Next Article in Journal
Antispasmodic Effect of Alstonia boonei De Wild. and Its Constituents: Ex Vivo and In Silico Approaches
Next Article in Special Issue
Aza Analogs of the TRPML1 Inhibitor Estradiol Methyl Ether (EDME)
Previous Article in Journal
The Effects of Honeysuckle (Lonicera caerulea L.) Berry Iridoid-Anthocyanin Extract on the Intestinal and Muscle Histopathology in Mice during Experimental Trichinellosis
Previous Article in Special Issue
Exploration of Binding Affinities of a 3β,6β-Diacetoxy-5α-cholestan-5-ol with Human Serum Albumin: Insights from Synthesis, Characterization, Crystal Structure, Antioxidant and Molecular Docking
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Montmorillonite Catalyzed Synthesis of Novel Steroid Dimers

by
Aneta M. Tomkiel
1,*,
Adam D. Majewski
1,2,
Leszek Siergiejczyk
1 and
Jacek W. Morzycki
1,*
1
Faculty of Chemistry, University of Bialystok, Ciołkowskiego 1K, 15-245 Bialystok, Poland
2
Doctoral School of Exact and Natural Sciences, University of Bialystok, Ciołkowskiego 1K, 15-245 Bialystok, Poland
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(20), 7068; https://doi.org/10.3390/molecules28207068
Submission received: 28 September 2023 / Revised: 10 October 2023 / Accepted: 10 October 2023 / Published: 13 October 2023
(This article belongs to the Special Issue Steroid Compounds with Potential Biological Activity)

Abstract

:
The reactions of sterols (androst-5-en-3β-ol-17-one, diosgenin, and cholesterol) and their tosylates with hydroquinone aimed at the synthesis of O,O-1,4-phenylene-linked steroid dimers were studied. The reaction course strongly depended on the conditions used. The study has shown that the major reaction products are the elimination products and unusual steroid dimers resulting from the nucleophilic attack of the hydroquinone C2 carbon atom on the steroid C3 position, followed by an intramolecular addition to the C5–C6 double bond. A different reaction course was observed when montmorillonite K10 was used as a catalyst. The reaction of androst-5-en-3β-ol-17-one under the promotion of this catalyst afforded the O,O-1,4-phenylene-linked steroid dimer in addition to the disteroidal ether. The formation of the latter compound was suppressed by using 3-tosylate as a substrate instead of the free sterol. The reactions of androst-5-en-3β-ol-17-one tosylate and cholesteryl tosylate with hydroquinone catalyzed by montmorillonite K10 carried out under optimized conditions afforded the desired dimers in 31% and 67% yield, respectively.

Graphical Abstract

1. Introduction

During the last few decades steroid dimers have emerged as an interesting new area of steroid chemistry. The first review article on steroid dimers was published by Li and Dias in 1997 [1]. Since then, numerous reports have appeared on these compounds, which have been reviewed in the book [2] and in several review articles by Nahar et al. [3,4]. In the beginning, steroid dimers, which were formed as byproducts of certain reactions, were considered mere curiosities. Later, steroid dimers were found in natural sources, e.g., japindine isolated from the root bark of Chonemorpha macrophylla [5], crellastatins from marine sponge Crella sp. [6], cephalostatins from tiny marine worm Cephalodiscus gilchristi [7], and ritterazines from tunicate Ritterella tokioka (Figure 1) [8]. The highly cytotoxic cephalostatins and ritterazines inspired chemists to attempt their chemical synthesis [9,10,11]. The synthesis of these dimeric steroidal pyrazine alkaloids and their analogs was particularly important because the natural sources of these compounds are extremely limited. These intensive studies were covered by several review articles [12,13,14,15]. Oligomeric steroids with or without spacer groups can be used as chiral building blocks to construct artificial receptors and as architectural components in biomimetic/molecular recognition chemistry. Davis has briefly reviewed his work in this area, directed toward the construction of enzyme mimics [16]. The study on dimerization and oligomerization of the cholic acid skeleton as an architectural building block was particularly intensive [17,18,19]. Various methods of synthesis of steroid dimers with linear [20] or cyclic [21] structures, e.g., Sonogashira coupling, Yamaguchi esterification, Wurtz reaction, ring closing metathesis, etc., were employed [22]. In the last decades, the interesting chemical, biological, and physical properties of the growing number of steroid dimers synthesized or isolated from living organisms have triggered increased activity in this field. The large number of steroid dimers described so far include compounds that have shown properties as catalysts [23], artificial receptors [24], molecular umbrellas [25], as well as activity as sulfatase inhibitors [26], antimalarials [27], and cytotoxic and antiproliferative agents [28,29], amongst others. Recent solid-state studies revealed that several crystalline steroid dimers act as molecular rotors [30].
Figure 1. Naturally occurring steroid dimers.
Figure 1. Naturally occurring steroid dimers.
Molecules 28 07068 g001
Figure 2. Target steroid dimers (13) and starting materials (46).
Figure 2. Target steroid dimers (13) and starting materials (46).
Molecules 28 07068 g002
The aim of this research was the synthesis of steroid dimers, in which steroid units are connected by an O,O-1,4-phenylene linker (Figure 2). These compounds (13) were selected as potential molecular rotors. Molecular motors are an important class of nanomolar-sized molecular machines that use different energy sources to generate unidirectional mechanical motion. In the crystal lattice, they contain both a static part and a fragment capable of free rotation [31]. Compound 1 has a carbonyl group at C17, which allows modification of the structure by functional group manipulation. Compounds 1 and 2 have not been described before, while compound 3 was previously obtained in low yield from cholesterol hydroquinone ether as an undesired product of an electrochemical reaction [32]. In the present study, a chemical procedure was elaborated for the efficient synthesis of dimers 13 from readily available sterols, androst-5-en-3β-ol-17-one 4a (dehydroepiandrosterone; DHEA), (25R)-spirost-5-en-3β-ol 5a (diosgenin), and cholesterol 6a, respectively.

