Next Article in Journal
Effect of Lacking ZKSCAN3 on Autophagy, Lysosomal Biogenesis and Senescence
Next Article in Special Issue
Injectable Hydrogel Guides Neurons Growth with Specific Directionality
Previous Article in Journal
Inflammaging and Immunosenescence as Part of Skin Aging—A Narrative Review
Previous Article in Special Issue
Water Decontamination from Cr(VI) by Transparent Silica Xerogel Monolith
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of Sulfur-Containing Functionalized Disiloxanes with Nonconventional Fluorescence by Thiol–Epoxy Click Reaction

1
Institute of Novel Semiconductors, State Key Laboratory of Crystal Materials, Shandong University, Jinan 250100, China
2
National Engineering Research Center for Colloidal Materials & Key Laboratory of Special Functional Aggregated Materials, Ministry of Education, Shandong Key Laboratory of Advanced Organosilicon Materials and Technologies, School of Chemistry and Chemical Engineering, Shandong University, Jinan 250100, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(9), 7785; https://doi.org/10.3390/ijms24097785
Submission received: 28 March 2023 / Revised: 22 April 2023 / Accepted: 23 April 2023 / Published: 24 April 2023

Abstract

:
Herein, a series of novel sulfur-containing functionalized disiloxanes based on a low-cost and commercially available material, i.e., 1,3-bis(3-glycidoxypropyl)-1,1,3,3-tetramethyldisiloxane, and various thiol compounds were prepared by thiol–epoxy click reaction. It was found that both lithium hydroxide (LiOH) and tetrabutylammonium fluoride (TBAF) have high catalytic activity after optimizing the reaction condition, and the reaction can be carried out with high yields, excellent regioselectivity, mild reaction condition, and good tolerance of functional groups. These compounds exhibit excellent nonconventional fluorescence due to the formation of coordination bonds between Si atoms and heteroatoms (e.g., S or N) and can emit blue fluorescence upon ultraviolet (UV) irradiation. These results demonstrate that the thiol–epoxy click reaction could promisingly act as an efficient organosilicon synthetic methodology to construct various organosilicon materials with novel structures and functionality, and thus their application scope will be significantly expanded.

Graphical Abstract

1. Introduction

In recent years, intense efforts have been devoted to developing new organosilicon materials because of their unique properties and indispensable roles in emerging areas, such as aerospace, new energy, electronics, and biomedicine [1,2,3,4,5]. However, the development lags far behind the increasing requirements. One crucial reason is that organosilicon synthetic methodologies are relatively few compared to numerous synthetic strategies to build carbon-based materials. Up to now, three traditional methodologies, including direct synthesis, hydrolysis and polycondensation, and hydrosilylation reactions, are still the most widely used techniques in the laboratory and in industry [6,7,8,9], although these traditional methodologies were found more than 80 years ago. Thus, developing new efficient synthetic methodologies to design and construct organosilicon materials is still highly desirable.
Thanks to the well-developed and emerging organic synthetic methodologies, diversified organosilicon materials with novel structures and functionality have been designed, prepared, and new applications found, and thus organosilicon synthetic methodologies are increasingly growing [10,11,12,13,14,15,16]. Click reactions have attracted specific attention because of their intriguing features, such as mild reaction conditions, fast speed, high yield, and brilliant tolerance of functional groups [17,18,19]. The azide–alkyne reaction [20], thiol–ene reaction [21,22], and aza-Michael addition reaction [23,24] with the “click” feature have all been widely used to prepare organosilicon compounds [21], functionalized polysiloxanes [20,25,26,27], and silicone elastomers [28] by us and other groups, and thus have become mature or semi-mature organosilicon synthetic methodologies. From the standpoint of existing raw materials, these strategies require azide- (or alkyl), thiol-, vinyl-, and amine-containing silanes or (poly)siloxanes, most of which have relatively high costs except vinyl-containing materials. However, epoxy-based materials (e.g., epoxy-containing disiloxane and polysiloxanes), which have a much lower cost than thiol- or amine-based ones, have not been fully exploited. It is of great interest to develop an organosilicon synthetic methodology using epoxy groups.
Herein, we introduce the thiol–epoxy reaction as the new organosilicon synthetic methodology and choose the preparation of functionalized disiloxanes as the representative examples. The selection of the thiol–epoxy reaction is motivated by its click feature and wide usage in the construction of functional molecules and polymers [29,30]. Although this reaction has been used to prepare some organosilicon materials [31], such as silsesquioxane-based materials [32,33,34,35,36], silicone/epoxy composites [37,38], and epoxy/silanes hybrid coating [39], exploration in this area is still limited and systematical investigation is rare. Moreover, the nonconventional properties of organosilicon materials constructed by this reaction have still not been studied despite the presence of unique S→Si coordination bonds within the structures. Thus, a systematical investigation on the synthesis of new sulfur-containing functionalized disiloxanes by the thiol–epoxy reaction is reported in the present study. The catalytic system and reaction condition for the synthesis of organosilicon compounds were optimized by studying the effects of catalysts, substituents, reaction temperature, and reaction times on the reaction efficiency. Using the optimal reaction, eleven new functionalized disiloxanes were prepared and their nonconventional fluorescent properties were explored.

2. Results and Discussion

2.1. Optimization of Reaction Condition

The thiol–epoxy reaction is considered a click reaction [40] when using the base as the catalyst, and previous reports proved that tetrabutylammonium fluoride (TBAF) and lithium hydroxide (LiOH) were the most efficient catalysts [41]. When this reaction was used to prepare organosilicon materials, the present sulfur-containing functionalized disiloxanes, the base catalyst is required to evaluate the suitability because of the sensitivity of Si–O or Si–C bonds under basic conditions. Thus a model reaction between 1,3-bis(3-glycidoxypropyl)tetramethyldisiloxane (E1) and 1-dodecanethiol (T1) was conducted to find out the optimal condition using LiOH or TBAF as catalysts under solvent-free condition or in tetrahydrofuran (THF) (Scheme 1) [41]. Table 1 demonstrates the thiol–epoxy reaction data of E1 and T1.
LiOH was initially used as the catalyst (5 mol%) without solvent and with a reaction time of 6 h at room temperature. The conversion rate of the epoxy group is 63.1% (Table 1, entry 1). It is a good sign that prolonging the reaction time to 12 h led to the quantitative conversion of the epoxy group (Table 1, entry 2). However, when the amount of catalyst was reduced to 3 mol% and 1 mol%, the conversion rates were significantly decreased to 80.3% and 39.8%, respectively (Table 1, entries 3–4). Then TBAF was used as a catalyst in THF. It was found that the conversion rate of the epoxy group was 81.3% with 8 mol% of TBAF and a reaction time of 12 h at room temperature (Table 1, entry 5). When the temperature was increased to 50 °C, the epoxy groups could be quantitatively consumed (Table 1, entry 6). Then, the amount of catalyst was reduced to 5 mol%, 3 mol%, 1 mol%, and 0.8 mol% under the same condition (Table 1, entries 7–10) and it was found that 3 mol% of catalyst could still provide high catalytic activity with a high conversion rate of the epoxy group (98.6%) (Table 1, entry 8). Based on this finding, the reaction time was further optimized with 3 mol% of catalyst loading (Table 1, entries 11–12). The results demonstrated that 6 h was sufficient to achieve almost quantitative conversion (Table 1, entry 11).
In addition, no obvious by-products, which may be generated by the cleavage of Si–O or Si–C bonds, were found in the final products using TBAF and LiOH as the catalysts, as evidenced by the high yields (98% and 96% for entries 2 and 11) after purification and the 29Si NMR spectra of the reaction mixture, in which no peaks in the ranges of −30 ppm to −20 ppm and 20 ppm to 40 ppm, which are assigned to the silicon atoms from Si–OH or Si-F bonds, respectively, were observed during the reaction (Figure S1). These results indicate that both TBAF and LiOH can act as efficient catalysts for the synthesis of new functionalized disiloxanes. It is known that LiOH is soluble in water but insoluble in most of organic solvents; thus, it is considered a heterogeneous catalyst [30]. Although it can efficiently catalyze this reaction of E1T1 without the solvent, its catalytic activity is relatively low when thiol compounds (e.g., methimazole) appear as the solid, and thus the condition (e.g., additional solvents) needs to be further optimized [30]. In contrast, this possibility will not occur when selecting TBAF in THF because THF can dissolve most thiol compounds. In addition, the alkalinity of TBAF is weaker than LiOH, and the residual LiOH in the system may cause the instability of the products (e.g., crosslinking networks) [42]. Thus, considering the universality of TBAF in THF, herein we choose it for further synthesis of various sulfur-containing functionalized disiloxanes.

