Next Article in Journal
Structural Characterization and Anti-Inflammatory Effects of 24-Methylcholesta-5(6), 22-Diene-3β-ol from the Cultured Marine Diatom Phaeodactylum tricornutum; Attenuate Inflammatory Signaling Pathways
Previous Article in Journal
Self-Assembly Nanostructure of Myristoylated ω-Conotoxin MVIIA Increases the Duration of Efficacy and Reduces Side Effects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Concise Diastereoselective Total Synthesis of α-Ambrinol

by
Josefa L. López-Martínez
1,
Irene Torres-García
1,
Irene Moreno-Gutiérrez
1,
Pascual Oña-Burgos
2,
Antonio Rosales Martínez
3,
Manuel Muñoz-Dorado
1,
Míriam Álvarez-Corral
1,* and
Ignacio Rodríguez-García
1,*
1
Organic Chemistry, University of Almería, CIAIMBITAL, 04120 Almería, Spain
2
Instituto de Tecnología Química, Universitat Politècnica de València-Consejo Superior de Investigaciones Científicas (UPV-CSIC), 46022 Valencia, Spain
3
Department of Chemical Engineering, Escuela Politécnica Superior, University of Sevilla, 41011 Sevilla, Spain
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2023, 21(4), 230; https://doi.org/10.3390/md21040230
Submission received: 10 March 2023 / Revised: 24 March 2023 / Accepted: 31 March 2023 / Published: 1 April 2023
(This article belongs to the Section Synthesis and Medicinal Chemistry of Marine Natural Products)

Abstract

:
(−)-cis-α-Ambrinol is a natural product present in ambergris, a substance of marine origin that has been highly valued by perfumers. In this paper, we present a new approach to its total synthesis. The starting material is commercially available α-ionone and the key step is an intramolecular Barbier-type cyclization induced by CpTiCl2, an organometallic compound prepared in situ by a CpTiCl3 reduction with Mn.

Graphical Abstract

1. Introduction

Ambergris is a solid that forms pathogenically in the digestive tract of sperm whales (Physeter macrocephalus). Although it has an unpleasant odor when fresh, as it dries and ages it emits a pleasant and subtle fragrance, which is the reason why it is highly prized in the perfume industry. On average, ambergris is present in one out of every hundred sperm whales of both sexes. It has been documented that of the 1933 sperm whales captured between 1934 and 1953, only 19 had ambergris, with a total weight of approximately 1155 kg [1]. Due to the prized properties of ambergris as a fixative and stabilizer in perfumery, its chemical composition has been the subject of numerous studies. Although in variable proportions, the main constituents include a group of steroids with a cholestane skeleton, and the squalene-derived triterpene ambrein (2). In fact, although ambrein (2) has been prepared in vitro by enzymatic cyclization of squalene (1) (Scheme 1a), 13C isotopic ratio studies suggest that its biosynthesis [2], unlike the co-occurring steroids, takes place with the participation of some bacteria and it is not exclusive to the sperm whale, a result which is in line with considering ambergris as a product resulting from a pathology of the marine mammal [3].
Although ambrein (2) is an odorless compound, the aroma of ambergris is due to products formed through a natural oxidative degradation, possibly as a result of its exposure to air, seawater and sunlight, catalyzed by the presence of copper. This metal could be present because it is in the hemocyanin of the blood of squids, the main food of sperm whales [1]. This oxidative degradation process generates highly appreciated products in perfumery known as ambroxides, substances that are usually obtained by distillation and that usually represent less than 1% by weight of the raw material. These include ambrafuran (3), also called ambroxide (ambroxTM), and other related compounds such as ambraoxide (4), ambracetal (5), and cis-α-ambrinol (6) (Scheme 1b) [4,5]. Solid phase micro extraction (SPME) and gas chromatography-mass spectrometry (GC-MS) have allowed the identification of some of the more volatile natural components of ambergris, such as pristane, a fully saturated acyclic odorless nor-diterpene and γ-dihydroionone (7) [6], which has a similar structure to ambraaldehyde (8) (Scheme 1b) [7].
The stereochemistry of ambrein (2) has been well established by chemical correlation methods [8]. In addition, the configuration of the tri- and tetracyclic natural products (35) derives from the bicyclic part or ambrein (2) (in red in Scheme 1), while bi- and monocycles (68) derive from the left-hand-side monocyclic part of ambrein (in blue) [9]. cis-α-Ambrinol (6) has long attracted the attention of synthetic chemists due to its odor, which has been related to damp earth with a crude civet subnote [9]. It can be easily prepared by acid-catalyzed cyclization of γ-ionone, although in the process, β-ambrinol (9) and other cyclization products (10) are formed (Scheme 2a) [10].
Enantiopure α-ambrinol has also been prepared using different approaches. In fact, both enantiomers of cis-α-ambrinol can be prepared by chemical resolution of the racemate through fractional crystallization of the camphanic derivatives (11a, 11b) (Scheme 2b) [11]. Both the natural (−)-(S,S)-cis-α-ambrinol ((−)-6) and the unnatural (+)-(R,R) enantiomer ((+)-6) have intense odor properties, although they are very different from each other [11]. In addition, (−)-cis-α-ambrinol ((−)-6) has also be prepared by acid-catalyzed cyclization of (+)-(S)-γ-dihydroionone ((+)-7) [10] (Scheme 2c). This enantiopure compound is another highly appreciated component of ambergris that can also be obtained by enzymatic resolution of a derivative of commercial racemic α-ionone [10]. More recently, a hydrogen-bond assisted Brønsted-acid biocatalysis in water of neryl acetone (13) has allowed the enantiopure preparation of the unnatural enantiomer, (−)-(R)-γ-dihydroionone ((−)-7), which was cyclized into unnatural (+)-α-ambrinol ((+)-6) (Scheme 2d) [12].
A different strategy, starting from commercially available geranylacetone (14), allowed the formation of γ-dihydroionone (7) through a Ti(III)-catalyzed radical cyclization (Scheme 3) [13]. The enantioselective process is based on the preparation of chiral epoxide (15) through a Jacobsen’s asymmetric epoxidation followed by a Cp2TiCl-catalyzed stereoselective cyclization, a process that furnishes (16) in a highly stereoselective way. Cp2TiCl is a single electron transfer reagent that promotes a radical epoxide opening of (15) which proceeds with the retention of the configuration at the epoxide chiral center, leading to the secondary alcohol (16). Enantiomeric enrichment through kinetic resolution furnished the acetyl derivative (17). Its deoxygenation led to the protected (+)-γ-dihydroionone (18). Acid-catalyzed deprotection with concomitant cyclization yielded the desired (−)-cis-α-ambrinol (−)-6) [13].