2. Results and Discussion

The synthetic plan of dimers 13 assumed the conversion of starting sterols 4a6a into the corresponding p-tosylates 4b6b and their reaction with hydroquinone. It is well known that upon leaving p-TsO the mesomerically stabilized homoallylic carbocation (Scheme 1) is generated, which can react with a nucleophilic reagent, e.g., hydroquinone, in the 3β position. For stereoelectronic reasons, only 3β-substituted steroid products can be formed. In fact, the nucleophile may also attack the 6β position, but the resulting product (so-called i-steroid) is thermodynamically less stable and is not formed as a product under acidic conditions. However, the reaction of androst-5-en-3β-ol-17-one tosylate 4b (2 equiv.) with hydroquinone (1 equiv.) did not afford the expected dimer 1. There was no reaction in boiling acetone (even in the presence of p-TsOH), while the analogous reaction carried out in dioxane (24 h at reflux) yielded mostly the nonpolar elimination products 7 and 8 (Table 1; run 1). The fast solvent−free reaction at 120 °C (run 2) gave a mixture of several products, which were carefully analyzed.
In addition to a mixture of dienes 7 and 8, androst-5-en-3β-ol-17-one hydroquinone ether 9 and unexpected products 10 and 11 as stereomeric mixtures were obtained. The compound 9 could be an intermediate in the synthesis of dimer 1, but the desired compound 1 was not found among the reaction products. Instead, the different dimers were isolated, 11a11c, and their likely precursors, 10a and 10b. The similar reaction products, but in a different ratio (more elimination products), were formed in xylenes at reflux for 3 h (run 3). The structures of the products prove that both nucleophilic sites of hydroquinone, the oxygen atom and the carbon atom C2 of the aromatic ring, take part in the reaction. Interestingly, during the reaction, the hydroquinone carbon atom binds to the steroid C3 position, while the oxygen atom binds to C5. The direction of the reaction was deduced from the analysis of 1H NMR spectra. The chemical shifts of protons at C3 were shifted upfield (compared to δ 4.34 in tosylate 4b) and appeared at δ below 3 ppm. Therefore, the oxygen atom in this compound cannot be attached to C3. It seems that compounds 10b and 11b resulted from an attack of the carbon atom of hydroquinone on the mesomeric carbocation, which is formed upon the leaving of p-TsO from 4b (Scheme 2).
Table 1. The results of tosylate 4b (2 equiv.) reaction with hydroquinone (1 equiv.) under different conditions.
Table 1. The results of tosylate 4b (2 equiv.) reaction with hydroquinone (1 equiv.) under different conditions.
Run No./
Reaction Conditions
Structure of ProductsConversion
Molecules 28 07068 i001Molecules 28 07068 i002Molecules 28 07068 i003Molecules 28 07068 i004
1: Dioxane, reflux, 24 h16%<5%--30%
2: 120 °C, 5 min23%12%10a: <5%,
10b: <5%
11a: 7%,
11b: 5%,
11c: 9%
100%
3: Xylenes, reflux, 3 h48%13%10a: <1%,
10b: <1%
11a: <5%,
11b: 5%,
11c: 5%
100%
4: Ball mill, 48 h7%19%10a: <1%,
10b: <1%
11a: <5%,
11b: <5%,
11c: 5%
50%
5: Acetone, MW, 110 °C, 6 h47%8%10a: <1%,
10b: <1%
11a: <1%,
11b: <5%,
11c: 5%
90%
6: Dioxane, US, 45 °C,
12 h
5%---15%
This attack is possible from the steroid β side only due to stereoelectronic reasons. In the next step, an intramolecular addition of phenol to the steroid C5–C6 double bond occurred, leading to a six-membered ring formation. Compound 10b obtained this way may undergo the same reaction sequence to provide the dimer 11b. The mechanism of the dimer 11a formation (via 10a) is less obvious (Scheme 3). Apparently, the attack of hydroquinone on steroid tosylate 4b occurred from the less hindered α side in this case. However, the reaction timing is not clear. It could be either a nucleophilic substitution of 3-tosylate with the carbon C2 of hydroquinone followed by an intramolecular addition to the C5–C6 double bond or a concerted attack of hydroquinone on both 3-tosylate and the double bond.
The dimers 11a, 11b, and 11c were separated by column chromatography since they differ in polarity. The Rf values of dimers determined by their migration on TLC plates three times developed in the solvent system hexane—ethyl acetate (74:26) were 0.77, 0.66, and 0.71, respectively. The C2 symmetrical dimers 11a and 11b were isolated in similar amounts, while the mixed unsymmetrical dimer 11c was obtained in larger quantities. Optimized structures of dimers 11a and 11b are presented in Figure 3. The 1H NMR spectra of 11a and 11b were very similar, except for protons in the A-ring region. Particularly, the chemical shifts of protons at C3 and C3′ in these dimers proved different. The two-proton signal in dimer 11a came out at δ 2.90 ppm, while an analogous two-proton signal in 11b appeared at δ 2.85 ppm. Of course, the 1H NMR spectrum of mixed dimer 11c, which consists of two different steroid units, showed both proton signals (δ 2.85 and 2.90 ppm) as presented in Figure 4. The difference of 0.05 ppm between the chemical shifts of protons at the C3 and C3′ positions in dimers 11a and 11b is comparable to the literature values described for analogous A/B ring systems. For example, the 3β-H signal in 3α-hydroxy-5α-androstane-17-one is located at 4.05 ppm [33], while the analogous signal (3α-H) in 3β-hydroxy-5β-androstane-17-one shows up at 4.11 ppm [34]. In addition, the chemical shifts of the C4 protons in dimer 11a appear at δ 1.40 and 2.08 ppm, compared to 1.43 and 2.48 ppm for the corresponding proton signals in compound 11b. The signal of the 4α proton in 11b is significantly (0.40 ppm) shifted downfield compared to the analogous signal in 11a. A similar difference (Δδ 0.43 ppm) between the chemical shifts of protons at C4 was reported in the literature for 5α- and 5β-androstan-17-ones [35]. The full assignment of signals in 1H and 13C NMR spectra was carried out based on different NMR techniques (DEPT, COSY, HMQC, and HMBC) and is presented in Table S1 (Supplementary Material).
A series of experiments were performed using mechanochemistry (Table 1; run 4), microwave (run 5), or ultrasound irradiation (run 6), but none of them produced reasonable yields of tosylate 4b with hydroquinone reaction products.
In the next series of experiments, montmorillonite was used as a catalyst. Montmorillonite is a very soft phyllosilicate group of minerals that form when they precipitate from water solution as microscopic crystals, known as clay. It is named after Montmorillon in France. Montmorillonite, a member of the smectite group, is a 2:1 clay, meaning that it has two tetrahedral sheets of silica sandwiching a central octahedral sheet of alumina. The acid-treated montmorillonite clays have been extensively used in various catalytic processes [36]. Other heterogeneous Brønsted and Lewis acid catalysts could be developed using the cation-exchange ability of the montmorillonite interlayer. The organic syntheses using the above acid catalysts have several advantages: ease of preparing the solid catalysts, high catalytic activities, wide applicability to large molecules, simple workup procedure, and reusability of the catalyst [37]. Montmorillonite was also used for the preparation of ethers from alcohols and phenols. A comprehensive review of the catalytic O-alkylation of phenol and hydroquinone has recently appeared [38]. Also, the preparation of cholesteryl ethers from cholesterol and various alcohols or phenols has been described [39].
Before we started studying reactions catalyzed with montmorillonite, the reaction of androst-5-en-3β-ol-17-one tosylate (4b) with hydroquinone in the presence of basic Al2O3 was carried out. Unfortunately, the reaction was dirty and led to i-steroid products (e.g., 3α,5α-cyclo-androst-6-en-17-one) predominately. Therefore, a conclusion has been drawn that basic conditions should be avoided, and further reactions were carried out with K10 (commercial montmorillonite) as a catalyst under different conditions (Table 2). For these reactions, a non-activated androst-5-en-3β-ol-17-one (4a) was used instead of tosylate 4b. In addition to the previously described products, new dimeric products, i.e., disteroidal ether 12 and the desired hydroquinone disteroidal diether 1, appeared among the reaction products. The synthesis of disteroidal ethers was previously described. They can be efficiently prepared from Δ5-steroids by treating them with montmorillonite in a dichloromethane solution [40] or by an electrochemical method [41]. In both cases, a mesomeric homoallylic carbocation (Scheme 1) is formed as an intermediate, which finally reacts with the starting sterol.
The reaction was optimized against the quantity of catalyst used. With 500 mg of K10 activated at 120 °C, a 10% yield of dimer 1 was achieved (Table 2; run 1); higher amounts of catalyst deteriorated (run 2) the reaction result. The temperature of the montmorillonite activation is very important. Montmorillonite changes its properties during activation due to the desorption of water, dehydration, and alteration of crystalline structure. The activation at 120 °C causes water desorption, which exposes acid groups. This increases the number and strength of acid centers. A higher temperature of activation results in partial dehydration of Si-OH and Al-OH groups. Then the number of Brønsted acid sites decreases, while the number of Lewis acid (e.g., Al3+) sites increases. The reactions of hydroquinone with 2.5 equiv. of DHEA (4a) in CHCl3 at reflux catalyzed with K10 (500 mg) activated at different temperatures were carried out (runs 3–6). The best result was obtained with K10 activated at 280 °C (run 5). The desired dimer 1 was obtained in 23% under optimized conditions. Further attempts to increase the yield of dimer 1 failed. The experiments presented in Table 2 were conducted in 5 mL of CHCl3 at reflux. The reaction carried out as in the case of run 5, but in a more diluted solution (10 mL of solvent), gave essentially the same result (23% of 1). The reaction performed in 1,2-dichloroethane at reflux (83 °C) provided only 13% of 1, while in other solvents at reflux (cyclohexane, toluene, xylenes, acetonitrile, acetone, THF, and dioxane), dimer 1 was not formed at all. Also, the attempts to carry out the reaction at room temperature were completely unsuccessful. In the next series of experiments, montmorillonite activated with hydrochloric acid was used. With this catalyst dried at 280 °C (H+-K10), up to a 57% yield of disteroidal ether 12 (Table 2; runs 7 and 8) could be achieved, but the yield of dimer 1 was negligible. It seems that increasing the number of Brønsted reactive sites deteriorates the yield of dimer 1. Some experiments were carried out with montmorillonite activated with the ions Ti4+ or Cu2+. However, this modification of the catalyst promoted only the elimination processes (runs 9 and 10).
The optimal reaction conditions (as in run 5) were applied to analogous reactions of other sterols—diosgenin 5a and cholesterol 6a. In the latter case, the yield of desired dimer 3 was slightly better (28%) than that achieved for DHEA (4a). The dimer 3 was accompanied by dicholesteryl ether (22%), hydroquinone mono cholesteryl ether (10%), and the elimination products (35%). The yield of diosgenin-derived dimer 2 was lower (12%) under the same conditions. The corresponding disteroidal ether was isolated in tiny amounts [42].
In the next experiments, androst-5-en-3β-ol-17-one tosylate (4b) and cholesteryl tosylate (6b) served as substrates. The advantage of using tosylates instead of free sterols is that disteroidal ethers cannot be formed under these conditions, which is important from the point of view of reaction product separation. Also, dimers 11ac (or analogous compounds derived from cholesterol) were formed only in negligible amounts. In addition, much milder reaction conditions were used, resulting in fewer by-products. The results of tosylate reactions are shown in Table 3. It should be noted that all reactions occurred at ambient temperature and required less catalyst. In the case of the androstane series (runs 1–3), the best yield of the desired dimer 1 (31%) was observed for the reaction catalyzed by Ti4+ activated montmorillonite K10 (run 3). Without Ti4+ activation, the yield of 1 was 23%, provided that K10 was calcined at 500 °C (run 2). The reactions of cholesteryl tosylate 6b with hydroquinone (runs 4–6) worked much better and provided dimer 3 in excellent yields. Especially the catalyst dried at 500 °C (run 5) was found to be very active in promoting the reaction, even in low quantities (50 mg). The likely explanation for the high activity of montmorillonite K10 against cholesterol 6a and its tosylate 6b is the lack of additional groups in these substrates capable of binding to the active catalyst sites.

