Abstract
SuFEx click chemistry demonstrates remarkable molecular assembly capabilities. However, the effective utilization of alkyl sulfonyl fluoride hubs in SuFEx chemistry, particularly in reactions with alcohols and primary amines, presents considerable challenges. This study pioneers an intramolecular chalcogen bonding activated SuFEx (S-SuFEx) click chemistry employing alkyl sulfonyl fluorides with γ-S as the activating group. The ChB-activated alkyl sulfonyl fluorides can react smoothly with phenols, alcohols, and amines, exhibiting enhanced reactivity compared to SO2F2. Excellent yields have been achieved with all 75 tested substrates. Pioneering the application of S-SuFEx chemistry, we highlight its immense potential in organic-inorganic linking, considering the critical role of interfacial covalent bonding in material fabrication. The S-SuFEx hub 1c, incorporating a trialkoxy silane group has been specifically designed and synthesized for organic-inorganic linking. In a simple step, 1c efficiently anchors various organic compounds onto surfaces of inorganic materials, forming functionalized surfaces with properties such as antibacterial activity, hydrophobicity, and fluorescence.
Similar content being viewed by others
Introduction
The important role of click chemistry in modern science has been recognized by the 2022 Nobel Prize in Chemistry1,2. Recently, a new generation of click chemistry, sulfur fluoride exchange (SuFEx) promoted by Sharpless, Moses and co-workers, has demonstrated impressive molecular assembly capabilities3,4,5,6,7,8,9,10,11,12,13,14,15,16,17. Benefited from the diverse collection of SuFExable hubs3,7,8,9,10,11,12,13,14, SuFEx chemistries offer tremendous scope for exploring chemical space and wide range of sulfuryl-heteroatom connectors to meet more diverse linking needs, thus enabling its rapid development as the technology of choice in organic synthesis18,19,20, drug discovery21,22,23, chemical biology24,25 and polymer chemistry26,27,28 (Fig. 1a). Typically, SuFEx chemistries require activated substrates3,29,30 or additional activators5,6,31 to broaden the SuFEx substrate scope. At the same time, intramolecular activation of sulfonyl fluorides has been relatively unexplored. Of note, the effective utilization of alkyl sulfonyl fluoride hubs in SuFEx chemistry remains elusive, particularly in their reactions with aliphatic alcohols and primary amines4,5,6. Alkyl sulfonyl fluorides are less electrophilic, which often requires high base loadings and extended reaction times to react with nucleophiles. However, alkyl sulfonyl fluorides are not well compatible with strong bases due to the facile deprotonation of α-H30,32. Therefore, significant challenges persist in effectively activating alkyl sulfonyl fluorides (Fig. 1b).
Non-covalent interactions represent a powerful and promising activation mode in organic synthesis33,34,35, among which chalcogen bonding (ChB) has been attracting more and more interest36,37,38,39,40,41. The low-lying C − S σ* called σ-holes is available for interaction with electron-donating atoms, particularly O and N41,42,43 (Fig. 1c). ChB interactions have gained the relatively wide applications in medicinal chemistry44, solid–state chemistry45, material science46, and anion recognition/transport47. Recently, ChB interactions have increasingly found applications in organic synthesis, regulating conformation and activating various functional groups, such as nitro derivatives36, carbonyl compounds48, imine groups49, and alkenes50. Notably, ChB activation provides an alternative that avoids the need for stoichiometric Lewis acids used in traditional synthetic methods38. Considering the validated use of stoichiometric Lewis acids to activate sulfonyl fluoride31, we postulated that leveraging intramolecular ChB interactions could enhance the efficiency of SuFEx chemistry.
In this study, alkyl sulfonyl fluorides bearing γ-S were designed and synthesized. γ-S will engage in intramolecular non-bonding 1,5-sulfur···fluorine interactions, enhancing the activity of sulfonyl fluoride group towards nucleophilic attack. These S-SuFEx hubs were synthesized through a straightforward thiol-ene reaction using thiols and ESF. Remarkably, our results revealed that S-SuFEx hubs did react smoothly with a wide variety of O, N-nucleophiles, exhibiting one of the broadest substrate scopes among all the reported SuFEx reactions5,6. The high reactivity of S-SuFEx chemistry will enable the facile connection between thiols and various O, N- nucleophiles using ESF as the linkage (Fig. 1d). Within the realm of potential applications of S-SuFEx chemistry, our initial emphasis was placed on the construction of organic-inorganic linkages. This focus arises from the recognized importance of interfacial covalent bonding, representing a pivotal approach for integrating the tailoring capabilities of organic molecules with the diverse properties of inorganic materials. This strategy has seen extensive utilization in the fabrication of functionalized materials51,52. Nevertheless, establishing efficient linkages between a wide range of functional organic molecules and inorganic materials remains a significant challenge. Herein, we introduced S-SuFEx hub 1c, distinguished by the presence of a trialkoxy silane group, designed and synthesized to facilitate efficient organic-inorganic linking. The sulfonyl fluoride group within 1c readily engages with a spectrum of O, N- nucleophiles through S-SuFEx chemistry. Simultaneously, its trialkoxy silane group adeptly anchors onto surfaces of inorganic materials. Employing 1c, we have achieved a one-step creation of functionalized self-assembled monolayers (SAMs) on glass surfaces. These monolayers exhibit diverse properties based on the covalently attached organic compounds, including antibacterial activity, hydrophobicity, and fluorescence. Ultimately, this protocol establishes a versatile clickable platform, offering an efficient and practical method for constructing a diverse range of customizable organic-inorganic hybrid materials (Fig. 1e).
