Next Article in Journal
Atomic and Molecular Layer Deposition as Surface Engineering Techniques for Emerging Alkali Metal Rechargeable Batteries
Next Article in Special Issue
Heterocyclization of Bis(2-chloroprop-2-en-1-yl)sulfide in Hydrazine Hydrate–KOH: Synthesis of Thiophene and Pyrrole Derivatives
Previous Article in Journal
Effect of Germination on the Avenanthramide Content of Oats and Their in Vitro Antisensitivity Activities
Previous Article in Special Issue
Scalable (Enantioselective) Syntheses of Novel 3-Methylated Analogs of Pazinaclone, (S)-PD172938 and Related Biologically Relevant Isoindolinones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iodine-Mediated Alkoxyselenylation of Alkenes and Dienes with Elemental Selenium

by
Evgeny O. Kurkutov
and
Bagrat A. Shainyan
*
A. E. Favorsky Irkutsk Institute of Chemistry, Siberian Branch, Russian Academy of Sciences, 1 Favorsky Str., 664033 Irkutsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(19), 6169; https://doi.org/10.3390/molecules27196169
Submission received: 22 August 2022 / Revised: 7 September 2022 / Accepted: 14 September 2022 / Published: 20 September 2022
(This article belongs to the Special Issue Feature Papers in Organic Chemistry)

Abstract

:
A one-pot synthesis of linear and cyclic β-alkoxyselenides is developed through the iodine-mediated three-component reaction of elemental selenium with alkenes (dienes) and alcohols. Selenylation of 1,5-hexadiene gives 2,5-di(methoxymethyl)tetrahydroselenophene and 2-methoxy-6-(methoxymethyl)tetrahydro-2H-selenopyran via the 5-exo-trig and 6-endo-trig cyclization. 1,7-Octadiene affords only linear 1:2 adduct with two terminal double bonds. 1,5-Cyclooctadiene results in one diastereomer of 2,6-dialkoxy-9-selenabicyclo [3.3.1]nonanes via 6-exo-trig cyclization. With 1,3-diethenyl-1,1,3,3-tetramethyldisiloxane, the first ring-substituted representative of a very rare class of heterocycles, 1,4,2,6-oxaselenadisilinanes, was obtained at a high yield.

1. Introduction

Organoselenium compounds are valuable reagents in organic and organoelement synthesis [1,2,3,4] and demonstrate diverse biological activities [5,6,7,8,9]. For example, 4-phenylselenyl-7-chloroquinoline (4-PSQ) exerts acute anti-inflammatory effects on apar with the widely used drug Meloxicam [10]. Some organoselenium compounds affect mood and exhibit high antidepressant activity [11,12]. They possess powerful radical scavenging and broad anti-HIV activity [13] and are capable of blocking the replication of SARS-CoV-2 [14,15,16]. One of the actively studied properties of organoselenium compounds is their glutathione peroxidase activity, [17,18] protecting living organisms against oxidative stress. All of this shows that organoselenium compounds are good candidates for the medicines of the future and promotes the search for new methods of their synthesis and investigation.
Nowadays, the iodine-mediated reactions of alkene functionalization are actively used in organic synthesis [19,20,21] since molecular iodine is non-toxic, inexpensive, safe, and easy to use in the handling of reagents. These reactions are employed in the synthesis of organoselenium compounds, for example, for selenylation of alkenes with organic diselenides via the formation of a Se–C bond, Scheme 1 [22,23,24].
The method is promising, although the scope of the substrates was limited to alkenes and did not include other unsaturated substrates. The drawbacks of this method are the impossibility to obtain the Se-containing heterocycles and the use, as a rule, of only diphenyl (or diaryl) diselenides. The main methods of the synthesis of symmetrical difunctionalized and cyclic selenides are based on the reactions of nucleophilic selenium reagents (selenolates and selenols) with electrophiles such as epoxides [25,26,27]. An alternative approach is the electrophilic addition of inorganic selenium halides to alkenes [28,29,30], leading to β-haloselenides (mustard gas analogs).
The drawbacks of such selenium-containing reagents as organic diselenides, organoselenyl halides, inorganic selenium halides, selenols, etc., are their high toxicity, low stability and inability to store for a long time. In contrast, elemental selenium is low-toxic and easy to store, transport and handle and thus has significant advantages as a reagent for the synthesis of organoselenium compounds [4,31].
Recently, we first reported the possibility of introducing a selenium atom into the organic chain by using molecular iodine in the reaction of hydroxyselenylation of alkenes with elemental selenium in an aqueous organic medium, resulting in the formation of bifunctionalized symmetrical β,β-dihydroxyselenides [32]. The reaction with cycloalkenes was shown to proceed stereoselectively as anti-addition, resulting exclusively in the trans,trans-diastereomers [32,33]. In distinction to that, in dry acetonitrile the iodine-mediated reaction of elemental selenium with alkenes proceeded as a Ritter-type reaction with an interception of the solvent (acetonitrile) and resulted in assembling the 1,3-selenazoline ring, Scheme 2 [34].
Here, we report a one-pot procedure for the three-component synthesis of cyclic and linear β,β-alkoxyselenides based on the iodine-mediated alkoxyselenylation of alkenes and dienes with elemental selenium and alcohols. The method takes advantage of using elemental selenium and alcohols as the reagents and allows the selective synthesis of selenium-containing compounds; possible structures are presented in Figure 1.
The novelty of the present study is not only in extending the reaction of hydroxyselenylation [32] to alkoxyselenylation but also in including, in addition to alkenes, the earlier not-studied dienes as the substrates, which can (and do) result in the formation of novel selenium heterocycles (vide infra).

