Next Article in Journal
Antimicrobial Activity and Spectral, Magnetic and Thermal Studies of Some Transition Metal Complexes of a Schiff Base Hydrazone Containing a Quinoline Moiety
Previous Article in Journal
Antioxidant Activity of a New Aromatic Geranyl Derivative of the Resinous Exudates from Heliotropium glutinosum Phil.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Coordination Compounds Based on 1,2,3,4-Tetrahydro-isoquinoline-3-carboxylic Acid

1
Department of Organic Chemistry, Faculty of Chemical Technology, University of Pardubice, Nám. Čs. legií 565, 53210 Pardubice, Czech Republic
2
Institute of Organic Chemistry and Biochemistry, Czech Academy of Sciences, Flemingovo nám. 2, 16610 Praha 6, Czech Republic
3
Department of Biochemical Sciences, Faculty of Pharmacy, Charles University, Heyrovského 1203, 50005 Hradec Králové, Czech Republic
*
Author to whom correspondence should be addressed.
Molecules 2007, 12(5), 1064-1079; https://doi.org/10.3390/12051064
Submission received: 27 March 2007 / Revised: 11 May 2007 / Accepted: 11 May 2007 / Published: 21 May 2007

Abstract

:
Syntheses of 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine and its coordination compounds with Cu2+, Co2+, Co3+, or Fe3+ are described. By means of 1H- and 13C-NMR spectra it was proved that 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine as well as its coordination compound with Co3+ exist in the form of a mixture of three conformers, differing in the conformations at the two amide groups present. The prepared coordination compounds were tested in the enantioselective catalysis of the nitroaldol addition of nitromethane with 2-nitrobenzaldehyde or 4-nitrobenzaldehyde, and in the Michael addition of ethyl 2-oxocyclohexanecarboxylate to but-3-en-2-one.

Introduction

The study of catalysis of enantioselective reactions continues to attract attention. Although the focus has shifted towards design of the optimum catalyst for carrying out a certain particular reaction, a number of papers are still being published, generally dealing with tests of chiral ligands with “potential ability” to catalyze enantioselective reactions. In particular, such ligands are taken from the “chiral pool” of natural homochiral amino acids, their derivatives and other compounds derived from them (chiral aminoalcohols, aminoamides etc.) [1,2,3,4]. The derivatives of (S)-1,2,3,4-tetrahydro- isoquinoline-3-carboxylic acid (Tic Acid) [5,6] (a chiral α-amino acid not found in nature) which structurally resemble anellated oxazolines [7,8] have not been studied yet in enantioselective catalytic reactions.