Results
Development of intramolecular chalcogen bonding activated SuFEx click chemistry
In order to improve the reactivity of alkyl sulfonyl fluoride towards nucleophilic attack through intramolecular non-bonding 1,5-S···F interactions53, we designed the alkyl sulfonyl fluorides bearing γ-S (S-SuFEx hubs). These S-SuFEx hubs were synthesized via the robust thiol-ene reaction between readily available thiols and ESF. To evaluate the reactivity of S-SuFEx hubs, alkyl sulfonyl fluoride 1a was used as the model compound, and its analog 2a/2b without γ-S was used as the control compound. Significantly, the SuFEx reaction of 1a with phenol 4a performed well in a near quantitative yield to give the desired product 5a using 0.5 equiv. of BTMG (2-tert-butyl-1,1,3,3-tetramethylguanidine) as the base (Fig. 2a, Eq. 1). Under identical reaction conditions, 2a/2b reacted with phenol very slowly, delivering sulfonate 6a/6b in only 23%/16% yield (Fig. 2a, Eq. 2). These results exhibited that the intramolecular ChB interaction has dramatically improved the efficiency of the SuFEx reaction. Changing the thiol precursor from 1a to 1b/1c leads to a similar high reactivity. Nor Me3Si- group or (EtO)3Si- group has an impact on the reactivity of S-SuFEx reaction, and (EtO)3Si- group was well retained under the reaction conditions (Fig. 2a, Eq. 3 and 4). To further prove the effect of ChB on this reaction, 1a and 1b were oxidized to the corresponding sulfones (3a and 3b), which hindered the formation of ChB. As expected, the SuFEx reaction between 3a and 3b with phenol gave sulfonates 7 in extremely low yield (Fig. 2a, Eq. 5 and 6).
Subsequently, we explored the activation of alkyl sulfonyl fluorides through intermolecular ChB interaction. Dibutyl sulfide (1.0 equiv.) was introduced into the SuFEx reaction to act as the intermolecular ChB donor. However, despite this addition, there was no observed enhancement in the reaction efficiency between 2b/3b and phenol (Fig. 2a, Eq. 7 and 8). This outcome unequivocally highlights that the efficiency of SuFEx reactions can only be augmented through intramolecular ChB interaction. To compare the reactivity of SO2F2, 1b, and 2b in SuFEx reactions, we monitored the variation in phenol concentration over time for these three reactions (Fig. 2b, Supplementary Table 2). The results indicated that 1b exhibited significantly higher reactivity compared to 2b and even surpassed SO2F2, widely acknowledged as one of the most efficient SuFEx hubs with excellent reactivity towards phenols3. These findings substantiate the effectiveness of ChB activation in S-SuFEx chemistry as well.
Design of S-SuFEx click hub for organic-inorganic linking
After establishing the S-SuFEx chemistry, we devised S-SuFEx hub 1c to enable efficient covalent organic-inorganic linkages. The chemical structure of S-SuFEx hub 1c comprises three functional moieties: a sulfonyl fluoride group, a ChB donor, and an anchoring group (Fig. 3). The sulfonyl fluoride group reacts with a variety of O, N-nucleophiles, forming a covalent bond that links the organic compound to the hub. Concurrently, the ChB donor enhances the reactivity of the sulfonyl fluoride through intramolecular non-bonding 1,5-S···F interactions, initiating the S-SuFEx reaction. The trialkoxy silane group can effectively anchor the organic moiety linked to the hub onto the inorganic surface by the formation of stable Si–O–X covalent bonds, as its capability to react with hydroxyl groups on the surface of inorganic materials has been extensively validated54. The synthesis of 1c has been achieved via a straightforward thiol-ene reaction between a prevalent silicon coupling agent, 3-mercaptopropyltriethoxysilane, and ESF.
We continued to explore the reactivity of 1c towards different O, N-nucleophiles. Given the high reactivity of trialkoxy silane group towards water, its analog 1b with a trimethyl silane group was employed as the model compound for the study to avoid the hydrolysis of the trialkoxy silane group during the aqueous workup process.
