Yus, M. A versatile Ru catalyst for the asymmetric transfer hydrogenation of both aromatic and aliphatic sulfinylimines. Chem. Eur. J. 2012, in press. [Google Scholar]
  42. Zhou, S.; Fleischer, S.; Junge, K.; Das, S.; Addis, D.; Beller, M. Enantioselective synthesis of amines: General, efficient iron-catalyzed asymmetric transfer hydrogenation of imines. Angew. Chem. Int. Ed. 2010, 49, 8121–8125. [Google Scholar]
  43. Zhang, G.; Wen, X.; Wang, Y.; Mo, W.; Ding, C. Sodium nitrite catalyzed aerobic oxidative deoximation under mild conditions. J. Org. Chem. 2011, 76, 4665–4668. [Google Scholar]
  44. Masumoto, S.; Usuda, H.; Suzuki, M.; Kanai, M.; Shibasaki, M. Catalytic enantioselective strecker reaction of ketoimines. J. Amer. Chem. Soc. 2003, 125, 5634–5635. [Google Scholar]
  45. Almansa, R.; Guijarro, D.; Yus, M. Enantioselective addition of dialkylzinc reagents to N-(diphenylphosphinoyl)imines catalyzed by β-aminoalcohols with the prolinol skeleton. Tetrahedron: Asymmetry 2007, 18, 2828–2840, (retention times of the two enantiomers of compounds 2a and 2b). [Google Scholar]
  46. See reference 28 for the retention times of the two enantiomers of compound 2c.
  47. Almansa, R.; Guijarro, D.; Yus, M. N-Benzylprolinol: An efficient catalyst for the enantioselective addition of dialkylzinc reagents to N-(diphenylphosphinoyl)imines. Tetrahedron: Asymmetry 2007, 18, 896–899. [Google Scholar] [CrossRef]
  48. Almansa, R.; Guijarro, D.; Yus, M. Enantioselective addition of dialkylzinc reagents to N-(diphenylphosphinoyl)imines catalyzed by β-amino alcohols with the prolinol skeleton. Tetrahedron: Asymmetry 2007, 18, 2828–2840. [Google Scholar] [CrossRef]
  49. Almansa, R.; Guijarro, D.; Yus, M. Microwave-accelerated enantioselective addition of dialkylzinc reagents to N-(diphenylphosphinoyl)imines catalyzed by β-amino alcohols with the prolinol skeleton. Tetrahedron: Asymmetry 2008, 19, 1376–1380. [Google Scholar] [CrossRef]
  50. For an excellent review on the use of cis-aminoindanol in asymmetric synthesis see: Gallou, I.; Senanayake, C.H. cis-1-Amino-2-indanol in drug design and applications to asymmetric processes. Chem. Rev. 2006, 106, 2843–2874. [Google Scholar]

Share and Cite

MDPI and ACS Style

Pablo, Ó.; Guijarro, D.; Yus, M. Chiral β-Amino Alcohols as Ligands for the Ruthenium-Catalyzed Asymmetric Transfer Hydrogenation of N-Phosphinyl Ketimines. Appl. Sci. 2012, 2, 1-12. https://doi.org/10.3390/app2010001

AMA Style

Pablo Ó, Guijarro D, Yus M. Chiral β-Amino Alcohols as Ligands for the Ruthenium-Catalyzed Asymmetric Transfer Hydrogenation of N-Phosphinyl Ketimines. Applied Sciences. 2012; 2(1):1-12. https://doi.org/10.3390/app2010001

Chicago/Turabian Style

Pablo, Óscar, David Guijarro, and Miguel Yus. 2012. "Chiral β-Amino Alcohols as Ligands for the Ruthenium-Catalyzed Asymmetric Transfer Hydrogenation of N-Phosphinyl Ketimines" Applied Sciences 2, no. 1: 1-12. https://doi.org/10.3390/app2010001

Article Metrics

Back to TopTop