Results and Discussion

2,6-bis[((3S)-3-(Methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine (2) was prepared by a reaction of hydrochloride of ester of Tic acid (1) and pyridine-2,6-bis-(carbonyl chloride) in the presence of triethylamine (Scheme 1).
Scheme 1.
Scheme 1.
Molecules 12 01064 g008
The structure of the obtained 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine (2) was investigated by (i) quantum chemical calculations at the HF/6-31G(d,p) level [9,10] and (ii), 1H- and 13C-NMR spectroscopy. Calculations showed that compound 2 can exist in three isomeric forms, namely in two symmetrical forms and one unsymmetrical one. This relatively extensive system had to be described using a simpler HF/6-31G(d,p) calculation model [9,10]. The calculation indicates the same probabilities of formation of rotameric forms A and B (Figure 1) during formation of amide bond.
Formation of two amide bonds results in the creation of forms AA, BB, AB and BA with comparable probability. Forms AB and BA are identical and will be referred to henceforth as form AB+BA. The forms discussed are depicted in Figure 1. The unsymmetrical form AB+BA is formed in a double amount as compared with forms AA or BB. From the standpoint of symmetry, form AB+BA belongs to the point group C1 and would exhibit two sets of signals in both 1H- and 13C-NMR spectra. The intensity of both sets of multiplets in the 1H NMR spectrum should be comparable with the intensity of multiplets of forms AA and BB, because forms AA and BB belong to point group C2 and will exhibit only one set of signals for each of the forms in 1H-NMR spectrum. According to the quantum-chemical simulations, the 1H-NMR spectrum should contain 4 sets of multiplets with similar intensities.
Figure 1. Structures of forms of compound 2 optimized at the HF/6-31G(d,p) level [9,10].
Figure 1. Structures of forms of compound 2 optimized at the HF/6-31G(d,p) level [9,10].
Molecules 12 01064 g001
Figure 2. 1H-NMR spectrum of compound 2 in CDCl3 (500 MHz).
Figure 2. 1H-NMR spectrum of compound 2 in CDCl3 (500 MHz).
Molecules 12 01064 g002
This prediction fully corresponds with the experimental NMR spectrum of compound 2 (Figure 2), in which there really are four sets of signals of comparable intensities for the individual protons. This situation can be easily observed on the multiplets of protons H(3) and H(1) of the tetrahydro-isoquinoline skeleton (δ 4.6–5.6) and the signals of the OCH3 groups (δ 3.4–3.7). No mutual transformation of individual forms on the NMR time scale was observed up to 50 °C.
The formation of rotamers due to hindered rotation around the amide bond C–N in derivatives of 1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid was also observed in the case of the corresponding N-chlorocarbonyl [5] and N-acetyl derivatives. Methyl N-acetyl-1,2,3,4-tetrahydroisoquinoline-3-carboxylate (3) was prepared from (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid, or from the racemic acid (Scheme 2).
Scheme 2.
Scheme 2.
Molecules 12 01064 g009
The optical purity of the non-racemic product determined by HPLC on chiral column by comparison with the racemic substance was 94.6 % (Figure 3). The proportion of conformers in the (S)-enantiomer is ca 5:2 according to the 1H-NMR spectrum (Figure 4).
Figure 3. Chiral HPLC separation of the enantiomers of acetyl derivative 3. Upper chromatogram represents separation of racemic mixture and lower the enantiomeric purity of the (S)-enantiomer 3a (e.e. 94.6%). For separation conditions see Experimental.
Figure 3. Chiral HPLC separation of the enantiomers of acetyl derivative 3. Upper chromatogram represents separation of racemic mixture and lower the enantiomeric purity of the (S)-enantiomer 3a (e.e. 94.6%). For separation conditions see Experimental.
Molecules 12 01064 g003
Figure 4. 1H-NMR spectrum of compound 3a in CDCl3 (500 MHz).
Figure 4. 1H-NMR spectrum of compound 3a in CDCl3 (500 MHz).
Molecules 12 01064 g004
The same conclusions were obtained based on the geometry optimization of acetyl derivative carried out at the B3LYP/TZVP level [11,12,13] (Figure 5). In order to achieve better correlation of the results with the 1H-NMR spectrum the calculation included the solvent (chloroform) effect by means of the polarised continuum method (PCM) [14]. Out of the pair of optimized structures (Figure 5) structure A was assigned to predominating rotameric form in NMR spectrum on the basis of calculated energies.
Figure 5. Structures of compound 3a optimized at B3LYP/TZVP + PCM [11,12,13,14] level.
Figure 5. Structures of compound 3a optimized at B3LYP/TZVP + PCM [11,12,13,14] level.
Molecules 12 01064 g005
In the case of compounds having an RNHCO– group attached to nitrogen atom of tetra-hydroisoquinoline skeleton [6] the 1H-NMR spectrum only exhibits the presence of only one of the two possible rotamers, due to the existence of strong intramolecular hydrogen bond (Figure 6).
Figure 6. Structure of methyl-(3S)-N-[(1S)-1-methylbenzyl]carbamoyl-1,2,3,4-tetra- hydroisoquinoline-3-carboxylate [6] optimized at B3LYP/TZVP + PCM [11,12,13,14] level.
Figure 6. Structure of methyl-(3S)-N-[(1S)-1-methylbenzyl]carbamoyl-1,2,3,4-tetra- hydroisoquinoline-3-carboxylate [6] optimized at B3LYP/TZVP + PCM [11,12,13,14] level.
Molecules 12 01064 g006
The quantum-chemical calculation results clearly indicate that the electron density at the tetra- hydroisoquinoline residue nitrogen atom is noticeably higher in the molecule of compound 2 than in that of the acetyl derivative 3a. Due to the steric demands of the Tic residues in the molecule of 2 these residues are deviated, which disturbs the planarity and leads to partial loss of conjugation in the N-CO-Py grouping. Hence, according to these calculations compound 2 could operate as a tridentate ligand and coordinate with transition metals. This presumption was later confirmed experimentally.
The coordination compounds were prepared by a reaction of (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid or 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]- pyridine (2) with a transition metal salt – Cu2+, Co2+, Co3+, Fe3+ (chlorides, acetates) – in dry methanol (Scheme 3) [15]. The stoichiometric composition of the coordination compounds was established on the basis of elemental analyses. (S)-1,2,3,4-Tetrahydroisoquinoline-3-carboxylic acid is an bidentate ligand and coordinates with Cu2+ or Co2+ at a ratio of 2:1. 2,6-bis[((3S)-3-(Methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)­carbonyl]pyridine (2) is an tridentate ligand, as described for 2,6-bis-(oxazolyl)pyridines [4,16] or 2,6-bis(imidazolyl)pyridines [17,18], and it coordinates with Cu2+, Co2+, Co3+, Fe3+ in a ratio of 1:1 (Scheme 3 and Scheme 4).
Scheme 3.
Scheme 3.
Molecules 12 01064 g010
Scheme 4.
Scheme 4.
Molecules 12 01064 g011
Of all the coordination compounds prepared only proved suitable for NMR measurements, namely the 2,6-bis-[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine Co3+ complex. Both its 1H- and 13C-NMR spectra were of adequate quality with slightly broadened signals (Figure 7). The 1H-NMR spectrum of this coordination compound resembles that of the free ligand. From the 13C-NMR spectrum it is obvious that it also corresponds to a mixture of three compounds of the types AA, BB, AB+BA, which are present at roughly equimolecular ratios (the tetrad of signals for corresponding carbons in the spectrum).
Figure 7. 500 MHz H-H COSY spectrum of the coordination compound 5d (Co3+ 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine) in DMSO-D6.
Figure 7. 500 MHz H-H COSY spectrum of the coordination compound 5d (Co3+ 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine) in DMSO-D6.
Molecules 12 01064 g007
The ability of the prepared coordination compounds to catalyze enantioselective reactions was tested on the Henry nitroaldol addition of nitromethane with 2-nitro- or 4-nitrobenzaldehyde (Scheme 5), under the conditions described in [19], and on the Michael addition reaction of ethyl 2-oxocyclo-hexane­carboxylate with but-3-en-2-one (Scheme 6), under the condition described in [20,21,22,23].
Scheme 5.
Scheme 5.
Molecules 12 01064 g012
In the nitroaldol addition of nitromethane with 4-nitrobenzaldehyde the highest enantioselective efficiency was observed with the coordination compound 5a containing Cu2+. Application of this substance at 0 °C gave (R)-2-nitro-1-(4-nitrophenyl)ethanol (6) in an enantiomeric excess up to 61.7 % (Table 1). (R)-2-Nitro-1-(2-nitrophenyl)ethanol (7) resulted in an enantiomeric excess of only 12.3 % when using the same catalyst. The sterically more demanding nitroaldol addition with 2-nitro-benzaldehyde obviously prevents the optimum steric interaction of the chiral catalyst with the substrate. The other coordination compounds prepared exhibited only very low levels of enantioselectivity (Table 1 and Table 2). While the reaction of 4-nitrobenzaldehyde with nitromethane catalyzed with coordination compounds 5a and 5c gave a reasonable excess of (R)-2-nitro-1-(4-nitrophenyl)ethanol, the catalysis with coordination compound 4b gave just a low excess, but of (S)-2-nitro-1-(4-nitrophenyl)ethanol.
Scheme 6.
Scheme 6.
Molecules 12 01064 g013
In the Michael addition reaction of ethyl 2-oxocyclohexanecarboxylate with but-3-en-2-one catalysed by coordination compound 5e the required product was obtained in high chemical yield but with an enantiomeric excess of only 7.2 % (Table 3). This lower efficiency is obviously due to the far higher steric demands in the surroundings of coordination centres than are those of, e.g., the catalysts based on 2,6-bis(oxazolyl)pyridines [15]. Another problem lies in the fact that the coordination compounds derived from substance 2 are present (obviously all of them) in three rotameric forms (Figure 1), and it is not quite clear whether these forms can be transformed into one another during the interaction with the molecules undergoing the catalysed reaction, neither is it known which of the forms is active in the catalysed reaction.
Table 1. Nitroaldol addition of nitromethane with 4-nitrobenzaldehyde catalysed by coordination compounds 5a, 5c and 4a, b.
Table 1. Nitroaldol addition of nitromethane with 4-nitrobenzaldehyde catalysed by coordination compounds 5a, 5c and 4a, b.
Coordination compoundReaction timeTemperatureConversionm.p. of productEnantiomer excess (R)
5a10 days20 °C70%82–84 °C54.5%
5a15 days0 °C30%82–84 °C61.7%
5c4 days20 °C100%81–83 °C6.1%
4a20 days20 °C0%
4b20 days20 °C50%81–83 °C7.3%*
*(S) enantiomer
Table 2. Nitroaldol addition of nitromethane with 2-nitrobenzaldehyde catalysed by coordination compound 5a.
Table 2. Nitroaldol addition of nitromethane with 2-nitrobenzaldehyde catalysed by coordination compound 5a.
Coordination compoundReaction timeTemperatureConversionm.p. of productEnantiomer excess (R)
5a15 days20 °C50%80–82 °C12.3%
Table 3. Michael addition reaction of ethyl 2-oxocyclohexanecarboxylate with but-3-en-2-onecatalysed by coordination compounds 5b and 5e.
Table 3. Michael addition reaction of ethyl 2-oxocyclohexanecarboxylate with but-3-en-2-onecatalysed by coordination compounds 5b and 5e.
Coordination compoundReaction timeTemperatureConversionEnantiomer excess ( S)
5b20 days20 °C0%-
5e1 day20 °C100%7.2%
5e20 days–25 °C60%7.1%

