Next Article in Journal
Processing of Polymer Nanocomposites Reinforced with Polysaccharide Nanocrystals
Previous Article in Journal
An Expeditious Synthesis of [1,2]Isoxazolidin-5-ones and [1,2]Oxazin-6-ones from Functional Allyl Bromide Derivatives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of a New Series of N,N’-Dimethyltetrahydrosalen (H2 [H2Me]salen) Ligands by the Reductive Ring-Opening of 3,3'-Ethylene-bis(3,4-dihydro-6-substituted-2H-1,3-benzoxazines)

by
Augusto Rivera
*,
Jicli José Rojas
,
Jairo Salazar-Barrios
,
Mauricio Maldonado
and
Jaime Ríos-Motta
Departamento de Química, Universidad Nacional de Colombia, Bogotá, Colombia, A.A. 14490, Colombia
*
Author to whom correspondence should be addressed.
Molecules 2010, 15(6), 4102-4110; https://doi.org/10.3390/molecules15064102
Submission received: 26 April 2010 / Revised: 7 May 2010 / Accepted: 11 May 2010 / Published: 7 June 2010

Abstract

:
A new series of N,N'-bis(2'-hydroxy-5'-substituted-benzyl)-N,N´-dimethylethane-1,2-diamines (N,N'-dimethyltetrahydrosalen) ligands were prepared in good yield by reduction of the respective 3,3'-ethylene-bis(3,4-dihydro-6-substituted-2H-1,3-benzoxazine) precursors with sodium borohydride . The ligands were characterized by IR, NMR, and elemental analysis, which showed the compounds to be consistent with the proposed structures. Ring-opening reactions of bis-1,3-benzoxazines in the presence of sodium borohydride to produce N,N’-dimethylated tetrahydrosalens (H2 [H2Me]salen) have not been reported in the literature.

1. Introduction

The salen-type class of ligands (H2 salen; N,N’-disalicylidene-1,2-diaminoethane; 1, Figure 1) has had an extensive and continuing history in transition metal chemistry. Hydrogenation of the imine bond of salen compounds produces a new tetradentate ligand, which is known generally as salan (H2[H4]salen; tetrahydrosalen; N,N’-bis(2-hydroxybenzyl)-1,2-diaminoethane; 2, Figure 1) [1]. While the salen ligands feature two sites capable of covalent bonding with an electropositive element, the H4 salan ligands contain four such sites, and are therefore ideally suited to bind multiple metals [2].
Figure 1. Structures of compounds 1-3.
Figure 1. Structures of compounds 1-3.
Molecules 15 04102 g001
Tetrahydrosalen-type ligands are intimately involved with a number of metal coordination complexes, which include those elements located in groups 12, 13 and 14 [3]. Some of them have been mostly studied in polymerization catalysis in the past ten years [4,5,6]. Interest in these tetradentate ligands, whose properties may be manipulated by changing the bridging unit between the two nitrogen atoms, the substituents on the amine group, or the substitution patterns on the phenols, has stimulated research efforts in developing synthetic procedures to obtain a variety of these compounds [7,8,9,10,11,12,13,14,15]. Tetrahydrosalen, N,N’-dimethylated tetrahydrosalen 3 and its derivatives have rarely been studied, and the most common approach for the preparation of this class of compounds has involved the isolation of the salan intermediate followed by additional substitution steps on the salan products [16,17], or condensation of salans with formaldehyde/acetic acid followed by in situ sodium borohydride reduction to give the N-methylated salans [18]. Other procedures employ the reductive amination of N,N'-dimethylethylene diamine with NaBH3(CN) [19,20]. Recently, Tshuva et al. [14] reported a single-step synthetic procedure enabling the preparation in high yield of a variety of salan compounds, including N,N-disubstituted salans, by a Mannich condensation of substituted phenols, formaldehyde and N,N’-substituted-diamines. In a series of earlier works, we reported on the successful synthesis of 3,3'-ethylene-bis(3,4-dihydro-6-substituted-2H-1,3-benzoxazines) (BISBOAs) thru the condensation of p-substituted phenols, formaldehyde and ethylenediamine [21,22,23]. Herein, we report on the usefulness of these compounds for the expedient synthesis of a new series of N,N’-dimethylated tetrahydrosalens.
Based on a comparison of the basicity of tetrahydrosalen and salen, where the basicity decreases, we expected that the methyl functionality in tetrahydrosalens would provide the best template for metal binding. On the other hand, it is well known that tetrahydrosalen associated with metal centers displays cis-octahedral coordination geometry, which can form two possible diastereomers (cis fac-mer and cis fac-fac) [24]. Each of these can exist as a pair of chiral-at-metal enantiomers [8].