2.2. Synthesis and Characterization of Sulfur-Containing Functionalized Disiloxanes

The capability of the optimized reaction condition, that is TBAF (3 mol%) in THF at 50 °C for 6 h, was tested by preparing sulfur-containing functionalized disiloxanes from E1 and various thiol compounds (Scheme 2). The results revealed that eleven new disiloxanes were successfully achieved with excellent conversion rates of epoxy groups (≥99%) and isolated yields (>89%) and high regioselectivity (Table 2, entries 1–11) regardless of aliphatic thiols (e.g., 1-butanethiol, a yield of 95.5%, Table 2, entry 2) or aromatic thiols (e.g., 3-methylbenzenethiol, a yield of 96.5%, Table 2, entry 9). As expected, in using aromatic thiols, the products can be generated with high yields, even shortening the reaction times due to their stronger nucleophile ability of the formed thiolate anion than aliphatic thiols [30,43]. For example, when 3-methylbenzenethiol (T9) reacted with E1, 3 h was sufficient to afford a nearly quantitative conversion of the epoxy group (Table 2, entry 16). In addition, the yield of E1T5 (Table 2, entry 5) was slightly lower than that of E1T2 (Table 2, entry 2), and this finding could be explained by higher steric hindrance of 3-methyl-2-butanethiol (T5) than 1-butanethiol (T2). This speculation can also explain the fact that E1T2 and E1T4 can also be afforded with high yields, even shortening the reaction time to 3 h or 5 h (Table 2, entries 12 and 14), in comparison to the requirement of 6 h for E1T1 (Table 2, entry 1), concerning the higher steric hindrance of T1 than T2 and T4. These results demonstrated that the electronic effect and steric hindrance of substituent groups influenced the activity of thiol groups, resulting in the different reactivity of thiol compounds and further different yields.
The successful achievement of these sulfur-containing functionalized disiloxanes can be determined by FT-IR, 1H/13C/29Si NMR, HR-MS, and elemental analysis. Considering their similar structures, E1T2 and E1T9 were selected as the representative examples (the original spectra of all the products can be found in the Figures S2−S19 of the Supporting Information). Figure 1a illustrates the IR spectrum of E1, E1T2, and E1T9. After the reaction, the peaks assigned to the epoxy group at 910 cm−1 disappeared in the products E1T2 and E1T9, while a new strong and broad peak attributed to the generated β-hydroxyl group at 3464 cm−1 was observed. Figure 1b illustrates the 1H NMR of E1, E1T2, and E1T9, and the assignments of the characteristic protons. The peaks assigned to epoxy protons at 2.55 ppm (Hc’), 2.74 ppm (Hc), and 3.09 ppm (Hb) in E1 completely disappeared in E1T2 and E1T9, while two new signals at 2.76 ppm and 3.82 ppm, which are attributed to the protons from β-hydroxyl groups (H4 in E1T2 and HD in E1T9) and −CH groups adjacent to it (H2 in E1T2 and HB in E1T9), were observed (Figure 1b). In the 29Si NMR, a single peak was found at 7.73 ppm and 7.75 ppm for E1T2 and E1T9, in comparison to a single peak of 7.46 ppm for E1, indicating that the chemical environment of the Si atom has been changed (Figure 1c). It was proven that the Si atom could interact with heteroatoms (e.g., N or S) by accepting a lone pair of p electrons, resulting in S→Si or N→Si coordination bonds [44]. Thus, the present variation is apparently due to the formation of S→Si coordination.

2.3. Fluorescent Properties

Previous reports have demonstrated that the existence of heteroatom in organosilicon molecules or polymers could exhibit interesting nonconventional fluorescence [44]. The present sulfur-containing functionalized disiloxanes expectedly show nonconventional fluorescence. All the samples in ethanol (EtOH) solutions emit similar and strong blue fluorescence with the maximum emission wavelength (λmax) at ca. 410 nm when excited by 365 nm (Figure 2). Moreover, this phenomenon can be visualized under 365 nm ultraviolet (UV) light (Figure 3). Similar to previous reports [44], this nonconventional fluorescence was generated by through-space charge transfer from lone pairs of electrons in S atoms to the 3d empty orbital of Si atoms and aggregation of chromophores, including ester groups in E1T6 and aromatic units in E1T7 to E1T11. It is worth noting that E1T11 has the highest fluorescent intensities. Previous reports demonstrated that benzothiazole and its derivatives have excellent fluorescence due to the large delocalized π bond and rigid planar construction. Hence, the aromatic π–π conjugation contributes to the blue emission region with the λmax at ca. 400 nm, in addition to the contribution from S→Si and N→Si coordination [45]. It is worth noting that the coordination is not only contributed by the thioether group but also by the S and N from thiazole units. These combined contributions lead to the highest fluorescence intensity. In addition, a weak emission peak at 497 nm was found. This finding is probably due to higher electronic conjugation induced by the π–π interaction between aromatic thiazole groups [46], while S→Si and N→Si coordination facilitates this interaction. The combined contribution can also explain higher fluorescence intensities found in E1T6 and E1T9 than in other samples, because additional aggregation of ester groups and N→Si coordination contributed by imidazole groups [24,47,48], exist within E1T6 and E1T9, respectively.
The nonconventional fluorescent properties of these polysiloxanes can also be determined by the fluorescence variation in different concentrations in solutions. Figure 4 shows the fluorescence spectra of E1T6 and E1T11 in ethanol at different concentrations as the examples (see Figures S20–S22 for other samples). Their fluorescence intensities were progressively enhanced with increasing concentrations. This phenomenon is a typical example of aggregation-induced emission enhancement (AIEE), which is apparently induced by the aggregation of nonconventional chromophores, including X→Si (X = S or N) coordination bonds in these compounds and ester groups in E1T6 [47,49]. In addition, a new peak at ca. 500 nm could be observed at a high concentration of E1T11, which could be assigned to the extended conjugation formed by the aggregation of the thiazole groups.
Generally, the fluorescence property depends on the environment. Therefore, the photophysical characteristics of these sulfur-containing functionalized disiloxanes were investigated in various organic solvents. Figure 5 shows the fluorescence spectra of E1T6 and E1T11 in different solvents excited at 365 nm (Figures S23–S25 for other samples). For E1T6, the emission regions were similar in different solvents, but their fluorescence intensity varied. The fluorescence emission intensity became weaker in the order of dimethylformamide (DMF), THF, dichloromethane (DCM), ethanol, and acetone. The same variation trend of emission intensity was also found in E1T11 (Figure 5b). This finding may be due to the participation of the electron-rich solvents (lone pairs) in the aggregation of nonconventional chromophores and consequently physically altered the extended electronic conjugation. In addition, the polarity of the solvents can also influence the fluorescence. As a result, these factors affect the emission behavior [50].
To further prove the contribution of S→Si or Si–O to the nonconventional fluorescence, a model compound E2T6, which has similar functional groups with E1T6 but without Si–O–Si linker, was synthesized by the thiol–epoxy reaction of diethylene glycol diglycidyl ether (E2) and ethyl thioglycolate (T6) (Scheme 3).
As illustrated in Table 3, although E2T6 can also emit blue fluorescence and have similar emission wavelengths to E1T6, its fluorescence intensity is much lower than that of E1T6 under the same excitation wavelength (Figure 6a), indicating the important role of Si–O–Si in the enhancement of fluorescence intensity. This Si-O-Si induced enhancement can be found even without direct introduction in the structures. As shown in Figure 6b, the fluorescence intensity of E2T6 was obviously enhanced after adding equimolar hexamethyldisiloxane (MM), although it was still lower than that of E1T6. These results indicate that the interaction of S→Si coordination bonds can increase the aggregation of ester groups and thus enhance the fluorescence intensity.