2. Results and Discussion

The investigated alkenes 1ae and dienes 2ad are depicted in Figure 2.
The reaction of alkenes 1ae with selenium in alcohols proceeded under mild conditions in one preparative step to give β,β-dialkoxyselenides 3ag at good yields, taking into account the conversion (Scheme 3, Table 1). The conversion was estimated by subtracting the amount of unreacted (filtered off) selenium from that taken in the reaction. The reaction was highly regioselective; with unsymmetrical alkenes 1ad, only selenides 3 with the selenium atom at the terminal olefinic carbons were formed. Selenides 3a–f have two chiral centers which were formed as mixtures of diastereomers in the ratio of ca. 1:1, as proved by the presence of two 77Se NMR signals of equal intensity and the doubled sets of signals in the 13C NMR spectra. With cyclohexene, only two diastereomers were formed in the ratio of 1:1, although molecule 3g has four asymmetric centers allowing four diastereomers to be formed. Based on our previous results on the exclusive formation of the trans-products of hydroselenylation of cycloalkenes, we assumed stereoselective anti-addition and formation of product 3c as the d,l and meso diastereomers with trans-arrangement of the methoxy groups and the selenium atom. The yields of compounds 3 are given in Figure 3. 1H NMR spectroscopy revealed the formation of trace amounts of by-products of iodomethoxylation 4, which can easily be separated by eluting with CCl4. Independently synthesized compounds 4 do not react with selenium under the conditions of the reaction; so, they are not intermediates on the way to the target products 3.
Optimization of the reaction conditions was performed by varying the solvent, the iodine source and its amount, temperature and reaction time for the reaction of heptene-1 with methanol (Table 1). With 1 or 1.5 equivalents of iodine, the conversion is incomplete (entries 1–4) but increases with time. With two equivalents of iodine, the conversion is complete in 40 h (entry 5) without an increase in the number of by-products 4. When using five equivalents of methanol (2.5 molar excess), selenide 3a is formed selectively at a 79% yield (entry 10), but the conversion is rather low. A high conversion of 86% was achieved when using methanol as the solvent (entry 4). The use of catalytic amounts of iodine in combination with DMSO is known [21,22] when iodine is recycled by oxidation of the formed hydrogen iodide with DMSO. However, in our case, the use of DMSO shows low conversion and selectivity affording a mixture of adducts (entry 9). The increase in the temperature to 50 °C lowers the yield (entry 7). Since molecular iodine acts as an oxidant, we tried to replace it with N-iodosuccinimide; indeed, the rate of the reaction increased, but simultaneously increased the number of by-products 4, up to 30% (NIS, entry 8). No reaction occurs in the absence of iodine or its substitute. To summarize, the most favorable condition is the use of a two-fold excess of iodine in methanol at room temperature. Alkoxyselenylation of alkenes 1ae was performed under the conditions providing the highest yield of the target selenides (entries 2 and 5) in spite of incomplete conversion, which can be increased by prolonging the reaction time (entry 3).
With the results for alkenes 1ae in mind, one can expect the formation of various products of methoxyselenylation of dienes 2ad via addition to one or both double bonds of the diene, or the reaction with two diene molecules, or even the products of cyclization or oligomerization. The unpredictability of the reaction is exacerbated by the fact that it is very poorly studied, and the nature of active intermediates is not clear, taking into account that elemental selenium exists as a mixture of single chains, Se8 cyclic molecules, or polymer nanoparticles [35] and can form a mixture of iodides of general formula SenI2. However, to our surprise and delight, the reaction of methoxyselenylation of 1,5-hexadiene 2a led selectively to the isomeric products of cyclization, 2,5-di(methoxymethyl)tetrahydroselenophene 5 and its six-membered isomer, 5-methoxy-2-(methoxymethyl)tetrahydro-2H-selenopyran 6 in the ratio of 1:2 and an 86% total yield, Scheme 4.
Compounds 5 and 6 are formed as two diastereomers in the ratio of ca. 1:1, as proved by the presence of doubled signals in the 13C and 77Se NMR spectra of the products. The products could not be completely separated into pure diastereomers, but we succeeded in the isolation of the fractions of compound 6 enriched with diastereomer 1 (80:20) or diastereomer 2 (35:65), which allowed the assignment of the signals to specific diastereomers (see spectra S26–S32 in the Supplementary Materials).
Note that no linear adducts were formed, even with a large excess of diene 2a. Remarkably, while the formation of products 3 in Scheme 1 obeys the Markovnikov rule (addition of MeO nucleophile to the internal olefinic carbon), the reaction with 1,5-hexadiene 2a proceeded as an anti-Markovnikov addition. Apparently, this is due to the effect of the second double bond, which can interact with the selenium atom in the intermediate seleniranium cation with a subsequent opening by methanol via the 5-exo-trig or 6-endo-trig mode, both of which are favorable according to Baldwin’s rules [36], Scheme 5.
In contrast, with a higher homologue of diene 2a and 1,7-octadiene 2b, the reaction proceeded with the formation of 1:2 linear adduct, 8,8′-selenobis(7-methoxyoct-1-ene) 7, with two terminal double bonds being retained, and 2:3 linear adduct, 3,16-dihex-5-en-1-yl-7,12-dimethoxy-2,17-dioxa-5,14-diselenaoctadecane 8. As with alkenes 1ad, the addition follows the Markovnikov rule, probably due to a more remote second double bond than in the case of diene 2a, Scheme 6.
No cyclic products were detected in the reaction mixture. With a large excess of the diene (Se:2b = 1:3), selenide 7 is practically the only product of the reaction. For the equimolar ratio, it still is the main product, but a fraction enriched with compound 8 was also isolated via column chromatography. The structure of 8 was validated through 1H NMR spectroscopy and mass spectrometry. The relative intensity of the methylene proton signals with respect to olefinic protons is notably larger than for monoselenide 7, and the mass spectrum contains a peak with m/z 614, corresponding to the molecular ion [C28H54O4Se2]+.
For comparison, the reaction with 1,5-cyclooctadiene was investigated. The reaction proceeds regioselectively to give only one diastereomer of 2,6-dialkoxy-9-selenabicyclo [3.3.1]nonanes 9ac, apparently via alkoxyselenylation at one double bond followed by the 6-exo-trig cyclization [36]. Based on the common mechanism of anti-addition via the ring opening of the intermediate seleniranium cation and the earlier proved trans structure of the products of hydroxyselenylation of cyclohexene [32,33], it is reasonable to assume that the selenium atom in 9 is in the trans position to the alkoxy groups, Scheme 7.
Under the same conditions, the iodine-mediated reaction of elemental selenium with 1,3-diethenyl-1,1,3,3-tetramethyldisiloxane and methanol affords 3,5-bis(methoxymethyl)-2,2,6,6-tetramethyl-1,4,2,6-oxaselenadisilinane 10, Scheme 8.
The structure of compound 10 and, hence, the anti-Markovnikov regioselectivity of addition is proved through the presence of SeCH signals at 22–25 ppm in the 13C j-mod NMR spectrum (Figure S47) as well as the value of the 1JC-Se direct constant of ~58 Hz on the satellites of this signal (Figure S48). Heterocycle 10 is representative of a very rare class of 1,4,2,6-oxaselenadisilinanes and is the first one with substituents in the ring. It is formed of a 2:1 mixture of diastereomers generated by the methanol attack on the re or si faces of the double bond during the 6-exo-trig cyclization. Judging from theoretical calculations, the prevalent diastereomer has both substituents equatorial and is more stable than the ax-eq diastereomer by 1.24 kcal/mol in total energy ΔE and by 1.52 kcal/mol in free energy ΔG. The formula of compound 10 is proved by the presence of a fragment ion with m/z 297 corresponding to a fragment ion [M + H − MeOH]+ formed by protonation and elimination of methanol from 10. This ion can exist as two diastereomers, each as a carbocation or seleniranium cation. Interestingly, while seleniranium cations 10a+ and 10b+ are the minima on the MP2 potential energy surface, the six-membered carbocations undergo ring expansion upon geometry optimization to 10c or 10d. The relative total and free energies of all cations are given in Figure 4 relative to the most stable seven-membered cyclic cation 10d [37].
This result raises the question of the manner and relative ability of different substituents to stabilize the adjacent carbocationic center; however, it is beyond the scope of our paper.
Unlike chlorides or bromides, selenium iodides SenI2 were not isolated as individuals, but registered in a solution via 77Se NMR spectroscopy, the major constituent being Se2I2 [38]. Based on these data, we assume that the mechanism of hydroxyselenylation is similar to that described in [32], which includes the electrophilic addition of selenium iodides, successive Se–Se bonds rupture via their iodination and, finally, the formation of selenides. This is proved through mass spectrometric detection of di- and triselenides 11 and 12 in the reaction mixture with cyclohexene, as shown in Figure 5.
Therefore, a variety of transformations occur in the system elemental selenium–iodine–alcohol–alkene(diene), opening the way to different selenium-containing linear and heterocyclic compounds, which is illustrated in Figure 6.