2. Results and Discussion

The overall procedure for the preparation of N,N'-bis(2'-hydroxy-5'-substituted-benzyl)-N,N´-dimethylethane-1,2-diamines 6a-h is depicted in Scheme 1. The 3,3'-ethylene-bis(3,4-dihydro-6-substituted-2H-1,3-benzoxazines) 5a-h used were prepared according to a previously reported procedure [21,22,23] that involves a one-pot condensation–cyclization reaction of the appropriate phenol 4a-h with an excess of 37% aqueous formaldehyde and ethylenediamine in a mixture of dioxane and water.
Scheme 1. Synthesis of the N,N´-dimethylsalan from the BISBOAs.
Scheme 1. Synthesis of the N,N´-dimethylsalan from the BISBOAs.
Molecules 15 04102 g002
Based on previous results reported for the reduction of naphtho-1,3-oxazines [25] and benzo-1,3-oxazines we anticipated that the reaction between compounds 5a-h and sodium borohydride would yield tetrahydrosalens 6a-h [26]. Additionally, the efficacy of sodium borohydride as a reducing agent should give the expected tetrahydrosalen products. In fact, the reduction with sodium borohydride of the appropriate BISBOAs (5a-h) to the respective N,N'-dimethylated tetrahydrosalens occurs readily and with good yields, ranging from 36% to 70% (Table 1). The structures of all the synthesized molecules were confirmed by elemental analysis and spectral (FT-IR, 1H-NMR, 13C-NMR) data. The FT-IR spectra of compounds 6a-h lack the characteristic absorption peaks of the O-CH2-N methylene group of the benzoxazine ring structure at 1,226 cm-1 (asymmetric stretching of C–O–C) and 1,035 cm-1 (symmetric stretching of C–O–C). The spectra did show, however, the presence of a OH group with absorptions near 3,400 cm-1. In the 1H-NMR, characteristic peaks of the 1,3-oxazine ring were not observed at ca. 5.0 ppm, but a new two-methyl singlet (6H) appeared with a chemical shift range of 2.21–2.30 ppm. This indicates that the double reduction of 5a-h with NaBH4 proceeds by the chemoselective cleavage of the O-CH2 bond of the N,O-acetal moiety of BISBOAs. This chemoselectivity may be related to the preference of the boron atom toward alkoxy complex formation, which is more favorable to a subsequent hydrolysis reaction than the aminoborane obtained by reductive cleavage of the CH2-N bond. A reduction mechanism in two steps is proposed in Scheme 2.
Table 1. Substrate scope of reduction of BISBOAs.
Table 1. Substrate scope of reduction of BISBOAs.
EntryCompoundRProductYield (%)
15aF 6a38
25bCl6b61
35cBr6c40
45dI6d59
55eCOOMe6e36
65fCOOEt6f66
75gCOOPr6g70
85hCOOBu6h68
Scheme 2. Mechanism of reduction of 5a-h with NaBH4.
Scheme 2. Mechanism of reduction of 5a-h with NaBH4.
Molecules 15 04102 g003

3. Experimental

3.1. General

The 3,3'-ethylene-bis(3,4-dihydro-6-substituted-2H-1,3-benzoxazines) 5a-h used were prepared according to the literature procedure [21,22,23]. Chemicals were used without further purification, and infrared spectra were recorded on a Perkin–Elmer FT-IR Paragon spectrometer with a KBr disk. 1H- and 13C-NMR spectra were measured on a Bruker Advance 400 MHz spectrometer in CDCl3 operating at 400 MHz and 100 MHz, respectively. Elemental analyses (C, H, N) were determined in a Carlo-Erba model 1106 analyzer. Melting points (uncorrected) were determined on an Electrothermal 9100 melting point apparatus.