2. Results and Discussion

As part of our continued effort in the preparation of marine natural products [14,15,16], and due to the fact that all the previously reported synthesis of cis-α-ambrinol (6) rely on an acid cyclization step that affords a mixture of regioisomers which is not easily purified by conventional chromatographic technics, we were interested in developing a concise diastereoselective total synthesis of cis-α-ambrinol (6).
The synthesis of compound 6 was planned according to the retrosynthesis depicted in Scheme 4 through two alternative pathways, both of them using commercially available α-ionone (19) as a starting material. The first approach has as a key step a diastereoselective cyclization of the allylic carbonate (20) using the bimetallic system Cp2TiIIICl/Pd0 (Scheme 4a). The second one (Scheme 4b), is based on a Barbier-type CpTiIIICl2-catalyzed intramolecular allylation of the chlorinated derivative (22).
The synthetic route based on the bimetallic system Ti(III)/Pd(0) is depicted in Scheme 5. The first step is the regioselective reduction of the conjugated double in α-ionone (19), following a modification of a previously described methodology [17], to give α-dihydroinone (12). Epoxidation of (12) with m-CPBA afforded the epoxide (21) as a diastereoselective mixture cis:trans (86:12) in an 98% yield. Acid treatment of the cis-epoxide (cis-21) gave 1-hydroxy-γ-dihydroionone (23) in a 62% yield. The formation of carbonate (20) was carried out using ethyl chloroformate under basic conditions, yielding the desired compound (20) in an 80% yield. The key step of this synthetic approach relies on a cooperative catalytic method [18,19]. We first tried the reaction using the combination Cp2TiCl/Ni(PPh3)2Cl2, a bimetallic system which has been described as an efficient promoter for the allylation of carbonyl compounds [19], although with little success in our case, as the process proved to be unproductive. However, treatment of allylic carbonate (20) with a source of Ti(III)/Pd(0) [18] (see Section 3 for details) gave a mixture of the desired natural cis-α-ambrinol (6) (37% yield), trans-α-ambrinol (24) (14% yield), and the monocyclic compound (25) (9% yield).
The relative stereochemistry of both isomers cis-α-ambrinol (6) and trans-α-ambrinol (24) was determined with the aid of NOE experiments (Figure 1). Especially relevant is the presence of correlations in (6) between the equatorial CH3-13 (δ = 1.24 ppm) and all four hydrogens in CH2-7 and CH2-9 while in (24), the observed correlations between the axial methyl CH3-13 (δ = 1.14 ppm) are only those with the β H (equatorial) of CH2-7 and CH2-9 in each case.
Although the yield of (6) was not completely satisfactory, the C-C bond formation reaction proceeds with high diastereoselectivity, and with simultaneous formation of the desired trisubstituted double bond. Spectroscopic data for synthetic cis-α-ambrinol (cis-6) were identical to those of the natural compound [20,21,22].
The experimental results obtained in the cyclization of (20) with the Ti(III)/Pd(0) bimetallic system can be fully explained by the mechanism tentatively proposed in Scheme 6. Initially, an oxidative addition of the allylic electrophile carbonate (20) to Pd0 would give the corresponding ƞ3-allyl-palladium intermediate I. Monoelectronic reduction of intermediate I by Cp2TiCl generates a ƞ3-allyl-palladium(II) intermediate, which could give the carbon-centered radical intermediate II, while the Pd0 complex is regenerated. This radical intermediate II could be trapped by a second molecule of Cp2TiCl to form two alkyl-TiIV species in metallotropic equilibrium (III and IV). β-Elimination of hydrogen in species III would account for the formation of monocyclic diene (25) and Cp2TiCl(H). It has been previously reported that this titanium hydride spontaneously decomposes to regenerate Cp2TiCl and molecular hydrogen [23]. On the other hand, the least sterically hindered alkyl-TiIV intermediate IV can evolve through two different rotational conformers, IVa and IVb. In one of them, IVa, the carbonyl group oxygen is arranged spatially close to the titanium atom (axial-like orientation), and therefore the intramolecular nucleophilic addition of the alkyl-TiIV to the ketone favors the diastereoselective formation of cis-α-ambrinol (6) as the main product. However, if the ketone oxygen is located at a greater distance from the titanium atom (equatorial-like orientation, IVb), the interaction between both atoms should be slightly weaker, resulting in a slower nucleophilic addition of the alkyl-TiIV intermediate (IVb) to the carbonyl group, and further leading to the formation of the diastereoisomer trans-α-ambrinol (24) as a minor product.
The synthetic advantage of this synthetic route is that it could lead to the enantiopure (−)-6 using an enantiopure alcohol (−)-23. This was previously prepared by Serra [10] using a lipase-mediated racemic resolution.
Our second synthetic approach is based on a Barbier-type Ti(III)-catalyzed intramolecular allylation of an allyl chloride, which is summarized in Scheme 7. In this case, we used a starting material α-dihydroionone (12), previously prepared in the other route.
Chlorination of 12 with NaClO afforded a diastereomeric mixture of allylic chlorides (22) in a 96% yield. With this substrate in hand, we first tried to induce the intramolecular allylation using an excess of Zn dust, which is known to react with allyl chlorides to form organometallic systems which can react with carbonyl groups. However, the reaction led to the formation of the bicyclic product (26), which originated as a result of the formation of a C-C bond between C1 and the carbonyl (C9). We next tried the allylation with two Ti(III) systems, the well-established single electron transfer reagent Cp2TiCl, and the half-sandwich titanocene CpTiCl2, both prepared by reduction with Mn of the appropriate Ti(IV) species.
The allylation was tested under catalytic and stoichiometric conditions for both systems. The results, summarized in Table 1, show a similar behavior in all cases, although CpTiCl3 seems to be superior both in terms of global yields and diastereoselectivity, particularly under stoichiometric conditions (Table 1, entry 3). In the light of these results, we decided to check whether the Ti(III)/Pd(0) combination strategy could be performed with the half-sandwich titanocene reagent CpTiCl2 using the ethylcarbonate (20) as a substrate. Indeed, the reaction proved to be successful (Table 1, entry 5), leading to a 73% global yield of the cyclic product, and with a 73:27 diastereoselectivity ratio using stoichiometric amounts of the Ti(III) source and catalytic of the Pd(0).
The diastereoselectivity observed in the cyclization of (22) to give (6) and (24) mediated or catalyzed by CpTiCl2 can be easily explained by a mechanism similar to the one discussed in Scheme 6, although in this case the radical intermediate II would be formed by the homolytic cleavage of the activated C-Cl bond present in (22).
In conclusion, we have proved that cis-α-ambrinol (6) can be prepared from commercial α-ionone (19) with an overall yield of 46% in only three steps using a stoichiometric amount of CpTiCl2 for the intramolecular Barbier-type allylation of the chloro-derivative (22). cis-α-Ambrinol (6) can also be prepared in five steps from the same starting material through a Ti(III)/Pd(0) cyclization of the carbonate intermediate (20) with a 35% global yield. Finally, it should be mentioned that this synthesis of α-ambrinol (6) constitutes a new application of the usefulness of CpTiCl2 as a new monoelectronic transfer reagent, as we [24,25,26] and others [27,28] have previously reported.