3. Materials and Methods

3.1. General Experimental Data

All solvents were freshly distilled prior to use. Anhydrous solvents were prepared by distillation over appropriate drying agents under an argon atmosphere. The stabilizer (ethanol) contained in commercially available chloroform was removed before use. The reactions were monitored by TLC on silica gel plates 60 F254, and spots were visualized either by a UV hand lamp or by charring with molybdophosphoric acid/cerium(IV) sulfate in H2SO4. The reaction products were isolated by chromatographic methods using silica gel pore size 40 Å (70–230 mesh). 1H and 13C NMR (400 and 100 MHz, respectively) spectra of all compounds were recorded using a Bruker Avance II spectrometer in a CDCl3 or CDCl3/MeOD mixture and referenced to TMS (0.0 ppm) and CDCl3 (77.0 ppm), respectively. Infrared spectra were recorded using Attenuated Total Reflectance (ATR) as solid samples with a Nicolet 6700 FT-IR spectrometer. Mass spectra were obtained with an Accurate-Mass Q-TOFLC/MS 6530 spectrometer with electrospray ionization (ESI). Melting points were determined on a Kofler bench melting point apparatus.
Tosylates 4b [43] and 6b [44] were prepared according to literature procedures.
Unmodified montmorillonite K10 or metal cation-exchanged montmorillonites: H+-K10, Ti4+-K10, and Cu2+-K10 were used as catalysts. Modified forms: H+-K10 [45], Ti4+-K10 [46], and Cu2+-K10 [47] were obtained according to the literature procedures. All catalysts were activated prior to use by heating at high temperatures (unmodified montmorillonite K10 at 120 °C, 200 °C, 280 °C, 400 °C, or 500 °C; H+-K10 at 280 °C; Ti4+-K10 at 280 °C or 500 °C; and Cu2+-K10 at 120 °C). Characterization data and NMR spectra are presented in the supporting information.