3.2. General procedure for synthesis of BISBOAs

To a stirred and cooled solution of formaldehyde 37% (2.8 mL, 37.4 mmol) in dioxane (40 mL) is added slowly dropwise ethylenediamine (0.65 mL, 9.36 mmol). After stirring for 15 min at 5 ºC a solution of respectively 4-substituted-phenol (18.7 mmol) in dioxane (17 mL) is added dropwise with stirring. The mixture is gently refluxed for 4–24 h. After cooling to room temperature, the solvent is removed in vacuo and the crude product is recrystallized from methanol.
3,3'-ethylene-bis(3,4-dihydro-6-chloro-2H-1,3-benzoxazine) (5b). m.p. 170.2 ºC (literature [22]: 170–173 ºC). 1H-NMR δ (ppm): 3.02 (s, 4H, NCH2CH2N), 3.99 (s, 4H, Ar-CH2-N), 4.87 (s, 4H, O-CH2-N), 6.75 (d, 2H, J = 8.66 Hz, Ar-H), 6.94 (d, 2H, J = 2.4 Hz, Ar-H), 7.10 (dd, 2H, J = 8.66 Hz, J = 2.4 Hz, Ar-H). 13C-NMR δ (ppm): 49.6, 50.3, 82.8, 117.9, 121.4, 125.3, 127.2, 127.8, 152.7. The spectral data were consistent with literature values [22].
3,3'-ethylene-bis(3,4-dihydro-6-bromo-2H-1,3-benzoxazine) (5c). m.p. 177.2 ºC (literature [23]: 178–179 ºC). 1H-NMR δ (ppm): 2.91 (s, 4H, NCH2CH2N), 3.98 (s, 4H, Ar-CH2-N), 4.85 (s, 4H, O-CH2-N), 6.65 (d, 2H, J = 8.66 Hz, Ar-H), 7.06 (d, 2H, J = 2.4 Hz, Ar-H), 7.19 (dd, 2H, J = 8.66 Hz, J = 2,4 Hz, Ar-H). 13C-NMR δ (ppm): 49.4, 51.2, 82.1, 110.6, 117.7, 121.0, 129.9, 131.4, 153.6. The spectral data were consistent with literature values [23].
Dimethyl 3,3'-(ethane-1,2-diyl)bis(3,4-dihydro-2H-benzo[e][1,3]oxazine-6-carboxylate) (5e). m.p. 152.2 ºC (literature [22]: 151–154 ºC). 1H-NMR δ (ppm): 2.96 (s, 4H, NCH2CH2N), 3.89 (s, 6H, CH3-O), 4.06 (s, 4H, Ar-CH2-N), 4.97 (s, 4H, O-CH2-N), 6.77 (d, 2H, J = 8.4 Hz, Ar-H), 7.69 (d, 2H, J = 2.0 Hz, Ar-H), 7.83 (dd, 2H, J = 8.4 Hz, J = 2.0 Hz, Ar-H). 13C-NMR δ (ppm): 49.6, 50.2, 51.8, 83.3, 116.4, 119.4, 122.8, 129.6, 129.7, 158.2, 162.2. The spectral data were consistent with literature values [22].
Diethyl 3,3'-(ethane-1,2-diyl)bis(3,4-dihydro-2H-benzo[e][1,3]oxazine-6-carboxylate) (5f). m.p. 142.5 ºC. 1H-NMR δ (ppm): 1.38 (t, J = 8.00 Hz, 6H, CH3-CH2-O), 2.95 (s, 4H, NCH2CH2N), 4.07 (s, 4H, Ar-CH2-N), 4.24 (q, J = 8.00 Hz, 4H, CH3-CH2-O), 4.97 (s, 4H, O-CH2-N), 6.78 (d, 2H, J = 8.4 Hz, Ar-H), 7.69 (d, 2H, J = 2.0 Hz, Ar-H), 7.81 (dd, 2H, J = 8.4 Hz, J = 2.0 Hz, Ar-H). 13C-NMR δ (ppm): 13.6, 49.6, 50.3, 51.8, 83.3, 116.4, 119.4, 122.8, 129.6, 129.7, 158.2, 162.2.
Dipropyl 3,3'-(ethane-1,2-diyl)bis(3,4-dihydro-2H-benzo[e][1,3]oxazine-6-carboxylate) (5g). m.p. 127.2–127.9 ºC. 1H-NMR δ (ppm): 1.02 (t, J = 7.5 Hz, 6H, CH3-CH2-O), 1.79 (m, J = 7.5 Hz, 4H, CH3-CH2-CH2-O), 2.95 (s, 4H, NCH2CH2N), 4.07 (s, 4H, Ar-CH2-N), 4.22 (t, J = 7.6 Hz, 4H, CH3-CH2-CH2-O), 4.95 (s, 4H, O-CH2-N), 6.78 (d, 2H, J = 8.4 Hz, Ar-H), 7.69 (d, 2H, J = 2.0 Hz, Ar-H), 7.81 (dd, 2H, J = 8.4 Hz, J = 2.0 Hz, Ar-H). 13C-NMR δ (ppm): 10.5, 22.2, 49.6, 50.3, 53.8, 83.3, 116.4, 119.4, 122.8, 129.6, 129.7, 158.2, 162.2.
Dibutyl 3,3'-(ethane-1,2-diyl)bis(3,4-dihydro-2H-benzo[e][1,3]oxazine-6-carboxylate) (5h). m.p. 89.2 ºC. 1H-NMR δ (ppm): 0.95 (t, J = 7.40 Hz, 6H, CH3-CH2-CH2-CH2-O), 1.49 (m, 4H,CH3-CH2-CH2-CH2-O), 1.72 (q, J = 6.63 Hz, 4H, CH3-CH2-CH2-CH2-O), 2.93 (s, 4H, NCH2CH2N), 4.07 (s, 4H, Ar-CH2-N), 4.32 (t, J = 6.61 Hz, 4H, CH3-CH2-CH2-CH2-O), 4.95 (s, 4H, O-CH2-N), 6.78 (d, 2H, J = 8.3 Hz, Ar-H), 7.68 (d, 2H, J = 2.0 Hz, Ar-H), 7.80 (dd, 2H, J = 8.3 Hz, J = 2.0 Hz, Ar-H). 13C-NMR δ (ppm): 13.8, 19.3, 30.9, 49.7, 50.3, 51.8, 83.4, 116.4, 119.5, 122.8, 129.6, 129.7, 158.2, 162.2.