3. Materials and Methods

3.1. General Details

THF was distilled from Na/benzophenone under argon, and in all experiments involving titanocene (III) was deoxygenated prior to use, and oven-dried glassware was used in all cases. NMR spectra were recorded on Bruker Nanobay Avance III HD 300 MHz, and Avance III HD 600 MHz spectrometers. Proton-decoupled 13C{1H} NMR and DEPT-135 were measured in all cases. When required, NOE 1D, COSY, HSQC and HMBC experiments were used for signal assignation. Chemical shifts (δ) are expressed in ppm and coupling constants (J) in hertzs (Hz). Chemical shifts are reported using CDCl3 as internal reference. IR Spectra were recorded with a Bruker Alpha spectrometer. Mass spectra were recorded in a Waters Xevo by LC-QTof-MS by electrospray ionization. All reactions were monitored by thin-layer chromatography (TLC) carried out on 0.2 mm DC-Fertigfolien Alugram® XtraSil G/UV254 silica gel plates. The TLC plates were visualized with UV light and 7% phosphomolybdic acid or KMnO4 in water/heat. Flash chromatography was performed on silicagel 60 (0.04–0.06 mm). Hard copies of NMR and IR spectra can be found as Supplementary Materials.

3.2. Synthesis of α-Dihydroionone (12)

To a solution of α-ionone (19) (864 mg, 4.5 mmol) in THF (10 mL) was added Ni-Raney (0.4 g). The mixture was stirred under H2 (1 atm) for 30 min at room temperature in a hydrogenation apparatus. The mixture was filtered through celite, and the solvent evaporated, yielding (12) (873 mg, 4.5 mmol, 100%) as a colorless oil. Spectroscopic data are in agreement with literature values [29].
IR (ATR) v (cm−1): 2954, 2915, 2870, 1714, 1449, 1361, 1252, 1218, 1159, 961, 944, 811, 555.
1H NMR (300 MHz, CDCl3) δ (ppm): 5.36 (1H, bs, H1), 2.50 (1H, dd, J = 3.1, 7.0 Hz H8a), 2.47 (1H, dd, J = 2.0, 6.7 Hz H8b), 2.16 (3H, s, H9), 1.98 (2H, m), 1.85–1.73 (1H, m), 1.69 (3H, q, J = 1.7 Hz H13), 1.67–1.59 (1H, m), 1.51–1.48 (1H, m), 1.45–1.38 (1H, m), 1.19–1.12 (1H, m), 0.94 (3H, s), 0.89 (3H, s).
13C NMR (75 MHz, CDCl3) δ (ppm): 209.2 (C, C9), 135.6 (C, C6), 121.1 (CH, C1), 48.5 (CH), 43.8 (CH2), 32.6 (C, C4), 31.5 (CH2), 30.0 (CH3), 27.7 (CH3), 27.6 (CH3), 24.4 (CH2), 23.5 (CH3), 23.0 (CH2).

3.3. Preparation of Epoxide 21

To a solution of α-dihydroionone (12) (3.05 g, 15.67 mmol) in anhydrous CH2Cl2 (60 mL) at 0 °C, MCPBA (4.25 g, 17.24 mmol) was added. The mixture was stirred under N2 and allowed to reach room temperature for 3 h. The reaction was quenched by stirring for 15 min. with saturated NaHCO3 (30 mL) and another 30 mL of Na2S2O3 (10% in water). The two phases were separated and the organic layer was washed with brine, dried over anhydrous MgSO4 and the solvent was removed in a vacuum to give (21) as a mixture 86:12 cis:trans (3.21 g, 98%). Compound (cis-21) was purified by column chromatography (hexane:EtOAc 9:1) (83% yield from (12)). Colorless oil. Spectroscopic data are in agreement with literature values [17].
IR (ATR) v (cm−1): 2962, 2932, 2871, 1713, 1449, 1363, 1233, 1181, 1160, 1097, 1040, 995, 899, 864, 747, 557, 526.
1H NMR (600 MHz, CDCl3) δ (ppm): 2.96 (1H, s, H1), 2.76 (1H, ddd, J = 16.2, 10.1, 5.3 Hz, H8a), 2.52 (1H, ddd, J = 16.2, 9.9, 5.8 Hz, H8b), 2.18 (3H, s, H10), 1.94 (1H, dd, J = 15.5, 6.0 Hz, H2a), 1,87–1.81 (1H, m, H2b), 1.75–1.70 (1H, m, H7a), 1.61–1.54 (1H, m, H7b), 1.40 (1H, dd, J = 10.7, 6.0 Hz, H5), 1.34 (3H, s, H13), 1.32–1.27 (1H, m, H3a), 0.90 (3H, s, H11), 0.85 (3H, s, H12), 0.83 (1H, m, H3b).
13C NMR (75 MHz, CDCl3) δ (ppm): 209.4 (C, C9), 60.1 (CH, C1), 59.3 (C, C6), 46.2 (CH, C5), 43.1 (CH2, C8), 31.6 (C, C4), 30.1 (CH3, C10), 27.9 (CH3, C12), 27.3 (CH3, C11), 27.2 (CH3, C13), 26.7 (CH2, C3), 22.1 (CH2, C2), 21.6 (CH2, C7).

3.4. Synthesis of 1-Hydroxy-γ-dihydroionone (23)

To a solution of compound (cis-21) (934 mg, 4.44 mmol) in CH2Cl2 (30 mL) at 0 °C, p-TSA (76 mg, 0.44 mmol) was added. The mixture was stirred for 12 h (0 °C), and then p-TSA (76 mg, 0.44 mmol) was again added and stirred for 4 h at room temperature. The mixture was washed with saturated NaHCO3 (10 mL × 3) and dried over anhydrous MgSO4. The solvent was removed in a vacuum and the residue purified by silica gel flash column chromatography (hexane/EtOAc, 8:2) to afford alcohol (23) (575 mg, 62%) as a colorless oil. Spectroscopic data are in agreement with literature values [30].
IR (ATR) v (cm−1): 3412, 2934, 2867, 1707, 1648, 1456, 1410, 1363, 1159, 1062, 1038, 898, 639, 584, 502.
1H NMR (600 MHz, CDCl3) δ (ppm): 5.18 (1H, s, H13a), 4.65 (1H, s, H13b), 3.96 (1H, dd, J = 4.8, 9.0 Hz, H1), 2.58 (1H, ddd, J = 5.4, 9.0, 18.0 Hz, H8a), 2.34 (1H, ddd, J = 7.8, 8.4, 18.0 Hz, H8b), 2.11 (3H, s, H10), 1.91 (1H, m, H2a), 1.85 (1H, m, H7a), 1.74–1.62 (3H, m, H7b, H5, OH), 1.50 (1H, m, H3a), 1.46 (1H, m, H2b), 1.38 (1H, m, 3b), 0.98 (3H, s, H11), 0.74 (3H, s, H12).
13C NMR (75 MHz, CDCl3) δ (ppm): 209.5 (C, C9), 150.3 (C, C6), 105.5 (CH2, C13), 73.6 (CH, C1), 51.3 (CH, C5), 42.5 (CH2, C8), 37.9 (CH2, C3), 35.8 (C, C4), 33.2 (CH2, C2), 30.0 (CH3, C10), 29.3 (CH3, C11), 21.0 (CH3, C12), 19.5 (CH2, C7).