3. Experimental Section

3.1. General Information

Commercially available chemicals and reagents were used without further purification. 1H (400.1 MHz), 13C (100.6 MHz) and 77Se (76.2 MHz) NMR spectra ((Bruker BioSpin GmbH, Rheinstetten, Germany) were recorded on a Bruker DPX-400 spectrometer for 5–10% solutions in CDCl3. 1H and 13C chemical shifts (δ) are reported in parts per million (ppm) relative to the residual solvent peak of CDCl3 (δ = 7.27 (1H) and 77.10 (13C) ppm, respectively). 77Se NMR chemical shifts (δ) are given relative to Me2Se. High-resolution mass spectra (HRMS) were measured at 70 eV on an Agilent 1200 HPLC chromatograph (Agilent, Santa Clara, CA, USA) with an Agilent 6210 mass spectrometer (HR-TOF-MS, ESI + ionization in acetonitrile, Agilent, Santa Clara, CA, USA). Mass spectra were obtained on a Shimadzu GCMSQP5050A instrument (Shimadzu Corp., Kyoto, Japan). Elemental compositions were determined with a Thermo Scientific Flash 2000 CHNS analyzer (Thermo Fisher Scientific Inc., Milan, Italy).

3.2. General Procedure for the Preparation of Selenides 3ag and 7

Alkene 1ae (or diene 2b) (3.5 mmol) and alcohol (1 mL) were added to a mixture of fine-ground elemental Se (0.079 g, 1 mmol) and I2 (0.508 g, 2 mmol for 3ac or 1 mmol, 0.254 g for 3dg and 7), and the mixture was stirred for 40 h at room temperature. Then, a saturated solution of Na2S2O3 (2 mL) was added, stirred for 10 min, the mixture was diluted with water (10 mL), and Et2O (10 mL) was added. The unreacted selenium was filtered off, the water layer extracted with Et2O (3 × 10 mL), and the extract was washed with H2O (2 × 5 mL), dried with CaCl2, and filtered. The solvent was removed in a vacuum to give the crude product, which was purified on silica gel via successive eluting with CCl4 and CHCl3 to give the corresponding selenide 3ag and 7.

3.3. General Procedure for the Preparation of Selenides 5, 6, 9ac, and 10

Diene 2a, 2c or 2d (1.5 mmol) and alcohol (1 mL) were added to a mixture of fine-ground elemental Se (0.079 g, 1 mmol) and I2 (0.254 g, 1 mmol), and the mixture was stirred for 40 h at room temperature. Then, a saturated solution of Na2S2O3 (2 mL) was added, stirred for 10 min, the mixture was diluted with water (5 mL), and Et2O (EtOAc for 5,6) (10 mL) was added. The unreacted selenium was filtered off, the water layer extracted with Et2O (EtOAc for 5,6) (3 × 10 mL), and the extract was washed with H2O (2 × 5 mL), dried with CaCl2, and filtered. The solvent was removed in a vacuum to give the crude product, which was purified on silica gel via successive eluting with CCl4 and CHCl3 to give products 9, 10. Isomers 5 and 6 were obtained at an 86% total yield (selenium conversion 79%) as mixtures of regioisomers in the ratio of 2:1.