3.3. General procedure for reduction of BISBOAs

Sodium borohydride (3.0 mmol, 0.11 g) was added to a solution of the appropriate benzoxazine (1 mmol) in ethanol (15 mL), and the mixture was stirred magnetically for 30 min at room temperature. The progress of the reaction was monitored by thin-layer chromatography (TLC). After completion of the reaction, the mixture was poured into ice-cold water, neutralized with ammonium chloride (12 mL), and extracted with CHCl3 (3 × 10 cm3). The combined extracts were dried over anhydrous Na2SO4 and evaporated. The solid obtained was purified by recrystallization from ethanol to the desired products 6a-h.
N,N´-bis(2-hydroxy-5-fluorobenzyl)-N,N´-dimethylethane-1,2-diamine (6a). White solid, yield 38%, m.p. 110–112 ºC. IR: 3432 cm-1 (O-H), 2849 cm-1 (N-CH3 str.) 1H-NMR δ (ppm): 2.28 (s, 6H, H3C-N), 2.65 (s, 4H, NCH2CH2N), 3.65 (s, 4H, Ar-CH2-N), 6.65 (dd, 2H, J = 2.0 Hz, J = 24.4 Hz, Ar-H), 6.76 (dd, 2H, J = 8.7 Hz, J = 4.7 Hz, Ar-H), 6.86 (dd, 2H, J = 8.4 Hz, J = 2.4 Hz, J = 16.8 Hz). 13C-NMR δ (ppm): 41.7, 53.9, 61.4, 114.9, 115.2, 116.9, 122.4, 153.6, 156.0. Elem. anal. calcd. for C18H22F2N2O2: C 70.58%, H 6.59%, N 8.33%; found C 70.39%, H 6.54%, N 8.39%.
N,N´-bis(2-hydroxy-5-chlorobenzyl)-N,N´-dimethylethane-1,2-diamine (6b). White solid, yield 61%, m.p. 172–174 ºC. IR: 3434 cm-1 (O-H), 2849 cm-1 (N-CH3 str.). 1H-NMR [400 MHz, δ (ppm), CDCl3]: 2.21 (s, 6H, H3C-N), 2.58 (s, 4H, NCH2CH2N), 3.59 (s, 4H, Ar-CH2-N), 6.70 (d, 2H, J = 8.8 Hz, Ar-H), 6.87 (d, 2H, J = 2.4 Hz, Ar-H), 7.05 (dd, 2H, J = 8.8 Hz, J = 2.4 Hz, Ar-H). 13C-NMR [100 MHz, δ(ppm), CDCl3]: 40.6, 52.8, 60.2, 116.5, 121.9, 122.6, 127.1, 127.7, 155.3. Elem. anal. calcd. for C18H22Cl2N2O2: C 58.54%, H 6.00%, N 7.59%; found C 58.29%, H 5.84%, N 7.63%.
N,N´-bis(2-hydroxy-5-bromobenzyl)-N,N´-dimethylethane-1,2-diamine (6c). White solid, yield 40%, m.p. 181–183 ºC. IR: 3432 cm-1 (O-H), 2850 cm-1 (N-CH3 str.). 1H-NMR δ (ppm): 2.23 (s, 6H, H3C-N), 2.65 (s, 4H, NCH2CH2N), 3.58 (s, 4H, Ar-CH2-N), 6.73 (d, 2H, J = 8.8 Hz, Ar-H), 7.08 (d, 2H, J = 2.4 Hz, Ar-H), 7.05 (dd, 2H, J = 8.8 Hz, J = 2,4 Hz, Ar-H). 13C-NMR δ (ppm): 41.5, 53.7, 61.1, 110.7, 117.9, 123.3, 130.9, 131.6, 156.8. Elem. anal. calcd. for C18H22Br2N2O2: C 47.18%, H 4.84%, N 6.11%; found C 47.19%, H 4.54%, N 6.08%.
N,N´-bis(2-hydroxy-5-iodobenzyl)-N,N´-dimethylethane-1,2-diamine (6d). White solid, yield 59%, m.p. 170–171 ºC. IR: 3431 cm-1 (O-H), 2849 cm-1 (N-CH3 str.). 1H-NMR δ (ppm): 2.27 (s, 6H, H3C-N), 2.64 (s, 4H, NCH2CH2N), 3.64 (s, 4H, Ar-CH2-N), 6.62 (d, 2H, J = 8.4 Hz, Ar-H), 7.12 (d, 2H, J = 2.0 Hz, Ar-H), 7.36 (dd, 2H, J = 8.4 Hz, J = 2,0 Hz, Ar-H). 13C-NMR δ (ppm): 41.5, 53.7, 60.9, 80.4, 118.5, 124.0, 136.8, 137.6, 157.6. Elem. anal. calcd. for C18H22I2N2O2: C 39.15%, H 4.02%, N 5.07%; found C 38.96%, H 3.99%, N 5.03%.