3.5. Synthesis of Ethyl Carbonate (20)

In an N2 atmosphere at 0 °C, ethyl chloroformate (0.74 mL, 7.59 mmol), pyridine (1.54 mL, 18.98) and DMAP (64 mg. 0.51 mmol) were added to a solution of (23) (532 mg, 2.53 mmol) in CH2Cl2 (40 mL). After 10 min, the cooling bath was removed, and the mixture was stirred for 20 h. TIt was then diluted with Et2O and washed with HCl (3%) and water. The organic layer was dried over anhydrous MgSO4, the solvent was removed in a vacuum and the residue purified by silica gel flash column chromatography (gradient hexane/EtOAc) to afford carbonate (20) (571 mg, 80%) as a colorless oil.
IR (ATR) v (cm−1): 2949, 2871, 1741, 1715, 1650, 1456, 1367, 1251, 1162, 1007, 903, 856, 790. HREIMS (m/z) calcd. for C16H26O4 282.1831 [M]+, found 282.1833.
1H NMR (300 MHz, CDCl3) δ (ppm): 5.17 (1H, s, H13a), 4.96 (1H, dd, J = 4.5, 8.2 Hz, H1), 4.73 (1H, s, H13b), 4.22 (1H, q, J = 7.1 Hz, OCH2), 2.54 (1H, ddd, J = 4.5, 9.7, 17.6 Hz, H8a), 2.31 (1H, m, H8b), 2.13 (3H, s, H10), 1.98–1.83 (2H, m, H2a, H7a), 1.75–1.57 (4H, m, H2b, H7b, H5, H3a), 1.41 (1H, dd, J = 4.5, 10.3 Hz, H3b), 1.34 (3H, t, J = 7.1 Hz), 1.00 (3H, s, H11), 0.82 (3H, s, H12).
13C NMR (75 MHz, CDCl3) δ (ppm): 209.2 (C, C9), 154.5 (C, OCOO), 144.8 (C, C6), 109.2 (CH2, C13), 78.5 (CH, C1), 63.8 (CH2, OCH2), 51.4 (CH, C5), 42.3 (CH2, C8), 35.3 (C, C4), 30.0 (CH3, C10), 29.4 (CH2), 28.8 (CH3), 22.6 (CH3), 20.0 (CH2), 14.3 (CH3, OCH2CH3).

3.6. Synthesis of Ambrinol Using Cp2TiCl2

Deoxygenated, dry THF (6.5 mL) was added to a mixture of Cp2TiCl2 (195 mg, 0.76 mmol), Pd(PPh3)2Cl2 (54 mg, 0.076 mmol), and Mn dust (169 mg, 3.04 mmol) in an Ar atmosphere and the red suspension was stirred at room temperature until it turned lime green (after about 15 min). Then, a solution of (20) (107 mg, 0.38 mmol) in THF (2 mL) was added dropwise and the mixture was stirred for four days. The reaction was diluted in EtOAc, washed with HCl 12% and brine, dried (anhydrous Na2SO4) and the solvent was removed. The residue was purified by flash chromatography (hexane/EtOAc 9:1) yielding cis-α-ambrinol (6) (27 mg, 37%), trans-α-ambrinol (24) (10 mg, 14%) and compound (25) (7 mg, 9%).
cis-α-ambrinol (6): Colorless oil; spectroscopic data are in agreement with literature values [12]. IR (ATR) v (cm−1): 3449, 2957, 2913, 2868, 2843, 1451, 1383, 1363, 1321, 1263, 1237, 1186, 1134, 1111, 1097, 1019, 995, 926, 912, 878, 800, 727, 501. HREIMS (m/z) calcd. for C13H22O 194.1671 [M]+, found 194.1670.
1H NMR (600 MHz, CDCl3) δ (ppm): 5.48 (1H, s, H1), 2.15 (1H, bd, J = 13.2 Hz, H9a), 2.10 (1H, dd, J = 13.2, 2.5 Hz, H9b), 2.04–2.00 (2H, m, H2), 1.87 (1H, s, OH), 1.76–1.69 (2H, m, H6a, H7eq), 1.52 (1H, bd, J = 12.6 Hz, H5), 1.47 (1H, td, J = 13.4, 4.2 Hz, H7ax), 1.39 (1H, dt, J = 13.0, 7.3 Hz, H3eq), 1.29 (1H, td, J = 13.0, 3.4 Hz, H6b), 1.24 (3H, s, H13), 1.21 (1H, m, H3ax), 0.94 (3H, s, H11), 0.89 (3H, s, H12).
13C NMR (151 MHz, CDCl3) δ (ppm): 137.5 (C, C10), 122.2 (CH2, C1), 70.4 (C, C8), 50.0 (CH2, C9), 47.4 (CH, C5), 39.1 (CH2, C7), 33.4 (CH2, C3), 31.2 (C, C4), 29.4 (CH3, C13), 28.2 (CH3, C11), 26.1 (CH3, C12), 25.2 (CH2, C6), 22.9 (CH2, C2).
trans-α-ambrinol (24): Colorless oil; IR (ATR) v (cm−1): 3368, 2923, 2869, 1710, 1664, 1451, 1364, 1260, 1198, 1119, 1062, 1020, 1010, 947, 919, 831, 820, 727, 682, 571. HREIMS (m/z) calcd. for C13H22O 194.1671 [M]+, found 194.167.
1H NMR (600 MHz, CDCl3) δ (ppm): 5.37 (1H, s, H1), 2.21 (1H, dd, J = 12.5, 2.6 Hz, H9eq), 2.13 (1H, bd, J = 12.5 Hz, H9ax), 1.97 (2H, m, H2), 1.81–1.75 (2H, m, H7eq, H6eq), 1.60 (1H, bd, J = 13.2 Hz, H5), 1.55 (1H, td, J = 13.3, 4.3 Hz, H7ax), 1.34 (1H, dt, J = 13.0, 6.1 Hz, H3eq), 1.25 (2H, m, H3ax. OH), 1.14 (3H, s, H13), 1.10 (1H, td, J = 6.5, 1.9 Hz, H6eq), 0.94 (3H, s, H11), 0.84 (3H, s, H12).
13C NMR (151 MHz, CDCl3) δ (ppm): 137.4 (C, C10), 120.9 (CH2, C1), 72.1 (C, C8), 50.9 (CH2, C9), 47.0 (CH, C5), 40.8 (CH2, C7), 35.2 (CH2, C3), 31.3 (C, C4), 28.7 (CH3, C11), 25.6 (CH3, C13), 25.3 (CH2, C6), 24.6 (CH3, C12), 22.8 (CH2, C2).
Compound (25): Colorless oil [31]; 1H NMR (300 MHz, CDCl3) δ 6.01 (1H, m, H1), 5.67 (1H, m, H2), 4.91 (1H, s, H13a), 4.70 (1H, s, H13b), 2.55–2.32 (2H, m), 2.14 (3H, s, H10), 1.84–1.70 (3H, m), 1.35–1.25 (2H, m), 1.02 (3H, s, H11), 0.88 (3H, s, H12).