Conclusions

Derivatives of (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid were prepared by N-acylation with acetic anhydride or 2,6-bis(chlorocarbonyl)pyridine. The N-acetyl derivative was obtained in an optical purity of 94.6 % (determined by HPLC). In solution (CDCl3) it exists in the form of two rotamers, which are not mutually interconverted on the NMR time scale up to ca 50 °C. On the basis of quantum-chemical calculations it was predicted (and then confirmed by both 1H- and 13C-NMR spectra) that 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine exists in solution as a mixture of three rotameric forms differing in the conformations at the two amide C–N bonds. The rotameric forms are not mutually interconverted on the NMR time scale, not even at 50 °C. Coordination compounds were prepared from (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid (4) or 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine with the cations Cu2+, Co2+, Co3+ and Fe3+ (5a-e). The stoichiometry of these substances was determined on the basis of elemental analyses. (S)-1,2,3,4-Tetrahydroisoquinoline-3-carboxylic acid is coordinated with metals at the ratio of 2:1 and bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydro- isoquinolin-2-yl)carbonyl]pyridine at the ratio of 1:1. The Co3+ complex of bis[((3S)-3-(methoxy-carbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine also exists in solution as a mixture of three rotamers. The ability of the prepared coordination compounds to catalyze the enantioselective Henry reaction and Michael addition was low. The highest enantiomeric excess of 61.7 % with a conversion of only 30 % was achieved in the reaction of nitromethane with 4-nitrobenzaldehyde catalyzed with the coordination compound 5a. Uncertainty exists as to which of the three forms of the catalyst is active, or whether the individual forms can be mutually interconverted during interaction with the substrate of reaction.