3. Materials and Methods

3.1. Materials

All reagents and solvents were obtained from commercial suppliers and used without further purification unless otherwise noted. 1,3-bis(3-glycidoxypropyl)tetramethy-ldisiloxane (≥98%) was supplied by Hangzhou Sloan Materials Technology Co., Ltd. (Hangzhou, China). Diethylene glycol diglycidyl ether (68%) was purchased from Jinan Jing Shui Biotechnology Co., Ltd. (Jinan, China). Tetrabutylammonium fluoride (TBAF), lithium hydroxide (LiOH), 1-dodecanethiol (≥98%), 2-mercaptoethanol (≥99%), 1-butanethiol (≥97%), 2-propenthiol (50%), 3-methyl-2-butanethiol (≥98%), ethyl thioglycolate (≥98%), furfuryl mercaptan (≥98%), methimazole (≥98%), 3-methylbenzenethiol (≥97%), benzyl mercaptan (≥98%), and 2-mercaptobenzothiazole (≥98%) were purchased from MACKLIN Chemical Reagent Co., Ltd. (Shanghai, China). Tetrahydrofuran (THF), dichloromethane (DCM), ethyl acetate (EA), petroleum ether (PE), acetone, ethanol, and dimethylformamide (DMF) all with Grade AR were purchased from Tianjin Fuyu Chemical Reagent Co., Ltd. (Tianjin, China).

3.2. Methods

Fourier transform infrared (FT-IR) spectra were conducted on a Bruker TENSOR-27 FT-IR spectrophotometer (Germany) using the KBr pellet technique and recorded in the range of 400 cm−1 to 4000 cm−1. 1H NMR, 13C NMR, and 29Si NMR were recorded on a Bruker AVANCE 400 NMR spectrometer (Rheinstetten, Germany) operating at 25 °C using CDCl3 as solvent and without tetramethylsilane as an interior label, and using Software MestReNova-14.0.0-23239 to process the data. High-resolution mass spectra (HR-MS) were recorded on UPLC–QTOF–MS (Bruker, Bremen, Germany) with ESI+, source spray voltage 2.6 kV and capillary temperature 180.00 °C, and data processing was performed using Bruker Compass Data Analysis 4.4.2. The fluorescent emission spectra of the samples were determined at room temperature with a F-7000 fluorescence spectrophotometer (Hitachi, Tokyo, Japan) using a monochromated Xe lamp as an excitation source. The contents of C, H, N, and S were characterized by the Elementar Vario EL III elemental analyzer (Munich, Germany).