3.7. Synthesis of Ambrinol Using CpTiCl3

Deoxygenated, dry THF (6 mL) was added to a mixture of CpTiCl3 (137 mg, 0.62 mmol), Pd(PPh3)2Cl2 (44 mg, 0.062 mmol), and Mn dust (138 mg, 2.48 mmol) in an Ar atmosphere red suspension was stirred at room temperature until it turned lime green (after about 15 min). Then, a solution of (20) (88 mg, 0.31 mmol) in THF (0.5 mL) was added dropwise and the mixture was stirred for 19 h. The reaction was quenched and purified as described above for Cp2TiCl2. Compound (6) (32 mg, 0.16 mmol, 53%) and compound (24) (12 mg, 0.06 mmol, 20%) were obtained.

3.8. Synthesis of 1-Chloro-γ-dihydroionone (22)

To a solution of α-dihydroionone (12) (455 mg, 2.32 mmol) in hexane (1.5 mL), NaClO (aqueous solution 15%) (3 mL, 12.9 mmol) was added. The mixture was cooled at 0 °C and H3PO4 (aqueous solution 40%) (0.4 mL, 1.72 mmol) was added. After stirring for 2 h at 0 °C, water (15 mL) was added, and the mixture extracted with Et2O (15 mL × 3). The combined organic layer was washed with brine and dried over anhydrous MgSO4. The solvent was removed in a vacuum to give 1-chloro-γ-dihydroionone (22) as a mixture of cis (R*S*): trans (R*R*) isomers (10:0.8 ratio); colourless oil (0.51 g, 96%). Spectroscopic data are in agreement with literature values [32].
IR (ATR) v (cm−1): 2950, 2869, 1714, 1649, 1454, 1419, 1388, 1364, 1234, 1158, 989, 909, 886, 760, 705, 602, 537.
1H NMR (300 MHz, CDCl3) δ (ppm) (R*S*) isomer signals: 5.34 (1H, s, H13a), 4.78 (1H, s, H13b), 4.48 (1H, dd, J = 4.8, 6.6 Hz, H1), 0.94 (3H, s), 0.86 (3H, s); (R*R*) isomer signals: 5.22 (1H, s, H13a), 4.72 (1H, s, H13b), 4.62 (1H, t, J = 5.0 Hz, H1), 1.01 (3H, s), 0.79 (3H, s); both isomer signals: 2.64–2.53 (1H, m), 2.38–2.27 (1H, m), 2.16–2.06 (1H, m), 2.12 (3H, s, H10), 1.92–1.69 (5H, m), 1.30 (1H, ddd, J = 4.3, 8.3, 13.0 Hz).
13C NMR (75 MHz, CDCl3) δ (ppm): 209.3 (C, C9), 145.8 (C, C6), 111.2 (CH2, C13), 63.0 (CH, C1), 52.2 (CH, C5), 42.2 (CH2), 35.2 (CH2), 35.0 (C, C1), 33.9 (CH2), 30.1 (CH3), 28.5 (CH3), 21.3 (CH2).

3.9. Zinc Cyclization of 1-Chloro-γ-dihydroionone (22)

Zn (161 mg, 2.4 mmol) and HCOONa (169 mg, 2.4 mmol) were added to a solution of (22) (138 mg, 0.6 mmol) in THF (1 mL) and EtOH (3 mL). The mixture was heated at 80 °C and stirred overnight. The reaction was cooled, diluted with Et2O (10 mL) and filtered through celite. The celite was then washed with H2O (10 mL), the two phases were separated, and the aqueous layer was extracted with Et2O (10 mL × 3). The combined organic layer was washed with saturated NaHCO3 (30 mL) and brine (30 mL), dried over anhydrous MgSO4 and the solvent was removed in a vacuum. The residue was purified by flash chromatography (hexane/Et2O 9:1) afforded compound (26) (30 mg, 26%) as a colorless oil.
IR (ATR) v (cm−1): 3473, 3065, 2969, 2932, 2867, 1717, 1653, 1479, 1452, 1384, 1364, 1246, 1211, 1168, 1106, 983, 916, 882, 693, 541. HREIMS (m/z) calcd. for C13H22O 194.1671 [M]+, found 194.1675.
1H NMR (300 MHz, CDCl3) δ (ppm): 4.84 (2H, m), 2.12 (1H, m), 2.00–1.93 (1H, m), 1.89 (2H, m), 1.81–1.75 (2H, m), 1.73–1.64 (2H, m), 1.54 (1H, dd, J = 14.0, 7.1 Hz), 1.25 (3H, s), 1.21–1.19 (1H, m), 1.00 (3H, s), 0.93 (3H, s).
13C NMR (75 MHz, CDCl3) δ (ppm): 152.1 (C), 108.5 (CH2), 73.8 (C), 51.2 (CH), 49.3 (CH), 35.3 (C), 35.3 (CH2), 33.9 (CH2), 29.0 (CH3), 28.9 (CH3), 26.8 (CH3), 26.0 (CH2), 25.1 (CH2).

3.10. CpTiCl3-Catalyzed Cyclization of 1-Chloro-γ-dihydroionone (22)

Deoxygenated, dry THF (3 mL) was added to a mixture of CpTiCl3 (11 mg, 0.052 mmol) and Mn dust (57 mg, 1.04 mmol) in an Ar atmosphere resulting in a green suspension. Me3SiBr (0.22 mL, 0.52 mmol) was then added, and the mixture turned turquoise. Subsequently, a solution of (22) (119 mg, 0.52 mmol) in THF (1 mL) was added dropwise and the mixture was stirred for 4 h. The reaction was filtered, diluted in Et2O, washed with HCl 3% and brine, dried (anhydrous MgSO4), and the solvent was removed. The residue was purified by flash chromatography (hexane/EtOAc 8:2) afforded cis-α-ambrinol (6) (42 mg, 42%) and trans-α-ambrinol (24) (18 mg, 18%).