3.4. Characterization Data of Products 3ag, 5, 6, 9ac and 10

2-Methoxy-1-[(2-methoxyheptyl)selanyl]heptane (3a):
A yellowish oil, yield 87% (294 mg). 1H NMR (CDCl3, 400 MHz): δ 3.37 (s, 6H), 3.32−3.29 (m, 2H), 2.77−2.64 (m, 4H), 1.59−1.53 (m, 4H), 1.37−1.27 (m, 12H), 0.89−0.86 (m, 6H). 13C NMR (CDCl3, 100 MHz): δ 81.45, 56.96, 33.95, 32.00, 28.69, 28.61, 25.07, 22.70, 14.12. 77Se NMR (CDCl3, 76 MHz): δ 103.46. HRMS, found, 339.1640; calcd for C16H34O2Se [M + H]+, 339.1797.
2-Ethoxy-1-[(2-ethoxyheptyl)selanyl]heptane (3b).
A yellowish oil, yield 66% (228 mg) (selenium conversion 95%). 1H NMR (CDCl3, 400 MHz): δ 3.57−3.51 (m, 2H), 3.48−3.40 (m, 2H), 3.40−3.35 (m, 2H), 2.75−2.60 (m, 4H), 1.57−1.52 (m, 4H), 1.48−1.38(m, 12H), 1.36 (t, 6H), 0.87−0.85(m, 6H). 13C NMR (CDCl3, 100 MHz): δ 79.70, 79.66, 64.65, 34.37, 31.90, 29.21, 29.13, 25.13, 22.61, 15.59, 14.02. 77Se NMR (CDCl3, 76 MHz): δ 101.76, 100.67. Anal., found: C, 59.52; H, 10.16; Se, 22.14. Calcd for C18H38O2Se: C, 59.16; H, 10.48; Se, 21.61.
2-(Prop-2-yloxy)-1-{[2-(prop-2-yloxy)heptyl]selanyl}heptane (3c).
A yellowish oil, yield 58% (211 mg) (selenium conversion 92%). 1H NMR (CDCl3, 400 MHz): δ 3.69−3.63 (m, 2H), 3.48−3.42 (m, 2H), 2.75−2.61 (m, 4H), 1.64−1.57 (m, 2H), 1.33−1.28(m, 14H), 1.18−1.12 (m, 12H), 0.90−0.87(m, 6H). 13C NMR (CDCl3, 100 MHz): δ 77.31, 77.24, 70.07, 34.96, 32.01, 30.30, 30.16, 25.30, 23.21, 22.71, 22.66, 14.12. 77Se NMR (CDCl3, 76 MHz): δ 100.70, 98.29. Anal., found: C, 61.11; H, 10.48; Se, 20.10; calcd for C20H42O2Se: C, 61.04; H, 10.76; Se, 20.07.
2-Methoxy-1-[(2-methoxyhexyl)selanyl]hexane (3d):
A yellowish oil, yield 85% (213 mg) (selenium conversion 81%).1H NMR (CDCl3, 400 MHz): 3.35 (s, 6H), 3.33−3.29 (m, 2H), 2.78−2.65 (m, 4H), 1.61−1.53 (m, 4H), 1.37−1.26 (m, 8H), 0.89 (t, 6H). 13C NMR (CDCl3, 100 MHz): δ 81.38, 56.89, 33.63, 28.62, 28.54, 27.52, 22.79, 14.05. 77Se NMR (CDCl3, 76 MHz): δ 97.97, 97.53. Anal. found: C, 53.99; H, 9.48; Se, 25.15. Calcd for C14H30O2Se: C, 54.36; H, 9.77; Se, 25.52.
1,1′-[Selanylbis(1-methoxyethane-2,1-diyl)]dicyclohexane (3e):
A yellowish oil, yield 92% (233 mg) (selenium conversion 70%). 1H NMR (CDCl3, 400 MHz): 3.40 (s, 6H), 3.14−3.10 (m, 2H), 2.77−2.74 (m, 4H), 1.81−1.61 (m, 12H), 1.27−1.00 (m, 10H). 13C NMR (CDCl3, 100 MHz): δ 86.06, 86.04, 58.24,29.06, 28.23, 26.61, 26.34, 26.30, 26.27. 77Se NMR (CDCl3, 76 MHz): δ 102.38, 101.77. HRMS, found, 363.1645; calcd for C18H34O2Se [M + H]+, 363.1797.
1,1′-[Selanylbis(1-methoxyethane-2,1-diyl)]dibenzene (3f):
A yellowish oil, yield 90% (201 mg) (selenium conversion 64%). 1H NMR (CDCl3, 400 MHz): δ 7.33−7.35 (m, 10H), 4.29−4.23 (m, 2H), 3.22−4.20 (m, 6H), 2.93−2.86 (m, 2H), 2.69−2.64 (m, 2H). 13C NMR (CDCl3, 100 MHz): δ 141.35, 128.53, 128.01, 126.80, 126.74, 84.55, 56.96, 56.93, 32.12, 31.96. 77Se NMR (CDCl3, 76 MHz): δ 146.62, 146.12. Anal., found: C, 61.43; H, 5.87, Se, 22.92; calcd for C18H22O2Se: C, 61.89; H, 6.35, Se, 22.60.
1,1′-Selanylbis(2-methoxycyclohexane) 3g.
A yellowish oil, yield 68% yield (201 mg) (selenium conversion 79%). 1H NMR (CDCl3, 400 MHz): δ 3.37 (s, 6H), 3.22−3.09 (m, 4H), 2.17−2.01 (m, 4H), 1.71−1.50 (m, 6H), 1.34−1.26 (m, 6H). 13C NMR (CDCl3, 100 MHz): δ 83.69, 83.50, 56.58, 56.45, 31.98, 31.75, 30.01, 25.60, 25.40, 23.30, 23.15. 77Se NMR (CDCl3, 76 MHz): δ 327.32. HRMS, found, 307.1178; calcd for C14H26O2Se [M + H]+, 307.1171.
2,5-Bis(methoxymethyl)tetrahydroselenophene 5:
1H NMR (CDCl3, 400 MHz): δ 3.79−3.75 (m, 2H), 3.62−3.57 (m, 2H), 3.37 (s, 6H), 2.18−2.06 (m, 2H), 1.99−1.72 (m, 2H). 13C NMR (CDCl3, 100 MHz): δ 77.34, 58.78, 43.14, 43.00, 34.04, 33.67. 77Se NMR (CDCl3, 76 MHz): δ 312.40, 311.12. GS-MS: m/z (rel. int., %) 224 [M](40), 192 (21), 160 (30), 147 (100).