3.3. General Procedure for the Synthesis of Sulfur-Containing Functionalized Disiloxanes E1T1 to E1T11

1,3-bis(3-glycidoxypropyl)-1,1,3,3-tetramethyldisiloxane (E1, 362.6 mg, 1 mmol), thiol compounds (Tn, 2.5 mmol), and tetrahydrofuran (THF, 0.6 mL) were placed into a 25 mL round-bottom flask with a magnetic stirrer. Then, tetrabutylammonium fluoride (TBAF, 7.9 mg, 0.03 mmol) was added to the mixture at 0 °C. After removing the cooling, the mixture was heated at 50 °C for 6 h under stirring. Then, the mixture was cooled to room temperature and dichloromethane DCM (100 mL) was added. The mixture was washed with water and the organic layer was separated and dried over anhydrous magnesium sulfate. After the filtration, the solvent was removed under reduced pressure to yield the crude product. The pure products E1T1 to E1T11 were then purified by column chromatography on silica gel using PE/EA as an eluent.
1,3-bis(3-(dodecylthiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T1): The synthesis was performed according to general procedure using 1-dodecanethiol (T1, 506.0 mg) as a thiol compound. E1T1 was afforded as a colorless viscous oil (736.0 mg, yield: 96.0%). FT-IR (KBr pellet cm−1): ν = 3460, 2928, 2858, 1464, 1255, 1119, 1065, 843, 795. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.39–0.48 (t, 4H, −SiCH2), 0.83 (m, 6H, −CH2CH3), 1.20–1.32 (m, 36H, −(CH2)9CH3), 1.53 (m, 8H, −SiCH2CH2, −SCH2CH2), 2.44–2.71 (m, 10H, −CH(OH)CH2SCH2), 3.31–3.51 (m, 8H, −CH2OCH2), 3.81 (m, 2H, −CH(OH)). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.83, 13.91, 22.38, 23.09, 28.49–29.47, 31.61, 32.30, 35.50, 68.85, 73.08, 73.95. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.70. HR-MS (FAB) calcd for C40H86O5S2Si2 (MH+): 767.5533; Found: 767.5535. Anal. Calcd for C40H86O5S2Si2 (%): C, 62.61; H, 11.30; S, 8.36. Found: C, 61.69; H, 10.18; S, 9.52.
1,3-bis(3-(butylthiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T2): The synthesis was performed according to general procedure using 1-butanethiol (T2, 225.5 mg) as a thiol compound. E1T2 was obtained as a colorless viscous oil (518.0 mg). Yield: 95.5%. FT-IR (KBr pellet cm−1): ν = 3456, 2949, 2868, 1464, 1418, 1259, 1188, 1121, 1063, 845, 795. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.40–0.49(t, 4H, −SiCH2), 0.86 (t, 6H, −CH2CH3), 1.28–1.42 (m, 4H, −CH2CH3), 1.46–1.60 (m, 8H, −SiCH2CH2, −SCH2CH2), 2.45–2.68 (m, 8H, −CH2SCH2), 2.76 (s, 2H, −CH(OH)), 3.33–3.50 (m, 8H, −CH2OCH2), 3.77–3.87 (m, 2H, −CH(OH)). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.11, 13.69, 16.44, 22.00, 24.63, 32.07, 32.20, 36.21, 70.16, 73.49, 74.09. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.73. HR-MS (FAB) calcd for C24H54O5S2Si2 (MH+): 543.3029; Found: 543.3031. Anal. Calcd for C24H54O5S2Si2 (%): C, 53.09; H, 10.02; S, 11.81. Found: C, 52.58; H, 10.03; S, 12.08.
1,3-bis(3-((2-hydroxyethyl)-thiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T3): The synthesis was performed according to general procedure using 2-mercaptoethanol (T3, 195.0 mg) as a thiol compound. E1T3 was obtained as a colorless viscous oil (494.4 mg). Yield: 95.4%. FT-IR (KBr pellet cm−1): ν = 3410, 2941, 2869, 1651, 1410, 1256, 1182, 1115, 1057, 845, 791. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.39–0.48 (m, 4H, −SiCH2), 1.54 (dq, 4H, −SiCH2CH2), 2.52–2.74 (m, 8H, −CH2SCH2), 3.29–3.47 (m, 8H, −CH2OCH2), 3.56 (s, 4H, −CH2OH, −CH(OH)), 3.70 (t, 4H, −CH2OH), 3.86 (m, 2H, −CH(OH)). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.81, 22.98, 35.39, 35.41, 60.87, 69.49, 73.03, 73.90. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.70. HR-MS (FAB) calcd for C20H46O7S2Si2 (MH+): 519.2302; Found: 519.2305. Anal. Calcd for C20H46O7S2Si2 (%): C, 46.30; H, 8.94; S, 12.36. Found: C, 45.64; H, 8.93; S, 12.40.
1,3-bis(3-(allylthiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T4): The synthesis was performed according to general procedure using 2-propenthiol (T4, 370.0 mg) as a thiol compound. E1T4 was obtained as a colorless viscous oil (480.1 mg). Yield: 94.2%. FT-IR (KBr pellet cm−1): ν = 3456, 2924, 2868, 1639, 1414, 1256, 1188, 1121, 1065, 835, 791. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.40–0.49 (m, 4H, −SiCH2), 1.48–1.60 (m, 4H, −SiCH2CH2), 2.44–2.65 (m, 4H, −CH(OH)CH2SCH2), 2.85 (s, 2H, −CH(OH)), 3.10 (dd, 4H, −CH(OH)CH2SCH2), 3.30–3.50 (m, 8H, −CH2OCH2), 3.82 (q, 2H, −CH(OH)), 4.99–5.16 (m, 4H, −CH=CH2), 5.66–5.81 (m, 2H, −CH=CH2). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.90, 23.07, 33.79, 34.78, 68.87, 73.07, 73.90, 117.11, 133.80. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.73. HR-MS (FAB) calcd for C22H46O5S2Si2 (MH+): 511.2403; Found: 511.2402. Anal. Calcd for C22H46O5S2Si2 (%): C, 51.72; H, 9.08; S, 12.55. Found: C, 51.27; H, 8.47; S, 13.05.
1,3-bis(3-((3-methylbutan-2-methyl)-thiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T5): The synthesis was performed according to general procedure using 3-methyl-2-butanethiol (T5, 260.3 mg) as a thiol compound. E1T5 was obtained as a colorless viscous oil (514.4 mg). Yield: 90.2%. FT-IR (KBr pellet cm−1): ν = 3456, 2964, 2869, 1462, 1379, 1249, 1109, 1061, 843, 795. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.39–0.51 (m, 4H, −SiCH2), 0.91 (dd, 12H, −CH(CH3)2), 1.17 (d, 6H, −SCH(CH3)), 1.48–1.61 (m, 4H, −SiCH2CH2), 1.72–1.86 (m, 2H, −CH(CH3)2), 2.52–2.71 (m, 6H, −CH2SCH), 2.85 (m, 2H, −CH(OH)), 3.32–3.50 (m, 8H, −CH2OCH2), 3.80 (m, 2H, −CH(OH)). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.01, 13.93, 17.35, 18.14, 19.55, 23.09, 32.58, 34.58, 47.22, 69.07, 73.08, 73.95. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.74. HR-MS (FAB) calcd for C26H58O5S2Si2 (MH+): 571.3342; Found: 571.3345. Anal. Calcd for C26H58O5S2Si2 (%): C, 54.69; H, 10.24; S, 11.23. Found: C, 52.85; H, 10.09; S, 9.99.
1,3-bis(3-(ethyl-3-acetate-thiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T6): The synthesis was performed according to general procedure, ethyl thioglycolate (T6, 300.0 mg) as a thiol compound. E1T6 was obtained as a colorless viscous oil (572.1 mg). Yield: 95.0%. FT-IR (KBr pellet cm−1): ν = 3469, 2939, 2870, 1738, 1466, 1408, 1265, 1123, 1042, 843, 787. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.44 (m, 4H, −SiCH2), 1.23 (m, 6H, −CH2CH3), 1.53 (m, 4H, −SiCH2CH2), 2.62–2.83 (m, 4H, −SCH2CH(OH)), 2.88 (s, 2H, −CH(OH)), 3.25 (t, 4H, −SCH2C(O)), 3.31–3.51 (m, 8H, −CH2OCH2), 3.88 (m, 2H, −CH(OH)), 4.14 (m, 4H, −OCH2CH3). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.83, 13.88, 23.05, 33.94, 36.14, 61.18, 69.10, 72.94, 73.93, 170.47. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.65. HR-MS (FAB) calcd for C24H50O9S2Si2 (MH+): 603.2513; Found: 603.2520. Anal. Calcd for C24H50O9S2Si2 (%): C, 47.81; H, 8.36; S, 10.63. Found: C, 46.86; H, 8.40; S, 10.68.
1,3-bis(3-((furanyl-methyl)-thiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T7): The synthesis was performed according to general procedure using furfuryl mercaptan (T7, 285.0 mg) as a thiol compound. E1T7 was obtained as a pale-yellow viscous oil (559.0 mg). Yield: 94.7%. FT-IR (KBr pellet cm−1): ν = 3452, 3119, 2930, 2870, 1506, 1414, 1254, 1119, 1065, 868, 796. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.01 (s, 12H, −SiCH3), 0.39–0.49 (m, 4H, −SiCH2), 1.47–1.60 (m, 4H, −SiCH2CH2), 2.50–2.68 (m, 4H, −SCH2CH(OH)), 2.81 (t, 2H, −CH(OH)), 3.31–3.48 (m, 8H, −CH2OCH2), 3.68–3.74 (m, 4H, −CHCH2SCH2), 3.75–3.84 (m, 2H, −CH(OH)), 6.11–6.27 (m, 4H, −OCH=CHCH=C, Furan ring), 7.27–7.33 (m, 2H, −OCH=CHCH=C, Furan ring). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.88, 23.05, 28.37, 34.71, 69.03, 72.97, 73.87, 107.41, 110.10, 141.86, 151.05. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.74. HR-MS (FAB) calcd for C26H46O7S2Si2 (MH+): 591.2297; Found: 591.2292. Anal. Calcd for C26H46O7S2Si2 (%): C, 52.85; H, 7.85; S, 10.85. Found: C, 51.16; H, 7.23; S, 10.13.
1,3-bis(3-((1-methylimidazol)-thiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T8): The synthesis was performed according to general procedure, methimazole (T8, 285.0 mg) as a thiol compound. E1T8 was obtained as a pale-yellow viscous oil (545.0 mg). Yield: 92.3%. FT-IR (KBr pellet cm−1): ν = 3390, 3198, 3117, 2937, 2865, 1674, 1514, 1466, 1258, 1117, 1055. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.39–0.49 (m, 4H, −SiCH2), 1.46–1.59 (m, 4H, −SiCH2CH2), 3.05–3.27 (m, 4H, −SCH2CH(OH)), 3.34–3.52 (m, 8H, −CH2OCH2), 3.54 (s, 6H, −Imidazole−CH3), 4.05 (m, 2H, −CH(OH)), 5.43–6.04 (s, 2H, -CH(OH)), 6.85 (m, 4H, −Imidazole). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.90, 23.08, 33.16, 37.25, 70.25, 72.89, 73.86, 121.83, 127.73, 142.69. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.73. HR-MS (FAB) calcd for C24H46N4O5S2Si2 (MH+): 591.2526; Found: 591.2534. Anal. Calcd for C24H46N4O5S2Si2 (%): C, 48.78; H, 7.85; N, 9.48; S, 10.85. Found: C, 47.94; H, 8.07; N, 10.43; S, 10.39.
1,3-bis(3-(m-tolyl-thiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T9): The synthesis was performed according to general procedure using 3-methylbenzenethiol (T9, 310.0 mg) as a thiol compound. E1T9 was obtained as a colorless viscous oil (589.0 mg). Yield: 96.5%. FT-IR (KBr pellet cm−1): ν = 3454, 3051, 2939, 2864, 1587, 1475, 1412, 1258, 1180, 1111, 1053. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.39–0.50 (m, 4H, −SiCH2), 1.46–1.58 (m, 4H, −SiCH2CH2), 2.24 (s, 6H, −Phenyl−CH3), 2.91–3.08 (m, 6H, −SCH2CH(OH)), 3.32–3.48 (m, 8H, −CH2OCH2), 3.82 (m, 2H, −CH(OH)), 6.91–7.15 (m, 8H, −Phenyl). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.82, 20.93, 22.99, 36.85, 68.68, 72.62, 73.76, 125.96, 126.64, 128.39, 129.56, 135.27, 138.18. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.74. HR-MS (FAB) calcd for C30H50O5S2Si2 (MH+): 611.2716; Found: 611.2725. Anal. Calcd for C30H50O5S2Si2 (%): C, 58.97; H, 8.25; S, 10.49. Found: C, 58.38; H, 8.31; S, 10.49.
1,3-bis(3-(benzyl-thiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T10): The synthesis was performed according to general procedure using, benzyl mercaptan (T10, 310.0 mg) as a thiol compound. E1T10 was obtained as a colorless viscous oil (572.4 mg). Yield: 93.8%. FT-IR (KBr pellet cm−1): ν = 3447, 3022, 2928, 2868, 1603, 1495, 1416, 1258, 1123, 1049, 706. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.38–0.48 (m, 4H, −SiCH2), 1.46–1.58 (m, 4H, −SiCH2CH2), 2.50 (m, 4H, −SCH2CH(OH)), 2.68–2.74 (m, 2H, −CH(OH)), 3.30–3.44 (m, 8H, −CH2OCH2), 3.68 (s, 4H, -CHCH2SCH2), 3.79 (m, 2H, −CH(OH)), 7.14–7.28 (m, 10H, −Phenyl). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.84, 23.01, 34.36, 36.19, 68.84, 72.99, 73.80, 126.68, 128.13, 128.55, 137.79. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.78. HR-MS (FAB) calcd for C30H50O5S2Si2 (MH+): 611.2716; Found: 611.2729. Anal. Calcd for C30H50O5S2Si2 (%): C, 58.97; H, 8.25; S, 10.49. Found: C, 57.95; H, 8.32; S, 10.65.
1,3-bis(3-(benzothiazolyl-thiopropoxy)-2-ol-propyl)-1,1,3,3-tetramethyldisiloxane (E1T11): The synthesis was performed according to general procedure using 2-mercaptobenzothiazole (T11, 417.5 mg) as a thiol compound. E1T11 was obtained as a yellow viscous oil (622.3 mg). Yield: 89.4%. FT-IR (KBr pellet cm−1): ν = 3418, 3068, 2939, 2872, 1568, 1462, 1425, 1256, 1186, 1065, 845. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 0.00 (s, 12H, −SiCH3), 0.40–0.50 (m, 4H, −SiCH2), 1.54 (dq, 4H, −SiCH2CH2), 3.34–3.55 (dt, 12H, −CH2OCH2CH(OH)CH2), 4.14 (m, 2H, −CH(OH)), 4.64 (d, 2H, −CH(OH)), 7.15–7.69 (m, 8H, −Phenyl). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 0.00, 13.86, 23.05, 37.04, 69.54, 72.68, 73.88, 120.62, 120.87, 124.09, 125.78, 134.96, 152.10, 167.45. 29Si NMR (79 MHz, CDCl3): [δ, ppm] = 7.80. HR-MS (FAB) calcd for C30H44N2O5S4Si2 (MH+): 697.1750; Found: 697.1760. Anal. Calcd for C30H44N2O5S4Si2 (%): C, 51.69; H, 6.36; N, 4.02; S, 18.40. Found: C, 49.95; H, 6.39; N, 4.99; S, 17.65.

3.4. Synthesis of bis(2-(ethyl-3-acetate-thiopropoxy)-2-ol-ethyl)-ether (E2T6)

The synthesis of E2T6 was performed according to general procedure based on diethylene glycol diglycidyl ether (E2, 218.1 mg, l mmol) and ethyl thioglycolate (T6, 300.0 mg, 2.5 mmol). E2T6 was obtained as a colorless viscous oil (435.3 mg) with the same synthesis procedure of E1T1. Yield: 95.0%. FT-IR (KBr pellet cm−1): ν = 3446.67, 2912.41, 2881, 1732.01, 1467.78, 1284.55, 1132.17, 1031.88. 1H NMR (400 MHz, CDCl3): [δ, ppm] = 1.23 (t, 6H, −CH2CH3), 2.71 (m, 4H, −SCH2CH(OH)), 3.25 (d, 4H, −SCH2COO), 3.43–3.66 (m, 14H, −OCH2CH2OCH2CH(OH)), 3.85–3.93 (m, 2H, −CH(OH)), 4.14 (m, 4H, −CH2CH3). 13C NMR (101 MHz, CDCl3): [δ, ppm] = 14.15, 34.30, 35.75, 61.46, 69.50, 70.41, 70.53, 74.28, 170.71.

4. Conclusions

In summary, we developed an efficient synthetic strategy, i.e., a thiol–epoxy reaction, for the synthesis of novel sulfur-containing functionalized disiloxanes from 1,3-bis(3-glycidoxypropyl)-1,1,3,3-tetramethyldisiloxane and different thiol compounds. The study found that both lithium hydroxide (LiOH) and tetrabutylammonium fluoride (TBAF) have high catalytic activity, and TBAF in THF is universal for the synthesis of these compounds because THF can dissolve most thiol compounds. The reactions can be conducted under mild reaction conditions with high reaction yields and regioselectivity, no by-products, and good functional group tolerance. Furthermore, these new disiloxanes exhibit typical aggregation-induced emission enhancement (AIEE) characteristics and can emit strong blue fluorescence under the action of the unconventional fluorescent chromophores S→Si or N→Si coordination bonds. In addition, the S→Si coordination bonds can also increase the aggregation of the nonconventional fluorescent clusters and thus enhance the fluorescence intensity. These results demonstrate that the thiol–epoxy reaction could be developed as an efficient organosilicon synthetic methodology for constructing various organosilicon compounds or polymers with new structures and functionalities from the low-cost epoxy-containing organosilicon raw materials. It is worth noting that although TBAF is an efficient catalyst for the synthesis of functionalized disiloxanes, it may be not adaptable for the synthesis of other organosilicon compounds, especially silsesquioxanes, because fluoride ions could react rapidly with silicon atoms and it may be a cause of complicated and unpredictable synthetic results. Further synthesis of other organosilicon materials and the reaction methods may be also optimized. Our group is currently utilizing this reaction to construct various novel organosilicon polymers and materials and find their applications in emerging fields, such as sensing and catalysis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24097785/s1.

Author Contributions

Conceptualization, D.W.; methodology, J.T. and D.W.; validation, J.T. and D.W.; investigation, J.T. and D.W.; writing—original draft preparation, J.T. and D.W.; writing—review and editing, J.T., D.W. and S.F.; supervision, D.W.; project administration, D.W.; funding acquisition, D.W. and S.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (22271175 and 52173102), Fluorine Silicone Materials Collaborative Fund of Shandong Provincial Natural Science Foundation (ZR2020LFG011 and ZR2021LFG001), Basic Research Foundation of Institute of Silicon-Based High-End New Materials (YJG202103), and Young Scholars Program of Shandong University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yilgör, E.; Yilgör, I. Silicone containing copolymers: Synthesis, properties and applications. Prog. Polym. Sci. 2014, 39, 1165–1195. [Google Scholar] [CrossRef]
  2. Rücker, C.; Kümmerer, K. Environmental Chemistry of Organosiloxanes. Chem. Rev. 2015, 115, 466–524. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, J.; Yao, Y.; Li, X.; Zhang, Z. Fabrication of advanced polydimethylsiloxane-based functional materials: Bulk modifications and surface functionalizations. Chem. Eng. J. 2021, 408, 127262. [Google Scholar] [CrossRef]
  4. Cazacu, M.; Dascalu, M.; Stiubianu, G.-T.; Bele, A.; Tugui, C.; Racles, C. From passive to emerging smart silicones. Rev. Chem. Eng. 2022. [Google Scholar] [CrossRef]
  5. Qi, D.; Zhang, K.; Tian, G.; Jiang, B.; Huang, Y. Stretchable Electronics Based on PDMS Substrates. Adv. Mater. 2021, 33, 2003155. [Google Scholar] [CrossRef]
  6. Wang, D.; Klein, J.; Mejía, E. Catalytic Systems for the Cross-Linking of Organosilicon Polymers. Chem. Asian J. 2017, 12, 1180–1197. [Google Scholar] [CrossRef]
  7. Brook, M.A. Silicon in Organic, Organometallic, and Polymer Chemistry; Wiley: New York, NY, USA, 2000; Volume 123. [Google Scholar]
  8. Pawlenko, S. Organosilicon Chemistry; De Gruyter: Berlin, Germany, 1986. [Google Scholar]
  9. Chojnowski, J.; Marciniec, B. Progress in Organosilicon Chemistry; Gordon and Breach: Philadelphia, PA, USA, 1995. [Google Scholar]
  10. Wang, D.; Cao, J.; Hang, D.; Li, W.; Feng, S. Novel Organosilicon Synthetic Methodologies. Prog. Chem. 2019, 31, 110–120. [Google Scholar]
  11. Jones, R.; Ando, W.; Chojnowski, J. The Science and Technology of Their Synthesis and Applications. Silicon-Containing Polymers; Springer: Berlin/Heidelberg, Germany, 2000. [Google Scholar]
  12. LaRonde, F.J.; Brook, M.A.; Hu, G. Amino acid and peptide chemistry on silicones. Silicon Chem. 2002, 1, 215–222. [Google Scholar] [CrossRef]
  13. Fan, X.; Zhang, M.; Gao, Y.; Zhou, Q.; Zhang, Y.; Yu, J.; Xu, W.; Yan, J.; Liu, H.; Lei, Z.; et al. Stepwise on-demand functionalization of multihydrosilanes enabled by a hydrogen-atom-transfer photocatalyst based on eosin Y. Nat. Chem. 2023. [Google Scholar] [CrossRef]
  14. Fuchise, K.; Sato, K.; Igarashi, M. Precise Synthesis of Linear Polysiloxanes End-Functionalized with Alkynylsilyl Groups by Organocatalytic Ring-Opening Polymerization of Cyclotrisiloxanes. Macromolecules 2021, 54, 5765–5773. [Google Scholar] [CrossRef]
  15. Siripanich, P.; Bureerug, T.; Chanmungkalakul, S.; Sukwattanasinitt, M.; Ervithayasuporn, V. Mono and Dumbbell Silsesquioxane Cages as Dual-Response Fluorescent Chemosensors for Fluoride and Polycyclic Aromatic Hydrocarbons. Organometallics 2022, 41, 201–210. [Google Scholar] [CrossRef]
  16. Yi, B.; Wang, S.; Hou, C.; Huang, X.; Cui, J.; Yao, X. Dynamic siloxane materials: From molecular engineering to emerging applications. Chem. Eng. J. 2021, 405, 127023. [Google Scholar] [CrossRef]
  17. Golas, P.L.; Matyjaszewski, K. Marrying click chemistry with polymerization: Expanding the scope of polymeric materials. Chem. Soc. Rev. 2010, 39, 1338–1354. [Google Scholar] [CrossRef]
  18. Qin, A.; Lam, J.W.Y.; Tang, B.Z. Click polymerization. Chem. Soc. Rev. 2010, 39, 2522–2544. [Google Scholar] [CrossRef] [PubMed]
  19. Franc, G.; Kakkar, A.K. “Click” methodologies: Efficient, simple and greener routes to design dendrimers. Chem. Soc. Rev. 2010, 39, 1536–1544. [Google Scholar] [CrossRef]
  20. Rambarran, T.; Bertrand, A.; Gonzaga, F.; Boisson, F.; Bernard, J.; Fleury, E.; Ganachaud, F.; Brook, M.A. Sweet supramolecular elastomers from α,ω-(β-cyclodextrin terminated) PDMS. Chem. Commun. 2016, 52, 6681–6684. [Google Scholar] [CrossRef]
  21. Tucker-Schwartz, A.K.; Farrell, R.A.; Garrell, R.L. Thiol–ene Click Reaction as a General Route to Functional Trialkoxysilanes for Surface Coating Applications. J. Am. Chem. Soc. 2011, 133, 11026–11029. [Google Scholar] [CrossRef]
  22. Zuo, Y.; Cao, J.; Feng, S. Sunlight-Induced Cross-Linked Luminescent Films Based on Polysiloxanes and d-Limonene via Thiol–ene “Click” Chemistry. Adv. Funct. Mater. 2015, 25, 2754–2762. [Google Scholar] [CrossRef]
  23. Genest, A.; Binauld, S.; Pouget, E.; Ganachaud, F.; Fleury, E.; Portinha, D. Going beyond the barriers of aza-Michael reactions: Controlling the selectivity of acrylates towards primary amino-PDMS. Polym. Chem. 2017, 8, 624–630. [Google Scholar] [CrossRef]
  24. Lu, H.; Feng, L.; Li, S.; Zhang, J.; Lu, H.; Feng, S. Unexpected Strong Blue Photoluminescence Produced from the Aggregation of Unconventional Chromophores in Novel Siloxane–Poly(amidoamine) Dendrimers. Macromolecules 2015, 48, 476–482. [Google Scholar] [CrossRef]
  25. Zuo, Y.; Zhang, Y.; Gou, Z.; Lin, W. Facile construction of imidazole functionalized polysiloxanes by thiol-ene “Click” reaction for the consecutive detection of Fe3+ and amino acids. Sensors Actuat. Chem. 2019, 291, 235–242. [Google Scholar] [CrossRef]
  26. Bezlepkina, K.A.; Ardabevskaia, S.N.; Klokova, K.S.; Ryzhkov, A.I.; Migulin, D.A.; Drozdov, F.V.; Cherkaev, G.V.; Muzafarov, A.M.; Milenin, S.A. Environment Friendly Process toward Functional Polyorganosiloxanes with Different Chemical Structures through CuAAC Reaction. ACS Appl. Polym. Mater. 2022, 4, 6770–6783. [Google Scholar] [CrossRef]
  27. Milenin, S.A.; Drozdov, F.V.; Bezlepkina, K.A.; Majorov, V.Y.; Muzafarov, A.M. Acid-Catalyzed Rearrangement of Azidopropyl-Siloxane Monomers for the Synthesis of Azidopropyl-Polydimethylsiloxane and Their Carboxylic Acid Derivatives. Macromolecules 2021, 54, 2921–2935. [Google Scholar] [CrossRef]
  28. Zuo, Y.; Gou, Z.; Zhang, C.; Feng, S. Polysiloxane-Based Autonomic Self-Healing Elastomers Obtained through Dynamic Boronic Ester Bonds Prepared by Thiol–Ene “Click” Chemistry. Macromol. Rapid Commun. 2016, 37, 1052–1059. [Google Scholar] [CrossRef] [PubMed]
  29. Xu, Q.; Huang, T.; Li, S.; Li, K.; Li, C.; Liu, Y.; Wang, Y.; Yu, C.; Zhou, Y. Emulsion-Assisted Polymerization-Induced Hierarchical Self-Assembly of Giant Sea Urchin-like Aggregates on a Large Scale. Angew. Chem. 2018, 130, 8175–8179. [Google Scholar] [CrossRef]
  30. Stuparu, M.C.; Khan, A. Thiol-epoxy “click” chemistry: Application in preparation and postpolymerization modification of polymers. J. Polym. Sci. Polym. Chem. 2016, 54, 3057–3070. [Google Scholar] [CrossRef]
  31. Hoang, M.V.; Chung, H.-J.; Elias, A.L. Irreversible bonding of polyimide and polydimethylsiloxane (PDMS) based on a thiol-epoxy click reaction. J. Microme. Microeng. 2016, 26, 105019. [Google Scholar] [CrossRef]
  32. Shevchenko, V.; Bliznyuk, V.; Gumenna, M.; Klimenko, N.; Stryutsky, A.; Wang, J.; Binek, C.; Chernyakova, M.; Belikov, K. Coordination Polymers Based on Amphiphilic Oligomeric Silsesquioxanes and Transition Metal Ions (Co2+, Ni2+): Structure and Stimuli-Responsive Properties. Macromol. Mater. Eng. 2021, 306, 2100085. [Google Scholar] [CrossRef]
  33. Shibasaki, S.; Sasaki, Y.; Nakabayashi, K.; Mori, H. Synthesis and metal complexation of dual-functionalized silsesquioxane nanoparticles by sequential thiol–epoxy click and esterification reactions. React. Funct. Polym. 2016, 107, 11–19. [Google Scholar] [CrossRef]
  34. Lin, H.; Chen, L.; Ou, J.; Liu, Z.; Wang, H.; Dong, J.; Zou, H. Preparation of well-controlled three-dimensional skeletal hybrid monoliths via thiol–epoxy click polymerization for highly efficient separation of small molecules in capillary liquid chromatography. J. Chromatog. A 2015, 1416, 74–82. [Google Scholar] [CrossRef]
  35. Lungu, A.; Ghitman, J.; Cernencu, A.I.; Serafim, A.; Florea, N.M.; Vasile, E.; Iovu, H. POSS-containing hybrid nanomaterials based on thiol-epoxy click reaction. Polymer 2018, 145, 324–333. [Google Scholar] [CrossRef]
  36. Bekin Acar, S.; Ozcelik, M.; Uyar, T.; Tasdelen, M.A. Polyhedral oligomeric silsesquioxane-based hybrid networks obtained via thiol-epoxy click chemistry. Iran. Polym. J. 2017, 26, 405–411. [Google Scholar] [CrossRef]
  37. Lin, G.; Yin, J.; Lin, Z.; Zhu, Y.; Li, W.; Li, H.; Liu, Z.; Xiang, H.; Liu, X. Facile thiol-epoxy click chemistry for transparent and aging-resistant silicone/epoxy composite as LED encapsulant. Prog. Org. Coat. 2021, 156, 106269. [Google Scholar] [CrossRef]
  38. Murphy, Z.; Kent, M.; Freeman, C.; Landge, S.; Koricho, E. Halloysite nanotubes functionalized with epoxy and thiol organosilane groups to improve fracture toughness in nanocomposites. SN Appl. Sci. 2020, 2, 213. [Google Scholar] [CrossRef]