The S-SuFEx chemistry with phenols
We initiated our exploration by examining the reaction between S-SuFEx hub 1b with phenols. Upon optimizing the reaction conditions (see Supplementary Section 2.1.5), it was determined that the addition of 0.5 equiv. of BTMG represented the most favorable condition for triggering the reaction. No supplementary silicon additives or elevated temperatures, typically employed in accelerated SuFEx click chemistry, were deemed necessary. The required amount of BTMG could be further reduced by employing a sacrificial base, such as triethylamine (see Supplementary Section 2.1.8). The substrate scope investigation revealed the successful reactivity of the S-SuFEx hub with a wide range of phenols (Fig. 4). Phenols, whether containing electron-donating or electron-withdrawing groups on the aromatic ring, demonstrated comparable reactivity and were smoothly converted into sulfonates 8. Various functional groups, including fluoride (8d), chloride (8e), hydroxy (8j), amide (8k, 8q), aldehyde (8 l), carboxylate (8 m, 8w), ester (8o), ketone (8 v), and alkenyl (8w), at the p, o, m-positions, were well-tolerated in the reaction. Significantly, this reaction also highlights the exceptional chemoselectivity of phenols compared to anilines (8k, 8t) and alcohols (8j). Regarding bisphenols (8z, 8aa, 8ag, 8ah), the precise control of monosulfonate or bissulfonate formation is achievable by adjusting the amount of sulfonyl fluoride added. Moreover, in order to highlight the potential applicability of this synthetic protocol, fluorescein (8af) and a selection of bioactive phenols, such as acetaminophen (8n), triclosan (8ab), vitamin E (8ac), and estrone (8ae) were subjected to the S-SuFEx reaction with 1b, resulting in the production of the desired products with excellent yields ranging from 81% to 99%. Additionally, the reaction between 1b and 4a using DCM/H2O (v/v, 4/1) as the solvent produced an 85% yield of 8a. Even when employing a PBS buffer solution (pH = 7.4) as the solvent for the reaction between 1b and 4a, a moderate yield of 48% was still achieved (see Supplementary Section 2.1.9).
The S-SuFEx chemistry with aliphatic alcohols
The SuFEx reactions involving alkyl sulfonyl fluorides and alkyl alcohols present a more significant challenge than those with phenols, which have been notably elusive within the reported SuFEx chemistry. However, this investigation explored the compatibility of S-SuFEx chemistry with alcohols. Remarkably, the S-SuFEx reactions of 1b with various alcohols exhibited excellent reactivity (Fig. 5). Primary alcohols (10a–10d) and more sterically hindered secondary alcohols (10e) all gave excellent yields ranging from 72% to 97%. Notably, primary alcohols exhibited relatively higher reactivity compared to the more hindered secondary and tertiary alcohols (see Supplementary Section 7.8). Cyclic enols (10 f, 10i) also reacted smoothly under the reaction conditions. The excellent chemoselectivity of S-SuFEx chemistry was once again exemplified by its remarkable tolerance of various functional groups (10c, 10e, 10 f) and the distinct selectivity of alcohols over primary amines (10b). In the case of amino alcohol 9b, only the target product 10b derived from the reaction between the hydroxyl group and sulfonyl fluoride was detected, with the remaining mass balance accounted for by the unreacted starting material.
The S-SuFEx chemistry with amines
N-nucleophiles constitute a wide spectrum of bioactive compounds. Harmonious with N-nucleophiles, click chemistry will offer substantial potential for the progression of functionalized materials and novel drugs. In contrast to the enduring challenges posed by conventional SuFEx chemistry when dealing with N-nucleophiles, particularly primary amines, our investigation revealed the remarkable efficacy of the innovative S-SuFEx chemistry in facilitating smooth reactions with primary, secondary, and cyclic amines, anilines, as well as diverse N-heteroaromatic compounds (Fig. 6). Generally, aliphatic amines showed higher reactivity compared to anilines, with secondary amines proving to be the most reactive N-nucleophiles in S-SuFEx chemistry. High yields were achieved with both linear (12a-12c) and cyclic (12d–12i) secondary amines. In the case of amino alcohol (12 f), S-SuFEx chemistry exhibited higher selectivity toward cyclic secondary amines compared to secondary alcohols, resulting in a moderate yield of 64%. Although primary amines displayed slightly lower reactivity than secondary amines, the S-SuFEx reactions with primary amines, including sterically hindered primary amines (12k) and benzyl amines (12l–12q), also gave high yields by extending the reaction time to 12 h. Anilines, while less nucleophilic than aliphatic amines, also proved effective in S-SuFEx chemistry when the reaction time was further extended to 24 h\. Anilines with electron-donating MeO- groups (12w) yielded a perfect 94%, while anilines with electron-withdrawing groups (12x, 12 y) resulted in slightly lower yields. Additionally, N-heteroaromatic compounds, such as pyrrole (12aa), indoles (12ab), imidazoles (12ad) and benzimidazole (12ae), also exhibited high compatibility with S-SuFEx chemistry. To further underscore the potential applications of S-SuFEx chemistry in constructing bioactive functionalized materials, we extended our testing to a series of drugs and pesticides, including amlodipine (12 v), melatonin (12ac), and carbendazim (12ae). Remarkably, all of them yielded perfect sulfonylation products, with sensitive functional groups like ester and amide remaining entirely unaffected. These results demonstrate the potential of S-SuFEx chemistry for broad synthetic applications.