N,N´-bis(2-hydroxy-5-methoxycarbonylbenzyl)-N,N´-dimethylethane-1,2-diamine (6e). White solid, yield 36%, m.p. 157–159 ºC. IR: 3433 cm-1 (O-H), 2848 cm-1 (N-CH3 str.), 1704 cm-1 (C=O). 1H-NMR δ (ppm): 2.29 (s, 6H, CH3-N), 2.68 (s, 4H, NCH2CH2N), 3.75 (s, 4H, Ar-CH2-N), 3.87 (s, 6H, CH3-O), 6.85 (d, J = 8.50 Hz, 2H, Ar-H), 7.70 (d, J = 1.48 Hz, 2H, Ar-H), 7.88 (dd, J = 8.47 Hz, J = 1.94 Hz, 2H, Ar-H). 13C-NMR δ (ppm): 41.6, 51.8, 53.8, 61.4, 116.2, 121.0, 121.2, 130.5, 131.1, 162.4, 166.9. Elem. anal. calcd. for C22H28N2O6: C 63.45%, H 6.78%, N 6.73%; found C 63.37%, H 6.54%, N 6.63%.
N,N´-bis(2-hydroxy-5-ethoxycarbonylbenzyl)-N,N´-dimethylethane-1,2-diamine (6f). White solid, yield 66%, m.p. 114 ºC. IR: 3449 cm-1 (O-H), 2849 cm-1 (N-CH3 str.), 1703 cm-1 (C=O). 1H-NMR δ (ppm): 1.39 (t, J = 8.00 Hz, 6H, CH3-CH2-O), 2.31 (s, 6H, CH3-N), 2.70 (s, 4H, NCH2CH2N), 3.77 (s, 4H, Ar-CH2-N), 4.34 (q, J = 8.00 Hz, 4H, CH3-CH2-O), 6.87 (d, J = 8.00 Hz, 2H, Ar-H), 7.72 (d, J = 2.00 Hz, 2H, Ar-H), 7.91 (dd, J = 8.00 Hz, J = 2.00 Hz, 2H, Ar-H). 13C-NMR δ (ppm): 13.6, 41.6, 53.8, 60.6, 61.5, 116.1, 121.1, 121.4, 130.4, 131.0, 162.2, 166.4. Elem. anal. calcd. for C24H32N2O6: C 64.85%, H 7.26%, N 6.30%; found C 64.72%, H 7.15%, N 6.08%.
N,N´-bis(2-hydroxy-5-propoxycarbonylbenzyl)-N,N´-dimethylethane-1,2-diamine (6g). White solid, yield 70%, m.p. 123–125 ºC. IR: 3450 cm-1 (O-H), 2844 cm-1 (N-CH3 str.), 1700 cm-1 (C=O). 1H-NMR δ (ppm): 1.02 (t, J = 7.42 Hz, 6H, CH3-CH2-CH2-O), 1.77 (m, J = 7.5 Hz, 4H, CH3-CH2-CH2-O), 2.29 (s, 6H, CH3-N), 2.69 (s, 4H, NCH2CH2N), 3.76 (s, 4H, Ar-CH2-N), 4.23 (t, J = 6.66 Hz, 4H, CH3-CH2-CH2-O), 6.85 (d, J = 8.50 Hz, 2H, Ar-H), 7.71 (d, J = 1.48 Hz, 2H, Ar-H), 7.90 (dd, J = 8.49 Hz, J = 2.02 Hz, 2H, Ar-H). 13C-NMR δ (ppm): 10.5, 22.2, 41.6, 53.8, 61.5, 66.2, 116.1, 121.0, 121.5, 130.4, 131.1, 162.3, 166.5. Elem. anal. calcd. for C26H36N2O6: C 66.08%, H 7.68%, N 5.93%; found C 65.89%, H 7.54%, N 5.78%.
N,N´-bis(2-hydroxy-5-butoxycarbonylbenzyl)-N,N´-dimethylethane-1,2-diamine (6h). White solid, yield 68%, m.p. 117–119 ºC. IR: 3449 cm-1 (O-H), 2850 cm-1 (N-CH3 str.), 1704 cm-1 (C=O). 1H-NMR δ (ppm): 0.97 (t, J = 7.38 Hz, 6H, CH3-CH2-CH2-CH2-O), 1.47 (m, 4H, CH3-CH2-CH2-CH2-O), 1.73 (q, J = 6.73 Hz, 4H, CH3-CH2-CH2-CH2-O), 2.30 (s, 6H, CH3-N), 2.68 (s, 4H, NCH2CH2N), 3.76 (s, 4H, Ar-CH2-N), 4.28 (t, J = 6.61 Hz, 4H, CH3-CH2-CH2-CH2-O), 6.85 (d, J = 8.50 Hz, 2H, Ar-H), 7.70 (d, J = 1.48 Hz, 2H, Ar-H), 7.89 (dd, J = 8.49, J = 2.02 Hz, 2H, Ar-H). 13C-NMR δ (ppm): 13.8, 19.3, 30.9, 41.6, 53.8, 61.5, 64.5, 116.1, 121.1, 121.5, 130.4, 131.1, 162.3, 166.5. Elem. anal. calcd. for C28H40N2O6: C 67.18%, H 8.05%, N 5.60%; found C 67.02%, H 7.94%, N 5.43%.