5-Methoxy-2-(methoxymethyl)tetrahydro-2H-selenopyran (6), diastereomer 1:
1H NMR (CDCl3, 400 MHz): δ 3.50−3.36 (m, 3H), 3.37 (s, 3H), 3.34 (s, 3H), 3.21−3.14 (m, 1H), 2.83−2.79 (m, 1H), 2.61−2.56 (m, 1H), 2.40−2.34 (m, 1H), 2.11−2.06 (m, 1H), 1.77 −1.66 (m, 1H), 1.32−1.22 (m, 1H). 13C NMR (CDCl3, 100 MHz): δ 78.88, 76.06, 58.89, 55.94, 35.79, 32.64, 32.52, 22.44. 77Se NMR (CDCl3, 76 MHz): δ 201.52. GS-MS: m/z (rel. int., %) 224 [M](27), 192 (5), 179 (8), 147 (100).
5-Methoxy-2-(methoxymethyl)tetrahydro-2H-selenopyran (6), diastereomer 2:
1H NMR (CDCl3, 400 MHz): δ 3.62−3.58 (m, 3H), 3.38 (s, 3H), 3.37 (s, 3H), 3.09−3.02 (m, 1H), 2.81−2.76 (m, 1H), 2.71−2.66 (m, 1H), 2.32−2.24 (m, 1H), 2.03−1.95 (m, 1H), 1.75−1.69 (m, 2H). 13C NMR (CDCl3, 100 MHz): δ 76.18, 75.01, 58.94, 55.84, 32.14, 28.52, 28.14, 20.87. 77Se NMR (CDCl3, 76 MHz): δ 165.06.
7-Methoxy-8-[(2-methoxyoct-7-en-1-yl)selanyl]oct-1-ene (7):
A light yellow oil, yield 69% yield (158 mg) (selenium conversion 63%). 1H NMR (CDCl3, 400 MHz): δ 5.86−5.76 (m, 2H), 5.03−4.93 (m, 4H), 3.37−3.32 (m, 8H), 2.79−2.66 (m, 4H), 2.09−2.04 (m, 4H), 1.61−1.58 (m, 4H), 1.44−1.38 (m, 4H). 13C NMR (CDCl3, 100 MHz): δ 138.92, 114.45, 81.33, 56.97, 33.79, 33.75, 29.03, 28.62, 28.55, 24.85, 23.15 77Se NMR (CDCl3, 76 MHz): δ 97.94, 97,64. GS-MS: m/z (rel. int., %) 362 [M](1), 221 (14), 189 (21), 95 (100). Anal., found: C, 59.82; H, 9.48, Se, 21.85; calcd for C18H34O2Se: C, 60.19; H, 9.55, Se, 22.30.
2,6-Dimethoxy-9-selenabicyclo [3.3.1]nonane (9a):
Colorless oil, yield 96% (141 mg) (selenium conversion 59%). 1H NMR (CDCl3, 400 MHz): δ 3.89−3.83 (m, 2H), 3.33 (s, 6H), 3.00−3.99 (m, 2H), 2.62−2.57 (m, 2H), 2.20−2.11 (m, 2H), 2.03−1.96 (m, 2H), 1.81−1.70 (m, 2H). 13C NMR (CDCl3, 100 MHz): δ 81.00, 55.82, 28.85, 28.02, 27.85. 77Se NMR (CDCl3, 76MHz): δ 287.69. GS-MS: m/z (rel. int., %) 250 [M](37), 218 (10), 179 (24), 71 (100). Anal., found: C, 48.60; H, 7.25, Se, 31.87; calcd for C10H18O2Se: C, 48.20; H, 7.28, Se, 31.68.
2,6-Dimethoxy-9-selenabicyclo [3.3.1]nonane (9b):
Colorless oil, yield 91% (189 mg) (selenium conversion 75%). 1H NMR (CDCl3, 400 MHz): δ 3.97−3.92 (m, 2H), 3.60−3.52 (m, 2H), 3.46−3.39 (m, 2H), 2.95−2.94 (m, 2H), 2.65−2.59 (m, 2H), 2.19−2.09 (m, 2H), 1.98−1.91 (m, 2H), 1.17−1.13 (m, 6H). 13C NMR (CDCl3, 100 MHz): δ 79.34, 63.39, 29.31, 28.03, 15.74. 77Se NMR (CDCl3, 76MHz): δ 287.80. Anal., found: C, 51.77; H, 7.89, Se, 28.54; calcd for C12H22O2Se: C, 51.98; H, 8.00, Se, 28.48.
2,6-Bis(propan-2-yloxy)-9-selenabicyclo [3.3.1]nonane (9c):
Colorless oil, yield 89% (196 mg) (selenium conversion 72%). 1H NMR (CDCl3, 400 MHz): δ 4.04−3.99 (m, 2H), 3.73−3.66 (m, 2H), 3.88−3.85 (m, 2H), 2.71−2.66 (m, 2H), 2.20 −2.11 (m, 2H), 1.90−1.82 (m, 2H), 1.15−1.11 (m, 12H). 13C NMR (CDCl3, 100 MHz): δ 77.18, 68.92, 30.04, 29.64, 28.33, 23.57, 22.46. 77Se NMR (CDCl3, 76MHz): δ 291.20. Anal., found: C, 55.06; H, 8.58, Se, 25.90; calcd for C14H26O2Se: C, 55.07; H, 8.58, Se, 25.86.
3,5-Dimethoxy-2,2,6,6-tetramethyl-1,4,2,6-oxaselenadisilinane (10):
A light yellow oil, yield 78% (140 mg) (selenium conversion 60%). 1H NMR (CDCl3, 400 MHz): δ 3.71−3.52 (m, 4H), 3.34 (s, 3H), 3.33 (s, 3H), 2.39−2.33 (m, 2H), 0.21−0.19 (m, 12H). 13C NMR (CDCl3, 100 MHz): δ 74.15, 73.31, 58.54, 25.30, 22.23, 0.75, 0.62, −1.73, −2.29. 77Se NMR (CDCl3, 76 MHz): δ 160.43, 134.88. 29Si NMR (CDCl3, 79.5 MHz): δ 5.71, 5.46. HRMS, found, 297.0240; calcd for C9H21O2SeSi2 [M + H − MeOH]+, 297.0240. Anal., found: C, 36.77; H, 7.22, Se, 31.87; calcd for C10H24O3SeSi2: C, 36.68; H, 7.39.