Experimental

General

The NMR spectra were measured at 298 K with a Bruker AVANCE 500 spectrometer equipped with 5 mm broadband probe at the frequencies of 500.13 MHz (1H) and 125.77 MHz (13C) and with a Bruker AMX 360 spectrometer at the frequencies 360.14 MHz (1H) and 90.57 MHz (13C) in CDCl3 and DMSO-D6 respectively. Spectra were calibrated on TMS (in CDCl3) or on the central signal of the solvent multiplet in DMSO-D6 (δ 2.55, and 39.6 respectively). J values are given in Hertz. The 13C NMR spectra were measured in standard way and by means of the APT pulse sequence. The proton signals were assigned with the help of H-H COSY pulse sequence. Optical purity was determined by chiral HPLC. HPLC system consisted of a Spectra Series P200 gradient pump (Fremont, CA, USA), a HP 1100 Series autosampler, a HP 1100 Series thermostated column compartment from Hewlett Packard (Waldbronn, Germany), and a SPD-10AVP UV-Vis detector from Shimadzu (Prague, Czech Republic). The enantiomers of the compound 3 were measured at 209 nm (Figure 6). Data from chromatographic runs were processed using a chromatography station for Windows CSW (version 1.7) software from DataApex (Prague, Czech Republic). Separation of the respective enantiomers was performed using a 250 ° 4.6 mm OD-R Chiralcel column from Daicel Chemical Industries (Tokyo, Japan). The mobile phase was prepared by mixing buffer (0.3 M sodium perchlorate, pH 3.0 set by HClO4) with acetonitrile 50/50 (v/v). HPLC separation was performed at 25°C with a flow rate of 0.8 mL/min. Melting points were determined with a Kofler hot stage microscope and were not corrected. The microanalyses were performed on a FISONS EA 1108 CHNS automatic analyser. Optical rotations were measured on PERKIN ELMER 341 Polarimeter at λ 589.3 nm and 298 K, concentration c is given in g/100 mL. The starting material (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid (CMS Chemicals LTD), mp 331 °C (decomp) was 99.1% pure, according to titration with HClO4, and had Molecules 12 01064 i001 = –175.8° (c 1.1N NaOH, aq) [ref. [24] gives Molecules 12 01064 i001 = –177.4° (c 1.1N NaOH, aq)].

Hydrochloride of methyl (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylate (1)

This compound [5] was prepared from (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid and SOCl2 in dry methanol by a known procedure [24,25] in practically quantitative yield; m.p. 248–250 °C (decomp), from methanol-diethyl ether; ref. [24] gives m.p. 250-255 °C (decomp.) from the same solvent mixture. The product recrystallized from a mixture of chloroform-diethyl ether melts at 261–263 °C (decomp.); Molecules 12 01064 i001 = –155.1° (c 1, CHCl3), Molecules 12 01064 i001 = –128.2° (c 1, CH3OH); ref. [24] gives Molecules 12 01064 i001 = –104.1° (c 1, CH3OH).

Methyl (3S)-N-acetyl-1,2,3,4-tetrahydroisoquinoline-3-carboxylate (3a)

