Next Article in Journal
Antiulcerogenic Potential of the Ethanolic Extract of Ceiba speciosa (A. St.-Hil.) Ravenna Evaluated by In Vitro and In Vivo Studies
Next Article in Special Issue
Estradiol as the Trigger of Sirtuin-1-Dependent Cell Signaling with a Potential Utility in Anti-Aging Therapies
Previous Article in Journal
Glycogen Synthase Kinase 3β Is a Key Regulator in the Inhibitory Effects of Accumbal Cocaine- and Amphetamine-Regulated Transcript Peptide 55–102 on Amphetamine-Induced Locomotor Activity
Previous Article in Special Issue
Breed and Feeding System Impact the Bioactive Anti-Inflammatory Properties of Bovine Milk
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Click Chemistry of Selenium Dihalides: Novel Bicyclic Organoselenium Compounds Based on Selenenylation/Bis-Functionalization Reactions and Evaluation of Glutathione Peroxidase-like Activity

by
Maxim V. Musalov
* and
Vladimir A. Potapov
A. E. Favorsky Irkutsk Institute of Chemistry, Siberian Division of The Russian Academy of Sciences, 1 Favorsky Str., 664033 Irkutsk, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(24), 15629; https://doi.org/10.3390/ijms232415629
Submission received: 18 November 2022 / Revised: 5 December 2022 / Accepted: 7 December 2022 / Published: 9 December 2022
(This article belongs to the Special Issue Molecular Advances in Age-Related Diseases)

Abstract

:
A number of highly efficient methods for the preparation of novel derivatives of 9-selenabicyclo[3.3.1]nonane in high yields based on selenium dibromide and cis,cis-1,5-cyclooctadiene are reported. The one-pot syntheses of 2,6-diorganyloxy-9-selenabicyclo[3.3.1]nonanes using various O-nucleophiles including alkanols, phenols, benzyl, allyl, and propargyl alcohols were developed. New 2,6-bis(1,2,3-triazol-1-yl)-9-selenabicyclo[3.3.1]nonanes were obtained by the copper-catalyzed 1,3-dipolar cycloaddition of 2,6-diazido-9-selenabicyclo[3.3.1]nonane with unsubstituted gaseous acetylene and propargyl alcohol. The synthesis of 2,6-bis(vinylsulfanyl)-9-selenabicyclo[3.3.1]nonane, based on the generation of corresponding dithiolate anion from bis[amino(iminio)methylsulfanyl]-9-selenabicyclo[3.3.1]nonane dibromide, followed by the nucleophilic addition of the dithiolate anion to unsubstituted acetylene, was developed. The glutathione peroxidase-like activity of the obtained water-soluble products was estimated and compounds with high activity were found. Overall, 2,6-Diazido-9-selenabicyclo[3.3.1]nonane exhibits the highest activity among the obtained compounds.

1. Introduction

Selenium and selenium-containing compounds were considered poisons for many years, until this element was identified as a micronutrient for humans and mammals [1]. Since then, interest in the synthesis and properties of organoselenium compounds has increased significantly [2,3,4,5,6].
Organoselenium compounds, and particularly selenium heterocycles, exhibit various kinds of biological activity, including antitumor, antiviral, antibacterial, anti-inflammatory, antiproliferative, antifungal, and glutathione peroxidase-like properties [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33].
The glutathione peroxidase-like activity is perhaps the most important biological property of organoselenium compounds. It is known that a number of organoselenium compounds exhibit glutathione peroxidase-like activity and, in trace amounts, can potentially act in the human body as catalysts for the reduction in peroxides and other reactive oxygen species, preventing lipid peroxidation and other undesirable processes. Some examples of known functionalized organoselenium compounds with glutathione peroxidase-like activity, including selenides containing the hydroxyl group, are presented in Figure 1 [26,27,28,29,30,31,32,33]. The presence of the hydroxyl group increases the solubility in water, which is considered a desirable property of compounds with glutathione peroxidase-like activity [29,30,31,32,33].
Currently, selenium is recognized as an essential trace element for humans and plays an important role in the human body via selenium-containing enzymes. These enzymes (glutathione peroxidase, thioredoxin reductase, methionine sulfoxide reductase, etc.) are involved in redox regulation in the body, reducing hydrogen peroxide and lipid peroxide species and maintaining antioxidant activity [34,35,36]. Sufficient intake of selenium is very important for the human body. It is known that low or suboptimal levels of selenium intake are associated with a wide range of human diseases such as heart disease, stroke, arthritis, cystic fibrosis, and even several types of cancer [34,35,36]. Selenium supplementation in the elderly is an important strategy for prevention of age-related diseases.
A number of diseases and degenerative conditions are accompanied by particularly high levels of peroxide formation and oxidative stress, which suppresses the protective glutathione peroxidase effect. For example, ischemic reperfusion of infarction in stroke patients often results in cardiovascular and neurological injury from the damaging effects of peroxides and other reactive oxygen species released by neutrophils during the reperfusion process [33,37]. The well-known selenium heterocycle ebselen is used to treat cardiovascular diseases and to prevent ischemic stroke and acute stroke [23,24,25,26,27]. Ebselen is a novel anti-inflammatory drug exhibiting glutathione peroxidase-like and neuroprotective properties. This organoselenium compound has been studied in phase three clinical trials for its cardiovascular and neuroprotective effects [23,24,25,26,27,33]. Moreover, ebselen has recently entered clinical trials in COVID-19 patients as this compound was found to inhibit CoV2 activity and viral replication [24,25].
Aging is an inevitable process and is always accompanied by age-related diseases. Reactive oxygen species are important initial factors in aging and age-related diseases. Selenium contributes to reducing inflammation mediated by reactive oxygen species, reducing DNA damage, and plays an important role in the fight against aging and the prevention of age-related diseases [38,39,40]. Thus, there is a growing need to discover new classes of non-toxic water-soluble organoselenium GPx mimetics with high catalytic activity and improved properties.
Organoselenium compounds have proven themselves as versatile and efficient intermediates and synthons for modern organic synthesis [1,2,3,4,5,6,7,8,9,10,11,41,42,43,44,45,46,47,48,49,50]. The selenium-containing reagents and selenium-mediated reactions are used in the synthesis of many useful products, including the total synthesis of important biologically active molecules [1,2,3,4,5,6,7,8,9,10,11,41,42,43,44,45,46,47,48,49,50].
Earlier in this laboratory, selenium dichloride and dibromide were introduced for the first time into the synthesis of organoselenium compounds [51]. The use of these reagents in organic synthesis made it possible to obtain novel classes of organoselenium and heterocyclic compounds [51,52,53,54,55,56,57,58,59,60,61].
The transannular addition of selenium dichloride and dibromide to cis,cis-1,5-cyclooctadiene afforded 2,6-dichloro-9-selenabicyclo[3.3.1]nonane (1) and 2,6-dibromo-9-selenabicyclo[3.3.1]nonane (2) in near quantitative yields [56,57,58]. The compound 1 was used in studies of the anchimeric assistance effect of the selenium and sulfur atoms, quantified by the rates of nucleophilic substitution reactions. The anchimeric assistance effect of the selenium atom was found to be approximately two orders of magnitude higher than that of the sulfur atom [56]. Thus, the compounds 1 and 2 are very reactive in nucleophilic substitution reactions and useful reagents for click chemistry.
A number of efficient syntheses of novel organoselenium compounds were developed based on 2,6-dibromo-9-selenabicyclo[3.3.1]nonane (2) [56,57,58,59,60,61]. Inter alia, bis-pyridinium salt 3 was obtained by the reaction of nonane 2 with pyridine at room temperature in acetonitrile (Scheme 1).
Joint research of this laboratory with the Irkutsk research anti-plague institute established that compound 3 is a promising drug for metabolic correction during the vaccination process [61]. The introduction of compound 3 into the body of experimental animals significantly decreases the development of pathological reactions under the action of the tularemia vaccine and reduces the reactogenicity of the brucellosis vaccine by one order of magnitude [61]. Moreover, compound 3 does not show toxicity and may act as a catalyst for the decomposition of peroxides in the body, exhibiting glutathione peroxidase-like activity.
It is worthwhile to note that the undesirable post-vaccination reaction of the body is oxidative stress, which develops as a result of the increased generation of reactive oxygen species by cells [61]. It can also lead to inflammatory and allergic reactions, which are based on the process of lipid peroxidation. The development of new drugs with glutathione peroxidase-like activity for metabolic correction is an urgent task, considering the need to vaccinate the population against coronavirus and other diseases. The use of metabolic correction drugs can significantly reduce the side effects that occur during vaccination.
The recent award of the Nobel Prize to Sharpless, the founder of click chemistry, demonstrates the great importance of this field of organic chemistry [62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89]. The term “click chemistry” was introduced by Sharpless [62] and is widely used today. Click chemistry reactions should give very high yields of desired products, be broad in scope, and produce only harmless by-products. Available starting materials and reagents, simple reaction conditions, high selectivity, and convenient product isolation procedures are also important features of click chemistry. The classic example of click chemistry is the copper-catalyzed azide-alkynes 1,3-dipolar cycloaddition reaction. The products of this reaction, the 1,2,3-triazole derivatives, exhibit a variety of biological activities [62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89].
The application of organoselenium compounds in click chemistry reactions and the combination of the advantages of selenium-containing reagents with the copper-catalyzed click chemistry of azide-alkynes 1,3-dipolar cycloaddition reactions can give a new impetus to the development of organoselenium chemistry and the synthesis of new useful compounds with high biological activity [90,91].
We recently developed the efficient synthesis of bis-1,2,3-triazole derivatives of 9-selenabicyclo[3.3.1]nonane in high yields, combining selenium dihalide click chemistry with the click chemistry of copper-catalyzed azide-alkyne 1,3-dipolar cycloaddition reactions [59]. The copper-catalyzed cycloaddition reaction of 2,6-diazido-9-selenabicyclo[3.3.1]nonane with terminal acetylenes proceeded in a regioselective fashion, affording a number of 2,6-bis(4-organyl-1,2,3-triazole)-9-selenabicyclo[3.3.1]nonanes in high yields.
Thus, the development of efficient and selective methods for the synthesis of new classes of non-toxic water-soluble organoselenium compounds with high glutathione peroxidase-like activity based on the principles of click chemistry is an urgent task for chemists.