4. Conclusions

In summary, we have found a novel synthetic approach for the synthesis of tetrahydrosalens. The features of the present method include the ready availability of the starting materials, the mild reaction conditions, and the simplicity of the workup. Because the substitution pattern on the phenols may be varied, this simple methodology should be useful for the preparation of a variety of N,N´-dimethylated-tetrahydrosalens. Furthermore, the simplicity of the operations involved represents a good prerequisite for large scale applications.

Acknowledgements

We acknowledge the Dirección de Investigaciones Sede Bogotá (DIB) of Universidad Nacional de Colombia for its financial support of this work.
  • Sample Availability: Samples of the compounds 5a-f and 6a-f are available from the authors.

References and Notes

  1. Atwood, D.A.; Remington, M.P.; Rutherford, D. Use of the salan ligands to form bimetallic Aluminum Complexes. Organometallics 1996, 15, 4763–4769. [Google Scholar] [CrossRef]
  2. Wei, P.; Atwood, D.A. Syntheses and structures of products resulting from oxygen insertion and hydrolysis of salan-Al compounds. Polyhedron 1999, 18, 641–646. [Google Scholar]
  3. Atwood, D.A. Salan complexes of the group 12, 13 and 14 elements. Coord. Chem. Rev. 1997, 165, 267–296. [Google Scholar]
  4. Tshuva, E.Y.; Goldberg, I.; Kol, M. Isospecific living polymerization of 1-hexene by a readily available nonmetallocene C2-symmetrical Zirconium catalyst. J. Am. Chem. Soc. 2000, 122, 10706–10707. [Google Scholar]
  5. Hormnirun, P.; Marshall, E.L.; Gibson, V.C.; White, A.J.P.; Williams, D.J. Remarkable stereocontrol in the polymerization of racemic lactide using aluminum initiators supported by tetradentate aminophenoxide ligands. J. Am. Chem. Soc. 2004, 126, 2688–2689. [Google Scholar]
  6. Busico, V.; Cipullo, R.; Friederichs, N.; Ronca, S.; Talarico, G.; Togrou, M.; Wang, B. Block copolymers of highly isotactic polypropylene via controlled Ziegler−Natta polymerization. Macromolecules 2004, 37, 8201–8203. [Google Scholar] [CrossRef]
  7. Xiong, Y.; Wang, F.; Huang, X.; Wen, Y.; Feng, X. A new copper(I)–tetrahydrosalen-catalyzed asymmetric Henry reaction and its extension to the synthesis of (S)-norphenylephrine. Chem. Eur. J. 2007, 13, 829–833. [Google Scholar]
  8. MacMillan, S.N.; Jung, C.F.; Shalumova, T.; Tanski, J.M. Chiral-at-metal tetrahydrosalen complexes of resolved titanium(IV) sec-butoxides: Ligand wrapping and multiple asymmetric catalytic induction. Inorg. Chim. Acta 2009, 362, 3134–3146. [Google Scholar] [CrossRef]
  9. Huang, J.; Yu, J.F.; Wu, G.M.; Sun, W.L.; Shen, Z.Q. In situ generated, tetrahydrosalen stabilized yttrium borohydride complex: Efficient initiator for the ring-opening polymerization of ε-caprolactone. Chin. Chem. Lett. 2009, 20, 1357–1360. [Google Scholar]
  10. Matsumoto, K.; Sawada, Y.; Katsuki, T. Asymmetric epoxidation of olefins catalyzed by Ti(salan) complexes using aqueous hydrogen peroxide as the oxidant. Pure Appl. Chem. 2008, 80, 1071–1077. [Google Scholar] [CrossRef]
  11. Mitra, A.; Harvey, M.J.; Proffitt, M.K.; DePue L.J.; Parkin, S.; Atwood, A.D. Binuclear salan borate compounds with three-coordinate boron atoms. J. Organomet. Chem. 2006, 691, 523–528. [Google Scholar]
  12. Soto-Garrido, G.; Salas-Reyes, V. Salen and tetrahydrosalen nickel(II) and copper(II) complexes. Transit. Metal Chem. 2000, 25, 192–195. [Google Scholar] [CrossRef]