3.2. General Experimental Procedure for the Formation of All New Compounds

3.2.1. Solvent-Free Reaction of Tosylate 4b with Hydroquinone (Table 1, Run No. 2)

The mixture of tosylate 4b (100 mg. 0.2 mmol) and hydroquinone (11 mg. 0.1 mmol) was stirred and heated to 120 °C for 5 min under an argon atmosphere. Then, the reaction mixture was cooled to room temperature and subjected to silica gel column chromatography, which resulted in the separation of compounds: 7 and 8 (eluted with hexane in 23% yield), 11a (eluted with hexane/ethyl acetate 97:3 mixture in 7% yield), 11c (eluted with hexane/ethyl acetate 95:5 mixture in 9% yield), 10a (eluted with hexane/ethyl acetate 93:7 mixture in <5% yield), 11b (eluted with hexane/ethyl acetate 93:7 mixture in 5% yield), 9 (eluted with hexane/ethyl acetate 87:13 mixture in 12% yield), and 10b (eluted with hexane/ethyl acetate 87:13 mixture in <5% yield).
  • Compound 7: white solid (hexane/CH2Cl2); mp 85–87 °C; Rf = 0.32 (hexane/ethyl acetate 9:1); IR (ATR) νmax 2912, 2856, 1739 cm–1; 1H NMR (CDCl3, 400 MHz) δ 5.95 (1H, d, J = 9.9 Hz, H-4), 5.63 (1H, m, H-3), 5.42 (1H, m, H-6), 2.48 (1H, dd, J = 19.2 Hz, J = 8.8 Hz, H-16β), 0.99 (3H, s, H-19), 0.93 (3H, s, H-18); 13C NMR (CDCl3, 100 MHz) δ 221.1 (C), 141.6 (C), 128.7 (CH), 125.4 (CH), 122.1 (CH), 51.9 (CH), 48.5 (CH), 47.7 (C), 35.8 (CH2), 35.3 (C), 33.7 (CH2), 31.44 (CH2), 31.41 (CH), 30.6 (CH2), 23.0 (CH2), 21.8 (CH2), 20.3 (CH2), 18.8 (CH3), 13.7 (CH3); HRMS m/z 271.2054 (calcd for C19H27O+, 271.2056).
  • Compound 8: white solid; Rf = 0.36 (hexane/ethyl acetate 9:1); IR (ATR) νmax 2916, 1736, 1506 cm–1; 1H NMR (CDCl3, 400 MHz) δ 5.99 (1H, dd, J = 9.8 Hz, J = 2.5 Hz, H-4), 5.56 (1H, d, J = 9.8 Hz, H-7), 5.47 (1H, m, H-3), 0.98 (3H, s, H-19), 0.96 (3H, s, H-18); 13C NMR (CDCl3, 100 MHz) δ 220.7 (C), 142.2 (C), 129.9 (CH), 125.6 (CH), 124.7 (CH), 51.6 (CH), 49.6 (CH), 48.3 (C), 36.5 (CH), 35.8 (CH2), 34.8 (C), 34.5 (CH2), 31.6 (CH2), 25.3 (CH2), 21.5 (CH2), 20.2 (CH2), 18.4 (CH2), 18.3 (CH3), 13.8 (CH3); HRMS m/z 271.2061 (calcd for C19H27O+, 271.2056).
  • Compound 10a: white solid; Rf = 0.66 (3 × hexane/ethyl acetate 74:26); IR (ATR) νmax 3325, 2927, 1737, 1241, 1206, 1190, 1153, 1077, 1053, 813, 785 cm–1; 1H NMR (CDCl3, 400 MHz) δ 6.67 (1H, d, J = 8.6 Hz, H-Ar), 6.60 (1H, dd, J = 8.6 Hz, J = 3.0 Hz, H-Ar), 6.48 (1H, d, J = 3.0 Hz, H-Ar), 4.26 (1H, bs, -OH), 2.87 (1H, m, H-3β), 2.46 (1H, dd, J = 18.9 Hz, J = 8.3 Hz, H-16β), 1.05 (3H, s, H-19), 0.88 (3H, s, H-18); 13C NMR (CDCl3, 100 MHz) δ 221.4 (C), 150.1 (C), 147.9 (C), 128.0 (C), 115.3 (CH), 114.2 (CH), 113.9 (CH), 78.5 (C), 51.2 (CH), 47.8 (C), 46.2 (CH), 42.1 (C), 35.9 (CH2), 34.4 (CH), 33.6 (CH2), 32.7 (CH), 31.8 (CH2), 31.5 (CH2), 29.41 (CH2), 29.36 (CH2), 24.8 (CH2), 21.7 (CH2), 20.0 (CH2), 16.2 (CH3), 13.8 (CH3); HRMS m/z 381.2428 (calcd for C25H33O3+, 381.2424).
  • Compound 11a: white solid; Rf = 0.77 (3 × hexane/ethyl acetate 74:26); IR (ATR) νmax 2915, 2853, 1737, 1242, 1205, 1151, 1001, 813, 785 cm–1; 1H NMR (CDCl3, 400 MHz) δ 6.41 (2H, s, H-Ar), 2.85 (2H, m, H-3β), 2.46 (2H, dd, J = 18.9 Hz, J = 8.2 Hz, H-16β), 1.04 (6H, s, H-19), 0.88 (6H, s, H-18); 13C NMR (CDCl3, 100 MHz) δ 221.5 (2 × C), 148.4 (2 × C), 126.1 (2 × C), 112.8 (2 × CH), 78.1 (2 × C), 51.2 (2 × CH), 47.8 (2 × C), 46.1 (2 × CH), 42.1 (2 × C), 35.9 (2 × CH2), 34.4 (2 × CH), 33.7 (2 × CH2), 32.6 (2 × CH), 32.2 (2 × CH2), 31.6 (2 × CH2), 29.5 (2 × CH2), 29.3 (2 × CH2), 24.8 (2 × CH2), 21.7 (2 × CH2), 20.0 (2 × CH2), 16.1 (2 × CH3), 13.8 (2 × CH3); HRMS m/z 651.4408 (calcd for C44H59O4+, 651.4408).
  • Compound 11b: white solid; Rf = 0.66 (3 × hexane/ethyl acetate 74:26); IR (ATR) νmax 2914, 2853, 1737, 1242, 1205, 813, 785 cm–1; 1H NMR (CDCl3, 400 MHz) δ 6.43 (2H, s, H-Ar), 2.90 (2H, m, H-3α), 2.51–2.43 (4H, m, H-4α and H-16β), 1.03 (6H, s, H-19), 0.88 (6H, s, H-18); 13C NMR (CDCl3, 100 MHz) δ 221.0 (2 × C), 149.1 (2 × C), 125.4 (2 × C), 113.1 (2 × CH), 78.4 (2 × C), 51.6 (2 × CH), 47.9 (2 × C), 43.8 (2 × CH), 42.9 (2 × C), 35.9 (2 × CH2), 34.6 (2 × CH), 34.4 (2 × CH2), 32.6 (2 × CH), 31.6 (2 × CH2), 30.8 (2 × CH2), 28.7 (2 × CH2), 27.9 (2 × CH2), 27.1 (2 × CH2), 21.8 (2 × CH2), 20.6 (2 × CH2), 17.6 (2 × CH3), 13.8 (2 × CH3); HRMS m/z 651.4401 (calcd for C44H59O4+, 651.4408).
  • Compound 11c: white solid; Rf = 0.71 (3 × hexane:ethyl acetate 74/26); IR (ATR) νmax 2929, 2870, 1737, 1241, 1194, 1152, 785 cm–1; 1H NMR (CDCl3, 400 MHz) δ 6.43 (2H, s, H-Ar), 2.91 (1H, m, H-3α), 2.85 (1H, m, H-3β), 1.03 (3H, s, H-19), 1.01 (3H, s, H-19′), 0.88 (s, 6H, H-18 and H-18′); 13C NMR (CDCl3, 100 MHz) δ 221.4 (C), 220.9 (C), 149.0 (C), 148.7 (C), 126.1 (C), 125.2 (C), 113.0 (CH), 112.7 (CH), 78.3 (C), 78.1 (C), 51.6 (CH), 51.2 (CH), 47.8 (2 × C), 46.1 (CH), 43.9 (CH), 42.9 (C), 42.1 (C), 35.9 (2 × CH2), 34.5 (CH), 34.4 (CH), 34.4 (CH2), 33.6 (CH2), 32.5 (2 × CH), 32.0 (CH2), 31.6 (CH2), 31.5 (CH2), 30.7 (CH2), 29.5 (CH2), 29.4 (CH2), 28.8 (CH2), 27.9 (CH2), 27.0 (CH2), 24.8 (CH2), 21.8 (CH2), 21.7 (CH2), 20.6 (CH2), 20.0 (CH2), 17.6 (CH3), 16.2 (CH3), 13.8 (2 × CH3); HRMS m/z 651.4403 (calcd for C44H59O4+, 651.4408).

3.2.2. Optimal Procedure for Preparation of Dimer 1 in the Montmorillonite K10 Catalyzed Reaction of Androst-5-en-3β-ol-17-one (4a) with Hydroquinone (Table 2, Run No. 5)

A stirred mixture of androst-5-en-3β-ol-17-one (4a) (100 mg, 0.35 mmol), hydroquinone (15.4 mg, 0.14 mmol), and unmodified montmorillonite K10 activated at 280 °C (500 mg) in dry chloroform (5 mL) was gently refluxed under argon, and the reaction progress was monitored by TLC. After completion of the reaction (4 h), the suspension was filtered through a sintered glass funnel, and the precipitate was washed with a methanol/chloroform 2:8 mixture (3 × 50 mL). The filtrate was evaporated in a vacuum. The residue was subjected to column chromatography on silica gel, which resulted in the separation of compounds 7 and 8 (eluted with hexane in 28% yield), 12 (eluted with hexane/ethyl acetate 95:5 mixture in 7% yield), 1 (eluted with hexane/ethyl acetate 94:6 mixture in 23% yield), 9 (eluted with hexane/ethyl acetate 87:13 mixture in 12% yield), and small amounts of compounds 11a, 11b, and 11c (total <5%).
  • Compound 1: colorless crystals (CH2Cl2/ethyl acetate); mp 252–254 °C; Rf = 0.50 (3 × benzene/ethyl acetate 94:6); IR (ATR) νmax 2938, 2907, 1744, 1731, 1501, 1214, 1043, 1029, 815 cm–1; 1H NMR (CDCl3, 400 MHz) δ 6.82 (4H, s, H-Ar), 5.42 (2H, m, H-6), 4.00 (2H, m, H-3α), 1.09 (6H, s, H-19), 0.91 (6H, s, H-18); 13C NMR (CDCl3, 100 MHz) δ 221.0 (2 × C), 151.8 (2 × C), 140.7 (2 × C), 121.4 (2 × CH), 117.4 (4 × CH), 77.9 (2 × CH), 51.8 (2 × CH), 50.3 (2 × CH), 47.5 (2 × C), 38.8 (2 × CH2), 37.1 (2 × CH2), 37.0 (2 × C), 35.8 (2 × CH2), 31.5 (2 × CH), 31.4 (2 × CH2), 30.8 (2 × CH2), 28.3 (2 × CH2), 21.9 (2 × CH2), 20.4 (2 × CH2), 19.4 (2 × CH3), 13.5 (2 × CH3); HRMS m/z 651.4409 (calcd for C44H59O4+, 651.4408).
  • Compound 9: colorless crystals (CH2Cl2/ethyl acetate); mp 277–279 °C; Rf = 0.48 (3 × hexane/ethyl acetate 74:26); IR (ATR) νmax 3311, 2948, 2864, 1710, 1505, 1211, 1028, 819 cm−1; 1H NMR (CDCl3/MeOD, 400 MHz) δ 6.76 (2H, d, J = 19.2 Hz, H-Ar), 6.71 (2H, d, J = 19.1 Hz, H-Ar), 5.37 (1H, m, H-6), 3.93 (1H, m, H-3α), 1.05 (3H, s, H-19), 0.87 (3H, s, H-18); 13C NMR (CDCl3/MeOD, 100 MHz) δ 221.9 (C), 150.8 (C), 150.7 (C), 140.7 (C), 121.3 (CH), 117.9 (2 × CH), 115.8 (2 × CH), 78.3 (CH), 51.7 (CH), 50.2 (CH), 47.6 (C), 38.8 (CH2), 37.0 (CH2), 36.9 (C), 35.8 (CH2), 31.4 (CH), 31.3 (CH2), 30.7 (CH2), 28.2 (CH2), 21.8 (CH2), 20.3 (CH2), 19.4 (CH3), 13.5 (CH3); HRMS m/z 381.2438 (calcd for C25H33O3+, 381.2424).
  • Compound 12: colorless crystals (hexane/CH2Cl2); mp 268–269 °C; Rf = 0.45 (3 × benzene/ethyl acetate 94:6); IR (ATR) νmax 2931, 2895, 1731, 1094, 1058, 1005 cm–1; 1H NMR (CDCl3, 400 MHz) δ 5.38 (2H, m, H-6), 3.30 (2H, m, H-3α), 2.47 (2H, dd, J = 19.2 Hz, J = 8.6 Hz, H-16β), 1.04 (6H, s, H-19), 0.89 (6H, s, H-18); 13C NMR (CDCl3, 100 MHz) δ 221.2 (2 × C), 141.5 (2 × C), 120.6 (2 × CH), 76.2 (2 × CH), 51.8 (2 × CH), 50.3 (2 × CH), 47.5 (2 × C), 40.0 (2 × CH2), 37.3 (2 × CH2), 37.0 (2 × C), 35.8 (2 × CH2), 31.5 (2 × CH), 31.4 (2 × CH2), 30.8 (2 × CH2), 29.3 (2 × CH2), 21.9 (2 × CH2), 20.3 (2 × CH2), 19.4 (2 × CH3), 13.5 (2 × CH3).
Analogously, steroid dimers 2 (12%) and 3 (28%) were obtained from diosgenin (5a) and cholesterol (6a), respectively, according to the procedure described above for androst-5-en-3β-ol-17-one (4a). The elimination products (~35% in both cases), disteroidal ethers (12% and 22%, respectively), and hydroquinone mono steroidal ethers (12% and 10%, respectively) were also formed.
  • Compound 2: colorless crystals (hexane/CH2Cl2); mp 307–309 °C; Rf = 0.38 (3 × benzene/ethyl acetate 94:6); IR (ATR) νmax 2925, 1502, 1225, 1050, 1016, 809 cm–1; 1H NMR (CDCl3, 400 MHz) δ 6.82 (4H, s, H-Ar), 5.38 (2H, m, H-6), 4.43 (2H, m, H-16), 3.98 (2H, m, H-3α), 3.49 (2H, m, H-26β), 3.39 (2H, t, J = 10.9 Hz, H-26α), 1.08 (6H, s, H-19), 0.99 (6H, d, J = 6.9 Hz, H-21), 0.81 (6H, s, H-18), 0.80 (6H, d, J = 5.0 Hz, H-27); 13C NMR (CDCl3, 100 MHz) δ 151.8 (2 × C), 140.5 (2 × C), 121.9 (2 × CH), 117.4 (4 × CH), 109.3 (2 × C), 80.8 (2 × CH), 78.0 (2 × CH), 66.8 (2 × CH2), 62.1 (2 × CH), 56.5 (2 × CH), 50.1 (2 × CH), 41.6 (2 × CH), 40.3 (2 × C), 39.8 (2 × CH2), 38.8 (2 × CH2), 37.2 (2 × CH2), 37.0 (2 × C), 32.1 (2 × CH2), 31.8 (2 × CH2), 31.43 (2 × CH), 31.38 (2 × CH2), 30.3 (2 × CH), 28.8 (2 × CH2), 28.3 (2 × CH2), 20.8 (2 × CH2), 19.4 (2 × CH3), 17.1 (2 × CH3), 16.3 (2 × CH3), 14.5 (2 × CH3); HRMS m/z 903.6502 (calcd for C60H87O6+, 903.6497).
  • Diosgenin derived hydroquinone mono steroidal ether: colorless crystals (hexane/CH2Cl2); mp 165–166 °C; Rf = 0.35 (hexane/ethyl acetate 88:12); IR (ATR) νmax 3309, 2929, 1507, 1210, 1046, 1012, 830 cm–1; 1H NMR (CDCl3, 400 MHz) δ 6.80 (2H, d, J = 9.0 Hz, H-Ar), 6.75 (2H, d, J = 9.0 Hz, H-Ar), 5.38 (1H, m, H-6), 4.76 (1H, bs, -OH), 4.43 (1H, m, H-16), 3.97 (1H, m, H-3α), 3.49 (1H, m, H-26β), 3.39 (1H, t, J = 10.9 Hz, H-26α), 1.07 (3H, s, H-19), 0.99 (3H, d, J = 6.9 Hz, H-21), 0.81 (3H, s, H-18), 0.80 (3H, d, J = 6.4 Hz, H-27); 13C NMR (CDCl3, 100 MHz) δ 151.6 (C), 149.8 (C), 140.5 (C), 121.9 (CH), 117.8 (CH), 116.0 (CH), 109.4 (C), 80.8 (CH), 78.3 (CH), 66.9 (CH2), 62.1 (CH), 56.5 (CH), 50.1 (CH), 41.6 (CH), 40.3 (C), 39.8 (CH2), 38.8 (CH2), 37.1 (CH2), 37.0 (C), 32.1 (CH2), 31.8 (CH2), 31.42 (CH), 31.37 (CH2), 30.3 (CH), 28.8 (CH2), 28.3 (CH2), 20.9 (CH2), 19.4 (CH3), 17.1 (CH3), 16.3 (CH3), 14.5 (CH3); HRMS m/z 507.3476 (calcd for C33H47O4+, 507.3469).