  39. Bera, S.; Rout, T.K.; Udayabhanu, G.; Narayan, R. Water-based & eco-friendly epoxy-silane hybrid coating for enhanced corrosion protection & adhesion on galvanized steel. Prog. Org. Coat. 2016, 101, 24–44. [Google Scholar]
  40. De, S.; Khan, A. Efficient synthesis of multifunctional polymersviathiol–epoxy “click” chemistry. Chem. Commun. 2012, 48, 3130–3132. [Google Scholar] [CrossRef]
  41. Gadwal, I.; Stuparu, M.C.; Khan, A. Homopolymer bifunctionalization through sequential thiol–epoxy and esterification reactions: An optimization, quantification, and structural elucidation study. Polym. Chem. 2015, 6, 1393–1404. [Google Scholar] [CrossRef]
  42. From, M.; Larsson, P.T.; Andreasson, B.; Medronho, B.; Svanedal, I.; Edlund, H.; Norgren, M. Tuning the properties of regenerated cellulose: Effects of polarity and water solubility of the coagulation medium. Carbohyd. Polym. 2020, 236, 116068. [Google Scholar] [CrossRef]
  43. Azizi, N.; Khajeh-Amiri, A.; Ghafuri, H.; Bolourtchian, M. LiOH-Catalyzed simple ring opening of epoxides under solvent-free conditions. Phosphorus Sulfur. 2010, 185, 1550–1557. [Google Scholar] [CrossRef]
  44. Zuo, Y.; Gou, Z.; Quan, W.; Lin, W. Silicon-assisted unconventional fluorescence from organosilicon materials. Coord. Chem. Rev. 2021, 438, 213887. [Google Scholar] [CrossRef]
  45. Huang, Z.; Zhou, C.; Chen, W.; Li, J.; Li, M.; Liu, X.; Mao, L.; Yuan, J.; Tao, L.; Wei, Y. A polymerizable Aggregation Induced Emission (AIE)-active dye with remarkable pH fluorescence switching based on benzothiazole and its application in biological imaging. Dye. Pigment. 2021, 196, 109793. [Google Scholar] [CrossRef]
  46. Razi, S.S.; Gupta, R.C.; Ali, R.; Dwivedi, S.K.; Srivastava, P.; Misra, A. A new D–π–A type intramolecular charge transfer Dyad System to detect F: Anion induced CO2 sensing. Sensors Actuat. Chem. 2016, 236, 520–528. [Google Scholar] [CrossRef]
  47. Cao, J.; Zuo, Y.; Lu, H.; Yang, Y.; Feng, S. An unconventional chromophore in water-soluble polysiloxanes synthesized via thiol-ene reaction for metal ion detection. J. Photo. Photo. A Chem. 2018, 350, 152–163. [Google Scholar] [CrossRef]
  48. Lu, H.; Zhang, J.; Feng, S. Controllable photophysical properties and self-assembly of siloxane-poly(amidoamine) dendrimers. Phys. Chem. Chem. Phys. 2015, 17, 26783–26789. [Google Scholar] [CrossRef]
  49. Fu, P.-Y.; Li, B.-N.; Zhang, Q.-S.; Mo, J.-T.; Wang, S.-C.; Pan, M.; Su, C.-Y. Thermally Activated Fluorescence vs Long Persistent Luminescence in ESIPT-Attributed Coordination Polymer. J. Am. Chem. Soc. 2022, 144, 2726–2734. [Google Scholar] [CrossRef]
  50. Lu, H.; Hu, Z.; Feng, S. Nonconventional Luminescence Enhanced by Silicone-Induced Aggregation. Chem. Asian J. 2017, 12, 1213–1217. [Google Scholar] [CrossRef]
Scheme 1. The model reaction between 1,3-bis(3-glycidoxypropyl)tetramethyldisiloxane (E1) and 1-dodecanethiol (T1).
Scheme 1. The model reaction between 1,3-bis(3-glycidoxypropyl)tetramethyldisiloxane (E1) and 1-dodecanethiol (T1).
Ijms 24 07785 sch001
Scheme 2. Synthetic routes of sulfur-containing functionalized disiloxanes from 1,3-bis(3-glycidoxypropyl)tetramethyldisiloxane (E1) and various thiol compounds (Tn).
Scheme 2. Synthetic routes of sulfur-containing functionalized disiloxanes from 1,3-bis(3-glycidoxypropyl)tetramethyldisiloxane (E1) and various thiol compounds (Tn).
Ijms 24 07785 sch002
Figure 1. (a) FT-IR spectra of epoxy-containing groups disiloxane E1 and functionalized disiloxanes E1T2, E1T9. (b) 1H NMR spectrum of E1, E1T2, and E1T9 in CDCl3. (c) 29Si NMR spectrum of E1, E1T2, and E1T9 in CDCl3.
Figure 1. (a) FT-IR spectra of epoxy-containing groups disiloxane E1 and functionalized disiloxanes E1T2, E1T9. (b) 1H NMR spectrum of E1, E1T2, and E1T9 in CDCl3. (c) 29Si NMR spectrum of E1, E1T2, and E1T9 in CDCl3.
Ijms 24 07785 g001
Figure 2. Fluorescence emission spectra of functionalized disiloxanes E1T1 to E1T11 in ethanol solutions excited at 365 nm (0.01 mol L−1).
Figure 2. Fluorescence emission spectra of functionalized disiloxanes E1T1 to E1T11 in ethanol solutions excited at 365 nm (0.01 mol L−1).
Ijms 24 07785 g002
Figure 3. The photographs of E1T1 to E1T11 and E2T6 in ethanol solutions under 365 nm UV light.
Figure 3. The photographs of E1T1 to E1T11 and E2T6 in ethanol solutions under 365 nm UV light.
Ijms 24 07785 g003
Figure 4. (a) Fluorescence emission spectra of E1T6 with different concentrations excited at 365 nm. (b) Fluorescence emission spectra of E1T11 with different concentrations excited at 365 nm.
Figure 4. (a) Fluorescence emission spectra of E1T6 with different concentrations excited at 365 nm. (b) Fluorescence emission spectra of E1T11 with different concentrations excited at 365 nm.
Ijms 24 07785 g004
Figure 5. (a) Fluorescence emission spectra of E1T6 in different solvents excited at 365 nm (0.01 mol/L). (b) Fluorescence emission spectra of E1T11 in different solvents excited at 365 nm (0.01 mol/L).
Figure 5. (a) Fluorescence emission spectra of E1T6 in different solvents excited at 365 nm (0.01 mol/L). (b) Fluorescence emission spectra of E1T11 in different solvents excited at 365 nm (0.01 mol/L).
Ijms 24 07785 g005
Scheme 3. Synthesis route of the model compound E2T6.
Scheme 3. Synthesis route of the model compound E2T6.
Ijms 24 07785 sch003
Figure 6. (a) The relative fluorescence intensity of E2T6 and E1T6 in THF solution (0.01 mol L−1). (b) Fluorescence emission spectra of E1T6, E2T6, and E2T6/MM mixture in THF solution excited at 365 nm (MM was insoluble in ethanol at room temperature, 0.01 mol L−1).
Figure 6. (a) The relative fluorescence intensity of E2T6 and E1T6 in THF solution (0.01 mol L−1). (b) Fluorescence emission spectra of E1T6, E2T6, and E2T6/MM mixture in THF solution excited at 365 nm (MM was insoluble in ethanol at room temperature, 0.01 mol L−1).
Ijms 24 07785 g006
Table 1. Optimization of the reaction condition of the thiol–epoxy reaction based on E1 and T1.
Table 1. Optimization of the reaction condition of the thiol–epoxy reaction based on E1 and T1.
Entry aCatalystCat. Loading (mol%)Time (h)Temp (°C)Conversion (%) c
1LiOH562563.1
2LiOH5122599.5
3LiOH3122580.3
4LiOH1122539.8
5TBAF b8122581.3
6TBAF b8125099.5
7TBAF b5125099.4
8TBAF b3125098.6
9TBAF b1125092.3
10TBAF b0.8125080.0
11TBAF b365096.7
12TBAF b335091.8
a Reaction conditions: E1 (1 mmol), T1 (1.25 mmol), and catalyst. The amount of catalyst was calculated based on epoxy group. b The reactions were carried out in THF. c Determined by 1H NMR.
Table 2. The performance of optimized conditions in the thiol–epoxy reaction of E1 and various thiol compounds (Tn).
Table 2. The performance of optimized conditions in the thiol–epoxy reaction of E1 and various thiol compounds (Tn).
Entry aSubstratesProductsTime (h)Conversion (%) bYields (%) c
1T1Ijms 24 07785 i001E1T1Ijms 24 07785 i002696.096.0
2T2Ijms 24 07785 i003E1T2Ijms 24 07785 i004699.995.5
3T3Ijms 24 07785 i005E1T3Ijms 24 07785 i006699.995.4
4T4Ijms 24 07785 i007E1T4Ijms 24 07785 i008699.994.3
5T5Ijms 24 07785 i009E1T5Ijms 24 07785 i010699.090.3
6T6Ijms 24 07785 i011E1T6Ijms 24 07785 i012698.095.0
7T7Ijms 24 07785 i013E1T7Ijms 24 07785 i014699.994.7
8T8Ijms 24 07785 i015E1T8Ijms 24 07785 i016699.992.3
9T9Ijms 24 07785 i017E1T9Ijms 24 07785 i018699.996.5
10T10Ijms 24 07785 i019E1T10Ijms 24 07785 i020699.993.8
11T11Ijms 24 07785 i021E1T11Ijms 24 07785 i022698.089.3
12T2E1T2399.995.5
13T3E1T3599.995.3
14T4E1T4599.993.8
15T8E1T8599.992.1
16T9E1T9399.996.4
a Reaction conditions: E1 (1 mmol), substrates (1.25 mmol), TBAF (3 mol%), THF (0.6 mL), 50 °C, 6 h. b The conversion rate of epoxy groups determined by 1H NMR. c Isolated yields and confirmed by FT-IR, NMR, HR-MS, and elemental analysis.
Table 3. The excitation and emission of E2T6 and E1T6.
Table 3. The excitation and emission of E2T6 and E1T6.
EntryEx/nmE2T6E1T6
Em/nmIntensityEm/nmIntensity
1310345299.13498069
23203572453635824
3330370212.33663534
4340381184.83862730
5350392147.94101959
6360404139.74092310
7370417131.84091951
83804301024121700
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tang, J.; Feng, S.; Wang, D. Facile Synthesis of Sulfur-Containing Functionalized Disiloxanes with Nonconventional Fluorescence by Thiol–Epoxy Click Reaction. Int. J. Mol. Sci. 2023, 24, 7785. https://doi.org/10.3390/ijms24097785

AMA Style

Tang J, Feng S, Wang D. Facile Synthesis of Sulfur-Containing Functionalized Disiloxanes with Nonconventional Fluorescence by Thiol–Epoxy Click Reaction. International Journal of Molecular Sciences. 2023; 24(9):7785. https://doi.org/10.3390/ijms24097785

Chicago/Turabian Style

Tang, Jing, Shengyu Feng, and Dengxu Wang. 2023. "Facile Synthesis of Sulfur-Containing Functionalized Disiloxanes with Nonconventional Fluorescence by Thiol–Epoxy Click Reaction" International Journal of Molecular Sciences 24, no. 9: 7785. https://doi.org/10.3390/ijms24097785

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