  13. Rivera, A.; Quevedo, R.; Navarro, M.A.; Maldonado, M. Efficient tetrahydrosalen synthesis from Mannich type bases. Synth. Commun. 2004, 34, 2479–2485. [Google Scholar] [CrossRef]
  14. Tshuva, E.Y.; Gendeziuk, N.; Kol, M. Single-step synthesis of salans and substituted salans by Mannich condensation. Tetrahedron Lett. 2001, 42, 6405–6407. [Google Scholar] [CrossRef]
  15. Cohen, A.; Yeori, A.; Kopilov, J.; Goldberg, I.; Kol, M. Construction of C1-symmetric Zirconium complexes by the design of new salan ligands. Coordination chemistry and preliminary polymerisation catalysis studies. Chem. Commun. 2008, 2149–2151. [Google Scholar]
  16. Klement, R.; Stock, F.; Elias, H.; Paulus, H.; Pelikán, P.; Valko, M.; Mazúr, M. Copper(II) complexes with derivatives of salen and tetrahydrosalen: A spectroscopic, electrochemical and structural study. Polyhedron 1999, 18, 3617–3628. [Google Scholar] [CrossRef]
  17. Elias, H.; Stock, F.; Röhr, C.A. Dioxomolybdenum(VI) Complex with a new enantiomerically pure tetrahydrosalen ligand. Acta Cryst. 1997, C53, 862–864. [Google Scholar]
  18. Balsells, J.; Carroll, P.J.; Walsh, P.J. Achiral tetrahydrosalen ligands for the synthesis of C2-symmetric Titanium complexes: A structure and diastereoselectivity study. Inorg. Chem. 2001, 40, 5568–5574. [Google Scholar] [CrossRef]
  19. Hinshaw, C.J.; Peng, G.; Singh, R.; Spence, J.T.; Enemark, J.H.; Bruck, M.; Kristofzski, J.; Merbs, S.L.; Ortega, R.B.; Wexler, P.A. Molybdenum (VI)-dioxo complexes with linear and tripodal tetradentate ligands: models for the molybdenum(VI/V) centers of the molybdenum hydroxylases and related enzymes. 1. Syntheses and structures. Inorg. Chem. 1989, 28, 4483–4491. [Google Scholar]
  20. Wong, Y.-L.; Yan, Y.; Chan, E.S.H.; Yang, Q.; Mak, T.C.W.; Ng, D.K.P. Cis-Dioxo-tungsten(VI) and -molybdenum(VI) complexes with N2O2 tetradentate ligands: synthesis, structure, electrochemistry and oxo-transfer properties. J. Chem. Soc., Dalton Trans. 1 1998, 3057–3064. [Google Scholar]
  21. Rivera, A.; Aguilar, Z.; Clavijo, D.; Joseph-Nathan, P. Synthesis and characterization of two new 6-halosubstituted bis benzoxazines. Anal. Quim. 1989, 85, 9–10. [Google Scholar]
  22. Rivera, A.; Gallo, G.I.; Gayón, M.E.; Joseph-Nathan, P. 1,3-bis-(2´-hydroxybenzil)imidazolidines as novel precursors of 3,3´-ethylene-bis)-3,4-dihydro-2H-1,3-benzoxazine. Synth. Commun. 1994, 24, 2081–2089. [Google Scholar] [CrossRef]
  23. Rivera, A.; Sánchez, A.; Ospina, E.; Joseph-Nathan, P. Síntesis de nuevas bis-benzoxazinas del tipo 3,3´-etilen-bis-(3,4-dihidro-6-R-2H-1,3-benzoxazina), elucidación de sus estructuras y comparación de sus espectros de RMN-1H con espectros teóricamente calculados (parte II). Rev. Colomb. Quim. 1985, 14, 77–80. [Google Scholar]
  24. García-Zarracino, R.; Ramos-Quiñones, J.; Höpfl, H. Preparation and structural characterization of six new diorganotin(IV) complexes of the R2Sn(SaleanH2) and R2Sn(SalceanH2) type (R = Me, n-Bu, Ph). J. Organomet. Chem. 2002, 664, 188–200. [Google Scholar] [CrossRef]
  25. Cardellicchio, C.; Ciccarella, G.; Naso, F.; Perna, F.; Tortorella, P. Use of readily available chiral compounds related to the Betti base in the enantioselective addition of ethylzinc to aryl aldehydes. Tetrahedron 1999, 55, 14685–14692. [Google Scholar]
  26. Cimarelli, C.; Palmieri, G.; Volpini, E. Ready N-alkylation of enantiopure aminophenols: Synthesis of aminotertiary aminophenols. Tetrahedron 2001, 57, 6089–6096. [Google Scholar]