3.2.3. Optimal Procedure for Preparation of Dimer 12 in the Montmorillonite K10 Catalyzed Reaction of Androst-5-en-3β-ol-17-one (4a) (Table 2, Run No. 8)

A stirred mixture of androst-5-en-3β-ol-17-one (4a) (100 mg, 0.35 mmol), hydroquinone (15.4 mg, 0.14 mmol), and modified montmorillonite H+-K10 activated at 280 °C (500 mg) in dry chloroform (5 mL) was heated to 50 °C for 24 h under argon, and the reaction progress was monitored by TLC. After completion of the reaction, the suspension was filtered through a sintered glass funnel, and the precipitate was washed with a methanol/chloroform 2:8 mixture (3 × 50 mL). The filtrate was evaporated in a vacuum. The residue was subjected to silica gel column chromatography, which resulted in separation of compounds 7 and 8 (eluted with hexane in 13% yield), 12 (eluted with hexane/ethyl acetate 95:5 mixture in 57% yield), 1 (eluted with hexane/ethyl acetate 94:6 mixture in 1% yield), and 9 (eluted with hexane/ethyl acetate 87:13 mixture in 7% yield).

3.2.4. Optimal Procedure for Preparation of Dimer 1 in the Montmorillonite K10 Catalyzed Reaction of Tosylate 4b with Hydroquinone (Table 3, Run No. 3)

A mixture of tosylate 4b (88.5 mg; 0.2 mmol), hydroquinone (8.4 mg; 0.076 mmol), and modified montmorillonite Ti4+-K10 activated at 280 °C (300 mg) in dry chloroform (5 mL) was stirred under argon, and the reaction progress was monitored by TLC. After completion of the reaction (3 days), the suspension was filtered through a sintered glass funnel, and the precipitate was washed with a methanol/chloroform 2:8 mixture (3 × 50 mL). The filtrate was evaporated in a vacuum. The residue was subjected to column chromatography on silica gel, which resulted in the separation of compounds 7 and 8 (eluted with hexane in 28% yield), 1 (eluted with hexane/ethyl acetate 94:6 mixture in 31% yield), and 9 (eluted with hexane/ethyl acetate 87:13 mixture in 19% yield).
The reaction of tosylate 6b with hydroquinone was carried out according to the procedure described above for tosylate 4b, but using montmorillonite K10 activated at 500 °C (50 mg) as a catalyst (Table 3, Run No. 5). Dimer 3 was obtained with a 67% yield. Small amounts of elimination products (10%) were also formed.

4. Conclusions

The reactions of sterols and their tosylates with hydroquinone were studied. They provided different steroid dimers depending on the reaction conditions. The solvolytic reactions of DHEA tosylate afforded the elimination products in addition to three stereoisomeric steroid dimers 11a11c, which resulted from the nucleophilic attack of the C2 carbon atom of hydroquinone on the C3 position of the steroid, followed by an intramolecular addition to the C5–C6 double bond. The cause of reactions catalyzed by montmorillonite was different. The major reaction products were the steroid dimers with a 3,3′-O,O-1,4-phenylene linker 13 and the disteroidal ethers 12. The formation of the latter compounds was suppressed by using sterol tosylates for the montmorillonite-catalyzed reactions with hydroquinone. As a result, an excellent yield (67%) of hydroquinone dicholesteryl diether (dimer 3) was achieved.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28207068/s1, Table S1: 1H NMR and 13C NMR signal assignments for dimers 11a and 11b, 1H and 13C NMR spectra of all synthesized compounds.

Author Contributions

Conceptualization, A.M.T. and J.W.M.; methodology, A.M.T. and J.W.M.; validation, J.W.M.; formal analysis, A.M.T. and L.S.; investigation, A.D.M., A.M.T. and L.S.; resources, J.W.M.; data curation, A.M.T. and A.D.M.; writing—original draft preparation, J.W.M.; writing—review and editing, J.W.M. and A.M.T.; visualization, A.M.T.; supervision, J.W.M.; project administration, A.M.T.; funding acquisition, J.W.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All generated data will be available from the authors upon request.