4. Conclusions

To summarize, a one-pot procedure for the synthesis of linear and cyclic alkoxy-functionalized organoselenium compounds is proposed, based on the iodine-mediated reaction of elemental selenium with alkenes and dienes in alcohols.
The method has significant advantages over the existing procedures, such as the use of simple and low-toxic reagents, mild conditions, and a diversity of synthesized products.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27196169/s1. Copies of the 1H NMR, 13C NMR, and 77Se spectra for compounds 3ag, 5, 6, 9ac and 10 can be found in Supplementary Materials.

Author Contributions

Conceptualization, E.O.K.; methodology, E.O.K.; software, B.A.S.; validation, E.O.K. and B.A.S.; formal analysis, E.O.K.; investigation, E.O.K.; resources, E.O.K.; data curation, E.O.K.; writing—original draft preparation, E.O.K.; writing—review and editing, B.A.S.; visualization, B.A.S.; supervision, B.A.S.; project administration, B.A.S.; funding acquisition, B.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was performed using the equipment of the Baikal Analytical Center for Collective Use of SB RAS.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. Wirth, T. (Ed.) Organoselenium Chemistry: Synthesis and Reactions; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  2. Wirth, T. (Ed.) Organoselenium Chemistry. Encyclopedia of Inorganic and Bioinorganic Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2018. [Google Scholar]
  3. Makhal, P.N.; Nandi, A.; Kaki, V.R. Insights into the recent synthetic advances of organoselenium compounds. Chem. Select 2021, 6, 663–679. [Google Scholar] [CrossRef]
  4. Azeredo, J.B.; Penteado, F.; Nascimento, V.; Sancineto, L.; Braga, A.L.; Lenardao, E.J.; Santi, C. Green Is the Color: An Update on Ecofriendly Aspects of Organoselenium Chemistry. Molecules 2022, 27, 1597. [Google Scholar] [CrossRef]
  5. Mugesh, G.; du Mont, W.W.; Sies, H. Chemistry of biologically important synthetic organoselenium compoundds. Chem. Rev. 2001, 101, 2125–2180. [Google Scholar] [CrossRef] [PubMed]
  6. Nogueira, C.W.; Zeni, G.; Rocha, J.B.T. Organoselenium and organotellurium compounds: Toxicology and pharmacology. Chem. Rev. 2004, 104, 6255–6286. [Google Scholar] [CrossRef] [PubMed]
  7. Jain, V.K.; Priyadarsini, K.I. (Eds.) Organoselenium Compounds in Biology and Medicine: Synthesis, Biological and Therapeutic Treatments; RSC: London, UK, 2017. [Google Scholar]
  8. Nogueira, C.W.; Barbosa, N.V.; Rocha, J.B.T. Toxicology and pharmacology of synthetic organoselenium compounds: An update. Arch. Toxicol. 2021, 95, 1179–1226. [Google Scholar] [CrossRef] [PubMed]
  9. Ruberte, A.C.; Sanmartin, C.; Aydillo, C.; Sharma, A.K.; Plano, D. Development and therapeutic potential of selenazo compounds. J. Med. Chem. 2019, 63, 1473–1489. [Google Scholar] [CrossRef] [PubMed]
  10. Pinz, M.; Reis, A.S.; Duarte, V.; da Rocha, M.J.; Goldani, B.S.; Alves, D.; Savegnago, L.; Luchese, C.; Wilhelm, E.A. 4-Phenylselenyl-7-chloroquinoline, a new quinoline derivative containing selenium, has potential antinociceptive and anti-inflammatory actions. Eur. J. Pharmacol. 2016, 780, 122–128. [Google Scholar] [CrossRef] [PubMed]
  11. Casaril, A.M.; de Lourenço, D.A.; Domingues, M.; Smaniotto, T.Â.; Birmann, P.T.; Vieira, B.; Sonego, M.S.; Seixas, F.K.; Collares, T.; Lenardão, E.J.; et al. Anhedonic- and anxiogenic-like behaviors and neurochemical alterations are abolished by a single administration of a selenium-containing compound in chronically stressed mice. Compr. Psychoneuroendocrinol. 2021, 6, 100054. [Google Scholar] [CrossRef]
  12. Casaril, A.M.; Domingues, M.; Bampi, S.R.; de Lourenço, D.A.; Smaniotto, T.Â.; Segatto, N.; Vieira, B.; Seixas, F.K.; Collares, T.; Lenardão, E.J.; et al. The antioxidant and immunomodulatory compound 3-[(4-chlorophenyl)selanyl]-1-methyl-1H-indole attenuates depression-like behavior and cognitive impairment developed in a mouse model of breast tumor. Brain Behav. Immun. 2020, 84, 229–241. [Google Scholar] [CrossRef]
  13. Sancineto, L.; Mariotti, A.; Bagnoli, L.; Marini, F.; Desantis, J.; Iraci, N.; Santi, C.; Pannecouque, C.; Tabarrini, O. Design and Synthesis of DiselenoBisBenzamides (DISeBAs) as Nucleocapsid Protein 7 (NCp7) Inhibitors with anti-HIV Activity. J. Med. Chem. 2015, 58, 9601–9614. [Google Scholar] [CrossRef]
  14. Jin, Z.; Du, X.; Yechun, X.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li, X.; Zhang, L.; Peng, C.; et al. Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors. Nature 2020, 582, 289–293. [Google Scholar] [CrossRef]