Share and Cite

MDPI and ACS Style

Rivera, A.; Rojas, J.J.; Salazar-Barrios, J.; Maldonado, M.; Ríos-Motta, J. Synthesis of a New Series of N,N’-Dimethyltetrahydrosalen (H2 [H2Me]salen) Ligands by the Reductive Ring-Opening of 3,3'-Ethylene-bis(3,4-dihydro-6-substituted-2H-1,3-benzoxazines). Molecules 2010, 15, 4102-4110. https://doi.org/10.3390/molecules15064102

AMA Style

Rivera A, Rojas JJ, Salazar-Barrios J, Maldonado M, Ríos-Motta J. Synthesis of a New Series of N,N’-Dimethyltetrahydrosalen (H2 [H2Me]salen) Ligands by the Reductive Ring-Opening of 3,3'-Ethylene-bis(3,4-dihydro-6-substituted-2H-1,3-benzoxazines). Molecules. 2010; 15(6):4102-4110. https://doi.org/10.3390/molecules15064102

Chicago/Turabian Style

Rivera, Augusto, Jicli José Rojas, Jairo Salazar-Barrios, Mauricio Maldonado, and Jaime Ríos-Motta. 2010. "Synthesis of a New Series of N,N’-Dimethyltetrahydrosalen (H2 [H2Me]salen) Ligands by the Reductive Ring-Opening of 3,3'-Ethylene-bis(3,4-dihydro-6-substituted-2H-1,3-benzoxazines)" Molecules 15, no. 6: 4102-4110. https://doi.org/10.3390/molecules15064102

Article Metrics

Back to TopTop