Organic-inorganic linking via S-SuFEx click chemistry
Upon establishing the pronounced reactivity of the S-SuFEx click hub with various O, N-nucleophiles, we proceeded to apply click hub 1c for the creation of functionalized SAMs on the surface of inorganic materials via organic-inorganic linking. Glass slides were selected as the exemplar inorganic substrates. To prepare functionalized SAM, we initiated the process by coupling 1c with the selected organic compounds using S-SuFEx chemistry. Notably, 1c displayed a slightly higher reactivity than its analog 1b, reacting smoothly with phenols, alcohols, and amines in the presence of 0.2 equiv. BTMG. The reactions between 1c and the 15 tested substrates gave 76–99% yields (Fig. 7a). In a typical approach for the preparation of SAM, the reaction mixture of 1c and organic compounds can be directly employed to react with the glass surface without the need for purification. Covalent linkage between the glass and selected organic compounds is achieved by simply immersing the glass slides in the hydrolysate of the reaction mixture for 24 h (Fig. 7b). The hydrolyzed trialkoxy silane group effectively anchors the organic moiety onto the glass surface, forming SAMs. In this study, we examined the linkage of four representative organic compounds to the glass surface, namely, triclosan (TC), fluorescein (FR), 1H,1H,2H,2H-perfluoro-1-decanol (PD), and nonylphenol (NP).
The glass slides linked with TC, FR, PD and NP displayed no significant changes in appearance, remaining colorless and transparent. However, the contact angles of water on their surfaces were markedly changed (Fig. 8 and Supplementary Fig. 8). Atomic force microscopy (AFM) analysis revealed that both the pristine slide and the TC linked slides exhibited a rough surface with irregularities and protuberances. The subsequent AFM-IR spectra provided further confirmation of the covalently linked TC molecules on the glass surfaces (Fig. 8b, c). TC is a widely recognized antimicrobial agent. Covalently linking TC to the glass surface was expected to render it antibacterial. Indeed, the antibacterial tests confirmed that the SAM exhibited excellent antibacterial efficacy against Staphylococcus aureus, with a remarkable reduction in the counts of S. aureus, resulting in an antibacterial efficacy (R) of 4.59 (Fig. 9a and Supplementary Fig. 9). Furthermore, linking FR to the glass surface using the S-SuFEx hub 1c resulted in a glass surface with fluorescence, as confirmed by fluorescence microscope images (Fig. 9b and Fig. Supplementary Fig. 10). This provides an effective method for preparing inorganic materials with fluorescence labeling. Functionalizing the glass surface with PD and NP rendered it hydrophobic. The highest hydrophobicity of 98.5 degree was achieved with PD functionalized glass surfaces (Fig. 9c). As expected, NP-functionalized glass surfaces also exhibited excellent antibacterial properties against S. aureus (Supplementary Fig. 9). Using FR-functionalized glass as the model, we tested the stability of this organic-inorganic linking system. The results indicated that the system exhibited excellent stability, with robust tolerance to acid (pH = 2) (Supplementary Fig. 11), high temperature (Supplementary Fig. 12) or flushing with organic solvents (Supplementary Fig. 13). However, the functionalized glass slides showed poor base tolerance, which was primarily caused by the dissolution of Si–O–Si network structure in a strong alkaline environment55. Furthermore, we specifically attempted to immobilize the peptide oxytocin on the glass surface using 1c (Supplementary Fig. 14). The successful outcomes of the peptide immobilization process suggested that this clickable platform may have significant application value in biological detection. In summary, the covalent linkage of various functional organic molecules and inorganic materials has been successfully established, showcasing the general versatility and simplicity of the organic-inorganic linking strategy using 1c. This approach generates a versatile platform for achieving any desired organic-inorganic linking.
Mechanistic investigations
A series of experiments were conducted to elucidate the mechanism of S-SuFEx click chemistry, using the SuFEx reaction between 1b and phenol 4a as the model reaction. Initial investigations focused on the base’s role in this process, as illustrated in Fig. 10. The absence of BTMG resulted in no observable reaction between 1b and 4a, underscoring the indispensable role of BTMG (Fig. 10a). With BTMG present, two potential reaction mechanisms were considered. The first hypothesis suggests that the base deprotonates the α-position of 1b to form intermediate 15, which then reacts with nucleophile 4a to yield the final product 8a through a deprotonation-elimination-addition mechanism, involving sulfene 16 as the reactive intermediate (Fig. 10d). Alternatively, BTMG might first interact with 4a to produce deprotonated phenol 17, which could then undergo an SN2 reaction with 1b to form the final product (Fig. 10e). To determine the prevailing mechanism, further experiments were conducted. Introduction of 0.5 equiv. of BTMG to 1b in MeCN-d3 revealed partial deprotonation at the α-position to give 15 through 1H NMR analysis (Fig. 10b and Supplementary Fig. 16), yet without evidence of S-F bond dissociation according to the 19F NMR (Fig. 10c), indicating the absence of sulfene 16 formation. Further adding 1.0 equiv. of phenol 4a to the deprotonated alkyl sulfonyl fluoride 15 and stirring for 2 hours led to no detectable product 8a, suggesting the reaction between alkyl sulfonyl fluorides and BTMG was a non-productive side pathway. Conversely, introducing BTMG to a MeCN-d3 solution of 4a induced a significant upfield shift in the aromatic region of the 1H NMR spectrum (Supplementary Fig. 17), indicative of the formation of 175. Addition of 1b to the solution of 17 resulted in a quantitative yield of the product 8a (Fig. 10b), strongly supporting the SN2 pathway between 1b and 17 as the primary reaction route. Considering the high yields achieved with most substrates, it is hypothesized that the side reaction between 1b and BTMG progresses much more slowly than the favorable SN2 reaction.