Acknowledgments

The authors thank the University of Bialystok for its continuous support of our research programs. The authors are grateful to Jadwiga Maj for her skillful technical assistance.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. Li, Y.; Dias, J.R. Dimeric and oligomeric steroids. Chem. Rev. 1997, 97, 283–304. [Google Scholar] [CrossRef] [PubMed]
  2. Nahar, L.; Sarker, S.D. Steroid Dimers—Chemistry and Applications in Drug Design and Delivery; John Wiley & Sons: Chichester, UK, 2012; ISBN 978-1-119-97284-6. [Google Scholar]
  3. Nahar, L.; Sarker, S.D.; Turner, A.B. A review on synthetic and natural steroid dimers: 1997–2006. Curr. Med. Chem. 2007, 14, 1349–1370. [Google Scholar] [CrossRef] [PubMed]
  4. Nahar, L.; Sarker, S.D. A review on steroid dimers: 2011–2019. Steroids 2020, 164, 108736. [Google Scholar] [CrossRef] [PubMed]
  5. Banerji, J.; Chatterjee, A.; Itoh, Y.; Kikuchi, T. New steroid alkaloids from Chonemorpha macrophylla G. don (Chonemorpha fragrans Moon Alston). Indian J. Chem. 1973, 11, 1056–1057, ISSN/ISBN 0019-5103. [Google Scholar]
  6. D’Auria, M.V.; Giannini, C.; Zampella, A.; Minale, L.; Debitus, C.; Roussakis, C. Crellastatin A:  A Cytotoxic Bis-Steroid Sulfate from the Vanuatu Marine Sponge Crella sp. J. Org. Chem. 1998, 63, 7382–7388. [Google Scholar] [CrossRef]
  7. Pettit, G.R.; Kamano, Y.; Dufresne, G.; Inoue, M.; Christie, N.; Schmidt, K.M.; Doubek, D.L. Isolation and structure of the unusual Indian Ocean Cephalodiscus gilchristi components, cephalostatins 5 and 6. Can. J. Chem. 1989, 67, 1509–1513. [Google Scholar] [CrossRef]
  8. Fuzukawa, S.; Matsunaga, S.; Fusetani, N. Ritterazine A, a highly cytotoxic dimeric steroidal alkaloid, from the tunicate Ritterella tokioka. J. Org. Chem. 1994, 59, 6164–6166. [Google Scholar] [CrossRef]
  9. LaCour, T.G.; Guo, C.; Bhandaru, S.; Boyd, M.R.; Fuchs, P.L. Interphylal product splicing: The first total syntheses of cephalostatin 1, the North hemisphere of ritterazine G, and the highly active hybrid analogue, ritterostatin GN1N1. J. Am. Chem. Soc. 1998, 120, 692–697. [Google Scholar] [CrossRef]
  10. Shi, Y.; Jia, L.; Xiao, Q.; Lan, Q.; Tang, X.; Wang, D.; Li, M.; Ji, Y.; Zhou, T.; Tian, W. A Practical Synthesis of Cephalostatin 1. Chem. Asian J. 2011, 6, 786–790. [Google Scholar] [CrossRef]
  11. Jautelat, R.; Müller-Fahrnow, A.; Winterfeldt, E. A novel oxidative cleavage of the steroidal skeleton. Chem. Eur. J. 1999, 5, 1226–1233. [Google Scholar] [CrossRef]
  12. Iglesias-Arteaga, M.A.; Morzycki, J.W. Cephalostatins and ritterazines. In The Alkaloids, Chemistry and Biology; Knölker, H.-J., Ed.; Academic Press: Cambridge, MA, USA, 2013; Volume 72, pp. 153–279. [Google Scholar] [CrossRef]
  13. Ganesan, A. The dimeric steroid-pyrazine marine alkaloids: Challenges for isolation, synthesis, and biological studies. Angew. Chem. Int. Ed. Engl. 1996, 35, 611–615. [Google Scholar] [CrossRef]
  14. Gryszkiewicz-Wojtkielewicz, A.; Jastrzębska, I.; Morzycki, J.W.; Romanowska, D.B. Approaches towards the synthesis of cephalostatins, ritterazines and saponins from Ornithogalum saundersiae—New natural products with cytostatic activity. Curr. Org. Chem. 2003, 7, 1257–1277. [Google Scholar] [CrossRef]
  15. Lee, S.; LaCour, T.G.; Fuchs, P.L. Chemistry of trisdecacyclic pyrazine antineoplastics: The cephalostatins and ritterazines. Chem. Rev. 2009, 109, 2275–2314. [Google Scholar] [CrossRef] [PubMed]
  16. Davis, A.P. Cholaphanes et al.; steroids as structural components in molecular engineering. Chem. Soc. Rev. 1993, 22, 243–253. [Google Scholar] [CrossRef]
  17. Bonar-Law, R.P.; Davis, A.P. Synthesis of steroidal cyclodimers from cholic acid; a molecular framework with potential for recognition and catalysis. J. Chem. Soc. Chem. Commun. 1989, 15, 1050–1052. [Google Scholar] [CrossRef]
  18. Joachimiak, R.; Paryzek, Z. Synthesis and alkaline metal ion binding ability of new steroid dimers derived from cholic and lithocholic acids. J. Inclusion Phenom. Macrocycl. Chem. 2004, 49, 127–132. [Google Scholar] [CrossRef]
  19. Virtanen, E.; Kolehmainen, E. Use of bile acids in pharmacological and supramolecular applications. Eur. J. Org. Chem. 2004, 16, 3385–3399. [Google Scholar] [CrossRef]
  20. Vazquez-Chavez, J.; Aguilar-Granda, A.; Iglesias Arteaga, M.A. Synthesis and characterization of a fluorescent steroid dimer linked through C-19 by a 1,4-Bis(phenylethynyl)phenylene fragment. Steroids 2022, 187, 109098. [Google Scholar] [CrossRef]
  21. Czajkowska, D.; Morzycki, J.W. Synthesis of cholaphanes by ring closing metathesis. Tetrahedron Lett. 2007, 48, 2851–2855. [Google Scholar] [CrossRef]
  22. Jurášek, M.; Džubák, P.; Sedlák, D.; Dvořáková, H.; Hajdúch, M.; Bartůnĕk, P.; Drašar, P. Preparation, preliminary screening of new types of steroid conjugates and their activities on steroid receptors. Steroids 2013, 78, 356–361. [Google Scholar] [CrossRef]
  23. Guthrie, J.P.; Cossar, J.; Darson, B.A. A water soluble dimeric steroid with catalytic properties. Rate enhancements from hydrophobic binding. Can. J. Chem. 1986, 64, 2456–2469. [Google Scholar] [CrossRef]
  24. Chattopadhyay, P.; Pandey, P.S. Synthesis and binding ability of bile acid-based receptors for recognition of flavin analogues. Tetrahedron 2006, 62, 8620–8624. [Google Scholar] [CrossRef]
  25. Janout, V.; Staina, I.V.; Bandyopadhyay, P.; Regen, S.L. Evidence for an umbrella mechanism of bilayer transport. J. Am. Chem. Soc. 2001, 123, 9926–9927. [Google Scholar] [CrossRef]
  26. Fournier, D.; Poirier, D. Estradiol dimers as a new class of steroid sulfatase reversible inhibitors. Bioorg. Med. Chem. Lett. 2009, 19, 693–696. [Google Scholar] [CrossRef] [PubMed]
  27. Opsenica, D.; Pocsfalvi, G.; Juranić, Z.; Tinant, B.; Declercq, J.-P.; Kyle, D.E.; Milhous, W.K.; Šolaja, B.A. Cholic acid derivatives as 1,2,4,5-tetraoxane carriers: Structure and antimalarial and antiproliferative activity. J. Med. Chem. 2000, 43, 3274–3282. [Google Scholar] [CrossRef]
  28. Moser, B.R. Review of cytotoxic cephalostatins and ritterazines: Isolation and synthesis. J. Nat. Prod. 2008, 71, 487–491. [Google Scholar] [CrossRef] [PubMed]
  29. Jurášek, M.; Černohorská, M.; Řehulka, J.; Spiwok, V.; Sulimenko, T.; Dráberová, E.; Darmostuk, M.; Gurská, S.; Frydrych, I.; Buriánová, R.; et al. Estradiol dimer inhibits tubulin polymerization and microtubule dynamics. J. Steroid Biochem. Mol. Biol. 2018, 183, 68–79. [Google Scholar] [CrossRef]
  30. Czajkowska-Szczykowska, D.; Rodríguez-Molina, B.; Magaña-Vergara, N.E.; Santillan, R.; Morzycki, J.W.; Garcia-Garibay, M.A. Macrocyclic Molecular Rotors with Bridged Steroidal Frameworks. J. Org. Chem. 2012, 77, 9970–9978. [Google Scholar] [CrossRef] [PubMed]
  31. Lino, R.; Kinbara, K.; Bryant, Z. Introduction: Molecular Rotors. Chem. Rev. 2020, 120, 1–4. [Google Scholar] [CrossRef]
  32. Tomkiel, A.M.; Kowalski, J.; Płoszyńska, J.; Siergiejczyk, L.; Łotowski, Z.; Sobkowiak, A.; Morzycki, J.W. Electrochemical synthesis of glycoconjugates from activated sterol derivatives. Steroids 2014, 82, 60–67. [Google Scholar] [CrossRef]