Prepared by a method analogous to one described in the literature [26]. A suspension of the hydrochloride of methyl (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylate (1) (5 g, 22.0 mmol) and anhydrous sodium acetate (1.25 g, 15.3 mmol) in acetic anhydride (11.35 mL, 120.4 mmol) was stirred and heated at 50-60 °C for 1 h, whereupon the reaction mixture was poured into water (50 mL) and immediately extracted with CHCl3 (3 × 25 mL). The chloroform extract was concentrated and the solution obtained was extracted with water (50 mL). Drying of the chloroform solution with anhydrous Na2SO4 and removal of the solvent by distillation gave an oily crude product, which was further purified by flash chromatography (CH3OH - silica gel, 60 μm). Recrystallization from cyclohexane with charcoal gave 4.6 g of a white crystalline solid (90% of theory), m.p. 93–96 °C; Molecules 12 01064 i001 = +35.6° (c = 1, CHCl3). The analysis by 1H-NMR showed that the product is a mixture of two isomers (according to the integral intensities, the proportion of isomers I/II in the isolated mixture is ca 5:2). Isomer I: 1H-NMR (CDCl3): 7.22-7.11, multiplet, 4H(arom.); 4.72 H(1a), 4.66 H(1b), AB quartet, 2J(H(1a),H(1b))=15.8 Hz, 2×1H; 5.49 H(3), 3.25 H(4a), 3.11 H(4b), AMX system, 3J(H(3),H(4a))=3.5 Hz, 3J(H(3),H(4b))=6.3 Hz, 2J(H(4a),H(4b))=15.9 Hz, 3×1H; 3.61, s, OCH3, 3H; 2.25, s, CH3, 3H; 13C-NMR (CDCl3): 171.39 a 170.61 (CO), 132.10 a 131.97 (arom, 2×Cq), 128.47, 127.15, 126.88, 126.00 (arom, 4×CH), 52.26 (CHCO), 51.06 (OCH3), 46.31 (ArCH2N), 30.80 (ArCH2C), 21.91 (CH3); Isomer II: 1H-NMR (CDCl3): 7.22-7.11, multiplet, 4H(arom.); 4.93 H(1a), 4.49 H(1b), AB quartet, 2J(H(1a),H(1b))=17.3 Hz, 2×1H; 4.78 H(3), 3.34 H(4a), 3.19 H(4b), AMX system, 3J(H(3),H(4a))=2.8 Hz, 3J(H(3),H(4b))=6.0 Hz, 2J(H(4a),H(4b))=15.6 Hz, 3×1H; 3.60, s, OCH3, 3H; 2.16, s, CH3, 3H; 13C-NMR (CDCl3): 170.94 a 170.54 (CO), 132.59 a 131.00 (arom, 2×Cq), 128.00, 127.06, 126.74, 126.58 (arom, 4×CH), 55.67 (OCH3), 52.61 (CHCO), 43.34 (ArCH2N), 31.82 (ArCH2C), 21.80 (CH3); Anal. calcd. for C13H15NO3 (233.27): C 66.94 H 6.48 N 6.00%, found: C 66.68 H 6.61 N 5.93%.

Methyl (3±)-N-acetyl-1,2,3,4-tetrahydroisoquinoline-3-carboxylate (3b)

Prepared by the same procedure as 3a from (3±)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid in a yield of 92%. After flash chromatography the product was isolated as an oily substance. Its NMR analysis showed that the product is a mixture of two isomers (according to the integral intensities, the proportion of isomers in the isolated mixture is ca 5:2). The NMR spectra of racemic compound are identical with those of methyl (3S)-N-acetyl-1,2,3,4-tetrahydroisoquinoline-3-carboxylate (3a).

2,6-bis[((3S)-3-(Methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine (2)

Prepared in analogy with the procedure described in ref. [27] for the reaction of pyridine-2,6-bis(carbonyl chloride) with amino acids. A suspension was prepared from dry CH2Cl2 (200 mL) and the hydrochloride of methyl (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylate (1), (2.27 g, 10 mmol). At a temperature of –25 °C the suspension was treated with dry NEt3 (1.01 g, 10 mmol). After 5 min stirring, the mixture was treated with a solution of pyridine-2,6-bis(carbonyl chloride) (1.02 g, 5 mmol) in dry CH2Cl2 (50 mL) gradually added dropwise with vigorous stirring at –20 °C, whereupon another dry NEt3 (1.01 g, 10 mmol) was added. The reaction mixture was stirred further at –20 °C for a period of 3 h, and then at r.t. for 12 h. The obtained yellowish solution was extracted with H2O (2 × 100 mL), then with 5% aqueous hydrochloric acid (50 mL) and finally with H2O (2 × 50 mL). After drying and removal of the solvent by distillation an oily product was isolated. Its recrystallisation from a cyclohexane-hexane mixture gave 2.19 g (86 %) of a white solid melting at 61–64 °C. Its NMR analysis showed that the product is a mixture of three isomeric forms – two symmetrical forms, and one unsymmetrical form present in a double amount, hence the NMR spectrum exhibits 4 sets of signals of comparable intensities (for the proton at the 4-position of pyridine ring only 3 multiplets for 4H). 1H- NMR (CDCl3) (see Figure 2): 8.02-7.78, multiplet (pyridine); 7.25-6.94, multiplet (arom.-isoquinoline);5.55-5.29, 4 × dd (H(3)); 5.22-4.68, 4 × AB q (H(1a), H(1b)); 3.70, 3.69, 3.51 a 3.41, 4 × s, 4×(OCH3), 3.36-3.23, multiplet (H(4a), H(4b)); 13C-NMR (CDCl3): 171.02, 171.00, 170.79 a 170.66 4×(CCOO), 168.12, 167.83, 167.69 a 167.53 4×(CCON), not given 35×C(arom.), 56.17, 55.88, 52.81 a 52.14 4×(OCH3), 52.47, 52.47, 52.43 a 52.43 4×(CHCO), 47.57, 47.43, 44.10 a 43.82 4×(ArCH2N), 31.49, 31.25, 30.73 a 30.64 4×(ArCH2C); Anal. calcd. for C29H27N3O6 (513.55): C 67.83 H 5.30 N 8.18%, found: C 67.56 H 5.45 N 7.95%.

Coordination compounds of (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid with transition metals

Coordination compounds 4a,b were prepared according to a procedure described in ref. [15]. A suspension was prepared in methanol from equimolecular amounts of (S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid and the transition metal salt. After 12 h vigorous stirring, the precipitate formed was collected by suction, and then washed with methanol and ether.
Table 4. Microanalysis data for compounds 4a,b
Table 4. Microanalysis data for compounds 4a,b
SaltMolecular formula (m.w)Elemental composition - Calculated / Found
C (%)H (%)N (%)m.p.(°C)
4aCu(OAc)2C20H20N2O4Cu (415.93)57.75/57.924.85/4.656.74/6.71360–362
4bCo(OAc)2C20H20N2O4Co (411.32)58.40/58.264.90/5.156.81/6.62360–361

Coordination compounds of 2,6-bis[(3S)-3-methoxycarbonyl-1,2,3,4-tetrahydroisoquinolin-2-yl)-carbonyl]pyridine with transition metals 5a-e

Equimolecular amounts of 2,6-bis[((3S)-3-(methoxycarbonyl)-1,2,3,4-tetrahydroisoquinolin-2-yl)carbonyl]pyridine and transition metal salt were dissolved in dry methanol. The corresponding coordination compound was obtained after evaporation of the solvent and washing with ether and hexane. Cobalt(III) acetate was prepared by oxidation of cobalt(II) acetate with aqueous solution of peroxyacetic acid [28].
Table 5. Microanalysis data for compounds 5a-e.
Table 5. Microanalysis data for compounds 5a-e.
SaltMolecular formula (m.w.)Elemental composition - Calculated / Found
C (%)H (%)N (%)Cl (%)m.p. (°C)
5aCu(OAc)2C33H33N3O10Cu (695.18)57.02/56.764.78/4.536.04/5.95235–238
5bCoCl2C29H27N3O6Cl2Co (643.39)54.14/54.494.23/4.626.53/6.3811.02/10.89149–151
5cCo(OAc)2C33H33N3O10Co (690.57)57.40/57.134.82/4.656.08/5.87198–201
5dCo(OAc)3C35H36N3O12Co (749.61)56.08/55.744.84/4.535.61/5.32218–220
5eFeCl3C29H27N3O6Cl3Fe (675.75)51.55/51.784.03/4.096.22/6.1715.74/15.67280–282
The structure of coordination compound 5d containing Co(III) can be studied by means of NMR. These spectra clearly show that the product is a mixture of three isomeric forms – two symmetrical ones, and one unsymmetrical, the latter being present in a double amount. Hence, the NMR spectrum exhibits 4 sets of signals of comparable intensities (for the proton at the 4-position of pyridine ring only 3 multiplets for 4 protons). 1H-NMR (DMSO-D6) (see Figure 7): 8.15-7.77, multiplet (pyridine); 7.37-7.16, multiplet (arom.-isoquinoline); 5.43-5.25, 4 × multiplet (H(3)); 5.13-4.57 4 × AB q (H(1a), H(1b)); 3.68, 3.67, 3.50 a 3.50, 4 × s, (OCH3), 3.39-3.20, multiplet (H(4a), H(4b)); 2.49, s, (CH3COO); 13C-NMR (DMSO-D6): 184.56, (CH3COO), 171.52, 171.49, 171.40 a 171.39 4×(CCOO), 167.88, 167.85, 167.82 a 167.68 4×(CCON), not given 35×C (arom.), 56.22, 56.12, 52.78 a 52.76 4×(OCH3), 53.24, 53.03, 52.83 a 52.76 4×(CHCO), 47.51, 47.20, 43.85 a 43.74 4×(ArCH2N), 31.53, 31.37, 30.74 a 30.71 4×(ArCH2C), 26.92 (CH3COO).

2-Nitro-1-(4-nitrophenyl)ethanol (6)

The reaction of nitromethane with 4-nitrobenzaldehyde catalyzed by the coordination compound (Scheme 5) was carried out by the known procedure [19]: a solution of nitromethane (1.4 mL, 25 mmol) and 4-nitrobenzaldehyde (0.38 g, 2.5 mmol) in dry ethanol (2 mL) was treated with coordination compound (5 mol %, 0.125 mmol). The reaction course was monitored by means of TLC (silica gel – ethyl acetate-hexane 1:4 by vol.). After keeping at the chosen temperature for a chosen time interval, the reaction was stopped by evaporating the solvent in vacuum without heating. The evaporation residue was treated with CoCl2 (0.05 g, 0.385 mmol) in ethanol (10 mL) in order to transform any possible uncoordinated ligand present into the coordination compound. Methanol was evaporated under vacuum, and the residue was dissolved in ether. The coordination compounds and the unreacted CoCl2 were removed from the ether solution by flash chromatography (silica gel 60 μm – ether). The ethereal filtrate was then extracted with 10% aqueous solution of sodium sulphite (2 × 20 mL) and with H2O (1 × 10 mL). This procedure removed all unreacted aldehyde. Drying and removal of ether by evaporation without heating gave pure 2-nitro-1-(4-nitrophenyl)ethanol; m.p. 82–84 °C. The enantiomeric excess was then calculated from chemical purity and optical rotation. The reaction of nitromethane with 2-nitrobenzaldehyde was carried out in the same way as that with 4-nitrobenzaldehyde above to afford 2-nitro-1-(2-nitrophenyl)ethanol (7); m.p. 80–82 °C, o.r. +31.4 (c = 1, CH2Cl2).