Despite the generally slow SN2 reactions observed with most alkyl sulfonyl fluorides, we propose that the presence of an intramolecular ChB could significantly increase the electrophilicity of 1b, thereby enhancing its reactivity towards nucleophiles. To validate our hypothesis, we undertook computational studies. The initial step involved determining the conformation of 1b using DFT computations at the M062X/6-311 G(d,p) level. Our findings indicated that the gauche conformation of 1b is only marginally less stable than its anti counterpart, with a free energy barrier of approximately 1.0 kcal/mol between them (Supplementary Fig. 18). This barrier is sufficiently low to allow for easy interconversion between these conformations at room temperature. Further analysis was focused on the electrostatic potential map of the gauche conformation (Fig. 10f), revealing distinct electrostatic interactions between the negatively charged fluorine and the positively charged sulfur atoms. We further computationally measured the distance and strength of the ChB between S and F in 1b using NBO analysis. The measured distance was 3.11 Å, which is less than the sum of their van der Waals radii (∑rvdW (S···F) = 3.27 Å)56. The strength of this interaction was measured to be 5.37 kcal/mol (Fig. 10g). These computational insights strongly advocate for the existence of an intramolecular ChB between sulfur and fluorine in the gauche conformation. The ChB will make the fluorine a better leaving group and thus accelerating the S-SuFEx chemistry.
Furthermore, inspired by the discovery of BTMG as an exceptional SuFEx catalyst by Moses et al.5, we hypothesized that the phenoxide guanidinium complex 17 might also facilitate sulfonyl fluoride group activation via hydrogen bonding with fluorine. Empirical validation involved direct interaction of 1b with sodium phenoxide 4a’, achieving an impressive 87% yield of 8a (Fig. 10h, Eq. 9), albeit slightly lower than the 99% yield obtained with 4a and BTMG (Fig. 10h, Eq. 3). Conversely, the reaction of 3b with sodium phenoxide yielded a mere 31% of 8a (Fig. 10h, Eq. 10), underscoring the intramolecular ChB as the predominant factor in alkyl sulfonyl fluoride activation, with the phenoxide guanidinium complex playing a supportive role in this activation process.
Given the established activation of alkyl sulfonyl fluorides through non-bonding 1,5-S···F interactions, we explored the possibility of achieving a similar activation effect via 1,6-S···F interactions. Accordingly, we synthesized compound 18, using 1-propanethiol, and 2-propene-1-sulfonyl fluoride as precursors and 2,2’-azobis(2-methylpropionitrile) as the reagent. The reaction between 18 and 4a, in the presence of 0.5 equiv. of BTMG, gave an exceptional 99% yield of 19 within 0.5 h (Fig. 7i, Eq. 11), demonstrating the feasibility of ChB activation by 1,6-S···F interactions. The Markovnikov addition product 20 and 4a also exhibited high reactivity towards nucleophile 4a, achieving a 99% yield using 0.5 equiv. of BTMG within 0.5 h (Fig. 10i, Eq. 12). Furthermore, the reactivity of phenyl fluorosulfonate could also be moderately enhanced by a 2-thio group via a 1,6-S···F interaction (see Supplementary Section 3.7). These outcomes collectively underscore the broader applicability of the intramolecular ChB-activated S-SuFEx reaction, demonstrating its potential for diverse synthetic applications.
Discussion
In summary, we present the intramolecular SuFExability activation strategy of alkyl sulfonyl fluorides using γ-S atom. The introduction of γ-S may take advantage of intramolecular chalcogen bonding to active sulfonyl fluoride unit and facilitate attack of O, N-nucleophiles. The activated SuFEx reactivity of alkyl sulfonyl fluorides is achieved with relatively low base loadings under mild and simple conditions while circumventing the need to prepare silyl-ether substrates or additional stoichiometric activators/additives. Thus, we realize the efficient SuFEx process between alkyl sulfonyl fluorides and O, N-nucleophiles with great abundance and structural diversity. This intramolecular activation strategy offers a potentially general and practically sustainable approach for straightforward transformation to useful molecules, which is a valuable addition to the synthetic toolkit. Significantly, the S-SuFEx hub 1c, which incorporates a trialkoxy silane group, has been established as a universal clickable platform for the highly efficient covalent linkage of organic compounds and inorganic materials. Its application was showcased in creating SAMs that impart glass surfaces with antibacterial activity, hydrophobicity, and fluorescence. In our ongoing investigation into the potential applications of S-SuFEx chemistry, we will focus on an in-depth examination of utilizing ESF as a linking agent for the efficient coupling of thiols with O- and N-nucleophiles. The strategy reported herein may open a new avenue in SuFEx chemistries.