2. Results and Discussion

The aim of this research is to develop efficient syntheses of novel derivatives of 9-selenabicyclo[3.3.1]nonane by selenenylation/bis-functionalization reactions and nucleophilic substitution with various O-centered nucleophiles (water, alkanols, phenols, benzyl, allyl, and propargyl alcohols) and to estimate glutathione peroxidase-like activity of the obtained water-soluble products.
A highly efficient synthesis of dihydroxyl derivative of 9-selenabicyclo[3.3.1]nonane 4 in 96% yield was developed through the selenenylation/bis-hydroxylation reaction of selenium dibromide with cis,cis-1,5-cyclooctadiene in acetonitrile in the presence of water and sodium bicarbonate at room temperature (Scheme 2).
The process was carried out by the addition of selenium dibromide to cis,cis-1,5-cyclooctadiene, followed by the addition of an aqueous solution of sodium bicarbonate to the reaction mixture. The product 4 is an odorless, water-soluble, white crystalline compound that is easy to handle.
The selenenylation/bis-methoxylation reaction proceeded very smoothly in a mixture of acetonitrile and methanol at room temperature. The yield of the product, 2,6-dimethoxy-9-selenabicyclo[3.3.1]nonane 5, was as high as 98% (Scheme 3).
The process was carried out by adding a solution of selenium dibromide to a mixture of acetonitrile and methanol containing cis,cis-1,5-cyclooctadiene, followed by stirring at room temperature.
In the case of other alcohols (ethanol, propanol, butanol and isobutanol), the addition of sodium bicarbonate to the reaction mixture containing alcohols was necessary for effectively conducting the selenenylation/bis-alkoxylation process and obtaining the target products, 2,6-dialkoxy-9-selenabicyclo[3.3.1]nonanes 6–9, in high yields (91–96%, Scheme 4).
Unsaturated alcohols, allyl and propargyl alcohols, were successfully involved in the selenenylation/bis-alkoxylation reaction. The nucleophilicity of allyl and propargyl alcohols seems to be somewhat lower than that of the corresponding saturated alcohol, propanol, and also methanol and ethanol. We found that heating (~40 °C) in methylene chloride in the presence of sodium bicarbonate is preferable for obtaining allyloxy and propargyloxy derivatives 10 and 11 in high yields (97% and 95%, respectively) (Scheme 4).
When using benzyl and 3,4,5-trimethoxybenzyl alcohols under the conditions similar to reactions of alkanols (Scheme 3), as well as to reactions of allyl and propargyl alcohols (Scheme 4), the product yields were not high enough. It was possible to obtain high yields of benzyloxy derivatives by carrying out the reaction in a solvent with a slightly higher boiling point than that of methylene chloride. Refluxing in chloroform in the presence of sodium bicarbonate made it possible to obtain benzyloxy derivatives 12 and 13 in 94% and 92% yields, respectively (Scheme 5).
It is known that phenols are less reactive in nucleophilic reactions than alkanols and usually, in this case, more stringent conditions are required. Indeed, under the conditions of the reactions of alcohols shown in the Scheme 4, phenols gave low yields of the products. The one-pot synthesis of bis(4-methoxyphenoxy) and bis(3,5-dimethylphenoxy) derivatives 14 and 15 in 82–85% yields was developed through the addition of selenium dibromide to cyclooctadiene in acetonitrile, followed by refluxing the reaction mixture in the presence of potassium carbonate (Scheme 6).
The reaction with unsubstituted phenol under the same conditions, as indicated in Scheme 6, was very sluggish. The use of dibromo derivative 2 as a starting material in the nucleophilic substitution reaction with phenol was chosen as the better approach to the target product. Heating (60–70 °C) compound 2 with phenol in a solution of DMF in the presence of potassium carbonate allowed us to obtain bis-phenolic derivative 16 in 80% yield (Scheme 7).
It is known that the 1,2,3-triazole derivatives are available by the click chemistry reaction of copper-catalyzed azide-alkynes 1,3-dipolar cycloaddition and this class of organic compounds exhibit a variety of biological activities [62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89]. However, relatively few cycloaddition reactions with unsubstituted gaseous acetylene under atmospheric pressure have been described in the literature.
We developed the synthesis of the new triazole derivative of selenabicyclo[3.3.1]nonane, 2,6-bis(1,2,3-triazol-1-yl)-9-selenabicyclo[3.3.1]nonane 18, from diazide 17 and unsubstituted acetylene by the copper-catalyzed azide-alkynes 1,3-dipolar cycloaddition reaction (Scheme 8).
The catalytic system of Cu(OAc)2·H2O and sodium ascorbate was used to carry out cycloaddition reactions of diazide 17 with acetylene and propargyl alcohol. This system was found to be very efficient in the reactions of selenium-containing organic azides with terminal acetylenes [90,91]. The active Cu(I) catalyst is generated in situ from the Cu(II) salt via the reduction in copper acetate with sodium ascorbate. The addition of a slight excess of sodium ascorbate prevents the formation of oxidative homocoupling products.
We involved unsubstituted gaseous acetylene in the copper-catalyzed azide-alkyne 1,3-dipolar cycloaddition reaction with diazido derivative 17 (Scheme 8). The process was carried out under atmospheric pressure by bubbling gaseous acetylene into the reaction mixture. The target product 18 was obtained in 72% yield.
The efficient synthesis of 2,6-bis(4-hydroxymethyl-1,2,3-triazol-1-yl)-9-selenabicyclo[3.3.1]nonane 19 in 90% yield was developed by the 1,3-dipolar cycloaddition reaction of diazido derivative 17 with propargyl alcohol (Scheme 9).
Along with the 1,3-dipolar cycloaddition reaction, unsubstituted acetylene was involved in the nucleophilic addition reaction of dithiolate anion 21 generated from bis-isothiuronium salt 20. We previously obtained compound 20 by refluxing dibromo derivative 2 with a high excess of thiourea in acetonitrile [60]. Here, we report the preparation of this compound in 95% yield, at room temperature, at a stoichiometric ratio of the reagents and the synthesis of bis(vinylsulfanyl) derivative of 9-selenabicyclo[3.3.1]nonane 22 in 81% yield (Scheme 10).
Thiourea is a sulfur-centered nucleophile. In the same period of the periodic table is phosphorus, the compounds of which are more nucleophilic than their sulfur counterparts. We carried out the reaction of dibromo derivative 2 with triphenyl phosphine, which proceeded smoothly with the formation of water-soluble bis-phosphonium salt 23 in quantitative yield (Scheme 11).
Of the products obtained, compounds 3, 4, 17, 20 and 23 were well water-soluble. These compounds were used for the estimation of glutathione peroxidase-like activity.
Previously, the glutathione peroxidase-like activity of compound 3 was not estimated. It is important that compound 3 is non-toxic [61] (and probably some other 9-selenabicyclo[3.3.1]nonane derivatives as well). It is also worth noting that 9-selenabicyclo[3.3.1]nonane derivatives have a relatively rigid configuration with a highly sterically accessible selenium atom that can act as an active center for the glutathione peroxidase-like catalysis.
The glutathione peroxidase-like activity of the obtained products was estimated using the model reaction of dithiothreitol oxidation by tert-butyl hydroperoxide in D2O, in the presence of synthesized compounds as a catalysts (10% mol, Scheme 12) [28,29,30,31,32,33]. The progress of this reaction was monitored by 1H NMR spectroscopy at room temperature (tert-butyl hydroperoxide, dithiothreitol, 0.025 mmol; tested product, 0.0025 mmol; D2O, 0.5 mL). The control experiment was conducted under the same reaction conditions, but in the absence of the catalyst.
It was found that diazido derivative 17 showed the best activity among the tested products (Figure 2). This compound is considerably superior to other products in activity. The second most active product is bis-pyridinium salt 3, which is considered a promising drug for metabolic correction during the vaccination process [61]. Compound 4, with two hydroxyl groups, is the third in activity. Isothiuronium and phosphonium salts 20 and 23 exhibit less activity than compounds 17, 3 and 4.
Compound 17, which is superior to other products in activity, has two very polar azido groups that have a linear configuration and are located in the relatively rigid molecule on the opposite side of the selenium atom, i.e., the azido groups do not sterically interfere with reactions at the selenium atom and the manifestation of glutathione peroxidase-like activity.
A supposed catalytic cycle to explain the catalytic effect of the obtained compounds with the regeneration of the catalyst is presented in Scheme 13. The reaction of the catalyst with tert-butyl hydroperoxide leads to corresponding selenoxides A, which form the heterocyclic intermediate B with dithiothreitol. The intermediate B undergoes conversion to the oxidized form of dithiothreitol with the regeneration of the catalyst. The intermediates with the sulfur−selenium bond are often considered as intermediates in the oxidation reactions of thiols with peroxides, catalyzed by organoselenium compounds [28,29,30,31,32,33].
The catalytic cycle of glutathione peroxidase in the human body also involves the formation of intermediates with the S−Se bond [92]. Glutathione peroxidase (GPx) contains a selenocysteine residue with the selenol function. The catalytic cycle involves oxidation of the selenol group of the selenocysteine residue by hydrogen peroxide. This process gives selenenic acid derivative (RSeOH), which reacts with glutathione (GSH) to form the intermediate with the S−Se bond (GS−SeR). A second glutathione molecule reduces the GS−SeR intermediate back to the selenol derivative, releasing the disulfide form of glutathione (GS−SG) [92].
Previously, we studied the exchange reactions of dialkyl disulfides with dialkyl diselenides [93]. It was found that the ease of the exchange reaction of dialkyl dichalcogenides (dialkyl disulfides, diselenides and ditellurides) generally rises with the increasing atomic number of the chalcogen and with the decreasing bulk of the alkyl moiety [93]. The equilibrium constants of the exchange reaction of dialkyl disulfides and dialkyl diselenides decrease with the increasing bulk of the alkyl moiety. A number of alkylselenenyl alkyl sulfides have been isolated from the exchange reaction and described for the first time [93].
Selenoxides A are considered to be intermediates in the catalytic cycle (Scheme 13). We attempted to obtain corresponding selenoxides from the compounds, which were used for studies of glutathione peroxidase-like activity (Figure 2). It is known that some organic selenoxides are unstable compounds. Pure selenoxides 24 and 25 were obtained in 92–94% yields by oxidation of bis-pyridinium salt 3 in water and dihydroxyl derivatives 4 in methylene chloride with tert-butyl hydroperoxide (Scheme 14).
The structural assignments of the synthesized compounds were made using 1H, 13C, 77Se and 31P-NMR spectroscopy, including two-dimensional HMBC experiments, and were confirmed by elemental analysis.
The signals of the carbon atoms of the CH group, which are bonded to the oxygen atom, are observed in the 79–81 ppm region in the 13C-NMR spectra of 2,6-diorganyloxy-9-selenabicyclo[3.3.1]nonanes 516. The 13C-NMR spectra of products 18 and 19 contain signals in the olefin region, which correspond to the C=C group of the triazole ring.
The obtained values of the 77Se-NMR chemical shifts for alkoxy, allyloxy and propargyloxy derivatives 5–13 and aryloxy products 14–16 are very close (~287–293 ppm). The selenium atom in azido compound 17 and the triazole derivatives 18 and 19 resonates at 331.1, 344.7, and 334.2 ppm, respectively. A high downfield shift of the selenium signals is observed for compounds containing positively charged atoms (382.2, 415.1 and 521.4 ppm for bis-isothiuronium 20, bis-pyridinium 3, and bis-phosphonium 23 salts, respectively). The obtained values of the 77Se-NMR chemical shifts for selenoxides (851.6 and 841.5 ppm for the products 24 and 25, respectively) are typical for this class of organoselenium compounds.
Characteristic fragment ions [M−R]+ in the mass spectra of all products and molecular ions in the mass spectra of organyloxy derivatives 5–11 are observed. The mass spectra of bis-aryloxy derivatives 14–16 show intense ions, which correspond to the elimination of one aryloxy fragment from the molecule.

3. Materials and Methods

3.1. General Information

The 1H (400.1 MHz), 13C (100.6 MHz), and 77Se (76.3 MHz) NMR spectra (the spectra can be found in the Supplementary Materials) were recorded on a Bruker DPX-400 spectrometer (Bruker BioSpin GmbH, Rheinstetten, Germany) and referred to the residual solvent peaks of CDCl3 (δ = 7.27 and 77.16 ppm in 1H- and 13C-NMR, respectively), DMSO (δ = 2.50 and 39.50 ppm for 1H- and 13C-NMR, respectively) or D2O (δ = 4.79 ppm for 1H-NMR) and dimethyl selenide (77Se-NMR).
The mass spectra were recorded on a Shimadzu GCMS-QP5050A (Shimadzu Corporation, Kyoto, Japan) with electron impact (EI) ionization (70 eV). The elemental analysis was performed on a Thermo Scientific Flash 2000 Elemental Analyzer (Thermo Fisher Scientific Inc., Milan, Italy). The melting points were determined on a Kofler Hot-Stage Microscope PolyTherm A apparatus (Wagner and Munz GmbH, München, Germany). The distilled organic solvents and degassed water were used in syntheses.

3.2. Synthesis of Compounds 4–9

2,6-Dihydroxy-9-selenabicyclo[3.3.1]nonane (4). A solution of selenium dibromide was prepared from elemental selenium (158 mg, 2 mmol) and bromine (320 mg, 2 mmol) in methylene chloride (1 mL). The solution of selenium dibromide was added dropwise to a solution of cyclooctadiene (216 mg, 2 mmol) in CH3CN (10 mL). The mixture was stirred for 4 h at room temperature and a solution of NaHCO3 (0.3 g, 3.6 mmol) in water (2 mL) was added. The reaction mixture was stirred overnight (18 h) at room temperature. The solvent was removed on a rotary evaporator, the residue was extracted with methylene chloride (3 × 10 mL). The organic phase was dried over CaCl2, the solvent was removed by a rotary evaporator and the residue was dried in vacuum giving the product (424 mg, 96% yield) as white crystals, mp 249–250 °C.
1H NMR (400 MHz, CDCl3) δ 1.73–1.81 (m, 2H, CH2), 1.85–1.92 (m, 2H, CH2), 2.04–2.14 (m, 2H, CH2), 2.62–2.70 (m, 4H, CH2, OH), 3.59–3.63 (m, 2H, SeCH), 4.29–4.34 (m, 2H, OCH).
13C NMR (100 MHz, CDCl3) δ 25.91, 29.50, 30.45, 69.43.
77Se NMR (76.3 MHz, CDCl3): 278.1.
MS (EI): m/z (%) = 222 (52, M+), 205 (5), 178 (20), 149 (15), 133 (17), 123 (27), 95 (58), 79 (69), 71 (49), 67 (41), 57 (31), 55 (58), 41 (100), 39 (51).
IR (KBr): λ = 873, 982, 1017, 2899, 2932, 3340 cm−1.
Found: C, 43.74; H, 6.56; Se, 35.49. Calc. for C8H14O2Se: C, 43.45; H, 6.38; Se 35.70.
2,6-Dimethoxy-9-selenabicyclo[3.3.1]nonane (5). A solution of selenium dibromide was prepared from elemental selenium (158 mg, 2 mmol) and bromine (320 mg, 2 mmol) in methylene chloride (1 mL). The solution of selenium dibromide was added dropwise to a solution of cyclooctadiene (216 mg, 2 mmol) in acetonitrile (10 mL). The mixture was stirred for 4 h at room temperature and methanol (2 mL) was added. The mixture was stirred overnight (20 h) at room temperature. The solvent was removed on a rotary evaporator and the residue was dried in vacuum giving the product (488 mg, 98% yield) as a light-yellow oil.
1H NMR (400 MHz, CDCl3) δ 1.78–1.82 (m, 2H, CH2), 2.00–2.04 (m, 2H, CH2), 2.17–2.19 (m, 2H, CH2), 2.60–2.64 (m, 2H, CH2), 3.02–3.03 (m, 2H, SeCH), 3.36 (s, 6H, CH3), 3.88–3.92 (m, 2H, OCH).
13C NMR (100 MHz, CDCl3) δ 27.87, 28.05, 28.87, 55.86, 81.04.
77Se NMR (76.3 MHz, CDCl3): 288.7
MS (EI): m/z (%) = 250 (30, M+), 218 (10), 179 (18), 137 (64), 105 (50), 79 (100), 71 (89), 45 (78), 41 (90).
IR (film): λ = 1086, 1153, 1186, 2817, 2922, 2977 cm−1.
Found: C, 47.89; H, 7.41; Se, 31.43. Calc. for C10H18O2Se: C, 48.20; H, 7.28; Se 31.68.
2,6-Diethoxy-9-selenabicyclo[3.3.1]nonane (6). A solution of selenium dibromide was prepared from elemental selenium (158 mg, 2 mmol) and bromine (320 mg, 2 mmol) in methylene chloride (1 mL). The solution of selenium dibromide was added dropwise to a solution of cyclooctadiene (216 mg, 2 mmol) in methylene chloride (10 mL). The mixture was stirred for 8 h at room temperature and ethanol (2 mL) and NaHCO3 (0.3 g, 3.6 mmol) were added. The mixture was stirred overnight (18 h) at room temperature. The mixture was filtered and the solvent was removed from the filtrate on a rotary evaporator. The residue was dried in vacuum giving the product (532 mg, 96% yield) as a light-yellow oil.
1H NMR (400 MHz, CDCl3): 1.20 (t, 6H, CH3), 1.80–1.91 (m, 2H, CH2), 1.96–2.03 (m, 2H, CH2), 2.14–2.24 (m, 2H, CH2), 2.64–2.69 (m, 2H, CH2), 2.98–3.01 (m, 2H, CHSe), 3.43–3.51 (m, 2H, CH2O), 3.57–3.65 (m, 2H, CH2O), 3.97–4.03 (m, 2H, CHO).
13C NMR (100 MHz,CDCl3): 15.9 (CH3), 28.2 (CH2), 28.9 (CHSe), 29.5 (CH2), 63.6 (CH2O), 79.5 (CHO).
77Se NMR (76.3 MHz, CDCl3): 288.9.
MS (EI): m/z (%) = 278 (40, M+), 232 (14), 193 (30), 151 (36), 123 (35), 105 (50), 85 (48), 79 (60), 57 (100), 41 (86).
IR (film): λ = 1020, 1082, 1159, 2887, 2922, 2971 cm−1
Anal. calcd for C12H22O2Se (277.26): C 51.98, H 8.00, O 11.54, Se 28.48%. Found: C 51.89, H 7.96, Se 28.64%.
2,6-Dipropoxy-9-selenabicyclo[3.3.1]nonane (7) was obtained under the same conditions as compound 6 in 94% yield using propanol.
1H NMR (400 MHz, CDCl3): 0.86 (t, 6H, CH3), 1.48–1.58 (m, 4H, CH2), 1.74–1.86 (m, 2H, CH2), 1.90–1.97 (m, 2H, CH2), 2.07–2.17 (m, 2H, CH2), 2.58–2.64 (m, 2H, CH2), 2.92–2.95 (m, 2H, CHSe), 3.29–3.38 (m, 2H, CH2O), 3.41–3.49 (m, 2H, CH2O), 3.89–3.94 (m, 2H, CHO).
13C NMR (100 MHz,CDCl3): 10.7 (CH3), 23.4 (CH2), 28.1 (CH2), 28.8 (CHSe), 29.3 (CH2), 69.9 (CH2O), 79.5 (CHO).
77Se NMR (76.3 MHz, CDCl3): 289.2.
MS (EI): m/z (%) = 306 (12, M+), 246 (10), 207 (11), 123 (30), 105 (34), 79 (36), 57 (50), 43 (100).
IR (film): λ = 1038, 1082, 1165, 2873, 2932, 2959 cm−1
Anal. calcd for C14H26O2Se (305.31): C 55.08, H 8.58, O 10.48, Se 25.86%. Found: C 54.98, H 8.56, Se 26.06%.
2,6-Dibutoxy-9-selenabicyclo[3.3.1]nonane (8) was obtained under the same conditions as compound 6 in 93% yield using butanol.
1H NMR (400 MHz, CDCl3): 0.91 (t, 6H, CH3), 1.33–1.41 (m, 4H, CH2), 1.50–1.57 (m, 4H, CH2), 1.81–1.89 (m, 2H, CH2), 1.94–2.01 (m, 2H, CH2), 2.13–2.21 (m, 2H, CH2), 2.62–2.68 (m, 2H, CH2), 2.97–2.99 (m, 2H, CHSe), 3.37–3.43 (m, 2H, CH2O), 3.50–3.55 (m, 2H, CH2O), 3.93–3.98 (m, 2H, CHO).
13C NMR (100 MHz,CDCl3): 14.0 (CH3), 19.5 (CH2), 28.2 (CH2), 28.9 (CHSe), 29.4 (CH2), 32.4 (CH2), 68.1 (CH2O), 79.7 (CHO).
77Se NMR (76.3 MHz, CDCl3): 288.9.
MS (EI): m/z (%) = 334 (8, M+), 260 (5), 221 (8), 165 (13), 123 (32), 105 (27), 79 (41), 57 (92), 41 (100).
IR (film): λ = 1047, 1087, 1148, 2868, 2931, 2966 cm−1.
Anal. calcd for C16H30O2Se (333.37): C 57.65, H 9.07, O 9.60, Se 23.69%. Found: C 57.67, H 9.09, Se 23.85%.
2,6-Diisobutoxy-9-selenabicyclo[3.3.1]nonane (9) was obtained under the same conditions as compound 6 in 91% yield using isobutanol.
1H NMR (400 MHz, CDCl3): 0.87–0.93 (m, 12H, CH3), 1.77–1.91 (m, 4H, CH2, CH), 1.96–2.03 (m, 2H, CH2), 2.12–2.22 (m, 2H, CH2), 2.64–2.69 (m, 2H, CH2), 2.97–3.00 (m, 2H, CHSe), 3.16–3.20 (m, 2H, CH2O), 3.26–3.30 (m, 2H, CH2O), 3.92–3.97 (m, 2H, CHO).
13C NMR (100 MHz,CDCl3): 19.5 (CH3), 19.6 (CH3), 28.2 (CH2), 28.9 (CHSe), 29.0 (CH2), 29.4 (CH2), 75.4 (CH2O), 79.8 (CHO).
77Se NMR (76.3 MHz, CDCl3): 286.9.
MS (EI): m/z (%) = 334 (7, M+), 260 (6), 221 (5), 165 (14), 123 (29), 105 (23), 79 (37), 57 (91), 41 (100).
IR (film): λ = 1043, 1085, 1139, 2863, 2934, 2959 cm−1.
Anal. calcd for C16H30O2Se (333.37): C 57.65, H 9.07, O 9.60, Se 23.69%. Found: C 57.74, H 9.11, Se 23.72%.