  15. Zhang, J.; Saad, R.; Taylor, E.W.; Rayman, M.P. Selenium and selenoproteins in viral infection with potential relevance to COVID-19. Redox Boil. 2020, 37, 101715. [Google Scholar] [CrossRef]
  16. Mangiavacchi, F.; Botwina, P.; Menichetti, E.; Bagnoli, L.; Rosati, O.; Marini, F.; Fonseca, S.F.; Abenante, L.; Alves, D.; Dabrowska, A.; et al. Seleno-Functionalization of Quercetin Improves the Non-Covalent Inhibition of Mpro and Its Antiviral Activity in Cells against SARS-CoV-2. Int. J. Mol. Sci. 2021, 22, 7048. [Google Scholar] [CrossRef]
  17. Santi, C.; Marini, F.; Lenardão, E.J. Synthetic Advances on Bioactive Selenium Compounds. Looking Beyond the Traditional Idea of Glutathione Peroxidase Mimics as Antioxidants. In Organoselenium Compounds in Biology and Medicine: Synthesis, Biological and Therapeutic Treatments; Chapter 2; Royal Society of Chemistry: London, UK, 2017; pp. 35–76. [Google Scholar]
  18. Sands, K.N.; Tuck, T.A.; Back, T.G. Cyclic seleninate esters, spirodioxyselenuranes and related compounds: New classes of biological antioxidants that emulate glutathione peroxidase. Chem. Eur. J. 2018, 24, 9714–9728. [Google Scholar] [CrossRef]
  19. Zhang, X.J.; Wang, Z.; Zhang, H.; Gao, J.J.; Yang, K.R.; Fan, W.Y.; Wu, R.X.; Feng, M.L.; Zhu, W.; Zhu, Y.P. Iodine-Mediated Domino Cyclization for One-Pot Synthesis of Indolizine-Fused Chromones via Metal-Free sp3 C–H Functionalization. J. Org. Chem. 2022, 87, 835–845. [Google Scholar] [CrossRef]
  20. Kitamura, T.; Komoto, R.; Oyamada, J.; Higashi, M.; Kishikawa, Y. Iodine-Mediated Fluorination of Alkenes with an HF Reagent: Regioselective Synthesis of 2-Fluoroalkyl Iodides. J. Org. Chem. 2021, 86, 18300–18303. [Google Scholar] [CrossRef]
  21. Cui, H.; Liu, X.; Wei, W.; Yang, D.; He, C.; Zhang, T.; Wang, H. Molecular iodine-mediated difunctionalization of alkenes with nitriles and thiols leading to β-acetamido sulfides. J. Org. Chem. 2016, 81, 2252–2260. [Google Scholar] [CrossRef]
  22. Vieira, A.A.; Azeredo, J.B.; Godoi, M.; Santi, C.; da Silva Junior, E.N.; Braga, A.L. Catalytic chalcogenylation under greener conditions: A solvent-free sulfur- and seleno-functionalization of olefins via I2/DMSO oxidant system. J. Org. Chem. 2015, 80, 2120–2127. [Google Scholar] [CrossRef]
  23. Liu, Y.; Li, C.; Mu, S.; Li, Y.; Feng, R.; Sun, K. Eco-Friendly Selenoamidation of Alkenes with Sulfamides and Organoseleniums: Rapid access to β-Amidoselenides. Asian J. Org. Chem. 2018, 7, 720–723. [Google Scholar] [CrossRef]
  24. Wang, X.L.; Li, H.J.; Yan, J. Iodine-mediated regioselective hydroxyselenenylation of alkenes: Facile access to β-hydroxy selenides. Chin. Chem. Lett. 2018, 29, 479–481. [Google Scholar] [CrossRef]
  25. Tanini, D.; Degl’Innocenti, A.; Capperucci, A. Bis (trimethylsilyl) selenide in the Selective Synthesis of β-Hydroxy, β-Mercapto, and β-Amino Diorganyl Diselenides and Selenides Through Ring Opening of Strained Heterocycles. Eur. J. Org. Chem. 2015, 2015, 357–369. [Google Scholar] [CrossRef]
  26. Leng, T.; Wu, G.; Zhou, Y.B.; Gao, W.; Ding, J.; Huang, X.; Liu, M.; Wu, H. Silver-Catalyzed One-Pot Three-Component Selective Synthesis of β-Hydroxy Selenides. Adv. Synth. Catal. 2018, 360, 4336–4340. [Google Scholar] [CrossRef]
  27. Arai, K.; Kumakura, F.; Takahira, M.; Sekiyama, N.; Kuroda, N.; Suzuki, T.; Iwaoka, M. Effects of ring size and polar functional groups on the glutathione peroxidase-like antioxidant activity of water-soluble cyclic selenides. J. Org. Chem. 2015, 80, 5633–5642. [Google Scholar] [CrossRef] [PubMed]
  28. Potapov, V.A.; Kurkutov, E.O.; Musalov, M.V.; Amosova, S.V. Reactions of selenium dichloride and dibromide with divinyl sulfone: Synthesis of novel four-and five-membered selenium heterocycles. Tetrahedron Lett. 2010, 51, 5258–5261. [Google Scholar] [CrossRef]
  29. Potapov, V.A.; Kurkutov, E.O.; Musalov, M.V.; Amosova, S.V. Synthesis of bis (2-haloethyl) selenides by reaction of selenium dihalides with ethylene. Russ. J. Org. Chem. 2014, 50, 291–292. [Google Scholar] [CrossRef]
  30. Lautenschlaeger, F.K. Reaction of selenium monochloride with diolefins. J. Org. Chem. 1969, 34, 4002–4006. [Google Scholar] [CrossRef]
  31. Ma, Y.-T.; Liu, M.-C.; Zhou, Y.-B.; Wu, H.-Y. Synthesis of organoselenium compounds with elemental selenium. Adv. Synth. Catal. 2021, 363, 5386–5406. [Google Scholar] [CrossRef]