Methods
Synthesis of alkyl sulfonyl fluorides bearing γ-S
A 50 mL round bottom flask was charged with ESF (2.2 g, 20 mmol), thiol (20 mmol, 1.0 equiv.) in THF (20 mL), then added triethylamine (0.1 g, 1 mmol) dropwise. The reaction mixture was stirred at room temperature for 3 h. After completion of the reaction, the solvent was removed directly under reduced pressure. The crude product was purified by a flash column chromatography eluted with petroleum ether PE/EA to afford the products 2-(thiol)ethane-1-sulfonyl fluoride 1.
General procedure for the sulfonylation with O, N-nucleophiles
A 10 mL round bottom flask was charged with 1 (0.3 mmol, 1 equiv.), O, N-nucleophile (0.3 mmol, 1 equiv.) in MeCN (1 mL), then added BTMG (0.5 equiv.). The reaction mixture was stirred at room temperature. When the reaction was completed, the solvent was concentrated in vacuo. The residue was purified via a flash chromatography to afford the pure product.
General procedure for covalently linking organic compounds and inorganic materials
A 10 mL round bottom flask was charged with 1c (0.3 mmol, 1 equiv.), 13 (0.3 mmol, 1 equiv.) in anhydrous MeCN (1 mL), then added BTMG (0.06 mmol, 0.2 equiv.). The reaction mixture was stirred at room temperature for 2 h. When the reaction was completed, 100 μL of reaction mixture was added to the hydrolysate (EtOH/H2O, 10:1 v/v, 10 mL), and the pH was adjusted to 4.0 by acetic acid. Subsequently, the hydroxylated glass substrates were immersed to the solution and incubated for 24 h at room temperature. The treated cover slides were sonicated for 10 min in deionized water and then in acetone, ethanol and dried under the Ar flow.
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Data availability
Experimental data as well as characterization data for all new compounds prepared in the course of these studies are provided in the Supplementary Information. The data supporting the findings of this study are available within the main text and the supplementary information. Source data are provided in this paper. All data are available from the corresponding author upon request. Source data are provided with this paper.
References
Kolb, H. C., Finn, M. G. & Sharpless, K. B. Click chemistry: diverse chemical function from a few good reactions. Angew. Chem. Int. Ed. 40, 2004–2021 (2001).
Finn, M. G., Kolb, H. C. & Sharpless, K. B. Click chemistry connections for functional discovery. Nat. Synth. 1, 8–10 (2022).
Dong, J., Krasnova, L., Finn, M. G. & Sharpless, K. B. Sulfur(VI) fluoride exchange (SuFEx): another good reaction for click chemistry. Angew. Chem. Int. Ed. 53, 9430–9448 (2014).
Lou, T. S. B. & Willis, M. C. Sulfonyl fluorides as targets and substrates in the development of new synthetic methods. Nat. Rev. Chem. 6, 146–162 (2022).
Smedley, C. J. et al. Accelerated SuFEx click chemistry for modular synthesis. Angew. Chem. Int. Ed. 61, e202112375 (2022).
Wei, M. et al. A broad-spectrum catalytic amidation of sulfonyl fluorides and fluorosulfates. Angew. Chem. Int. Ed. 60, 7397–7404 (2021).
Li, S., Wu, P., Moses, J. E. & Sharpless, K. B. Multidimensional suFEx click chemistry: sequential sulfur(VI) fluoride exchange connections of diverse modules launched from an SOF4 hub. Angew. Chem. Int. Ed. 56, 2903–2908 (2017).
Smedley, C. J. et al. 1-Bromoethene-1-sulfonyl fluoride (BESF) is another good connective hub for SuFEx click chemistry. Chem. Commun. 54, 6020–6023 (2018).
Guo, T. et al. A new portal to SuFEx click chemistry: a stable fluorosulfuryl imidazolium salt emerging as an “F-SO2 +” donor of unprecedented reactivity, selectivity, and scope. Angew. Chem. Int. Ed. 57, 2605–2610 (2018).
Smedley, C. J. et al. Diversity oriented clicking (DOC): divergent synthesis of SuFExable pharmacophores from 2-substituted-alkynyl-1-sulfonyl fluoride (SASF) hubs. Angew. Chem. Int. Ed. 59, 12460–12469 (2020).