Ethyl 2-oxo-1-(3-oxobutyl)cyclohexanecarboxylate (8)

The reaction of ethyl 2-oxocyclohexanecarboxylate with but-3-en-2-one catalyzed with the coordination compounds (Scheme 6) was carried out by known procedures [20,21,22,23]. A solution (or suspension) of ethyl 2-oxocyclohexanecarboxylate (0.16 mL, 1 mmol) and coordination compound (5 mol %, 0.05 mmol) in CH2Cl2 (1 mL) was vigorously stirred and treated with but-3-en-2-one (0.20 mL, 2 mmol). The reaction course was monitored by means of TLC (silica gel – ether-hexane 1:2 by vol.). After keeping at the chosen temperature for a chosen time interval, the reaction was stopped by evaporation of solvent without heating, whereupon the evaporation residue was dissolved in ether. The coordination compound was removed from the resulting ethereal solution by means of flash chromatography (silica gel 60 μm – ether). The ether solvent was evaporated without heating to give pure ethyl 2-oxo-1-(3-oxobutyl)cyclohexanecarboxylate as an oily substance. Its chemical purity was checked by means of liquid chromatography and its enantiomeric purity by measuring optical rotation of the product. The enantiomeric excess was then calculated from chemical purity and optical rotation.

Acknowledgments

The authors thank to the Ministry of Education, Youth and Sports of the Czech Republic for financial support; project MSM 0021627501 (V.M., P.J. and M.S.) and LC512 (P.N.).