3.3. Synthesis of Compounds 10–13

2,6-Diallyloxy-9-selenabicyclo[3.3.1]nonane (10). A solution of selenium dibromide was prepared from elemental selenium (158 mg, 2 mmol) and bromine (320 mg, 2 mmol) in methylene chloride (1 mL). The solution of selenium dibromide was added dropwise to a solution of cyclooctadiene (216 mg, 2 mmol) in a mixture of methylene chloride (10 mL). The mixture was stirred for 4 h at room temperature and allyl alcohol (2 mL) and NaHCO3 (0.3 g, 3.6 mmol) were added. The mixture was refluxed for 8 h. The mixture was filtered and the solvent was removed from the filtrate on a rotary evaporator. The residue was dried in vacuum, giving the product (584 mg, 97% yield) as a light-yellow oil.
1H NMR (400 MHz, CDCl3): 1.78–1.90 (m, 2H, CH2), 1.94–2.00 (m, 2H, CH2), 2.11–2.18 (m, 2H, CH2), 2.62–2.67 (m, 2H, CH2), 2.93–2.96 (m, 2H, CHSe), 3.92–4.06 (m, 6H, CHO, CH2O), 5.10–5.25 (dd, 4H, CH2=CH), 5.82–5.92 (m, 2H, CH2=CH).
13C NMR (100 MHz,CDCl3): 28.1 (CH2), 28.7 (CHSe), 29.1 (CH2), 69.2 (CH2O), 79.0 (CHO), 116.6 (CH2=CH), 135.3 (CH2=CH).
77Se NMR (76.3 MHz, CDCl3): 291.6.
MS (EI): m/z (%) = 302 (14, M+), 245 (22), 205 (12), 187 (16), 121 (21), 93 (31), 79 (45), 55 (39), 41 (100).
IR (film): λ = 1049, 1069, 1126, 1645, 2850, 2917, 2984 cm−1.
Anal. calcd for C14H22O2Se (301.28): C 55.81, H 7.36, O 10.62, Se 26.21%. Found: C 55.78, H 7.32, Se 26.31%.
2,6-Dipropargyloxy-9-selenabicyclo[3.3.1]nonane (11) was obtained under the same conditions as compound 10 in 95% yield using propagyl alcohol.
1H NMR (400 MHz, CDCl3): 1.80–1.91 (m, 2H, CH2), 1.99–2.06 (m, 2H, CH2), 2.16–2.26 (m, 2H, CH2), 2.42 (t, 2H, CCH), 2.65–2.71 (m, 2H, CH2), 3.01–3.04 (m, 2H, CHSe), 4.19 (d, 4H, OCH2C), 4.21–4.26 (m, 2H, CHO).
13C NMR (100 MHz,CDCl3): 28.3 (CH2), 28.4 (CH2), 29.0 (CHSe), 55.6 (CH2O), 74.2 (CCH), 79.0 (CHO), 80.3 (CCH).
77Se NMR (76.3 MHz, CDCl3): 295.8.
MS (EI): m/z (%) = 298 (21, M+), 259 (6), 243 (10), 203 (11), 161 (18), 133 (21), 107 (22), 91 (41), 79 (37), 55 (42), 39 (100).
IR (film): λ = 1017, 1069, 1160, 2115, 2851, 2918, 2984 cm−1.
Anal. calcd for C14H18O2Se (297.25): C 56.57, H 6.10, O 10.76, Se 26.56%. Found: C 56.63, H 6.06, Se 26.61%.
2,6-Dibenzyloxy-9-selenabicyclo[3.3.1]nonane (12) was obtained under the same conditions as compound 10 in 94% yield using benzyl alcohol and chloroform.
1H NMR (400 MHz, CDCl3): 1.94–2.09 (m, 4H, CH2), 2.18–2.25 (m, 2H, CH2), 2.73–2.79 (m, 2H, CH2), 3.05–3.08 (m, 2H, CHSe), 4.11–4.16 (m, 2H, CHO), 4.52–4.57 (m, 2H, CH2O), 4.58–4.66 (m, 2H, CH2O), 7.28–7.39 (m, 10H, CHAr).
13C NMR (100 MHz,CDCl3): 28.3 (CH2), 28.8 (CHSe), 29.4 (CH2), 70.3 (CH2O), 79.3 (CHO), 127.7 (CHAr), 127.7 (CHAr), 128.5 (CHAr), 139.0 (CAr).
77Se NMR (76.3 MHz, CDCl3): 291.3.
MS (EI): m/z (%) = 402 (5, M+), 203 (10), 105 (8), 91 (100), 79 (45), 65 (12), 41 (7).
IR (film): λ = 1027, 1063, 1087, 1363, 1454, 2860, 2919 cm−1.
Anal. calcd for C22H26O2Se (401.40): C 65.83, H 6.53, O 7.97, Se 19.67%. Found: C 65.88, H 6.54, Se 19.72%.
2,6-Bis(3,4,5-trimethoxybenzyloxy)-9-selenabicyclo[3.3.1]nonane (13) was obtained under the same conditions as compound 10 in 92% yield using 3,4,5-trimethoxybenzyl alcohol and chloroform.
1H NMR (400 MHz, CDCl3): 1.93–2.09 (m, 4H, CH2), 2.17–2.27 (m, 2H, CH2), 2.72–2.77 (m, 2H, CH2), 3.02–3.05 (m, 2H, CHSe), 3.82 (s, 6H, OCH3), 3.84 (s, 12H, OCH3), 4.08–4.14 (m, 2H, CHO), 4.42–4.48 (m, 2H, CH2O), 4.52–4.58 (m, 2H, CH2O), 6.55 (s, 4H, CHAr).
13C NMR (100 MHz,CDCl3): 28.3 (CH2), 28.7 (CHSe), 29.3 (CH2), 56.1 (OCH3), 60.8 (OCH3), 70.4 (CH2O), 79.2 (CHO), 104.6 (CHAr), 134.5 (CAr), 137.5 (CAr), 153.3 (CAr).
77Se NMR (76.3 MHz, CDCl3): 292.8.
MS (EI): m/z (%) = 293 (7), 195 (6), 181 (100), 79 (41), 65 (8), 41 (7).
Anal. calcd for C26H38O8Se (557.53): C 56.01, H 6.87, O 22.96, Se 14.16%. Found: C 55.94, H 6.81, Se 14.22%.

3.4. Synthesis of Phenol Derivatives 14–16

2,6-Bis(4-methoxyphenoxy)-9-selenabicyclo[3.3.1]nonane (14). The solution of selenium dibromide (1 mmol) was added dropwise to a solution of cyclooctadiene (108 mg, 1 mmol) in acetonitrile (5 mL). The mixture was stirred for 4 h at room temperature and 4-methoxyphenol (310 mg, 2.5 mmol) and powdered potassium carbonate (300 mg, 2.1 mmol) were added and the mixture was refluxed for 16 h. The mixture was cooled, diluted with cold water (20 mL) and extracted with ethyl acetate (3 × 15 mL). The combined organic phase was washed with water, dried over Na2SO4, and the solvent was removed on a rotary evaporator. The residue was subjected to column chromatography on silica gel (eluent: hexane/chloroform 7:1 → hexane/chloroform 1:5) giving the product (355 mg, 82% yield).
1H NMR (400 MHz, CDCl3): 2.17–2.34 (m, 6H, CH2), 2.82–2.87 (m, 2H, CH2), 3.06–3.10 (m, 2H, CHSe), 3.78 (s, 6H, OCH3), 4.87–4.93 (m, 2H, CHO), 6.82–6.86 (s, 4H, CHAr), 6.88–6.93 (m, 4H, CHAr).
13C NMR (100 MHz,CDCl3): 28.16(CH2), 28.48 (CHSe), 28.99 (CH2), 55.82 (OCH3), 79.36 (CHO), 114.89 (CHAr), 118.33 (CHAr), 151.20 (OCAr), 154.51 (OCAr).
77Se NMR (76.3 MHz, CDCl3): 293.2.
MS (EI): m/z (%) = 311 (79, M+–MeC6H4O), 205 (23), 187 (32), 161 (15), 137 (12), 123 (100), 105 (47), 79 (58), 67 (19), 41 (37).
IR (film): λ = 1017, 1037, 1103, 1215, 1442, 1453, 1504, 2851, 2918 cm−1.
Anal. calcd for C22H26O4Se (433.40): C 60.97, H 6.05, O 14.77, Se 18.22%. Found: C 61.02, H 6.04, Se 18.31%.
2,6-Bis(3,5-dimethylphenoxy)-9-selenabicyclo[3.3.1]nonane (15) was obtained under the same conditions as compound 14 in 85% yield.
1H NMR (400 MHz, CDCl3): 2.20–2.26 (m, 4H, CH2), 2.28–2.36 (m, 2H, CH2), 2.31 (s, 12H, CH3), 2.83–2.88 (m, 2H, CH2), 3.14–3.16 (m, 2H, CHSe), 5.00–5.05 (m, 2H, CHO), 6.59 (s, 4H, CHAr), 6.64 (s, 2H, CHAr).
13C NMR (100 MHz,CDCl3): 21.6 (CH3), 28.2 (CH2), 28.5 (CHSe), 28.9 (CH2), 77.7 (CHO), 114.3 (CHAr), 121.1 (CHAr), 139.5 (CAr), 157.3 (CAr).
77Se NMR (76.3 MHz, CDCl3): 291.5.
MS (EI): m/z (%) = 309 (97, M+–Me2C6H4O), 227 (8), 187 (48), 159 (21), 135 (32), 105 (92), 101 (52), 79 (100), 67 (25), 41 (33).
IR (film): λ = 1049, 1147, 1187, 1292, 1318, 1472, 1593, 2850, 2919, 2844 cm−1.
Anal. calcd for C24H30O2Se (429.45): C 67.12, H 7.04, O 7.45, Se 18.39%. Found: C 67.24, H 7.01, Se 18.55%.
2,6-Diphenoxy-9-selenabicyclo[3.3.1]nonane (16). Powdered potassium carbonate (300 mg, 2.1 mmol) was added to a mixture of compound 2 (248 mg, 1 mmol), phenol (282 mg, 3 mmol), and DMF (4 mL) and the mixture was heated at 60–70 C for 8 h. The mixture was cooled, diluted with cold water (20 mL) and extracted with ethyl acetate (3 × 15 mL). The combined organic phase was washed with water, dried over Na2SO4, and the solvent was removed on a rotary evaporator. The residue was subjected to column chromatography on silica gel (eluent: hexane/chloroform 9:1 → hexane/chloroform 1:5) giving the product (299 mg) in 80% yield.
1H NMR (400 MHz, CDCl3): 2.20–2.35 (m, 6H, CH2), 2.82–2.88 (m, 2H, CH2), 3.12–3.15 (m, 2H, CHSe), 5.02–5.07 (m, 2H, CHO), 6.93–6.99 (m, 6H, CHAr), 7.30 (t, 4H, CHAr).
13C NMR (100 MHz,CDCl3): 28.2 (CH2), 28.4 (CHSe), 28.8 (CH2), 78.0 (CHO), 116.7 (CHAr), 121.4 (CHAr), 129.8 (CHAr), 157.3 (CAr).
77Se NMR (76.3 MHz, CDCl3): 294.8.
MS (EI): m/z (%) = 281 (87, M+–C6H5O), 187 (41), 157 (15), 145 (28), 105 (73), 79 (100), 67 (50), 39 (58).
IR (film): λ = 1017, 1168, 1226, 1491, 1597, 2853, 2926 cm−1.
Anal. calcd for C20H22O2Se (373.35): C 64.34, H 5.94, O 8.57, Se 21.15%. Found: C 64.42, H 5.91, Se 21.20%.

3.5. Synthesis of Compounds 17–19

2,6-Diazido-9-selenabicyclo[3.3.1]nonane (17). A solution of sodium azide (1.8 g, 2.7 mmol) in water (14 mL) was added dropwise to a mixture of compound 2 (1 g, 2.87 mmol) and acetonitrile (24 mL) with stirring at room temperature. The reaction mixture was stirred overnight (20 h) at room temperature. Acetonitrile was removed by a rotary evaporator and the residue was extracted with methylene chloride (3 × 20 mL). The organic phase was dried over Na2SO4, methylene chloride was removed by a rotary evaporator and the residue was dried in vacuum giving a compound 3 (763 mg, 98% yield) as a grey oil. Spectral characteristics and elemental analysis data were reported [49].
77Se NMR (76.3 MHz, CDCl3): 331.1.
2,6-Bis(1,2,3-triazol-1-yl)-9-selenabicyclo[3.3.1]nonane (18). A solution of sodium ascorbate (84 mg, 0.42 mmol) in water (2 mL) was added to Cu(OAc)2·H2O (42 mg, 0.21 mmol) and the mixture was stirred for 5 min. A solution of compound 3 (189 mg, 0.7 mmol) in methanol (3 mL) was added to the reaction mixture. The reaction mixture was saturated with acetylene by bubbling for 8 h at room temperature with stirring. Then the bubbling of acetylene was stopped, the flask was closed and the mixture was stirred overnight (18 h) at room temperature. The reaction mixture was diluted with H2O (8 mL) and extracted with methylene chloride (3 × 10 mL). The organic phase was dried over Na2SO4, the solvent was removed by a rotary evaporator. The residue was subjected to column chromatography on silica gel (eluent: hexane → hexane/chloroform 1:1 → hexane/chloroform 1:9) giving the product (163 mg) in 72% yield as a white powder; mp 155–157 °C.
1H NMR (400 MHz, CDCl3): 2.28–2.36 (m, 2H, CH2), 2.46–2.58 (m, 4H, CH2), 3.09–3.19 (m, 2H, CH2), 3.35–3.40 (m, 2H, CHSe), 5.49–5.55 (m, 2H, CHN), 7.68 (s, 2H, CHCH), 7.77 (s, 2H, CHCH).
13C NMR (100 MHz,CDCl3): 27.43(CH2), 29.21 (CHSe), 30.36 (CH2), 63.04 (CHN), 122.41 (CHCH), 133.60 (CHCH).
77Se NMR (76.3 MHz, CDCl3): 344.7.
MS (EI): m/z (%) = 324 (5, M+), 255 (50), 186 (44), 145 (12), 105 (100), 79 (61), 67 (38), 41 (67).
IR (KBr): λ = 1025, 1070, 1113, 1486, 1562, 2851, 2921 cm−1.
Anal. calcd for C12H16N6Se (323.26): C 44.59, H 4.99, N 26.00, Se 24.43%. Found: C 44.82, H 5.18, N 25.76, Se 23.15%.
2,6-Bis(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)-9-selenabicyclo[3.3.1]nonane (19). A solution of sodium ascorbate (84 mg, 0.42 mmol) in water (3 mL) was added to Cu(OAc)2.H2O (42 mg, 0.21 mmol) and the mixture was stirred for 5 min. A solution of compound 2 (189 mg, 0.7 mmol) and propargyl alcohol (136 mg, 2 mmol) in methanol (3 mL) was added dropwise for 10 min. The reaction mixture was stirred for 24 h at room temperature. Methanol was distilled off by a rotary evaporator. The residue was extracted with methylene chloride (3 × 15 mL). The organic phase was dried over Na2SO4, the solvent and an excess of propargyl alcohol was removed by a rotary evaporator and by drying in vacuum. The product (241 mg, 90% yield) was obtained as a grey powder, mp 181–183 °C.
1H NMR (400 MHz, CDCl3): 2.14–2.31 (m, 6H, CH2), 2.80–2.87 (m, 2H, CH2), 3.07–3.11 (m, 2H, CHSe), 3.75 (s, 6H, OCH3), 4.86–4.93 (m, 2H, CHO), 6.83–6.92 (m, 8H, CHAr).
13C NMR (100 MHz,CDCl3): 26.3 (CH2), 28.7 (CHSe), 29.9 (CH2), 55.2 (CH2OH), 62.2 (CH2CHN), 121.6 (CHAr), 147.6 (CAr).
77Se NMR (76.3 MHz, CDCl3): 334.2.
MS (EI): m/z (%) = 384 (2, M+), 229 (16), 202 (18), 186 (20), 120 (36), 105 (68), 49 (41), 93 (70), 79 (68), 67 (78), 57 (73), 41 (100).
IR (KBr): λ = 1019, 1046, 1130, 1556, 2863, 2922 cm−1.
Anal. calcd for C14H20N6O2Se (383.31): C 43.87, H 5.26, N 21.92, O 8.35, Se 20.60%. Found: C 43.74, H 5.24, N 21.98, Se 20.76%.