Sun, S., Gao, B., Chen, J., Sharpless, K. B. & Dong, J. Fluorosulfuryl Isocyanate Enabled SuFEx Ligation of Alcohols and Amines. Angew. Chem. Int. Ed. 60, 21195–21199 (2021).
Li, B. Y. et al. Ex situ generation of thiazyl trifluoride (NSF3) as a gaseous SuFEx hub. Angew. Chem. Int. Ed. 62, e202305093 (2023).
Smedley, C. J., Giel, M. C., Fallon, T. & Moses, J. E. Ethene-1,1-disulfonyl difluoride (EDSF) for SuFEx click chemistry: synthesis of SuFExable 1,1-bissulfonylfluoride substituted cyclobutene hubs. Angew. Chem. Int. Ed. 62, e202303916 (2023).
Wang, P., Zhang, H., Nie, X., Xu, T. & Liao, S. Photoredox catalytic radical fluorosulfonylation of olefins enabled by a bench-stable redox-active fluorosulfonyl radical precursor. Nat. Commun. 13, 3370 (2022).
Zeng, D., Deng, W. P. & Jiang, X. Advances in the construction of diverse SuFEx linkers. Natl. Sci. Rev. 10, nwad123 (2023).
Homer, J. A. et al. Sulfur fluoride exchange. Nat. Rev. Methods Primers 3, 58 (2023).
Barrow, A. S. et al. The growing applications of SuFEx click chemistry. Chem. Soc. Rev. 48, 4731–4758 (2019).
Lee, C. et al. The emerging applications of sulfur(VI) fluorides in catalysis. ACS Catal 11, 6578–6589 (2021).
Nielsen, M. K., Ugaz, C. R., Li, W. & Doyle, A. G. PyFluor: a Low-cost, stable, and selective deoxyfluorination reagent. J. Am. Chem. Soc. 137, 9571–9574 (2015).
Nielsen, M. K., Ahneman, D. T., Riera, O. & Doyle, A. G. Deoxyfluorination with sulfonyl fluorides: navigating reaction space with machine learning. J. Am. Chem. Soc. 140, 5004–5008 (2018).
Meng, G. et al. Modular click chemistry libraries for functional screens using a diazotizing reagent. Nature 574, 86–89 (2019).
Brighty, G. J. et al. Using sulfuramidimidoyl fluorides that undergo sulfur(VI) fluoride exchange for inverse drug discovery. Nat. Chem. 12, 906–913 (2020).
Cheng, Y. et al. Diversity oriented clicking delivers β-substituted alkenyl sulfonyl fluorides as covalent human neutrophil elastase inhibitors. Proc. Natl. Acad. Sci. 119, e2208540119 (2022).
Zhao, Q. et al. Broad-spectrum kinase profiling in live cells with lysine-targeted sulfonyl fluoride probes. J. Am. Chem. Soc 139, 680–685 (2017).
Liu, F. et al. Biocompatible SuFEx click chemistry: thionyl tetrafluoride (SOF4)-derived connective hubs for bioconjugation to DNA and proteins. Angew. Chem. Int. Ed. 58, 8029–8033 (2019).
Li, S. et al. SuFExable polymers with helical structures derived from thionyl tetrafluoride. Nat. Chem. 13, 858–867 (2021).
Subramaniam, M., Ruggeri, F. S. & Zuilhof, H. Degradable click-reaction-based polymers as highly functional materials. Matter 5, 2490–2492 (2022).
Li, H. et al. High-performing polysulfate dielectrics for electrostatic energy storage under harsh conditions. Joule 7, 95–111 (2023).
Liang, D. D. et al. Silicon-free SuFEx reactions of sulfonimidoyl fluorides: scope, enantioselectivity, and mechanism. Angew. Chem. Int. Ed. 59, 7494–7500 (2020).
Gao, B. et al. Bifluoride-catalysed sulfur(VI) fluoride exchange reaction for the synthesis of polysulfates and polysulfonates. Nat. Chem. 9, 1083–1088 (2017).
Mukherjee, P. et al. Sulfonamide synthesis via calcium triflimide activation of sulfonyl fluorides. Org. Lett. 20, 3943–3947 (2018).
Shirota, Y., Nagai, T. & Tokura, N. Reactions of alkanesulphonic acid derivatives with organoalkali metal compounds: formation and reactivity of α-sulphonylcarbanions. Tetrahedron 25, 3193–3204 (1969).
Gutiérrez López, M. Á., Tan, M. L., Frontera, A. & Matile, S. The origin of anion-π autocatalysis. JACS Au 3, 1039–1051 (2023).
Banik, S. M., Levina, A., Hyde, A. M. & Jacobsen, E. N. Lewis acid enhancement by hydrogen-bond donors for asymmetric catalysis. Science 358, 761–764 (2017).
Sutar, R. L. & Huber, S. M. Catalysis of organic reactions through halogen bonding. ACS Catal 9, 9622–9639 (2019).
Wonner, P., Dreger, A., Vogel, L., Engelage, E. & Huber, S. M. Chalcogen bonding catalysis of a nitro-michael reaction. Angew. Chem. Int. Ed. 58, 16923–16927 (2019).