  32. Kurkutov, E.O.; Borodina, T.N. The iodine-assisted hydroxyselenenylation of alkenes with elemental selenium. Mendeleev Commun. 2020, 30, 760–761. [Google Scholar] [CrossRef]
  33. Shainyan, B.A.; Kurkutov, E.O.; Kleinpeter, E.A. low-temperature dynamic 1H, 13C, and 77Se NMR study of 2, 2′-selenodicyclohexanol. Magn. Reson. Chem. 2022, 60, 165–171. [Google Scholar] [CrossRef]
  34. Kurkutov, E.O.; Shainyan, B.A. One-pot assembling of selenazolines from elemental selenium, alkenes and acetonitrile. Mendeleev Commun. 2022, 32, 395–396. [Google Scholar] [CrossRef]
  35. Greenwood, N.N.; Earnshaw, A. Chemistry of the Elements, 2nd ed.; Chapter 16; Butterworth-Heinemann: Oxford, UK, 1997; p. 571. [Google Scholar]
  36. Baldwin, J.E. Rules for ring closure. J. Chem. Soc. Chem. Commun. 1976, 734–736. [Google Scholar] [CrossRef]
  37. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  38. Gopal, M.; Milne, J. Spectroscopic evidence for selenium iodides in carbon disulfide solution: Se3I2, Se2I2, and SeI2. Inorg. Chem. 1992, 31, 4530–4533. [Google Scholar] [CrossRef]
Scheme 1. Iodine-mediated selenylation of alkenes with organic diselenides.
Scheme 1. Iodine-mediated selenylation of alkenes with organic diselenides.
Molecules 27 06169 sch001
Scheme 2. Previous works [32,34].
Scheme 2. Previous works [32,34].
Molecules 27 06169 sch002
Figure 1. Possible products of alkoxyselenylation of alkenes and dienes.
Figure 1. Possible products of alkoxyselenylation of alkenes and dienes.
Molecules 27 06169 g001
Figure 2. The studied alkenes 1ae and dienes 2ad.
Figure 2. The studied alkenes 1ae and dienes 2ad.
Molecules 27 06169 g002
Scheme 3. Synthesis of selenides 3ag.
Scheme 3. Synthesis of selenides 3ag.
Molecules 27 06169 sch003
Figure 3. Scope of selenides 3 *. * A total of 3.5 mmol alkene 1ae, 1 mL of alcohol, 1 mmol of fine-ground Se and I2 (2 mmol for 3ac or 1 mmol for 3dg), r.t., 40 h.
Figure 3. Scope of selenides 3 *. * A total of 3.5 mmol alkene 1ae, 1 mL of alcohol, 1 mmol of fine-ground Se and I2 (2 mmol for 3ac or 1 mmol for 3dg), r.t., 40 h.
Molecules 27 06169 g003
Scheme 4. Synthesis of selenides 5 and 6.
Scheme 4. Synthesis of selenides 5 and 6.
Molecules 27 06169 sch004
Scheme 5. Proposed mechanism of forming heterocycles 5 and 6.
Scheme 5. Proposed mechanism of forming heterocycles 5 and 6.
Molecules 27 06169 sch005
Scheme 6. Synthesis of selenide 7.
Scheme 6. Synthesis of selenide 7.
Molecules 27 06169 sch006
Scheme 7. Synthesis of selenide 9.
Scheme 7. Synthesis of selenide 9.
Molecules 27 06169 sch007
Scheme 8. Synthesis of selenide 10.
Scheme 8. Synthesis of selenide 10.
Molecules 27 06169 sch008
Figure 4. The relative total and free energies of all cations 10 with respect to the most stable cation 10d.
Figure 4. The relative total and free energies of all cations 10 with respect to the most stable cation 10d.
Molecules 27 06169 g004
Figure 5. Di- and triselenides 11 and 12.
Figure 5. Di- and triselenides 11 and 12.
Molecules 27 06169 g005
Figure 6. Synthesized linear and cyclic alkoxyselenides.
Figure 6. Synthesized linear and cyclic alkoxyselenides.
Molecules 27 06169 g006
Table 1. Optimization of reaction conditions *.
Table 1. Optimization of reaction conditions *.
Molecules 27 06169 i001
EntryIodine Source (Equiv)Solvent (mL)Time, hT, °CYield, % **Se Conversion, %
11.0 I2MeCN:MeOH (1:0.1)40257154
21.0 I2MeOH (1)40259068
31.0 I2MeOH (1)120259788
41.5 I2MeOH (1)40258586
52.0 I2MeOH (1)402587~100
62.0 I2MeOH (1)20258869
71.0 I2MeOH (1)10507048
82.0 NISMeOH (1)20257063
90.2 I2DMSO:MeOH (0.15:0.1)2025Mixture of
adducts
27
102.0 I2MeOH (0.2)
(5 equiv)
40257957
11MeOH (1)402500
* Reaction conditions: elemental Se (1 mmol), heptene-1 (3.5 mmol) and stirring in MeOH or MeCN:MeOH at 25–50 °C for 10–120 h. ** Iodides 4 formed in the amounts from traces to ca. 10% were not isolated and proved by NMR spectroscopy and mass spectrometry.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kurkutov, E.O.; Shainyan, B.A. Iodine-Mediated Alkoxyselenylation of Alkenes and Dienes with Elemental Selenium. Molecules 2022, 27, 6169. https://doi.org/10.3390/molecules27196169

AMA Style

Kurkutov EO, Shainyan BA. Iodine-Mediated Alkoxyselenylation of Alkenes and Dienes with Elemental Selenium. Molecules. 2022; 27(19):6169. https://doi.org/10.3390/molecules27196169

Chicago/Turabian Style

Kurkutov, Evgeny O., and Bagrat A. Shainyan. 2022. "Iodine-Mediated Alkoxyselenylation of Alkenes and Dienes with Elemental Selenium" Molecules 27, no. 19: 6169. https://doi.org/10.3390/molecules27196169

Article Metrics

Back to TopTop