3.6. Synthesis of Compounds 20–23

2,6-Bis[amino(iminio)methylsulfanyl]-9-selenabicyclo[3.3.1]nonane dibromide (20). A solution of compound 2 (0.348 g, 1 mmol) in methylene chloride (5 mL) was added to a mixture of thiourea (0.152 g, 2 mmol) in acetonitrile (5 mL). The mixture was stirred at room temperature overnight (20 h). The formation of white precipitate was observed. Precipitated product was filtered, washed with cold hexane and dried in vacuum, giving bis-isothiouronium salt (0.475 g, 95% yield) as a white powder; mp 219–220 °C. Spectral characteristics and elemental analysis data were reported [50].
77Se NMR (76.3 MHz, CDCl3): 382.2.
2,6-Bis(vinylsulfanyl)-9-selenabicyclo[3.3.1]nonane (22). A solution of sodium hydroxide (80%, 1 g, 20 mmol) and sodium borohydride (0.38 g, 10 mmol) in ethanol (10 mL) was added dropwise to a solution of bis-isothiouronium salt (1 g, 2 mmol) in ethanol (20 mL). The mixture was heated at in a 1 L steel rotating autoclave at temperature 110–120 C for 5 h. Methylene chloride (20 mL) and cold water (120 mL) were added to the reaction mixture. The mixture was transferred to a separatory funnel and the organic layer was separated. The mixture was additionally extracted with methylene chloride (2 × 20 mL), the organic phase was dried over Na2SO4 and the solvent was removed by a rotary evaporator. The residue was subjected to column chromatography on silica gel (eluent: hexane → hexane/chloroform 15:1 → hexane/chloroform 2:1) giving the product (0.495 g, 81% yield).
1H NMR (400 MHz, CDCl3): 1.90–2.02 (m, 2H, CH2), 2.08–2.15 (m, 2H,CH2), 2.26–2.36 (m, 2H, CH2), 2.74–2.80 (m, 2H, CH2), 3.10–3.12 (m, 2H, CHSe), 3.94–4.00 (m, 2H, CHS), 5.22–5.29 (m, 4H, CH2=CHS), 6.30–6.38 (m, 2H, CH2=CHS).
13C NMR (100 MHz, CDCl3): 29.1 (CH2) 29.3 (CHSe), 30.0 (CH2), 49.8 (CHS), 114.5 (CH2=CH), 131.1 (CH2=CH).
77Se NMR (76.3 MHz, CDCl3): 285.3.
Anal. calcd for C12H18S2Se (305.36): C 47.20, H 5.94, S 21.00, Se 25.86%. Found: C 47.32, H 5.99, S 21.12, Se 26.04%.
2,6-Bis(triphenylphosphonium)-9-selenabicyclo[3.3.1]nonane dibromide (23). A solution of compound 2 (0.348 g, 1 mmol) in methylene chloride (5 mL) was added to
A solution of triphenyl phosphine (0.525 g, 2 mmol) in acetonitrile (5 mL) was added to a mixture of compound 2 (0.348 g, 1 mmol) in acetonitrile (5 mL). The mixture was refluxed for 8 h. The formation of white precipitate was observed. The precipitated product was filtered, washed with cold hexane and dried in vacuum, giving the product (0.837 g, 96% yield) as a white powder; mp 216–218 °C.
1H NMR (400 MHz, CDCl3): 1.43–1.55 (m, 2H, CH2), 1.82–1.91 (m, 4H, CH2), 2.60–2.71 (m, 2H, CH2), 3.59–3.66 (m, 2H, CHSe), 4.88–4.98 (m, 2H, CHP), 7.57–7.71 (m, 30H, CHAr).
13C NMR (100 MHz,CDCl3): 22.9 (CH2), 24.8 (CH2), 29.6 (CHSe), 29.7 (CHSe), 38.4 (CHP), 38.8 (CHP), 115.7 (CAr), 116.6 (CAr), 130.7 (CHAr), 130.9 (CHAr), 133.8 (CHAr), 133.9 (CHAr), 135.1 (CHAr), 135.2 (CHAr). 31P NMR (100 MHz, CDCl3): 23.86.
77Se NMR (76.3 MHz, CDCl3): 521.4.
MS (EI): m/z (%) = 384 (2, M+), 229 (16), 202 (18), 186 (20), 120 (36), 105 (68), 49 (41), 93 (70), 79 (68), 67 (78), 57 (73), 41 (100).
IR (KBr): λ = 521, 693, 1104, 1436, 1480, 2895, 2992, 3038 cm−1.
Anal. calcd for C44H42P2Br2Se (895.54): C 61.69, H 4.73, P 6.92, Br 17.84, Se 8.82%. Found: C 61.79, H 4.69, P 6.98, Br 17.99, Se 8.94%.

3.7. Synthesis of Selenoxides

2,6-Dipyridinium-9-selenobicyclo[3.3.1]nonane-9-oxide dibromide (24). A solution of tert-butyl hydroperoxide (70%, 2 mmol) was added dropwise to 0.505 g (1 mmol) 2,6-dipyridinium 9-selenobicyclo[3.3.1]nonane dibromide 3. The reaction mixture is stirred for 2 h at room temperature. The reaction mixture was washed with acetonitrile (10 mL), the precipitate was filtered off and dried under vacuum. The product was isolated as a light-yellow powder (0.49 g, 94% yield), mp 96–98 °C (decomp.).
1H NMR (400 MHz, CDCl3) δ 2.41–2.59 (m, 4H, OCHCH2, SeCHCH2), 2.66–2.79 (m, 2H, OCHCH2, SeCHCH2), 3.27–3.46 (m, 2H, SeCHCH2), 3.95–4.01 (m, 2H, SeCH), 5.69–5.76 (m, 2H, NCHCH2), 8.27–8.32 (m, 4H, CHAr), 8.72–8.78 (m, 2H, CHAr), 9.18–9.23 (m, 4H, CHAr).
13C NMR (100 MHz, CDCl3) δ 18.75, 22.02, 25.46, 26.15, 47.41, 49.36, 66.94, 70.57, 130.17, 144.56, 144.94, 148.16, 148.35.
77Se NMR (76.3 MHz, CDCl3): 851.6.
Found: C, 41.27; H, 4.36; N, 5.66; Br, 31.03; Se, 15.44. Calc. for C18H22N2Br2OSe: C, 41.48; H, 4.25; N, 5.38; Br, 30.66; O, 3.07; Se 15.15.
2,6-Dihydroxy-9-selenobicyclo[3.3.1]nonane-9-oxide (25). A solution (0.25 mL) of tert-butyl peroxide (70%, 2 mmol) was added dropwise to a solution of 0.221 g (1 mmol) of 2,6-hydroxy-9-selenobicyclo[3.3.1]nonane 4 in methylene chloride (10 mL). The reaction mixture was stirred for 12 h at 0 °C. The mixture was washed with water (5 × 10 mL), dry with CaCl2, and the solvent was removed on a rotary evaporator. The residue was dried in vacuum. The product was isolated as a white powder (0.218 g, 92% yield), mp 86–88 °C (decomp.).
1H NMR (400 MHz, CDCl3) δ 1.56–1.64 (m, 1H, SeCHCH2), 1.72–1.87 (m, 4H, OCHCH2, SeCHCH2) 2.14–2.17 (m, 2H, SeCHCH2), 2.31–2.37 (m, 1H, OCHCH2), 2.96–2.98 (m, 1H, SeCH), 2.99–3.01 (m, 1H, SeCH), 3.92 (s, 1H, OH), 4.22 (s, 1H, OH), 4.89–4.91 (m, 1H, OCH), 5.24–5.26 (m, 1H, OCH).
13C NMR (100 MHz, CDCl3) δ 16.53, 19.23, 29.56, 30.09, 46.66, 50.15, 62.44, 67.92.
77Se NMR (76.3 MHz, CDCl3): 841.5.
Found: C, 40.64; H, 6.01; Se, 32.79. Calc. for C8H12O3Se: C, 40.52; H, 5.95; O, 20.24; Se 33.30.

4. Conclusions

A set of highly efficient syntheses of novel derivatives of 9-selenabicyclo[3.3.1]nonane in high yields based on selenium dibromide and cis,cis-1,5-cyclooctadiene were developed. Various oxygen-centered nucleophiles were involved in the selenenylation/bis-oxylation reactions including alkanols, benzyl, allyl, and propargyl alcohols, and phenols. The copper-catalyzed 1,3-dipolar cycloaddition of 2,6-diazido-9-selenabicyclo[3.3.1]nonane with unsubstituted gaseous acetylene and propargyl alcohol was used for the preparation of novel 1,2,3-triazole derivatives of selenabicyclo[3.3.1]nonane.
Bis-isothiuronium salt was obtained in 95% yield, at room temperature, at a stoichiometric ratio of compound 2 and thiourea. This salt was used for the generation of the corresponding dithiolate anion under the action of sodium hydroxide, followed by the nucleophilic addition of the dithiolate anion to unsubstituted acetylene with the formation of bis(vinylsulfanyl) derivative of 9-selenabicyclo[3.3.1]nonane. The synthesis of water-soluble bis-phosphonium salt in quantitative yield was developed from dibromo derivative 2 and triphenyl phosphine.
The obtained water-soluble products were used for the estimation of glutathione peroxidase-like activity. It was found that diazido derivative 17 is considerably superior to other products in activity. The second most active product is bis-pyridinium salt 3, which is considered a promising drug for metabolic correction during vaccination process [61]. The selenoxides 24 and 25, which are supposed to be the catalytic cycle intermediates, were synthesized by the oxidation of compounds 3 and 4 with tert-butyl hydroperoxide.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232415629/s1.

Author Contributions

Conceptualization, M.V.M. and V.A.P.; methodology, M.V.M. and V.A.P.; validation, M.V.M. and V.A.P.; formal analysis, M.V.M. and V.A.P.; investigation, M.V.M.; data curation, V.A.P.; writing—original draft preparation, M.V.M.; writing—review and editing, V.A.P. 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

The authors thank Baikal Analytical Center SB RAS for providing the instrumental equipment for structural investigations. We are grateful to Svetlana A. Zhivet’eva and Tatyana I. Yaroshenko for experimental assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schwarz, K.; Foltz, C.M. Selenium as an integral part of factor 3 against dietary necrotic liver degeneration. J. Am. Chem. Soc. 1957, 79, 3292–3293. [Google Scholar] [CrossRef]
  2. Nicolaou, K.C.; Petasi, N.A. Selenium in Natural Products Synthesis; CIS: Philadelphia, PA, USA, 1984; 300p. [Google Scholar]
  3. Wirth, T. (Ed.) Organoselenium Chemistry: Synthesis and Reactions; Wiley-VCH: Weinheim, Germany, 2012; 462p. [Google Scholar]
  4. Wirth, T. Organoselenium Chemistry—Modern Developments in Organic Synthesis, Topics in Current Chemistry, 208; Spring: Heidelberg, Germany, 2000; 260p. [Google Scholar]
  5. Paulmier, C. Selenium Reagents and Intermediates in Organic Synthesis; Pergamon: Oxford, UK, 1986; 463p. [Google Scholar]
  6. Back, T.G. Organoselenium Chemistry: A Practical Approach; Oxford University Press: Oxford, UK, 1999; 295p. [Google Scholar]
  7. 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]
  8. Santi, C. (Ed.) Organoselenium Chemistry: Between Synthesis and Biochemistry; Bentham Science Publishers: Sharjah, United Arab Emirates, 2014; p. 563. [Google Scholar]
  9. Tiekink, E.R.T. Therapeutic potential of selenium and tellurium compounds: Opportunities yet unrealized. Dalton Trans. 2012, 41, 6390–6395. [Google Scholar] [CrossRef] [PubMed]
  10. Mugesh, G.; du Mont, W.W.; Sies, H. Chemistry of biologically important synthetic organoselenium compounds. Chem. Rev. 2001, 101, 2125–2179. [Google Scholar] [CrossRef] [PubMed]
  11. Braga, A.L.; Rafique, J. Synthesis of biologically relevant small molecules containing selenium. Part B. Anti-infective and anticancer compounds. In Patai’s Chemistry of Functional Groups. Organic Selenium and Tellurium Compounds; Rappoport, Z., Ed.; John Wiley and Sons: Chichester, UK, 2013; Volume 4, pp. 1053–1117. [Google Scholar]
  12. Al-Rubaie, A.Z.; Al-Jadaan, S.A.S.; Muslim, S.K.; Saeed, E.A.; Ali, E.T.; Al-Hasani, A.K.J.; Al-Salman, H.N.K.; Al-Fadal, S.A.M. Synthesis, characterization and antibacterial activity of some new ferrocenyl selenazoles and 3,5-diferrocenyl-1,2,4-selenadiazole. J. Organomet. Chem. 2014, 774, 43–47. [Google Scholar] [CrossRef]
  13. Dhau, J.S.; Singh, A.; Singh, A.; Sooch, B.S.; Brandão, P.; Félix, V. Synthesis and antibacterial activity of pyridylselenium compounds: Self-assembly of bis(3-bromo-2-pyridyl)diselenide via intermolecular secondary and π⋯π stacking interactions. J. Organomet. Chem. 2014, 766, 57–66. [Google Scholar] [CrossRef]
  14. Angeli, A.; Tanini, D.; Capperucci, A.; Supuran, C.T. Synthesis of novel selenides bearing benzenesulfonamide moieties as carbonic anhydrase I, II, IV, VII, and IX inhibitors. ASC Med. Chem. Lett. 2017, 8, 1213–1217. [Google Scholar] [CrossRef]
  15. Banerjee, B.; Koketsy, M. Recent developments in the synthesis of biologically relevant selenium-containing scaffolds. Coord. Chem. Rev. 2017, 339, 104–127. [Google Scholar] [CrossRef]
  16. Elsherbini, M.; Hamama, W.S.; Zoorob, H.H. Recent advances in the chemistry of selenium-containing heterocycles: Five-membered ring systems. Coord. Chem. Rev. 2016, 312, 149–177. [Google Scholar] [CrossRef]
  17. Ninomiya, M.; Garud, D.R.; Koketsu, M. Biologically significant selenium-containing heterocycles. Coord. Chem. Rev. 2011, 255, 2968–2990. [Google Scholar] [CrossRef]
  18. Sonawane, A.D.; Sonawane, R.A.; Ninomiya, M.N.; Koketsu, M. Synthesis of seleno-heterocycles via electrophilic/radical cyclization of alkyne containing heteroatoms. Adv. Synth. Catal. 2020, 362, 3485–3515. [Google Scholar] [CrossRef]
  19. Ruberte, A.C.; Sanmartin, C.; Aydillo, C.; Sharma, A.K.; Plano, D. Development and Therapeutic Potential of Selenazo Compounds. J. Med. Chem. 2020, 63, 1473–1489. [Google Scholar] [CrossRef] [PubMed]
  20. Yu, S.-C.; Kuhn, H.; Daniliuc, C.-G.; Ivanov, I.; Jones, P.G.; du Mont, W.-W. 5-Selenization of salicylic acid derivatives yielded isoform-specific 5-lipoxygenase inhibitors. Org. Biomol. Chem. 2010, 8, 828–834. [Google Scholar] [CrossRef] [PubMed]
  21. Koketsu, M.; Yang, H.; Kim, Y.M.; Ichihash, M.; Ishihara, H. Preparation of 1,4-Oxaselenin from AgNO3/LDA-Assisted Reaction of 3-Selena-4-pentyn-1-one as Potential Antitumor Agents. Org. Lett. 2001, 3, 1705–1707. [Google Scholar] [CrossRef] [PubMed]
  22. Jain, V.K.; Priyadarsini, K.I. (Eds.) Organoselenium Compounds in Biology and Medicine: Synthesis, Biological and Therapeutic Treatments; RSC: London, UK, 2017; 458p. [Google Scholar]
  23. Azad, G.K.; Tomar, R.S. Ebselen, a promising antioxidant drug: Mechanisms of action and targets of biological pathways. Mol. Biol. Rep. 2014, 41, 4865–4879. [Google Scholar] [CrossRef]
  24. Jin, Z.; Du, X.; Xu, Y.; 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] [Green Version]
  25. Weglarz-Tomczak, E.; Tomczak, J.M.; Talma, M.; Burda-Grabowska, M.; Giurg, M.; Brul, S. Identification of ebselen and its analogues as potent covalent inhibitors of papain-like protease from SARS-CoV-2. Sci. Rep. 2021, 11, 3640. [Google Scholar] [CrossRef]
  26. Selvakumar, K.; Shah, P.; Singh, H.B.; Butcher, R.J. Synthesis, Structure, and Glutathione Peroxidase-Like Activity of Amino Acid Containing Ebselen Analogues and Diaryl Diselenides. Chem.-Eur. J. 2011, 17, 12741–12755. [Google Scholar] [CrossRef]
  27. Sarma, B.K.; Mugesh, G. Glutathione Peroxidase (GPx)-like Antioxidant Activity of the Organoselenium Drug Ebselen:  Unexpected Complications with Thiol Exchange Reactions. J. Am. Chem. Soc. 2005, 127, 11477–11485. [Google Scholar] [CrossRef]
  28. Kumakura, F.; Mishra, B.; Priyadarsini, K.I.; Iwaoka, M. A Water-Soluble Cyclic Selenide with Enhanced Glutathione Peroxidase-Like Catalytic Activities. Eur. J. Org. Chem. 2010, 2010, 440–445. [Google Scholar] [CrossRef]
  29. Back, T.G.; Dyck, B.P. A Novel Camphor-Derived Selenenamide That Acts as a Glutathione Peroxidase Mimetic. J. Am. Chem. Soc. 1997, 119, 2079–2083. [Google Scholar] [CrossRef]
  30. Back, T.G.; Moussa, Z. Remarkable Activity of a Novel Cyclic Seleninate Ester as a Glutathione Peroxidase Mimetic and Its Facile in Situ Generation from Allyl 3-Hydroxypropyl. J. Am. Chem. Soc. 2002, 124, 12104–12105. [Google Scholar] [CrossRef] [PubMed]
  31. Back, T.G.; Moussa, Z. Diselenides and Allyl Selenides as Glutathione Peroxidase Mimetics. Remarkable Activity of Cyclic Seleninates Produced in Situ by the Oxidation of Allyl ω-Hydroxyalkyl Selenides. J. Am. Chem. Soc. 2003, 125, 13455–13460. [Google Scholar] [CrossRef] [PubMed]
  32. Braverman, S.; Cherkinsky, M.; Kalendar, Y.; Jana, R.; Sprecher, M.; Goldberg, I. Synthesis of water-soluble vinyl selenides and their high glutathione peroxidase (GPx)-like antioxidant activity. Synthesis 2014, 46, 119–125. [Google Scholar] [CrossRef] [Green Version]
  33. McNeil, N.M.R.; Press, D.J.; Mayder, D.M.; Garnica, P.; Doyle, L.M.; Back, T.G. Enhanced Glutathione Peroxidase Activity of Water-Soluble and Polyethylene Glycol-Supported Selenides, Related Spirodioxyselenuranes, and Pincer Selenuranes. J. Org. Chem. 2016, 81, 7884–7897. [Google Scholar] [CrossRef]
  34. Flohe, L.; Gunzler, W.A.; Schock, H.H. Glutathione peroxidase: A selenoenzyme. FEBS Lett. 1973, 32, 132–134. [Google Scholar] [CrossRef] [Green Version]
  35. Rotruck, J.T.; Pope, A.L.; Ganther, H.E.; Swanson, A.B.; Hafeman, D.G.; Hoekstra, W.G. Selenium: Biochemical role as a component of glutathione peroxidase. Science 1973, 179, 588–590. [Google Scholar] [CrossRef]
  36. Gladyshev, V.N.; Hatfield, D.L. Selenocysteine-Containing Proteins in Mammals. J. Biomed. Sci. 1999, 6, 151–160. [Google Scholar] [CrossRef]
  37. Dhalla, N.S.; Elmoselhi, A.B.; Hata, T.; Makino, N. Status of myocardial antioxidants in ischemia-reperfusion injury. Cardiovasc. Res. 2000, 47, 446–456. [Google Scholar] [CrossRef]
  38. Bjørklund, G.; Shanaida, M.; Lysiuk, R.; Antonyak, H.; Klishch, I.; Shanaida, V.; Peana, M. Selenium: An Antioxidant with a Critical Role in Anti-Aging. Molecules 2022, 27, 6613. [Google Scholar] [CrossRef]
  39. Cai, Z.; Zhang, J.; Li, H. Selenium, aging and aging-related diseases. Aging Clin. Exp. Res. 2019, 31, 1035–1047. [Google Scholar] [CrossRef] [PubMed]
  40. Alehagen, U.; Opstad, T.B.; Alexander, J.; Larsson, A.; Aaseth, J. Impact of Selenium on Biomarkers and Clinical Aspects Related to Ageing. A Review. Biomolecules 2021, 11, 1478. [Google Scholar] [CrossRef] [PubMed]
  41. Artem’ev, A.V.; Gusarova, N.K.; Malysheva, S.F.; Kraikivskii, P.B.; Belogorlova, N.A.; Trofimov, B.A. Efficient General Synthesis of Alkylammonium Diselenophosphinates via Multicomponent One-Pot Reaction of Secondary Phosphines with Elemental Selenium and Amines. Synthesis 2010, 21, 3724–3730. [Google Scholar] [CrossRef]
  42. Jimoh, Y.A.; LawalIge, A.O.; Kade, I.J.; Olatunde, D.M.; Oluwayomi, O. Diphenyl diselenide modulates antioxidant status, inflammatory and redox-sensitive genes in diesel exhaust particle-induced neurotoxicity. Chem.-Biol. Interact. 2022, 367, 110196. [Google Scholar] [CrossRef] [PubMed]
  43. Lorenzoni, S.; Cerra, S.; Angulo-Elizari, E.; Salamone, T.A.; Battocchio, C.; Marsotto, M.; Scaramuzzo, F.A.; Sanmartín, C.; Plano, D.; Fratoddi, I. Organoselenium compounds as functionalizing agents for gold nanoparticles in cancer therapy. Colloids Surf. B Biointerfaces 2022, 2019, 112828. [Google Scholar] [CrossRef] [PubMed]
  44. Malysheva, S.F.; Gusarova, N.K.; Artem’ev, A.V.; Belogorlova, N.A.; Albanov, A.I.; Borodina, T.N.; Smirnov, V.I.; Trofimov, B.A. Facile Non-Catalyzed Synthesis of Tertiary Phosphine Sulfides by Regioselective Addition of Secondary Phosphine Sulfides to Alkenes. Eur. J. Org. Chem. 2014, 2014, 2516–2521. [Google Scholar] [CrossRef]
  45. Chuai, H.; Zhang, S.-Q.; Bai, H.; Li, J.; Wang, Y.; Sun, J.; Wen, E.; Zhang, J.; Xin, M. Small molecule selenium-containing compounds: Recent development and therapeutic applications. Eur. J. Med. Chem. 2021, 223, 113621. [Google Scholar] [CrossRef]
  46. Sonawane, A.D.; Sonawane, A.R.; Ninomiya, M.; Koketsu, M. Diorganyl diselenides: A powerful tool for the construction of selenium containing scaffolds. Dalton Trans. 2021, 50, 12764–12790. [Google Scholar] [CrossRef]
  47. Trost, B.M.; Tang, W.; Toste, F.D. Divergent Enantioselective Synthesis of (−)-Galanthamine and (−)-Morphine. J. Am. Chem. Soc. 2005, 127, 14785–14803. [Google Scholar] [CrossRef]
  48. Shin, H.S.; Jung, Y.G.; Cho, H.K.; Park, Y.G.; Cho, C.G. Total Synthesis of (±)-Lycorine from the Endo-Cycloadduct of 3,5-Dibromo-2-pyrone and (E)-β-Borylstyrene. Org. Lett. 2014, 16, 5718–5720. [Google Scholar] [CrossRef]
  49. Gao, Y.; Wei, Y.; Ma, D. Synthetic Studies toward Plumisclerin A. Org. Lett. 2019, 21, 1384–1387. [Google Scholar] [CrossRef] [PubMed]
  50. Wirth, T.; Häuptli, S.; Leuenberger, M. Catalytic asymmetric oxyselenenylation–elimination reactions using chiral selenium compounds. Tetrahedron Asymmetry 1998, 9, 547–550. [Google Scholar] [CrossRef]
  51. Potapov, V.A.; Amosova, S.V. New Methods for Preparation of Organoselenium and Organotellurium Compounds from Elemental Chalcogens. Russ. J. Org. Chem. 2003, 39, 1373–1380. [Google Scholar] [CrossRef]
  52. Potapov, V.A.; Musalov, M.V.; Musalova, M.V.; Amosova, S.V. Recent Advances in Organochalcogen Synthesis Based on Reactions of Chalcogen Halides with Alkynes and Alkenes. Curr. Org. Chem. 2016, 20, 136–145. [Google Scholar] [CrossRef]
  53. Musalov, M.V.; Potapov, V.A. Selenium dihalides: New possibilities for the synthesis of selenium-containing heterocycles. Chem. Heterocycl. Comp. 2017, 53, 150–152. [Google Scholar] [CrossRef]
  54. Abakumov, G.A.; Piskunov, A.V.; Cherkasov, V.K.; Fedushkin, I.L.; Ananikov, V.P.; Eremin, D.B.; Gordeev, E.G.; Beletskaya, I.P.; Averin, A.D.; Bochkarev, M.N.; et al. Organoelement chemistry: Promising growth areas and challenges. Russ. Chem. Rev. 2018, 87, 393–507. [Google Scholar] [CrossRef]
  55. Musalov, M.V.; Yakimov, V.A.; Potapov, V.A.; Amosova, S.V.; Borodina, T.N.; Zinchenko, S.V. A novel methodology for the synthesis of condensed selenium heterocycles based on the annulation and annulation–methoxylation reactions of selenium dihalides. New J. Chem. 2019, 43, 18476–18483. [Google Scholar] [CrossRef]
  56. Accurso, A.A.; Cho, S.-H.; Amin, A.; Potapov, V.A.; Amosova, S.V.; Finn, M.G. Thia-, Aza-, and Selena[3.3.1]bicyclononane Dichlorides: Rates vs Internal Nucleophile in Anchimeric Assistance. J. Org. Chem. 2011, 76, 4392–4395. [Google Scholar] [CrossRef]
  57. Potapov, V.A.; Amosova, S.V.; Abramova, E.V.; Musalov, M.V.; Lyssenko, K.A.; Finn, M.G. 2,6-Dihalo-9-selenabicyclo[3.3.1]nonanes and their complexes with selenium dihalides: Synthesis and structural characterization. New J. Chem. 2015, 39, 8055–8059. [Google Scholar] [CrossRef]
  58. Abramova, E.V.; Sterkhova, I.V.; Molokeev, M.S.; Potapov, V.A.; Amosova, S.V. First coordination compounds of SeBr2 with selenium ligands: X-ray structural determination. Mendeleev Commun. 2016, 26, 532–534. [Google Scholar] [CrossRef]
  59. Potapov, V.A.; Musalov, M.V. Triple-Click Chemistry of Selenium Dihalides: Catalytic Regioselective and Highly Efficient Synthesis of Bis-1,2,3-Triazole Derivatives of 9-Selenabicyclo[3.3.1]nonane. Catalysts 2022, 12, 1032. [Google Scholar] [CrossRef]
  60. Musalov, M.V.; Potapov, V.A.; Amosova, S.V. Efficient Synthesis of a New Family of 2,6-Disulfanyl-9-Selenabicyclo[3.3.1]Nonanes. Molecules 2021, 26, 2849. [Google Scholar] [CrossRef] [PubMed]
  61. Yurieva, O.V.; Dubrovina, V.I.; Potapov, V.A.; Musalov, M.V.; Starovoitova, T.P.; Ivanova, T.A.; Gromova, A.V.; Shkaruba, T.T.; Balakhonov, S.V. Effect of Synthetic Organoselenium Drug on the Degree of Pathological Changes in the Organs of White Mice Immunized with Tularemia and Brucellosis Vaccines. Bull. Exp. Biol. Med. 2019, 168, 66–69. [Google Scholar] [CrossRef] [PubMed]
  62. Kolb, H.C.; Finn, M.G.; Sharpless, K.B. Click chemistry: Diverse chemical function from a few good reactions. Angew. Chem. Int. Ed. 2001, 40, 2004–2021. [Google Scholar] [CrossRef]
  63. Bozorov, K.; Zhao, J.; Aisa, H.A. 1,2,3-Triazole-containing hybrids as leads in medicinal chemistry: A recent overview. Bioorg. Med. Chem. 2019, 27, 3511–3531. [Google Scholar] [CrossRef]
  64. Jain, A.; Piplani, P. Exploring the Chemistry and Therapeutic Potential of Triazoles: A Comprehensive Literature Review. Mini Rev. Med. Chem. 2019, 19, 1298–1368. [Google Scholar] [CrossRef]
  65. Xu, M.; Peng, Y.; Zhu, L.; Wang, S.; Ji, J.; Rakesh, K.P. Triazole derivatives as inhibitors of Alzheimer’s disease: Current developments and structure-activity relationships. Eur. J. Med. Chem. 2019, 180, 656–672. [Google Scholar] [CrossRef] [PubMed]
  66. Agalave, S.G.; Maujan, S.R.; Pore, V.S. Click Chemistry: 1,2,3-Triazoles as Pharmacophores. Chem. Asian J. 2011, 6, 2696–2718. [Google Scholar] [CrossRef]
  67. Kashyap, S.J.; Garg, V.K.; Sharma, P.K.; Kumar, N.; Dudhe, R.; Gupta, J.K. Thiazoles: Having diverse biological activities. Med. Chem. Res. 2012, 21, 2123–2132. [Google Scholar] [CrossRef]
  68. Fabbrizzi, P.; Menchi, G.; Guarna, A.; Trabocchi, A. Use of Click-Chemistry in the Development of Peptidomimetic Enzyme Inhibitors. Curr. Med. Chem. 2014, 21, 1467–1477. [Google Scholar] [CrossRef]
  69. Bonandi, E.; Fumagalli, G.; Perdicchia, D.; Christodoulou, M.S.; Rastelli, G.; Passarella, D. The 1,2,3-triazole ring as a bioisostere in medicinal chemistry. Drug Discov. Today 2017, 22, 1572–1581. [Google Scholar] [CrossRef] [PubMed]
  70. Doiron, J.E.; Ody, B.K.; Brace, J.B.; Post, S.J.; Thacker, N.L.; Hill, H.M.; Breton, G.W.; Turlington, M.; Le Christina, A.; Aller, S.G.; et al. Evaluation of 1,2,3-Triazoles as Amide Bioisosteres in Cystic Fibrosis Transmembrane Conductance Regulator Modulators VX-770 and VX-809. Chem. Eur. J. 2019, 25, 3662–3674. [Google Scholar] [CrossRef] [PubMed]
  71. Kharb, R.; Sharma, P.C.; Yar, M.S. Pharmacological significance of triazole scaffold. J. Enzym. Inhib. Med. Chem. 2011, 26, 1–21. [Google Scholar] [CrossRef] [PubMed]
  72. Ferreira, S.B.; Sodero, A.C.; Cardoso, M.F.; Lima, E.S.; Kaiser, C.R.; Silva, F.P.; Ferreira, V.F. Synthesis, Biological Activity, and Molecular Modeling Studies of 1H-1,2,3-Triazole Derivatives of Carbohydrates as α-Glucosidases Inhibitors. J. Med. Chem. 2010, 53, 2364–2375. [Google Scholar] [CrossRef]
  73. Devender, N.; Gunjan, S.; Chhabra, S.; Singh, K.; Pasam, V.R.; Shukla, S.K.; Sharma, A.; Jaiswal, S.; Singh, S.K.; Kumar, Y.; et al. Identification of β-Amino alcohol grafted 1,4,5 trisubstituted 1,2,3-triazoles as potent antimalarial agents. Eur. J. Med. Chem. 2016, 109, 187–198. [Google Scholar] [CrossRef]
  74. Lee, T.; Cho, M.; Ko, S.-Y.; Youn, H.-J.; Baek, D.J.; Cho, W.-J.; Kang, C.-Y.; Kim, S. Synthesis and Evaluation of 1,2,3-Triazole Containing Analogues of the Immunostimulant α-GalCer. J. Med. Chem. 2007, 50, 585–589. [Google Scholar] [CrossRef] [PubMed]
  75. Singh, P.; Raj, R.; Kumar, V.; Mahajan, M.P.; Bedi, P.M.; Kaur, T.; Saxena, A.K. 1,2,3-Triazole tethered β-lactam-Chalcone bifunctional hybrids: Synthesis and anticancer evaluation. Eur. J. Med. Chem. 2012, 47, 594–600. [Google Scholar] [CrossRef]
  76. Ganesh, A. Potential biological activity of 1,4-sustituted-1H-[1,2,3]triazoles. Int. J. Chem. Sci. 2013, 11, 573–578. [Google Scholar]
  77. Dheer, D.; Singh, V.; Shankar, R. Medicinal attributes of 1,2,3-triazoles: Current developments. Bioorg. Chem. 2017, 71, 30–54. [Google Scholar] [CrossRef]
  78. Dhall, E.; Sain, S.; Jain, S.; Dwivedi, J. Synthesis of Triazole Derivatives Manifesting Antimicrobial and Anti-Tubercular Activities. Mini-Rev. Org. Chem. 2018, 15, 291–314. [Google Scholar] [CrossRef]
  79. Lauria, A.; Delisi, R.; Mingoia, F.; Terenzi, A.; Martorana, A.; Barone, G.; Almerico, A.M. 1,2,3-Triazole in Heterocyclic Compounds, Endowed with Biological Activity, through 1,3-Dipolar Cycloadditions. Eur. J. Org. Chem. 2014, 2014, 3289–3306. [Google Scholar] [CrossRef]
  80. Amblard, F.; Cho, J.H.; Schinazi, R.F. Cu(I)-catalyzed Huisgen azide-alkyne 1,3-dipolar cycloaddition reaction in nucleoside, nucleotide, and oligonucleotide chemistry. Chem. Rev. 2009, 109, 4207–4220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Librando, I.L.; Mahmoud, A.G.; Carabineiro, S.A.C.; Guedes da Silva, M.F.C.; Geraldes, C.F.G.C.; Pombeiro, A.J.L. The Catalytic Activity of Carbon-Supported Cu(I)-Phosphine Complexes for the Microwave-Assisted Synthesis of 1,2,3-Triazoles. Catalysts 2021, 11, 185. [Google Scholar] [CrossRef]
  82. Pucci, A.; Albano, G.; Pollastrini, M.; Lucci, A.; Colalillo, M.; Oliva, F.; Evangelisti, C.; Marelli, M.; Santalucia, D.; Mandoli, A. Supported Tris-Triazole Ligands for Batch and Continuous-Flow Copper-Catalyzed Huisgen 1,3-Dipolar Cycloaddition Reactions. Catalysts 2020, 10, 434. [Google Scholar] [CrossRef]
  83. Himo, F.; Lovell, T.; Hilgraf, R.; Rostovtsev, V.V.; Noodleman, L.; Sharpless, K.B.; Fokin, V.V. Copper(I)-Catalyzed Synthesis of Azoles. DFT Study Predicts Unprecedented Reactivity and Intermediates. J. Am. Chem. Soc. 2005, 127, 210–216. [Google Scholar] [CrossRef]
  84. Liang, L.; Astruc, D. The copper(I)-catalyzed alkyne-azide cycloaddition (CuAAC) “click” reaction and its applications. An overview. Coord. Chem. Rev. 2011, 255, 2933–2945. [Google Scholar] [CrossRef]
  85. Haldón, E.; Nicasio, M.C.; Pérez, P.J. Copper-catalysed azide–alkyne cycloadditions (CuAAC): An update. Org. Biomol. Chem. 2015, 13, 9528–9550. [Google Scholar] [CrossRef]
  86. Hein, J.E.; Fokin, V.V. Copper-catalyzed azide-alkyne cycloaddition (CuAAC) and beyond: New reactivity of copper(I) acetylides. Chem. Soc. Rev. 2010, 39, 1302–1315. [Google Scholar] [CrossRef]
  87. De Nino, A.; Maiuolo, L.; Costanzo, P.; Algieri, V.; Jiritano, A.; Olivito, F.; Tallarida, M.A. Recent Progress in Catalytic Synthesis of 1,2,3-Triazoles. Catalysts 2021, 11, 1120. [Google Scholar] [CrossRef]
  88. Saini, P.; Sonika; Singh, G.; Kaur, G.; Singh, J.; Singh, H. Robust and Versatile Cu(I) metal frameworks as potential catalysts for azide-alkyne cycloaddition reactions. Mol. Catal. 2021, 504, 111432. [Google Scholar] [CrossRef]
  89. Kalra, P.; Kaur, R.; Singh, H.; Singh, G.; Pawan; Kaur, G.; Singh, J. Metals as “Click” catalysts for alkyne-azide cycloaddition reactions: An overview. J. Organomet. Chem. 2021, 944, 121846. [Google Scholar] [CrossRef]
  90. Deobald, A.M.; Camargo, L.R.S.; Hörner, M.; Rodrigues, O.E.D.; Alves, D.; Braga, A.L. Synthesis of arylseleno-1,2,3-triazoles via copper-catalyzed 1,3-dipolar cycloaddition of azido arylselenides with alkynes. Synthesis 2011, 43, 2397–2406. [Google Scholar]
  91. Seus, N.; Saraiva, M.T.; Alberto, E.E.; Savegnago, L.; Alves, D. Selenium compounds in click chemistry: Copper catalyzed 1,3-dipolar cycloaddition of azidomethyl arylselenides and alkynes. Tetrahedron 2012, 68, 10419–10425. [Google Scholar] [CrossRef]
  92. Bhabak, K.P.; Mugesh, G. Functional Mimics of Glutathione Peroxidase: Bioinspired Synthetic Antioxidants. Accounts Chem. Res. 2010, 43, 1408–1419. [Google Scholar] [CrossRef] [PubMed]
  93. Potapov, V.A.; Amosova, S.V.; Petrov, P.A.; Romanenko, L.S.; Keiko, V.V. Exchange reactions of dialkyl dichalcogenides. Sulfur Lett. 1992, 15, 121–126. [Google Scholar]
Figure 1. Examples of known functionalized organoselenium compounds with glutathione peroxidase-like activity including selenides with the hydroxyl groups [26,27,28,29,30,31,32,33].
Figure 1. Examples of known functionalized organoselenium compounds with glutathione peroxidase-like activity including selenides with the hydroxyl groups [26,27,28,29,30,31,32,33].
Ijms 23 15629 g001
Scheme 1. Synthesis of bis-pyridinium salt 3 by the reaction of 2,6-dibromo-9-selenabicyclo[3.3.1]nonane 2 with pyridine.
Scheme 1. Synthesis of bis-pyridinium salt 3 by the reaction of 2,6-dibromo-9-selenabicyclo[3.3.1]nonane 2 with pyridine.
Ijms 23 15629 sch001
Scheme 2. The synthesis of dihydroxy derivative of 9-selenabicyclo[3.3.1]nonane 4.
Scheme 2. The synthesis of dihydroxy derivative of 9-selenabicyclo[3.3.1]nonane 4.
Ijms 23 15629 sch002
Scheme 3. The synthesis of 2,6-dialkoxy-9-selenabicyclo[3.3.1]nonane 5–9.
Scheme 3. The synthesis of 2,6-dialkoxy-9-selenabicyclo[3.3.1]nonane 5–9.
Ijms 23 15629 sch003
Scheme 4. The synthesis of allyloxy and propargyloxy derivatives 10 and 11.
Scheme 4. The synthesis of allyloxy and propargyloxy derivatives 10 and 11.
Ijms 23 15629 sch004
Scheme 5. The synthesis of benzyloxy derivatives 12 and 13.
Scheme 5. The synthesis of benzyloxy derivatives 12 and 13.
Ijms 23 15629 sch005
Scheme 6. The synthesis of bis(4-methoxyphenoxy) and bis(3,5-dimethylphenoxy) derivatives 14 and 15.
Scheme 6. The synthesis of bis(4-methoxyphenoxy) and bis(3,5-dimethylphenoxy) derivatives 14 and 15.
Ijms 23 15629 sch006
Scheme 7. The synthesis of bis(phenoxy) derivative 16 from compound 2 and phenol.
Scheme 7. The synthesis of bis(phenoxy) derivative 16 from compound 2 and phenol.
Ijms 23 15629 sch007
Scheme 8. The synthesis of 2,6-bis(1,2,3-triazol-1-yl)-9-selenabicyclo[3.3.1]nonane 18 by the copper-catalyzed azide-alkyne 1,3-dipolar cycloaddition reaction of acetylene with diazido derivative 17, which was obtained from compound 2 and sodium azide.
Scheme 8. The synthesis of 2,6-bis(1,2,3-triazol-1-yl)-9-selenabicyclo[3.3.1]nonane 18 by the copper-catalyzed azide-alkyne 1,3-dipolar cycloaddition reaction of acetylene with diazido derivative 17, which was obtained from compound 2 and sodium azide.
Ijms 23 15629 sch008
Scheme 9. The synthesis of compound 19 by the copper-catalyzed azide-alkyne 1,3-dipolar cycloaddition reaction of diazido derivative 17 with propagyl alcohol.
Scheme 9. The synthesis of compound 19 by the copper-catalyzed azide-alkyne 1,3-dipolar cycloaddition reaction of diazido derivative 17 with propagyl alcohol.
Ijms 23 15629 sch009
Scheme 10. The synthesis of bis(vinylsulfanyl) derivative 22 by the nucleophilic addition of dithiolate anion 21 generated from bis-isothiuronium salt 20, which was obtained from compound 2 and thiourea.
Scheme 10. The synthesis of bis(vinylsulfanyl) derivative 22 by the nucleophilic addition of dithiolate anion 21 generated from bis-isothiuronium salt 20, which was obtained from compound 2 and thiourea.
Ijms 23 15629 sch010
Scheme 11. The synthesis of bis-phosphonium salt 23 from compound 2 and triphenyl phosphine.
Scheme 11. The synthesis of bis-phosphonium salt 23 from compound 2 and triphenyl phosphine.
Ijms 23 15629 sch011
Scheme 12. The model reaction of dithiothreitol oxidation by tert-butyl hydroperoxide in D2O in the presence of synthesized compounds as catalysts (10% mol).
Scheme 12. The model reaction of dithiothreitol oxidation by tert-butyl hydroperoxide in D2O in the presence of synthesized compounds as catalysts (10% mol).
Ijms 23 15629 sch012
Figure 2. The evaluation of the glutathione peroxidase-like activity of the obtained water-soluble compounds 3, 4, 17, 20 and 23.
Figure 2. The evaluation of the glutathione peroxidase-like activity of the obtained water-soluble compounds 3, 4, 17, 20 and 23.
Ijms 23 15629 g002
Scheme 13. A supposed catalytic cycle with the regeneration of the catalyst.
Scheme 13. A supposed catalytic cycle with the regeneration of the catalyst.
Ijms 23 15629 sch013
Scheme 14. The synthesis of selenoxides 24 and 25 by oxidation of bis-pyridinium salt 3 in water and dihydroxyl derivatives 4 in methylene chloride with tert-butyl hydroperoxide.
Scheme 14. The synthesis of selenoxides 24 and 25 by oxidation of bis-pyridinium salt 3 in water and dihydroxyl derivatives 4 in methylene chloride with tert-butyl hydroperoxide.
Ijms 23 15629 sch014
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Musalov, M.V.; Potapov, V.A. Click Chemistry of Selenium Dihalides: Novel Bicyclic Organoselenium Compounds Based on Selenenylation/Bis-Functionalization Reactions and Evaluation of Glutathione Peroxidase-like Activity. Int. J. Mol. Sci. 2022, 23, 15629. https://doi.org/10.3390/ijms232415629

AMA Style

Musalov MV, Potapov VA. Click Chemistry of Selenium Dihalides: Novel Bicyclic Organoselenium Compounds Based on Selenenylation/Bis-Functionalization Reactions and Evaluation of Glutathione Peroxidase-like Activity. International Journal of Molecular Sciences. 2022; 23(24):15629. https://doi.org/10.3390/ijms232415629

Chicago/Turabian Style

Musalov, Maxim V., and Vladimir A. Potapov. 2022. "Click Chemistry of Selenium Dihalides: Novel Bicyclic Organoselenium Compounds Based on Selenenylation/Bis-Functionalization Reactions and Evaluation of Glutathione Peroxidase-like Activity" International Journal of Molecular Sciences 23, no. 24: 15629. https://doi.org/10.3390/ijms232415629

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