Liu, J., Zhou, M., Deng, R., Zheng, P. & Chi, Y. R. Chalcogen bond-guided conformational isomerization enables catalytic dynamic kinetic resolution of sulfoxides. Nat. Commun. 13, 4793 (2022).
Zhu, H., Zhou, P. P. & Wang, Y. Cooperative chalcogen bonding interactions in confined sites activate aziridines. Nat. Commun. 13, 3563 (2022).
Bitai, J., Nimmo, A. J., Slawin, A. M. Z. & Smith, A. D. Cooperative palladium/isothiourea catalyzed enantioselective formal (3+2) cycloaddition of vinylcyclopropanes and α,β-unsaturated esters. Angew. Chem. Int. Ed. 61, e202202621 (2022).
Lu, Y., Liu, Q., Wang, Z. X. & Chen, X. Y. Alkynyl sulfonium salts can be employed as chalcogen-bonding catalysts and generate alkynyl radicals under blue-light irradiation. Angew. Chem. Int. Ed. 61, e202116071 (2022).
Benz, S. et al. Catalysis with chalcogen bonds. Angew. Chem. Int. Ed. 56, 812–815 (2017).
Vogel, L., Wonner, P. & Huber, S. M. Chalcogen bonding: an overview. Angew. Chem. Int. Ed. 58, 1880–1891 (2019).
Pascoe, D. J., Ling, K. B. & Cockroft, S. L. The origin of chalcogen-bonding interactions. J. Am. Chem. Soc. 139, 15160–15167 (2017).
Beno, B. R., Yeung, K. S., Bartberger, M. D., Pennington, L. D. & Meanwell, N. A. A survey of the role of noncovalent sulfur interactions in drug design. J. Med. Chem. 58, 4383–4438 (2015).
Werz, D. B. et al. Self-organization of chalcogen-containing cyclic alkynes and alkenes to yield columnar structures. Org. Lett. 4, 339–342 (2002).
Gleiter, R., Haberhauer, G., Werz, D. B., Rominger, F. & Bleiholder, C. From noncovalent chalcogen-chalcogen interactions to supramolecular aggregates: experiments and calculations. Chem. Rev. 118, 2010–2041 (2018).
Biot, N. & Bonifazi, D. Chalcogen-bond driven molecular recognition at work. Coord. Chem. Rev. 413, 213243 (2020).
Wang, W. et al. Chalcogen-chalcogen bonding catalysis enables assembly of discrete molecules. J. Am. Chem. Soc. 141, 9175–9179 (2019).
He, X. et al. Bis-selenonium cations as bidentate chalcogen bond donors in catalysis. ACS Catal 11, 12632–12642 (2021).
Zhang, Q., Chan, Y. Y., Zhang, M., Yeung, Y. Y. & Ke, Z. Hypervalent chalcogenonium···π bonding catalysis. Angew. Chem. Int. Ed. 61, e202208009 (2022).
Al Zoubi, W. et al. Recent advances in hybrid organic-inorganic materials with spatial architecture for state-of-the-art applications. Prog. Mater. Sci. 112, 100663 (2020).
Zhou, C. et al. Hybrid organic-inorganic two-dimensional metal carbide MXenes with amido- and imido-terminated surfaces. Nat. Chem. 15, 1722–1729 (2023).
Young, C. M. et al. The importance of 1,5-oxygen···chalcogen interactions in enantioselective isochalcogenourea catalysis. Angew. Chem. Int. Ed. 59, 3705–3710 (2020).
Wang, L., Schubert, U. S. & Hoeppener, S. Surface chemical reactions on self-assembled silane based monolayers. Chem. Soc. Rev. 50, 6507–6540 (2021).
Dove, P. M., Han, N., Wallace, A. F. & De Yoreo, J. J. Kinetics of amorphous silica dissolution and the paradox of the silica polymorphs. Proc. Natl. Acad. Sci. 105, 9903–9908 (2008).
Ho, P. C., Wang, J. Z., Meloni, F. & Vargas-Baca, I. Chalcogen bonding in materials chemistry. Coordin. Chem. Rev. 422, 213464 (2020).
Acknowledgements
This work was financially supported by the National Key Research and Development Program of China (2022YFF0710402, J.A.).
Author information
Authors and Affiliations
Contributions
J.A. and M.W. designed the experiments. J.A. and F.R. directed the project. M.W., J.H., C.W., Q.D. and X.Z. conducted the experiments. H.D. performed the DFT calculations. M.W., H.D., Y.D., C.W., L.W. and K.N. contributed to the data analysis and visualization. J.A. and M.W. prepared the paper with contribution from all authors.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Jianguo Lin and the other anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Wang, M., Hou, J., Do, H. et al. Intramolecular chalcogen bonding activated SuFEx click chemistry for efficient organic-inorganic linking. Nat Commun 15, 6849 (2024). https://doi.org/10.1038/s41467-024-50922-9
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-024-50922-9
Comments
By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.