  33. Hunter, A.C.; Collins, C.; Dodd, H.T.; Dedi, C.; Koussoroplis, S.-J. Transformation of a series of saturated isomeric steroidal diols by Aspergillus tamarii KITA reveals a precise stereochemical requirement for entrance into the lactonization pathway. J. Steroid Biochem. Mol. Biol. 2010, 122, 352–358. [Google Scholar] [CrossRef]
  34. Shimizu, S.; Hagiwara, K.; Itoh, H.; Inoue, M. Unified total synthesis of five bufadienolides. Org. Lett. 2020, 22, 8652–8657. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, Y.; Haino, T.; Usui, S.; Fukazawa, Y. Shielding effect of ether C-O bond obtained from proton chemical shifts of 4-oxa-5α- and 4-oxa-5β-androstan-17-ones. Tetrahedron 1996, 52, 2325–2336. [Google Scholar] [CrossRef]
  36. Lloyd, L. Handbook of Industrial Catalysts; Springer: New York, NY, USA, 2011; pp. 181–182. [Google Scholar] [CrossRef]
  37. Kaneda, K. Cation-exchanged montmorillonites as solid acid catalysts for organic synthesis. Synlett 2007, 7, 0999–1015. [Google Scholar] [CrossRef]
  38. Bhongale, P.; Joshi, S.; Mali, N. A comprehensive review on catalytic O-alkylation of phenol and hydroquinone. Catal. Rev. 2023, 65, 455–500. [Google Scholar] [CrossRef]
  39. Lu, B.; Li, L.-J.; Li, T.-S.; Li, J.-L. Montmorillonite clay catalysis. Part 13. 1 Etherification of cholesterol catalysed by montmorillonite K-10. J. Chem. Res. Synop. 1998, 9, 604–605. [Google Scholar] [CrossRef]
  40. Li, T.; Li, H.; Guo, J.; Jin, T. Montmorillonite clay catalysis I: An efficient and convenient procedure for preparation of 5(6)/5′(6′)-unsaturated 3β-disteryl ethers. Synth. Commun. 1996, 26, 2497–2502. [Google Scholar] [CrossRef]
  41. Kowalski, J.; Morzycki, J.W.; Sobkowiak, A.; Wilczewska, A.Z. Unusual electrochemical oxidation of cholesterol. Steroids 2008, 73, 543–548. [Google Scholar] [CrossRef]
  42. Zmysłowski, A.; Sitkowski, J.; Bus, K.; Ofiara, K.; Szterk, A. Synthesis and search for 3β,3′β-disteryl ethers after high-temperature treatment of sterol-rich samples. Food Chem. 2020, 329, 127132. [Google Scholar] [CrossRef]
  43. Dhingra, N.; Bhardwaj, T.R.; Mehta, N.; Mukhopadhyay, T.; Kumar, A.; Kumar, M. Synthesis, antiproliferative activity, acute toxicity and assessment of the antiandrogenic activities of new androstane derivatives. Arch. Pharm. Res. 2011, 34, 1055–1063. [Google Scholar] [CrossRef]
  44. Shuping, W.; Zhiqin, J.; Heting, L.; Li, Y.; Daixun, Z. Sensitized photooxygenation of cholesterol and pseudocholesterol derivatives via singlet oxygen. Molecules 2001, 6, 52–60. [Google Scholar] [CrossRef]
  45. Wallis, P.J.; Gates, W.P.; Patti, A.F.; Scott, J.L.; Teoh, E. Assessing and improving the catalytic activity of K-10 montmorillonite. Green Chem. 2007, 9, 980–986. [Google Scholar] [CrossRef]
  46. Ebitani, K.; Kawabata, T.; Nagashima, K.; Mizugaki, T.; Kaneda, K. Simple and clean synthesis of 9,9-bis[4-(2-hydroxyethoxy)phenyl]fluorene from the aromatic alkylation of phenoxyethanol with fluoren-9-one catalysed by titanium cation-exchanged montmorillonite. Green Chem. 2000, 2, 157–160. [Google Scholar] [CrossRef]
  47. Joseph, T.; Shanbhag, G.V.; Halligudi, S.B. Copper(II) ion-exchanged montmorillonite as catalyst for the direct addition of N-H bond to CC triple bond. J. Mol. Catal. A Chem. 2005, 236, 139–144. [Google Scholar] [CrossRef]
Scheme 1. The reactions of androst-5-en-3β-ol-17-one tosylate with hydroquinone.
Scheme 1. The reactions of androst-5-en-3β-ol-17-one tosylate with hydroquinone.
Molecules 28 07068 sch001
Scheme 2. A presumable mechanism of 11b formation.
Scheme 2. A presumable mechanism of 11b formation.
Molecules 28 07068 sch002
Scheme 3. A presumable mechanism of 11a formation.
Scheme 3. A presumable mechanism of 11a formation.
Molecules 28 07068 sch003
Figure 3. Ball and stick presentation of optimized structures of dimers 11a and 11b.
Figure 3. Ball and stick presentation of optimized structures of dimers 11a and 11b.
Molecules 28 07068 g003
Figure 4. The diagnostic fragment of 1H NMR spectra of 11a, 11b, and 11c dimers.
Figure 4. The diagnostic fragment of 1H NMR spectra of 11a, 11b, and 11c dimers.
Molecules 28 07068 g004
Table 2. The results of androst-5-en-3β-ol-17-one 4a reaction with hydroquinone catalyzed with montmorillonite K10 under different conditions.
Table 2. The results of androst-5-en-3β-ol-17-one 4a reaction with hydroquinone catalyzed with montmorillonite K10 under different conditions.
Run No./
Reaction Conditions
Structure of ProductsConversion
Molecules 28 07068 i005Molecules 28 07068 i006Molecules 28 07068 i007Molecules 28 07068 i008Molecules 28 07068 i009
1 a: K10 (120 °C; 0.5 g)14%17%1%26%10%80%
2 a: K10 (120 °C; 1 g)19%11%1%26%9%80%
3 b: K10 (120 °C; 0.5 g)15%13%1%27%12%95%
4 b: K10 (200 °C; 0.5 g)36%8%<5%12%16%80%
5 b: K10 (280 °C; 0.5 g)28%12%<5%7%23%90%
6 b: K10 (400 °C; 0.5 g)19%17%1%22%6%80%
7 b: H+-K10 (280 °C; 0.5 g)26%16%<5%12%13%95%
8 c: H+-K10 (280 °C; 0.5 g)13%7%-57%1%99%
9 b: Ti4+-K10 (280 °C; 0.5 g)40%18%-22%1%95%
10 b: Cu2+-K10 (120 °C; 0.5 g)33%12%-9%12%95%
Reaction conditions: a DHEA (100 mg; 0.35 mmol), hydroquinone (19 mg; 0.175 mmol), CHCl3, reflux 4 h. b DHEA (100 mg; 0.35 mmol), hydroquinone (15.4 mg; 0.14 mmol), CHCl3, reflux 4 h. c DHEA (100 mg; 0.35 mmol), hydroquinone (15.4 mg; 0.14 mmol), CHCl3, 50 °C, 24 h.
Table 3. Reactions of tosylates 4b and 6b with hydroquinone catalyzed with montmorillonite K10 under different conditions.
Table 3. Reactions of tosylates 4b and 6b with hydroquinone catalyzed with montmorillonite K10 under different conditions.
Run No./
Reaction Conditions
Structure of ProductsConversion
Molecules 28 07068 i010Molecules 28 07068 i011Molecules 28 07068 i012Molecules 28 07068 i013
1 a: 4b, K10 (280 °C; 300 mg), r.t., 4 days22%19%1%15%80%
2 a: 4b, K10 (500 °C; 300 mg), r.t., 3 days21%14%1%23%100%
3 a: 4b, Ti4+-K10 (280 °C; 300 mg), r.t., 3 days28%19%-31%100%
4 b: 6b, K10 (280 °C; 50 mg), r.t., 5 days17%--32%70%
5 b: 6b, K10 (500 °C; 50 mg), r.t., 3 days10%--67%100%
6 b: 6b, Ti4+-K10 (500 °C; 50 mg), r.t., 3 days23%--42%80%
Reaction conditions: a DHEA tosylate (88.5 mg; 0.2 mmol), hydroquinone (8.4 mg; 0.076 mmol), CHCl3 (5 mL). b cholesteryl tosylate (100 mg; 0.2 mmol), hydroquinone (8.4 mg; 0.076 mmol), CHCl3 (5 mL).
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

Tomkiel, A.M.; Majewski, A.D.; Siergiejczyk, L.; Morzycki, J.W. Montmorillonite Catalyzed Synthesis of Novel Steroid Dimers. Molecules 2023, 28, 7068. https://doi.org/10.3390/molecules28207068

AMA Style

Tomkiel AM, Majewski AD, Siergiejczyk L, Morzycki JW. Montmorillonite Catalyzed Synthesis of Novel Steroid Dimers. Molecules. 2023; 28(20):7068. https://doi.org/10.3390/molecules28207068

Chicago/Turabian Style

Tomkiel, Aneta M., Adam D. Majewski, Leszek Siergiejczyk, and Jacek W. Morzycki. 2023. "Montmorillonite Catalyzed Synthesis of Novel Steroid Dimers" Molecules 28, no. 20: 7068. https://doi.org/10.3390/molecules28207068

Article Metrics

Back to TopTop