References

  1. Ojima, I. (Ed.) “Catalytic Asymmetric Synthesis”; Wiley: New York, 2000.
  2. Seyden-Penne, J. (Ed.) “Chiral Auxiliaries and Ligands in Asymmetric Synthesis”; Wiley: New York, 1995.
  3. Mulzer, J.; Waldmann, H. (Eds.) “Organic Synthesis Highlights III”; Wiley-VCH: Weinheim, 1998.
  4. Desimoni, G.; Faita, G.; Quadrelli, P. Pyridine-2,6-bis(oxazolines), helpful ligands for asymmetric catalysts. Chem. Rev. 2003, 103, 3119–3154. [Google Scholar] [CrossRef]
  5. Jansa, P.; Macháček, V.; Bertolasi, V. Derivatives of 1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid. Heterocycles 2006, 68, 59–69. [Google Scholar] [CrossRef]
  6. Jansa, P.; Wsól, V.; Bertolasi, V.; Macháček, V. Hydantoins and thiohydantoins derived from 1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid. Heterocycles 2006, 68, 2527–2547. [Google Scholar] [CrossRef]
  7. Davies, I.W.; Gerena, L.; Lu, N.; Larsen, R.D.; Reider, P.J. Concise synthesis of conformationally constrained pybox ligands. J. Org. Chem. 1996, 61, 9629–9630. [Google Scholar] [CrossRef]
  8. Jiang, M.; Dalgarno, S.; Kilner, C.A.; Halcrow, M.A.; Kee, T.P. Chiral bis(oxazoline) complexes. Synthesis, structure and application in catalytic phospho-transfer. Polyhedron 2001, 20, 2151–2162. [Google Scholar] [CrossRef]
  9. Roothan, C.C.J. New developments in molecular orbital theory. Rev. Mod. Phys. 1951, 23, 69–89. [Google Scholar] [CrossRef]
  10. Ditchfield, R.; Hehre, W.J.; Pople, J.A. Self-consistent molecular-orbital methods. IX. An extended Gaussian-type basis for molecular-orbital studies of organic molecules. J. Chem. Phys. 1971, 54, 724–728. [Google Scholar] [CrossRef]
  11. Lee, C.; Yang, W.; Parr, R. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B: Condens. Matter 1988, 37, 785–789. [Google Scholar] [CrossRef]
  12. Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  13. Schaefer, A.; Huber, C.; Ahlrichs, R. Fully optimized constracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829–5835. [Google Scholar]
  14. Cossi, M.; Barone, V.; Mennucci, B.; Tomasi, J. Ab initio study of ionic solutions by a polarizable continuum model. Chem. Phys. Lett. 1998, 286, 253–260. [Google Scholar] [CrossRef]
  15. Koch, D.; Polborn, K.; Sunkel, K.; Beck, W. N,O-chelate complexes of Ru(II), Ir(III), Pt(II) and Cu(II) with tetrahydroisoquinolinic acid. Molecular structures of [(Et3P)(Cl)M(OOC-CH-CH2-C6H4-CH2-NH], M = Pd, Pt. Inorg. Chim. Acta 2002, 334, 365–370. [Google Scholar] [CrossRef]
  16. Nishiyama, H.; Yamaguchi, S.; Kondo, M.; Itoh, K. Electronic substituent effect on nitrogen ligands in catalytic asymmetric hydrosilylation of ketones – chiral 4-substituted bis(oxazolinyl)pyridines. J. Org. Chem. 1992, 57, 4306–4309. [Google Scholar]
  17. Drabina, P.; Hanusek, J.; Jirásko, R.; Sedlák, M. Iron(II) complexes of 2,6-bis(5-alkyl-1,5-dimethyl-4,5-dihydro-1H-imidazol-4-on-2-yl)pyridine ligands. Synthesis, characterization and solvolytic stability. Trans. Metal. Chem. 2006, 31, 1052–1056. [Google Scholar] [CrossRef]
  18. Turský, M.; Nečas, D.; Drabina, P.; Sedlák, M.; Kotora, M. Rhodium catalyzed deallylation of allylmalonates and related compounds. Organometallics 2006, 25, 901–907. [Google Scholar] [CrossRef]
  19. Evans, A.D.; Seidel, D.; Rueping, M.; Lam, H.W.; Shaw, T.J.; Downey, C.W. A new copper acetate – bis(oxazoline) – catalyzed, enantioselective Henry reaction. J. Am. Chem. Soc. 2003, 125, 12692–12693. [Google Scholar]
  20. Christoffers, J.; Mann, A. New auxiliaries for copper – catalyzed asymmetric Michael reactions: Generation of quartenary stereocenter at room temperature. Chem. Eur. J. 2001, 7, 1014–1027. [Google Scholar] [CrossRef]
  21. Christoffers, J.; Mann, A. Novel diamino and diimino thioethers – Chiral tridentate ligands for asymmetric Michael reactions? Eur. J. Org. Chem. 1999, 6, 1475–1479. [Google Scholar] [CrossRef]
  22. Christoffers, J.; Mann, A.; Pickardt, J. Synthesis of chiral tridentate oxazolines with thioether and heteroaryl donor group and their application in the catalysis of asymmetric Michael reactions. Tetrahedron 1999, 55, 5377–5388. [Google Scholar]
  23. Bolm, C.; Legros, J.; LePaih, J.; Zani, L. Iron – catalyzed reactions in organic synthesis. Chem. Rev. 2004, 14, 6217–6254. [Google Scholar]
  24. Hayashi, K.; Ozaki, Y.; Nunami, K.; Yoneda, N. Facile preparation of optically pure (3R) and (3S)-1,2,3,4-tetrahydroisoquinoline-3-carboxylic acid. Chem. Pharm. Bull. 1983, 31, 312–314. [Google Scholar]
  25. Grunewald, G.L.; Sall, D.J.; Monn, J.A. Synthesis and evaluation of 3-substituted analogs of 1,2,3,4-tetrahydroisoquinoline as inhibitors of phenylethanolamine N-methyltransferase. J. Med. Chem. 1988, 31, 824–830. [Google Scholar]
  26. Severino, B.; Santagada, V.; Frecentese, F.; Perissutti, E.; Terracciano, F.; Cirillo, D.; Salvadori, S.; Balboni, G.; Caliendo, G. Synthesis of N-alpha-Fmoc N,N-bis-Boc-5-, 6- and 8-guanyl-1,2,3,4-tetrhadroisoquinoline-3-carboxylic acid (5-GTIC, 6-GTIC and 8-TIC). Synthesis 2004, 18, 3011–3016. [Google Scholar]
  27. Amr, A.E.-G.; El-Salam, A.O.I.; Attia, A.E.-H.; Stibor, I. Synthesis of new potentialbis-intercalators on chiral pyridine-2,6-dicarboxamides. Collect. Czech. Chem. Commun. 1999, 64, 288–298. [Google Scholar] [CrossRef]
  28. Lyčka, A.; Holeček, J. N-15, C-13 and H-1 NMR spectra of three 2:1 cobalt(III) complexes of 1-(2-carboxyphenyl)azo-2-naphthol. Dyes Pigments 2003, 57, 115–119. [Google Scholar] [CrossRef]
  • Sample availability: Contact the author (P.J.).

Share and Cite

MDPI and ACS Style

Jansa, P.; Macháček, V.; Nachtigall, P.; Wsól, V.; Svobodová, M. Coordination Compounds Based on 1,2,3,4-Tetrahydro-isoquinoline-3-carboxylic Acid. Molecules 2007, 12, 1064-1079. https://doi.org/10.3390/12051064

AMA Style

Jansa P, Macháček V, Nachtigall P, Wsól V, Svobodová M. Coordination Compounds Based on 1,2,3,4-Tetrahydro-isoquinoline-3-carboxylic Acid. Molecules. 2007; 12(5):1064-1079. https://doi.org/10.3390/12051064

Chicago/Turabian Style

Jansa, Petr, Vladimír Macháček, Petr Nachtigall, Vladimír Wsól, and Markéta Svobodová. 2007. "Coordination Compounds Based on 1,2,3,4-Tetrahydro-isoquinoline-3-carboxylic Acid" Molecules 12, no. 5: 1064-1079. https://doi.org/10.3390/12051064

Article Metrics

Back to TopTop