3.11. CpTiCl3-Induced Cyclization of 1-Chloro-γ-dihydroionone (22)

Deoxygenated, dry THF (3 mL) was added to a mixture of CpTiCl3 (105 mg, 0.48 mmol) and Mn dust (53 mg, 0.96 mmol) in an Ar atmosphere resulting in a green suspension. Subsequently, a solution of (22) (110 mg, 0.48 mmol) in THF (1 mL) was added dropwise and the mixture was stirred for 5 h. The reaction was quenched and purified as indicated for the catalytic version to give cis-α-ambrinol (6) (45 mg, 48%) and trans-α-ambrinol (24) (18 mg, 19%).

3.12. Cp2TiCl2 Cyclization of 1-Chloro-γ-dihydroionone (22)

For the catalytic cyclization, deoxygenated and dry THF (3 mL) was added to a mixture of Cp2TiCl2 (25 mg, 0.093 mmol) and Mn dust (160 mg, 2.85 mmol) in an Ar atmosphere and the red suspension was stirred at room temperature until it turned lime green (after about 15 min). Then, a mixture of 2,4,6-collidine (0.36 mL, 2.69 mmol) and Me3SiCl (0.2 mL, 1.54 mmol) in THF (1 mL) was added. Subsequently, a solution of (22) (110 mg, 0.48 mmol) in THF (1 mL) was added dropwise and the mixture was stirred for 1 h 45 min. The reaction was filtered, diluted in Et2O, washed with HCl 3% and brine, dried (anhydrous MgSO4), and the solvent was removed. The residue was purified by flash chromatography (hexane/EtOAc 8:2) affording cis-α-ambrinol (6) (31 mg, 33%) and trans-α-ambrinol (24) (16 mg, 17%).
With stoichiometric amounts of Cp2TiCl2, the reaction was carried out under the same conditions as above, employing Cp2TiCl2 (247 mg, 0.93 mmol), Mn dust (162 mg, 2.85 mmol) and (22) (100 mg, 0.44 mmol). The mixture was stirred for 2 h, quenched and purified as indicated above, and cis-α-ambrinol (6) (32 mg, 37%) and trans-α-ambrinol (24) (15 mg, 18%) were obtained.

4. Conclusions

Natural cis-α-ambrinol (6) can be diastereoselectively prepared by the intramolecular Barbier-type reaction of a carbonate derivative of 1-hydroxy-γ-dihydroionone (23) promoted by the bimetallic system based on Ti(III)/Pd(0) [Cp2TiCl, Pd(PPh3)2Cl2]. Although the yield is somewhat low, the process can lead to enantiopure cis-α-ambrinol (6) if compound (23) is resolved enzymatically. In addition, CpTiCl2-promoted cyclization of 1-chloro-γ-dihydroionone (22) leads to the formation of α-ambrinol in a 64% global yield as a separable mixture of cis and trans diastereomers in a 2.3:1 ratio, which accounts for a 46% global yield from commercial α-ionone for the natural diastereoisomer cis-α-ambrinol (6).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/md21040230/s1, Figures S1–S47: NMR and IR spectra of compounds (6), (12) and (21)(26).

Author Contributions

Conceptualization, M.Á.-C. and I.R.-G.; investigation, J.L.L.-M., I.T.-G., I.M.-G., M.Á.-C. and I.R.-G.; writing—review and editing, J.L.L.-M., I.T.-G., P.O.-B., A.R.M., M.M.-D., M.Á.-C. and I.R.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the University of Almería and Junta de Andalucía (Conserjería de Transformación Económica, Industria, Conocimiento y Universidades) and Fondo Europeo de Desarrollo Regional (FEDER) for the Projects UALFEDER 2020-FQM-B1989, project PY20_01027 and project CEIA3 PYC20 RE 060 UAL, and also for the Horizon 2020-Research and Innovation Framework Programme of the European Commission for the project 101022507 LAURELIN and also by the University of Seville, through the Vicerrectorado de Investigación (Projects 2020/00001014 and 2021/00000422: Ayudas a Consolidación de Grupos de la Junta de Andalucía and Project Politec-Biomat: Red de Biomateriales en la Universidad de Sevilla).

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Acknowledgments

We thank the University of Almería and the University of Seville for the administrative work on funding and projects.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Clarke, R. The origin of ambergris. Lat. Am. J. Aquat. Mamm. 2006, 5, 7–21. [Google Scholar] [CrossRef]
  2. Rowland, S.J.; Sutton, P.A.; Wolff, G.A. Biosynthesis of ambrein in ambergris: Evidence from isotopic data and identification of possible intermediates. Nat. Prod. Res. 2021, 35, 1235–1241. [Google Scholar] [CrossRef] [PubMed]
  3. Olimat, S. A review on ambergris perspective and modern chemical composition and pharmacology. Acad. J. Med. Plants 2020, 8, 96–101. [Google Scholar]
  4. Riad, N.; Zahi, M.R.; Bouzidi, N.; Daghbouche, Y.; Touafek, O.; El Hattab, M. Occurrence of Marine Ingredients in Fragrance: Update on the State of Knowledge. Chemistry 2021, 3, 1437–1463. [Google Scholar] [CrossRef]
  5. Mookherjee, B.D.; Patel, R.R. Isolation and identification of volatile constituents of tincture ambergris. In Proceedings of the VII International Congress of Essential Oils, Kyoto, Japan, 7–11 October 1979. [Google Scholar]
  6. Wilde, M.J.; Robson, W.J.; Sutton, P.A.; Rowland, S.J. Volatile and semi-volatile components of jetsam ambergris. Nat. Prod. Res. 2020, 34, 3048–3053. [Google Scholar] [CrossRef]
  7. Jegou, E.; Polonsky, J.; Lederer, E.; Schulte-Elte, K.H.; Egger, B.; Ohloff, G. Ambergris revisited. Isolation of volatile constituents; identification and synthesis of ambra-aldehyde C14H22O. Nouv. J. Chim. 1977, 1, 529. [Google Scholar]
  8. Ohloff, G.; Vial, C. Permanganate Oxidation of Ambrein and the Absolute Configuration of Dihydro-γ-ionone (Supplement and Rectification). Helv. Chim. Acta 1977, 60, 2767–2769. [Google Scholar] [CrossRef]
  9. Ohloff, G. The fragrance of ambergris. In Fragrance Chemistry: The Science and the Sense of Smell; Theimer, E.T., Ed.; Academic Press: Rumson, NJ, USA, 1982; pp. 535–574. [Google Scholar]
  10. Serra, S. An expedient preparation of enantio-enriched ambergris odorants starting from commercial ionone alpha. Flavour Fragr. J. 2013, 28, 46–52. [Google Scholar] [CrossRef]
  11. Naegeli, P.; Wirz-Habersack, Y. Enantiodifferentiation of odor perception of α-ambrinols. Tetrahedron Asymmetry 1992, 3, 221. [Google Scholar] [CrossRef]
  12. Schneider, A.; Jegl, P.; Hauer, B. Stereoselective Directed Cationic Cascades Enabled by Molecular Anchoring in Terpene Cyclases. Angew. Chem. Int. Ed. 2021, 60, 13251–13256. [Google Scholar] [CrossRef]
  13. Justicia, J.; Campana, A.G.; Bazdi, B.; Robles, R.; Cuerva, J.M.; Oltra, J.E. Titanium-catalyzed enantioselective synthesis of α-ambrinol. Adv. Synth. Catal. 2008, 350, 571–576. [Google Scholar] [CrossRef]
  14. Rosales, A.; López-Sánchez, C.; Álvarez-Corral, M.; Muñoz-Dorado, M.; Rodríguez-García, I. Total synthesis of (±)-euryfuran through Ti(III) catalyzed radical cyclization. Lett. Org. Chem. 2007, 4, 553–555. [Google Scholar] [CrossRef]
  15. Rosales, A.; Muñoz-Bascón, J.; Roldán-Molina, E.; Rivas-Bascón, N.; Padial, N.M.; Rodríguez-Maecker, R.; Rodríguez-García, I.; Oltra, J.E. Synthesis of (±)-Aureol by Bioinspired Rearrangements. J. Org. Chem. 2015, 80, 1866–1870. [Google Scholar] [CrossRef] [PubMed]
  16. Rosales, A.; Foley, L.A.R.; Padial, N.M.; Muñoz-Bascón, J.; Sancho-Sanz, I.; Roldan-Molina, E.; Pozo-Morales, L.; Irías-Álvarez, A.; Rodríguez-Maecker, R.; Rodríguez-García, I.; et al. Diastereoselective Synthesis of (±)-Ambrox by Titanium(III)-Catalyzed Radical Tandem Cyclization. Synlett 2016, 27, 369–374. [Google Scholar] [CrossRef]
  17. Serra, S.; Lissoni, V. First Enantioselective Synthesis of Marine Diterpene Ambliol-A. Eur. J. Org. Chem. 2015, 2015, 2226–2234. [Google Scholar] [CrossRef]
  18. Millán, A.; Campaña, A.G.; Bazdi, B.; Miguel, D.; Álvarez de Cienfuegos, L.; Echavarren, A.M.; Cuerva, J.M. Ti/Pd Bimetallic Systems for the Efficient Allylation of Carbonyl Compounds and Homocoupling Reactions. Chem. Eur. J. 2011, 17, 3985. [Google Scholar] [CrossRef]
  19. Martinez-Peragon, A.; Millan, A.; Campana, A.G.; Rodriguez-Marquez, I.; Resa, S.; Miguel, D.; Alvarez de Cienfuegos, L.; Cuerva, J.M. Ti/Ni-Based Multimetallic System for the Efficient Allylation of Carbonyl Compounds. Eur. J. Org. Chem. 2012, 2012, 1499. [Google Scholar] [CrossRef]
  20. Ohloff, G.; Schulte-Elte, K.H.; Mueller, B.L. Formation of ambergris odorants from ambrein under simulated natural conditions. Helv. Chim. Acta 1977, 60, 2763. [Google Scholar] [CrossRef]
  21. Stoll, M.; Hinder, M. Odor and constitution. XV. Cyclization of dihydro-γ-ionone. Helv. Chim. Acta 1955, 38, 1593. [Google Scholar] [CrossRef]
  22. Stoll, M.; Seidel, C.F.; Willhalm, B.; Hinder, M. Odor and constitution XVI. The constitution of ambrinol (Δ4-, Δ9-, and Δ5,10-1,1,6-hydroxytrimethyl-6-hydroxyoctahydronaphthalene). Helv. Chim. Acta 1956, 39, 183. [Google Scholar] [CrossRef]
  23. Gordon, J.; Hildebrandt, S.; Dewese, K.R.; Klare, S.; Gansauer, A.; RajanBabu, T.V.; Nugent, W.A. Demystifying Cp2Ti(H)Cl and Its Enigmatic Role in the Reactions of Epoxides with Cp2TiCl. Organometallics 2018, 37, 4801–4809. [Google Scholar] [CrossRef] [PubMed]
  24. López-Martínez, J.L.; Torres-García, I.; Rodríguez-García, I.; Muñoz-Dorado, M.; Álvarez-Corral, M. Stereoselective Barbier-Type Allylations and Propargylations Mediated by CpTiCl3. J. Org. Chem. 2019, 84, 806–816. [Google Scholar] [CrossRef] [PubMed]
  25. Torres-García, I.; López-Martínez, J.L.; Martínez-Martínez, R.; Oltra, J.E.; Muñoz-Dorado, M.; Rodríguez-García, I.; Álvarez-Corral, M. The half-sandwich titanocene CpTiIIICl2 as efficient system for the preparation of 2,5-dihydrofurans via α-allenols. Appl. Organomet. Chem. 2020, 34, e5244. [Google Scholar] [CrossRef]
  26. Torres-García, I.; López-Martínez, J.L.; López-Domene, R.; Muñoz-Dorado, M.; Rodríguez-García, I.; Álvarez-Corral, M. Enantioselective total synthesis of putative dihydrorosefuran, a monoterpene with an unique 2,5-dihydrofuran structure. Beilstein J. Org. Chem. 2022, 18, 1264–1269. [Google Scholar] [CrossRef]
  27. Roldán-Molina, E.; Padial, N.M.; Lezama, L.; Oltra, J.E. CpTiCl2, an Improved Titanocene(III) Catalyst in Organic Synthesis. Eur. J. Org. Chem. 2018, 2018, 5997–6001. [Google Scholar] [CrossRef]
  28. Roldán-Molina, E.; Nievas, M.M.; Navarro, J.A.R.; Oltra, J.E. CpTiCl2-Catalyzed Cross-Coupling between Internal Alkynes and Ketones: A Novel Concept in the Synthesis of Halogenated, Conjugated Dienes. Chem. Eur. J. 2020, 26, 8296–8301. [Google Scholar] [CrossRef]
  29. Fernandez-Mateos, A.; Madrazo, S.E.; Teijon, P.H.; Clemente, R.R.; Gonzalez, R.R.; Gonzalez, F.S. Synthesis of the BCDE Molecular Fragment of Azadiradione Mediated by Titanocene(III). J. Org. Chem. 2013, 78, 9571–9578. [Google Scholar] [CrossRef]
  30. Serra, S.; Fuganti, C.; Brenna, E. Two easy photochemical methods for the conversion of commercial ionone alpha into regioisomerically enriched γ-ionone and γ-dihydroionone. Flavour Fragr. J. 2007, 22, 505–511. [Google Scholar] [CrossRef]
  31. Serra, S.; Fuganti, C.; Brenna, E. Synthesis, olfactory evaluation, and determination of the absolute configuration of the 3,4-didehydroionone stereoisomers. Helv. Chim. Acta 2006, 89, 1110–1122. [Google Scholar] [CrossRef]
  32. Willis, B.J.; Christenson, P.A.; Mack, R. Halogen-Containing Ionone Derivatives and Compositions Containing Them. U.S. Patent 4,272,412, 9 June 1981. [Google Scholar]
Scheme 1. (a) Squalene cyclization into ambrein; (b) Main odourous ambergris constituents.
Scheme 1. (a) Squalene cyclization into ambrein; (b) Main odourous ambergris constituents.
Marinedrugs 21 00230 sch001
Scheme 2. Representative racemic or enantioselective syntheses of α-ambrinol from γ-dihydroionone. (a) Acid-catalyzed synthesis [10]; (b) resolution by derivatization [11]; (c) resolution with lipase [11]; (d) biocatalysis and terpene cyclase [12].
Scheme 2. Representative racemic or enantioselective syntheses of α-ambrinol from γ-dihydroionone. (a) Acid-catalyzed synthesis [10]; (b) resolution by derivatization [11]; (c) resolution with lipase [11]; (d) biocatalysis and terpene cyclase [12].
Marinedrugs 21 00230 sch002
Scheme 3. Enantioselective syntheses of cis-α-ambrinol through a Ti(III)-catalyzed radical process [13].
Scheme 3. Enantioselective syntheses of cis-α-ambrinol through a Ti(III)-catalyzed radical process [13].
Marinedrugs 21 00230 sch003
Scheme 4. Retrosynthetic approaches (a) or (b) of cis-α-ambrinol (6) from α-ionone (19) using Ti(III).
Scheme 4. Retrosynthetic approaches (a) or (b) of cis-α-ambrinol (6) from α-ionone (19) using Ti(III).
Marinedrugs 21 00230 sch004
Scheme 5. Synthesis of cis-α-ambrinol (6) from α-ionone (19) through a Ti(III)/Pd(0) C-C-forming step. (a) H2, Ni-Raney (cat), 100%; (b) m-CPBA, 83%; (c) p-TSA, 62%; (d) EtOCOCl, pyridine, DMAP 80%; (e) Cp2TiCl2 (2 equiv), Mn (8 equiv), Pd(PPh3)2Cl2 (0.2 equiv).
Scheme 5. Synthesis of cis-α-ambrinol (6) from α-ionone (19) through a Ti(III)/Pd(0) C-C-forming step. (a) H2, Ni-Raney (cat), 100%; (b) m-CPBA, 83%; (c) p-TSA, 62%; (d) EtOCOCl, pyridine, DMAP 80%; (e) Cp2TiCl2 (2 equiv), Mn (8 equiv), Pd(PPh3)2Cl2 (0.2 equiv).
Marinedrugs 21 00230 sch005
Figure 1. Diagnostic NOE in cis-α-ambrinol (6) and trans-α-ambrinol (24).
Figure 1. Diagnostic NOE in cis-α-ambrinol (6) and trans-α-ambrinol (24).
Marinedrugs 21 00230 g001
Scheme 6. Mechanistic proposal for the Cp2TiIIICl/Pd0-promoted diastereoselective cyclization of carbonate 20.
Scheme 6. Mechanistic proposal for the Cp2TiIIICl/Pd0-promoted diastereoselective cyclization of carbonate 20.
Marinedrugs 21 00230 sch006
Scheme 7. Second synthesis of cis-α-ambrinol (6) from α-dihydroionone (12). (a) NaClO (aq.), H3PO4, 96%; (b) Zn, HCOONa, 26%; (c) CpTiCl3 (0.1 equiv), Mn (2 equiv), Me3SiBr (1 equiv); (d) CpTiCl3 (1 equiv), Mn (2 equiv).
Scheme 7. Second synthesis of cis-α-ambrinol (6) from α-dihydroionone (12). (a) NaClO (aq.), H3PO4, 96%; (b) Zn, HCOONa, 26%; (c) CpTiCl3 (0.1 equiv), Mn (2 equiv), Me3SiBr (1 equiv); (d) CpTiCl3 (1 equiv), Mn (2 equiv).
Marinedrugs 21 00230 sch007
Table 1. Ti(III)-mediated intramolecular Barbier-type cyclizations.
Table 1. Ti(III)-mediated intramolecular Barbier-type cyclizations.
EntrySubstrateTi(III)ConditionsCombined Yield
6 + 24
Cis:Trans Ratio
6:24
122Cp2TiClstoichiometric a55%68:32
222Cp2TiClcatalytic b50%65:35
322CpTiCl2stoichiometric c67%72:28
422CpTiCl2catalytic d60%70:30
520CpTiCl2e73%73:27
(a) 2 eq Cp2TiCl2, 6 eq Mn; (b) 0.2 eq Cp2TiCl2, 6 eq Mn, 6 eq collidine, 3 eq TMSCl; (c) 1 eq CpTiCl3, 2 eq Mn; (d) 0.1 eq CpTiCl3, 2 eq Mn, 1 eq TMSBr; (e) 2 eq CpTiCl3, 8 eq Mn, 0.2 eq Pd(PPh3)2Cl2.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

López-Martínez, J.L.; Torres-García, I.; Moreno-Gutiérrez, I.; Oña-Burgos, P.; Rosales Martínez, A.; Muñoz-Dorado, M.; Álvarez-Corral, M.; Rodríguez-García, I. A Concise Diastereoselective Total Synthesis of α-Ambrinol. Mar. Drugs 2023, 21, 230. https://doi.org/10.3390/md21040230

AMA Style

López-Martínez JL, Torres-García I, Moreno-Gutiérrez I, Oña-Burgos P, Rosales Martínez A, Muñoz-Dorado M, Álvarez-Corral M, Rodríguez-García I. A Concise Diastereoselective Total Synthesis of α-Ambrinol. Marine Drugs. 2023; 21(4):230. https://doi.org/10.3390/md21040230

Chicago/Turabian Style

López-Martínez, Josefa L., Irene Torres-García, Irene Moreno-Gutiérrez, Pascual Oña-Burgos, Antonio Rosales Martínez, Manuel Muñoz-Dorado, Míriam Álvarez-Corral, and Ignacio Rodríguez-García. 2023. "A Concise Diastereoselective Total Synthesis of α-Ambrinol" Marine Drugs 21, no. 4: 230. https://doi.org/10.3390/md21040230

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop