Next Article in Journal
Investigation of Photosystem II Functional Size in Higher Plants under Physiological and Stress Conditions Using Radiation Target Analysis and Sucrose Gradient Ultracentrifugation
Next Article in Special Issue
Microwave-Accelerated McKenna Synthesis of Phosphonic Acids: An Investigation
Previous Article in Journal
Fast and Sensitive HPLC-ESI-MS/MS Method for Etoricoxib Quantification in Human Plasma and Application to Bioequivalence Study
Previous Article in Special Issue
Photoinduced Bisphosphination of Alkynes with Phosphorus Interelement Compounds and Its Application to Double-Bond Isomerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Ortho-Phosphinoarenesulfonamide-Mediated Staudinger Reduction of Aryl and Alkyl Azides

1
PCFM Lab and GDHPRC Lab, School of Chemistry, Sun Yat-sen University, Guangzhou 510275, China
2
Center for Nephrology and Clinical Metabolomics and Division of Nephrology and Rheumatology, Shanghai Tenth People’s Hospital, School of Medicine, Tongji University, Shanghai 200072, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(17), 5707; https://doi.org/10.3390/molecules27175707
Submission received: 29 July 2022 / Revised: 19 August 2022 / Accepted: 22 August 2022 / Published: 5 September 2022
(This article belongs to the Special Issue Modern Organophosphorus Chemistry)

Abstract

:
Conventional Staudinger reductions of organic azides are sluggish with aryl or bulky aliphatic azides. In addition, Staudinger reduction usually requires a large excess of water to promote the decomposition of the aza-ylide intermediate into phosphine oxide and amine products. To overcome the challenges above, we designed a novel triaryl phosphine reagent 2c with an ortho-SO2NH2 substituent. Herein, we report that such phosphine reagents are able to mediate the Staudinger reduction of both aryl and alkyl azides in either anhydrous or wet solvents. Good to excellent yields were obtained in all cases (even at a diluted concentration of 0.01 M). The formation of B-TAP, a cyclic aza-ylide, instead of phosphine oxide, eliminates the requirement of water in the Staudinger reduction. In addition, computational studies disclose that the intramolecular protonation of the aza-ylide by the ortho-SO2NH2 group is kinetically favorable and responsible for the acceleration of Staudinger reduction of the aryl azides.

1. Introduction

Staudinger reduction [1,2,3] is one of the most common transformations used to prepare amino compounds in organic synthesis. Recently, the biorthogonal property [4,5,6,7,8,9,10,11] of organic azides [12] has significantly enhanced the visibility of the Staudinger reaction in the context of chemical biology and medicinal chemistry. The classic Staudinger protocol employs trivalent phosphines, predominately triphenylphosphine, to mediate the reduction of organic azides under aqueous conditions via an aza-ylide (or iminophosphorane) intermediate [13,14] (Scheme 1A). Although thermodynamically the formation of triphenylphosphine oxide provides a sufficient driving force for the Staudinger reduction, reactions are sluggish when the iminophosphorane intermediates are kinetically stable [15,16,17]. For example, iminophosphorane intermediates formed from aryl azides can be quite stable under neutral conditions, thus requiring either acids or bases to mediate their hydrolysis [18,19]. In addition, the amount of water in the system has a significant impact on the rate of Staudinger reduction, particularly for highly hydrophobic substrates. Recently, Ito, Abe, and coworkers [20] reported that o-(diphenylphosphino) benzamide accelerated the reduction of aryl and alkyl azides towards the corresponding amines (Scheme 1B). In their report, aqueous conditions were still required for the reduction of aromatic azides. The authors indicated that a significant amount of by-product was formed without the addition of water. It therefore remains highly desirable to design novel phosphorus reagents that are able to mediate a broad-spectrum Staudinger reduction without the restriction of an aqueous solvent. Very recently, our group prepared a novel class of cyclic iminophosphoranes, namely (benzo[d])-1,2,3-thiazaphosphole (B-TAP) [21]. The improved stability of the B-TAP heterocycle towards hydrolysis suggests that the stronger P=N of B-TAP bond may drive a Staudinger reduction as the P=O of phosphine oxide did. Herein, we report the use of ortho-phosphinoarenesulfonamides as novel reagents to mediate the Staudinger reduction of both aryl and alkyl azides under either anhydrous or aqueous conditions (Scheme 1C).

2. Results and Discussions

The B-TAP heterocycle was originally synthesized in our laboratory by the annulation of ortho-phosphinoarenesulfonyl fluorides with commercially available trimethylsilyl azide [21]. Further studies indicated that B-TAP can also be prepared by the oxidative condensation of ortho-phosphinoarenesulfonamide. We speculated that the formation of B-TAP could be achieved by the use of an organic azide as an oxidant. This hypothesis led us to explore ortho-phosphinoarenesulfonamide-mediated Staudinger reductions. We proposed that the introduction of an acidic sulfonamide group at the ortho position would facilitate the decomposition step of the original aza-ylide and thus the addition of water would not be necessary. In this way, the formation of a B-TAP heterocycle with a more stable P=N bond would provide the driving force for Staudinger reduction.
We selected methyl 4-azidobenzoate (1a) as a model substrate (Table 1). As expected, when 2a was applied the major product was the triphenylphosphine-derived aza-ylide while the characteristic signals of methyl 4-aminobenzoate (3a) were not observed in the 1H NMR (entry 1, Table 1). Interestingly, ortho-phosphinoarenesulfonamide (2b and 2c)-mediated reductions were both complete within 3 h. Reductions at lower concentration (0.01 M) with shorter reaction times indicated that 2c gave faster rates than 2b (entries 4–5, Table 1) and therefore 2c was chosen for further investigation. The 2c-mediated Staudinger reduction was found compatible with a broad range of organic solvents (entries 8–12, Table 1). The rate of 2c-mediated Staudinger reduction in aqueous solvent (entry 13, Table 1) was similar to anhydrous solvents. This suggested that the addition of water did not affect the Staudinger reduction. This is in sharp contrast to other known staudinger reduction processes where water accelerates the hydrolysis step. Notably, the majority of B-TAP 4c (>90%) survived in aqueous THF and was not hydrolyzed to the corresponding phosphine oxide.
Next, we explored the substrate scope of the 2c-mediated Staudinger reduction (Scheme 2). Applying our standardized conditions, all aromatic azides evaluated gave excellent yields (1a1m, Scheme 2). Reductions of aliphatic azides were relatively slow at 30 °C and required elevated temperatures and longer reaction times for reaction completion, likely attributed to the decreased electrophilicity of alkyl azides. Gratifyingly, the reduction of aliphatic substrates 1n1v proceeded smoothly under the modified conditions, with satisfactory yields being obtained in all cases (Scheme 2). Indeed, 2c-mediated Staudinger reduction tolerated a broad range of functional groups and various substitution patterns. For instance, substrates with electron-withdrawing groups (1a1c, 1h, and 1i), electron-donating groups (1d and 1f) or reactive functionalities (1e, 1g, and 1j) are all fully compatible with the current protocol. Moreover, sterically-encumbered amines such as 3k, 3l, 3n, and 3o were obtained in high yields when ortho-substituted or branched substrates were used.
To demonstrate the synthetic potential of this methodology in a more complex setting, we explored the synthesis of a highly functional and sterically-hindered cinchonine-derived amine (3w). Staudinger reductions of cinchona alkaloid-derived azides have been reported to be problematic in some cases [22]. Their corresponding aza-ylide intermediates are reluctant to hydrolyze and are even able to be isolated by flash column chromatography. We proposed that accelerated hydrolysis could occur for 2c-mediated Staudinger reduction since the ortho-SO2NH2 may facilitate the decomposition of the aza-ylide through intramolecular protonation. To our delight, the above proposal was verified by experiments with 2c promoting an efficient Staudinger reduction of azide 1w. The reduction was conducted on a 5.00-mmol scale and 1.21 g of product 3w was isolated in 84% yield (Scheme 3).
To elucidate the mechanism of the 2c-mediated Staudinger reduction, we examined both the 2c- and PPh3-mediated Staudinger reductions by DFT calculation (Figure 1) [23]. Both reductions share a three-step reaction pathway. In the initial step, the nucleophilic attack of either of the phosphine reagents on the azide generates a betaine intermediate. Then, this intermediate loses one equivalent of nitrogen gas to form an aza-ylide intermediate. Finally, the decomposition of the aza-ylide yields the amine product. In the conventional PPh3-mediated Staudinger reduction, the rate-determining step was the step from Int-2′ to TS-3′. The high activation barrier (33.4 kcal/mol) was consistent with the fact that Int-2′ was resistant to hydrolysis. In the 2c-mediated Staudinger reduction, the rate-determining step was the step from Int-1 to TS-2 with a moderate activation barrier (24.7 kcal/mol). In addition, several mechanistic insights have been disclosed by comparing both reaction pathways. The activation barrier for the formation of aza-ylide in the 2c-mediated process (24.7 kcal/mol) was slightly higher than the one in PPh3-mediated processes (21.7 kcal/mol). However, the free energy barrier for the decomposition of aza-ylide Int-2 (18.8 kcal/mol) was significantly lower than the one for the decomposition of aza-ylide Int-2′ (33.4 kcal/mol). Notably, the difference in the enthalpy barriers between both processes was much smaller (18.5 kcal/mol vs. 21.1 kcal/mol). Apparently, the proximal SO2NH2 group in 2c plays an important role in the protonation of aza-ylide Int-2. Its significant contribution to entropy greatly lowers the free energy barrier. We also note that the PPh3-mediated reduction is more thermodynamically favorable than the 2c-mediated Staudinger reduction. Therefore, it is most likely that the 2c-mediated process is largely controlled by favorable kinetics. The rate acceleration in the 2c-mediated reduction probably originated from the rapid intramolecular proton transfer in the aza-ylide decomposition step.

3. Materials and Methods

3.1. Reagents and General Methods

All anhydrous solvents except toluene and dichloromethane were obtained from common suppliers and used as received. Toluene and dichloromethane were purified according to standard procedures [24]. Analytical thin-layer chromatography was performed on 20 × 50 mm silica gel 60 GF254 plates (Leyan, China). Visualization was accomplished with UV light, potassium permanganate, or ninhydrin stain followed by heating. Flash column chromatography was performed on 200−300 mesh silica gel (Leyan, China). Unless otherwise stated, all reagents were purchased from commercial sources and used without further purification. All reactions were conducted under an atmosphere of nitrogen in oven-dried glassware unless otherwise noted. 2-Iodo-5-methylbenzenesulfonamide was prepared according to the literature methods [25,26]. Bis(4-methoxyphenyl)phosphane [27,28] and organic azides except 1j, 1s, and 1w were prepared according to the literature methods (for 1a [29], 1b-i, 1k-m, 1o, 1p, 1r, 1t-v [30], 1n and 1q [31]).
High-resolution mass spectra (HRMS) were recorded on a Thermo Fisher Scientific’s Q Exactive UHMR Hybrid Quadrupole-Orbitrap Mass Spectrometer LC/MS (ESI); melting points were obtained with INESA WRS-3 apparatus; GC/MS spectra were recorded using a gas chromatograph mass spectrometer (GCMS-QP2010 SE; Shimadzu Corp.) with the electron impact ionization (EI) mode; 1H, 13C, 19F, and 31P NMR spectra were recorded on a Bruker AVANCE III 400 MHz spectrometer at 298 K and referenced to residual protium in the NMR solvent (CHCl3, δ 7.26 in 1H NMR) and the carbon resonances of the solvent (CDCl3, δ 77.16 in 13C NMR). Chemical shifts were reported in parts per million (ppm, δ) downfield from tetramethylsilane. NMR peaks are described as singlet (s), doublet (d), triplet (t), multiplet (m), complex (comp), approximate (app), and broad (br).

3.2. Synthetic Procedures

3.2.1. Synthesis of 2-(Diarylphosphaneyl)-5-methylbenzenesulfonamide

General Procedure: Under a N2 atmosphere, 2-iodo-5-methylbenzenesulfonamide (12.0 g, 40.0 mmol, 1.0 equiv), diarylphosphine (48.0 mmol, 1.2 equiv), PdCl2(dppf) (296.4 mg, 0.40 mmol, 0.01 equiv), diisopropylethylamine (DIPEA, 10.3 g, 80.0 mmol, 2.0 equiv), and anhydrous N,N-dimethylformamide (40 mL) were placed into a heavy-walled Schlenk tube. The tube was sealed and stirred at 120 °C in an oil bath for 12 h. After completion, the mixture was diluted with dichloromethane (200 mL) and washed with water (200 mL). The separated organic layer was further washed with water (100 mL), dried over anhydrous sodium sulfate, filtered, and concentrated in vacuo. The crude product was purified by flash column chromatography to afford the pure product.

2-(Diphenylphosphaneyl)-5-methylbenzenesulfonamide (2b) [32]

Following the General Procedure, compound 2b was obtained as a white solid (11.4 g, 80%). mp 212–213 °C (dichloromethane). Rf = 0.45 (petroleum ether/ethyl acetate, 2:1 v/v). 1H NMR (400 MHz, CDCl3) δ 7.99 (dd, J = 4.2, 1.8 Hz, 1H), 7.41–7.31 (comp, 6H), 7.31–7.21 (comp, 5H), 7.11 (dd, J = 7.7, 3.8 Hz, 1H), 5.51 (br s, 2H), 2.41 (s, 3H). 13C{1H} NMR (CDCl3, 101 MHz): δ 147.1 (d, JC-P = 26.2 Hz), 140.4, 136.7, 135.7 (d, JC-P = 5.9 Hz, 2C), 133.7 (d, JC-P = 19.4 Hz, 4C), 133.2, 131.9 (d, JC-P = 20.3 Hz), 129.2 (2C), 128.8 (d, JC-P = 7.1 Hz, 4C), 128.7 (d, JC-P = 7.8 Hz), 21.3. 31P NMR (162 MHz, CDCl3): δ −13.2. HRMS-ESI (m/z) for C19H18NO2PS [M+H]+: calcd 356.0869, found 356.0861.

2-(Bis(4-methoxyphenyl)phosphaneyl)-5-methylbenzenesulfonamide (2c)

Following the General Procedure, the reaction was conducted on a 2.00-mmol scale with higher loading of palladium catalyst (0.02 equiv) instead. Compound 2c was obtained as a white solid (423.7 mg, 51%). mp 90–93 °C (dichloromethane). Rf = 0.35 (petroleum ether/ethyl acetate, 2:1 v/v). 1H NMR (400 MHz, CDCl3) δ 7.96 (dd, J = 3.9, 1.9 Hz, 1H), 7.31–7.23 (m, 1H), 7.19 (app t, J = 8.2 Hz, 4H), 7.10 (dd, J = 7.8, 3.8 Hz, 1H), 6.88 (app d, J = 7.9 Hz, 4H), 5.47 (br s, 2H), 3.80 (s, 6H), 2.40 (s, 3H). 13C{1H} NMR (CDCl3, 101 MHz): δ 160.6 (2C), 146.7 (d, JC-P = 25.7 Hz), 140.0, 136.2, 135.2 (d, JC-P = 20.7 Hz, 4C), 133.1 (d, JC-P = 21.0 Hz), 133.1, 128.7 (d, JC-P = 4.4 Hz), 126.6 (d, JC-P = 2.6 Hz, 2C), 114.5 (d, JC-P = 8.2 Hz, 4C), 55.4 (2C), 21.3. 31P NMR (162 MHz, CDCl3): δ −15.9. HRMS-ESI (m/z) for C21H22NO4PS [M+H]+: calcd 416.1080, found 416.1074.

3.2.2. Synthesis of Organic Azides

Organic azides except 1j, 1s, and 1w were prepared from the literature methods and their analytical data were consistent with the literature data (see Supplementary Materials). Azides 1j, 1s and 1w were prepared by the procedures below.

Perfluorophenyl 4-azidobenzoate (1j)

A solution of 4-azidobenzoic acid (163.0 mg, 1.00 mmol, 1.0 equiv) and 2,3,4,5,6-pentafluorophenol (220.8 mg, 1.20 mmol, 1.2 equiv) in dichloromethane (5 mL) was cooled down to 0 °C. To this solution was added 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide hydrochloride (287.6 mg, 1.50 mmol, 1.5 equiv) and 4-dimethylaminopyridine (12.2 mg, 0.10 mmol, 0.1 equiv). The mixture was allowed to warm up to room temperature and stirred for 12 h. The solvent was concentrated in vacuo and the crude product was purified by flash column chromatography to afford the product 1j as a white solid (255.8 mg, 78%). mp 78–79 °C (dichloromethane). Rf = 0.46 (petroleum ether/ethyl acetate, 100:1 v/v). 1H NMR (400 MHz, CDCl3): δ 8.19 (d, J = 8.8 Hz, 2H), 7.17 (d, J = 8.8 Hz, 2H). 13C{1H} NMR (CDCl3, 101 MHz): δ 161.9, 146.9, 143.0–140.1 (m, 2C), 141.3–138.1 (m), 139.7–136.6 (m, 2C), 132.8 (2C), 125.9–125.2 (m), 123.4, 119.5 (2C). 19F NMR (377 MHz, CDCl3): δ −152.42 to −152.57 (m, 2F), −157.83 (t, J = 21.7 Hz), −162.16 to −162.37 (m, 2F).

2-Azido-N-(4-bromobenzyl)acetamide (1s)

N-(4-bromobenzyl)-2-chloroacetamide (7.88 g, 30.0 mmol, 1.0 equiv), sodium fluoride (3.15 g, 75.0 mmol, 2.5 equiv) and trimethylsilyl azide (5.18 g, 45.0 mmol, 1.5 equiv) were dissolved in THF/water (75 mL, 3:1 v:v). The mixture was heated at 85 °C for 12 h. The reaction mixture was concentrated in vacuo and extracted with ethyl acetate (3 × 80 mL). The organic layer was separated, washed with saturated brine, dried over anhydrous sodium sulfate, filtered, and concentrated in vacuo. The crude mixture was purified by flash column chromatography to afford the product 1s as a white solid (6.78 g, 84%). mp 93.0 °C (dichloromethane). Rf = 0.23 (petroleum ether/ethyl acetate, 3:1 v/v). 1H NMR (400 MHz, CDCl3): δ 7.47 (d, J = 8.4 Hz, 2H), 7.16 (d, J = 8.4 Hz, 2H), 6.62 (br, 1H), 4.43 (d, J = 6.0 Hz, 2H), 4.05 (s, 2H). 13C{1H} NMR (CDCl3, 101 MHz): δ 166.6, 136.7, 132.1 (2C), 129.7 (2C), 121.9, 52.9, 43.0. HRMS-ESI (m/z) for C9H11BrN2O [M−H]-: calcd 266.9886 (79Br), 268.9867 (81Br), found 266.9984 (79Br), 268.9864 (81Br).

(1S,2R,4S,5R)-2-((R)-Azido(quinolin-4-yl)methyl)-5-vinylquinuclidine (1w)

Following the literature method [33], cinchonine (2.94 g, 10.0 mmol, 1.0 equiv) and triphenylphosphine (3.26 g, 12.0 mmol, 1.2 equiv) were dissolved in 100 mL anhydrous tetrahydrofuran under a N2 atmosphere. The mixture was cooled down to 0 °C and then diisopropyl azodicarboxylate (DIAD, 2.32 mL, 12.0 mmol, 1.2 equiv) and a solution of diphenylphosphoryl azide (DPPA, 2.56 mL, 12.0 mmol, 1.2 equiv) in anhydrous tetrahydrofuran (20 mL) were added dropwise. The mixture was stirred at room temperature for 12 h and at 50 °C for another 2 h. The mixture was concentrated in vacuo and the residue was dissolved in dichloromethane and 10% hydrochloric acid (1:1, 100 mL). The aqueous phase was separated and further washed with dichloromethane (4 × 50 mL). Then the aqueous phase was made alkaline (pH ≥ 9) with an excess of concentrated aqueous ammonia (22–25%) and was extracted with dichloromethane (4 × 50 mL). The combined organic phases were dried over anhydrous sodium sulfate and concentrated. The crude product was purified by flash column chromatography and the title compound 1w was obtained as a yellowish viscous oil (2.24 g, 70%). Rf = 0.14 (ethyl acetate). 1H NMR (400 MHz, CDCl3): δ 8.92 (dd, J = 4.5, 1.4 Hz, 1H), 8.21 (d, J = 8.5 Hz, 1H), 8.16 (d, J = 8.4 Hz, 1H), 7.79–7.69 (m, 1H), 7.68–7.55 (m, 1H), 7.38 (dd, J = 4.5, 1.3 Hz, 1H), 5.81–5.65 (m, 1H), 5.11 (d, J = 10.6 Hz, 1H), 5.03–4.86 (comp, 2H), 3.43–3.11 (comp, 3H), 2.96–2.74 (comp, 2H), 2.32–2.15 (m, 1H), 1.65–1.59 (m, 1H), 1.59–1.46 (comp, 2H), 1.45–1.29 (m, 1H), 0.77–0.63 (m, 1H). 13C{1H} NMR (CDCl3, 101 MHz): δ 150.1, 148.9, 142.4, 141.4, 130.8, 129.6, 127.3, 126.7, 123.1, 120.4, 114.7, 62.7, 59.7, 56.1, 41.0, 39.5, 28.0, 27.3, 26.2. HRMS-ESI (m/z) for C19H21N5 [M+H]+: calcd 320.1870, found 320.1863.

3.2.3. Synthesis of Amines

General Procedure A (for the substrates 1a–1m): Under a N2 atmosphere, the corresponding organic azide (1.00 mmol, 1.0 equiv) was dissolved in anhydrous tetrahydrofuran (10 mL), followed by the addition of 2c (1.10 mmol, 1.1 equiv). The mixture was stirred at 30 °C for 3 h. Upon completion, the mixture was concentrated in vacuo and the crude product was purified by flash column chromatography.
General Procedure B (for the substrates 1n–1w): Under a N2 atmosphere, the corresponding organic azide (1.00 mmol, 1.0 equiv) was dissolved in anhydrous tetrahydrofuran (10 mL), followed by the addition of 2c (1.10 mmol, 1.1 equiv). The mixture was stirred at 60 °C for 15 h. Upon completion, the mixture was concentrated in vacuo and the crude product was purified by flash column chromatography. For 1w, 5.00 mmol of azide was used instead.
All amines except 3j and 3s are known compounds and their analytical data were consistent with the literature data (see Supplementary Materials). Their isolated yields were reported in Scheme 2.

Perfluorophenyl 4-aminobenzoate (3j)

Following General Procedure A, the title compound 3j was obtained as a white solid (275.7 mg, 91%) from the azide 1j (1.00 mmol, 329.0 mg). mp 129 °C (dichloromethane). Rf = 0.44 (petroleum ether/ethyl acetate, 3:1 v/v). 1H NMR (400 MHz, CDCl3): δ 7.99 (d, J = 8.7 Hz, 2H), 6.69 (d, J = 8.7 Hz, 2H), 4.29 (br, 2H). 13C{1H} NMR (CDCl3, 101 MHz): δ 162.7, 152.6, 143.3–140.1 (m, 2C), 141.0–137.8 (m), 139.7–136.4 (m, 2C), 133.2 (2C), 126.2–125.6 (m), 115.8, 114.0 (2C). 19F NMR (377 MHz, CDCl3): δ −152.56 to −152.76 (m, 2F), −158.88 (t, J = 21.6 Hz), −162.87 (td, J = 22.4, 5.0 Hz, 2F). GC-MS (EI, 70 eV) m/z: 303 [M]+.

2-Amino-N-(4-bromobenzyl)acetamide (3s)

Following General Procedure B, the title compound was obtained as a white solid (178.0 mg, 82%) from the azide 1s (0.50 mmol, 121.5 mg). mp 93 °C (dichloromethane). Rf = 0.19 (ethyl acetate/methanol/aqueous ammonia, 100:10:1 v/v/v). 1H NMR (400 MHz, CDCl3): δ 7.65 (br, 1H), 7.44 (d, J = 8.4 Hz, 2H), 7.16 (d, J = 8.4 Hz, 2H), 4.41 (d, J = 6.1 Hz, 2H), 3.39 (s, 2H), 1.61 (br, 2H, overlapped water). 13C{1H} NMR (101 MHz, CDCl3): δ 172.8, 137.6, 131.9, 129.6, 121.4, 44.8, 42.5. HRMS-ESI (m/z) for C9H11BrN2O [M+H]+: calcd 243.0128 (79Br), 245.0107 (81Br), found 243.0124 (79Br), 245.0104 (81Br).

3.3. Computational Methods

The calculations were carried out with the Gaussian 09 software package [34]. The structures were optimized by the density functional theory (DFT) [35] with the B3LYP functional [36,37] with basis set 6-31G(d) [38,39] in the gas phase. Frequency analysis was conducted at the same level of theory to verify the stationary points to be real minima or saddle points and to obtain the thermodynamic energy corrections at 298.15 K. Intrinsic reaction coordinate (IRC) [40,41,42] calculations were performed to confirm the connection between two correct minima for a transition state. More accurate electronic energy results were refined by calculating the single-point energy at the B3LYP-D3(BJ) [43]/6-311++G(2df, 2p) [38,39] level of theory with the SMD model [44] (solvent = THF).

4. Conclusions

In conclusion, we developed the first ortho-phosphinoarenesulfonamide-mediated Staudinger reduction without the need for water. Computational studies suggest that the ortho-SO2NH2 substituent of the phosphine reagent is significant for favorable reaction kinetics. We are currently investigating other aza-ylide-driven Staudinger-type transformations which will be reported in due course.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27175707/s1, the corresponding Cartesian coordinates in DFT studies, copies of NMR spectra for all new compounds, and a reference list for the characterization data of all known compounds [45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64].

Author Contributions

Conceptualization, L.L.; methodology, X.L., and Z.W. (Zhenguo Wang); validation, X.L., Z.W. (Zhenguo Wang), Z.W. (Zixu Wang), and W.L.; investigation, X.L., Z.W. (Zhenguo Wang), Z.W. (Zixu Wang), K.Y., L.L.; resources, L.L.; writing—original draft preparation, X.L.; writing—review and editing, X.L., W.L., Z.W. (Zixu Wang), L.L.; supervision, L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 21502241 and the Natural Science Foundation of Guangdong Province, grant number 2016A030313290.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Supercomputing Center in Shenzhen (Shenzhen Cloud Computer Center) and Sun Yat-sen University for providing computing resources. Le Li thanks Xiaohui Cao and Pyh Li for helpful discussions and support. All authors thank Christopher M. Plummer for proofreading.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of all of the compounds are available from the authors.

References

  1. Staudinger, H.; Meyer, J. Über neue organische Phosphorverbindungen III. Phosphinmethylenderivate und Phosphinimine. Helv. Chim. Acta 1919, 2, 635–646. [Google Scholar] [CrossRef]
  2. Gololobov, Y.G.; Kasukhin, L.F. Recent Advances in the Staudinger Reaction. Tetrahedron 1992, 48, 1353–1406. [Google Scholar] [CrossRef]
  3. Gololobov, Y.G.; Zhmurova, I.N.; Kasukhin, L.F. Sixty Years of Staudinger Reaction. Tetrahedron 1981, 37, 437–472. [Google Scholar] [CrossRef]
  4. Saxon, E.; Bertozzi, C.R. Cell Surface Engineering by a Modified Staudinger Reaction. Science 2000, 287, 2007–2010. [Google Scholar] [CrossRef]
  5. Kiick, K.L.; Saxon, E.; Tirrell, D.A.; Bertozzi, C.R. Incorporation of Azides into Recombinant Proteins for Chemoselective Modification by The Staudinger Ligation. Proc. Natl. Acad. Sci. USA 2002, 99, 19–24. [Google Scholar] [CrossRef]
  6. Heiss, T.K.; Dorn, R.S.; Prescher, J.A. Bioorthogonal Reactions of Triarylphosphines and Related Analogues. Chem. Rev. 2021, 121, 6802–6849. [Google Scholar] [CrossRef]
  7. Bednarek, C.; Wehl, I.; Jung, N.; Schepers, U.; Bräse, S. The Staudinger Ligation. Chem. Rev. 2020, 120, 4301–4354. [Google Scholar] [CrossRef]
  8. Darrah, K.; Wesalo, J.; Lukasak, B.; Tsang, M.; Chen, J.K.; Deiters, A. Small Molecule Control of Morpholino Antisense Oligonucleotide Function through Staudinger Reduction. J. Am. Chem. Soc. 2021, 143, 18665–18671. [Google Scholar] [CrossRef]
  9. Wesalo, J.S.; Luo, J.; Morihiro, K.; Liu, J.; Deiters, A. Phosphine-Activated Lysine Analogues for Fast Chemical Control of Protein Subcellular Localization and Protein SUMOylation. ChemBioChem 2020, 21, 141–148. [Google Scholar] [CrossRef]
  10. Lukasak, B.; Morihiro, K.; Deiters, A. Aryl Azides as Phosphine-Activated Switches for Small Molecule Function. Sci. Rep. 2019, 9, 1470. [Google Scholar] [CrossRef] [Green Version]
  11. Luo, J.; Liu, Q.; Morihiro, K.; Deiters, A. Small-Molecule Control of Protein Function through Staudinger Reduction. Nat. Chem. 2016, 8, 1027–1034. [Google Scholar] [CrossRef] [PubMed]
  12. Bräse, S.; Gil, C.; Knepper, K.; Zimmermann, V. Organic Azides: An Exploding Diversity of a Unique Class of Compounds. Angew. Chem. Int. Ed. 2005, 44, 5188–5240. [Google Scholar] [CrossRef]
  13. Kürti, L.; Czakó, B. Strategic Applications of Named Reactions in Organic Synthesis; Elsevier: Dorecht, Netherlands, 2005. [Google Scholar]
  14. Li, J.J. Name Reactions: A Collection of Detailed Reaction Mechanisms, 2nd ed.; Springer: Heidelberg, Germany, 2003. [Google Scholar]
  15. Meguro, T.; Terashima, N.; Ito, H.; Koike, Y.; Kii, I.; Yoshida, S.; Hosoya, T. Staudinger Reaction Using 2,6-Dichlorophenyl Azide Derivatives for Robust aza-Ylide Formation Applicable to Bioconjugation in Living Cells. Chem. Commun. 2018, 54, 7904–7907. [Google Scholar] [CrossRef] [PubMed]
  16. Sundhoro, M.; Jeon, S.; Park, J.; Ramström, O.; Yan, M. Perfluoroaryl Azide Staudinger Reaction: A Fast and Bioorthogonal Reaction. Angew. Chem. Int. Ed. 2017, 56, 12117–12121. [Google Scholar] [CrossRef] [PubMed]
  17. Xu, S.D.; Fang, C.H.; Tian, G.X.; Chen, Y.; Dou, Y.H.; Wu, X.H. Reduction of 4-Azidonaphthalimide with Different Phosphine Ligands and Exploration of Their Spectroscopic Properties. J. Mol. Struct. 2015, 1102, 197–204. [Google Scholar] [CrossRef]
  18. Meguro, T.; Yoshida, S.; Hosoya, T. Aromatic Azido-selective Reduction via the Staudinger Reaction Using Tri-n-butylphosphonium Tetrafluoroborate with Triethylamine. Chem. Lett. 2017, 46, 473–476. [Google Scholar] [CrossRef]
  19. Leffler, J.E.; Temple, R.D. Staudinger Reaction Between Triarylphosphines and Azides. A Study of the Mechanism. J. Am. Chem. Soc. 1967, 89, 5235–5246. [Google Scholar] [CrossRef]
  20. Saneyoshi, H.; Ochikubo, T.; Mashimo, T.; Hatano, K.; Ito, Y.; Abe, H. Triphenylphosphinecarboxamide: An Effective Reagent for the Reduction of Azides and its Application to Nucleic Acid Detection. Org. Lett. 2014, 16, 30–33. [Google Scholar] [CrossRef]
  21. Luo, W.J.; Wang, Z.G.; Cao, X.X.; Liang, D.C.; Wei, M.J.; Yin, K.S.; Li, L. Construction of Benzo-1,2,3-thiazaphosphole Heterocycles by Annulations of ortho-Phosphinoarenesulfonyl Fluorides with Trimethylsilyl Azide. J. Org. Chem. 2020, 85, 14785–14794. [Google Scholar] [CrossRef]
  22. Wang, Y.; Milkiewicz, K.L.; Kaufman, M.L.; He, L.; Landmesser, N.G.; Levy, D.V.; Allwein, S.P.; Christie, M.A.; Olsen, M.A.; Neville, C.J.; et al. Plant Process for the Preparation of Cinchona Alkaloid-Based Thiourea Catalysts. Org. Process Res. Dev. 2017, 21, 408–413. [Google Scholar] [CrossRef]
  23. Tian, W.Q.; Wang, Y.A. Mechanisms of Staudinger Reactions within Density Functional Theory. J. Org. Chem. 2004, 69, 4299–4308. [Google Scholar] [CrossRef]
  24. Armarego, W.L.F.; Perrin, D.D. Purification of Laboratory Chemicals, 4th ed.; Butterworth-Heinemann: Oxford, UK, 1997. [Google Scholar]
  25. Tan, C.-M.; Chen, G.S.; Chen, C.-S.; Chang, P.-T.; Chern, J.-W. Design, Synthesis And Biological Evaluation of Benzo[1.3.2]Dithiazolium Ylide 1,1-Dioxide Derivatives as Potential Dual Cyclooxygenase-2/5-Lipoxygenase Inhibitors. Bioorg. Med. Chem. 2011, 19, 6316–6328. [Google Scholar] [CrossRef] [PubMed]
  26. Li, J.; Shi, H.W.; Zhang, S.; Rudolph, M.; Rominger, F.; Hashmi, A.S.K. Switchable Divergent Synthesis in Gold-Catalyzed Difunctionalizations of o-Alkynylbenzenesulfonamides with Aryldiazonium Salts. Org. Lett. 2021, 23, 7713–7717. [Google Scholar] [CrossRef] [PubMed]
  27. Busacca, C.A.; Lorenz, J.C.; Grinberg, N.; Haddad, N.; Hrapchak, M.; Latli, B.; Lee, H.; Sabila, P.; Saha, A.; Sarvestani, M.; et al. A Superior Method for the Reduction of Secondary Phosphine Oxides. Org. Lett. 2005, 7, 4277–4280. [Google Scholar] [CrossRef] [PubMed]
  28. Stankevic, M.; Pietrusiewicz, M. An Expedient Reduction of sec-Phosphine Oxides to sec-Phosphine-Boranes by BH3·SMe2. Synlett 2003, 1012–1016. [Google Scholar] [CrossRef]
  29. Cornali, B.M.; Kimani, F.W.; Jewett, J.C. Cu-Click Compatible Triazabutadienes to Expand the Scope of Aryl Diazonium Ion Chemistry. Org. Lett. 2016, 18, 4948–4950. [Google Scholar] [CrossRef]
  30. Babin, V.; Sallustrau, A.; Loreau, O.; Caillé, F.; Goudet, A.; Cahuzac, H.; Del Vecchio, A.; Taran, F.; Audisio, D. A General Procedure for Carbon Isotope Labeling of Linear Urea Derivatives with Carbon Dioxide. Chem. Commun. 2021, 57, 6680–6683. [Google Scholar] [CrossRef]
  31. Büll, C.; Heise, T.; van Hilten, N.; Pijnenborg, F.A.J.; Bloemendal, V.R.L.J.; Gerrits, L.; Kers-Rebel, E.D.; Ritschel, T.; den Brok, M.H.; Adema, G.J.; et al. Steering Siglec–Sialic Acid Interactions on Living Cells using Bioorthogonal Chemistry. Angew. Chem. Int. Ed. 2017, 56, 3309–3313. [Google Scholar] [CrossRef]
  32. Li, L.; Luo, W.J.; Liang, D.C.; Wang, Z.G. Rigid Heterocyclic Compound, Preparation Method and Application Thereof, Phosphine-Containing Sulfonamide Compound and Preparation Method Thereof. Chinese Patent (Sun Yat-sen University). CN 111574567 A, 25 August 2020. [Google Scholar]
  33. Vakulya, B.; Varga, S.; Csámpai, A.; Soós, T. Highly Enantioselective Conjugate Addition of Nitromethane to Chalcones Using Bifunctional Cinchona Organocatalysts. Org. Lett. 2005, 7, 1967–1969. [Google Scholar] [CrossRef]
  34. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  35. Koch, W.; Holthausen, M.C. A Chemist’s Guide to Density Functional Theory; Wiley: New York, NY, USA, 2001; ISBN 9783527303724. [Google Scholar]
  36. Becke, A.D. Becke’s Three Parameter Hybrid Method Using the LYP Correlation Functional. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  37. Lee, C.T.; Yang, W.T.; Parr, R.G. Development of the ColleSalvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B Condens. Matter Mater. Phys. 1988, 37, 785–789. [Google Scholar] [CrossRef] [PubMed]
  38. Hariharan, P.C.; Pople, J.A. The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies. Theoret. Chimica Acta 1973, 28, 213–222. [Google Scholar] [CrossRef]
  39. Rassolov, V.; Pople, J.A.; Ratner, M.; Windus, T.L. 6-31G* Basis Set for Atoms K through Zn. J. Chem. Phys. 1998, 109, 1223–1299. [Google Scholar] [CrossRef]
  40. Fukui, K. Formulation of the Reaction Coordinate. J. Phys. Chem. 1970, 74, 4161–4163. [Google Scholar] [CrossRef]
  41. Gonzalez, C.; Schlegel, H.B. An Improved Algorithm for Reaction Path Following. J. Chem. Phys. 1989, 90, 2154–2161. [Google Scholar] [CrossRef]
  42. Gonzalez, C.; Schlegel, H.B. Reaction Path Following in MassWeighted Internal Coordinates. J. Phys. Chem. 1990, 94, 5523–5527. [Google Scholar] [CrossRef]
  43. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  44. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  45. Wang, J.; Liang, D.; Feng, J.; Tang, X. Multicolor Cocktail for Breast Cancer Multiplex Phenotype Targeting and Diagnosis Using Bioorthogonal Surface-Enhanced Raman Scattering Nanoprobes. Anal. Chem. 2019, 91, 11045–11054. [Google Scholar] [CrossRef]
  46. Barrow, A.S.; Moses, J.E. Synthesis of Sulfonyl Azides via Lewis Base Activation of Sulfonyl Fluorides and Trimethylsilyl Azide. Synlett 2016, 27, 1840–1843. [Google Scholar] [CrossRef]
  47. Barral, K.; Moorhouse, A.D.; Moses, J.E. Efficient Conversion of Aromatic Amines into Azides: A One-Pot Synthesis of Triazole Linkages. Org. Lett. 2007, 9, 1809–1811. [Google Scholar] [CrossRef] [PubMed]
  48. Xie, S.; Zhang, Y.; Ramström, O.; Yan, M. Base-Catalyzed Synthesis of Aryl Amides from Aryl Azides and Aldehydes. Chem. Sci. 2016, 7, 713–718. [Google Scholar] [CrossRef] [PubMed]
  49. Alt, I.T.; Guttroff, C.; Plietker, B. Iron-Catalyzed Intramolecular Aminations of C(Sp3)−H Bonds in Alkylaryl Azides. Angew. Chem. Int. Ed. 2017, 56, 10582–10586. [Google Scholar] [CrossRef] [PubMed]
  50. Lenstra, D.C.; Lenting, P.E.; Mecinović, J. Sustainable Organophosphorus-Catalysed Staudinger Reduction. Green Chem. 2018, 20, 4418–4422. [Google Scholar] [CrossRef]
  51. Bow, J.-P.J.; Adami, V.; Marasco, A.; Grønnevik, G.; Rivers, D.A.; Alvaro, G.; Riss, P.J. A Direct Fixation of CO2 for Isotopic Labelling of Hydantoins Using Iodine–Phosphine Charge Transfer Complexes. Chem. Commun. 2022, 7546–7549. [Google Scholar] [CrossRef]
  52. vanderKnaap, M.; Lageveen, L.T.; Busscher, H.J.; Mars-Groenendijk, R.; Noort, D.; Otero, J.M.; Llamas-Saiz, A.L.; vanRaaij, M.J.; vanderMarel, G.A.; Overkleeft, H.S.; et al. Evaluation of Readily Accessible Azoles as Mimics of the Aromatic Ring of D-Phenylalanine in the Turn Region of Gramicidin S. ChemMedChem 2011, 6, 840–847. [Google Scholar] [CrossRef]
  53. Shi, H.; Kwok, R.T.K.; Liu, J.; Xing, B.; Tang, B.Z.; Liu, B. Real-Time Monitoring of Cell Apoptosis and Drug Screening Using Fluorescent Light-up Probe with Aggregation-Induced Emission Characteristics. J. Am. Chem. Soc. 2012, 134, 17972–17981. [Google Scholar] [CrossRef]
  54. Jang, M.; Lim, T.; Park, B.Y.; Han, M.S. Metal-Free, Rapid, and Highly Chemoselective Reduction of Aromatic Nitro Compounds at Room Temperature. J. Org. Chem. 2022, 87, 910–919. [Google Scholar] [CrossRef]
  55. Lu, H.; Geng, Z.; Li, J.; Zou, D.; Wu, Y.; Wu, Y. Metal-Free Reduction of Aromatic Nitro Compounds to Aromatic Amines with B2pin2 in Isopropanol. Org. Lett. 2016, 18, 2774–2776. [Google Scholar] [CrossRef]
  56. Vo, G.D.; Hartwig, J.F. Palladium-Catalyzed Coupling of Ammonia with Aryl Chlorides, Bromides, Iodides, and Sulfonates: A General Method for the Preparation of Primary Arylamines. J. Am. Chem. Soc. 2009, 131, 11049–11061. [Google Scholar] [CrossRef] [Green Version]
  57. Xie, S.; Fukumoto, R.; Ramström, O.; Yan, M. Anilide Formation from Thioacids and Perfluoroaryl Azides. J. Org. Chem. 2015, 80, 4392–4397. [Google Scholar] [CrossRef] [PubMed]
  58. Zhao, H.; Leonori, D. Minimization of Back-Electron Transfer Enables the Elusive Sp3 C−H Functionalization of Secondary Anilines. Angew. Chem. Int. Ed. 2021, 60, 7669–7674. [Google Scholar] [CrossRef] [PubMed]
  59. Kita, Y.; Kuwabara, M.; Yamadera, S.; Kamata, K.; Hara, M. Effects of Ruthenium Hydride Species on Primary Amine Synthesis by Direct Amination of Alcohols over a Heterogeneous Ru Catalyst. Chem. Sci. 2020, 11, 9884–9890. [Google Scholar] [CrossRef] [PubMed]
  60. Murai, N.; Miyano, M.; Yonaga, M.; Tanaka, K. One-Pot Primary Aminomethylation of Aryl and Heteroaryl Halides with Sodium Phthalimidomethyltrifluoroborate. Org. Lett. 2012, 14, 2818–2821. [Google Scholar] [CrossRef] [PubMed]
  61. Amberchan, G.; Snelling, R.A.; Moya, E.; Landi, M.; Lutz, K.; Gatihi, R.; Singaram, B. Reaction of Diisobutylaluminum Borohydride, a Binary Hydride, with Selected Organic Compounds Containing Representative Functional Groups. J. Org. Chem. 2021, 86, 6207–6227. [Google Scholar] [CrossRef]
  62. Demong, D.E.; Ng, I.; Miller, M.W.; Stamford, A.W. A Novel Method for the Preparation of 4-Arylimidazolones. Org. Lett. 2013, 15, 2830–2833. [Google Scholar] [CrossRef]
  63. Taher, A.; Lee, D.J.; Lee, I.M.; Rahman, M.L.; Sarker, M.S. A Novel Graphene-Supported Palladium Catalyst for Suzuki-Miyaura Coupling of Less Reactive Heteroaryl Halides in Water. Bull. Korean Chem. Soc. 2016, 37, 1478–1485. [Google Scholar] [CrossRef]
  64. Aw, J.; Widjaja, F.; Ding, Y.; Mu, J.; Liang, Y.; Xing, B. Enzyme-Responsive Reporter Molecules for Selective Localization and Fluorescence Imaging of Pathogenic Biofilms. Chem. Commun. 2017, 53, 3330–3333. [Google Scholar] [CrossRef]
Scheme 1. Phosphine-mediated Staudinger reductions of organic azides.
Scheme 1. Phosphine-mediated Staudinger reductions of organic azides.
Molecules 27 05707 sch001
Scheme 2. Substrate scope of 2c-mediated Staudinger reduction a. a Reactions were run on a 1.00 mmol scale in 10 mL of anhydrous THF under a N2 atmosphere. All yields were isolated yields. b Reactions were run at 60 °C for 15 h instead. c Reactions were run on a 0.50 mmol scale instead.
Scheme 2. Substrate scope of 2c-mediated Staudinger reduction a. a Reactions were run on a 1.00 mmol scale in 10 mL of anhydrous THF under a N2 atmosphere. All yields were isolated yields. b Reactions were run at 60 °C for 15 h instead. c Reactions were run on a 0.50 mmol scale instead.
Molecules 27 05707 sch002
Scheme 3. Gram-scale Staudinger reduction of cinchonine-derived azide (1w).
Scheme 3. Gram-scale Staudinger reduction of cinchonine-derived azide (1w).
Molecules 27 05707 sch003
Figure 1. DFT calculated reaction pathways: ortho-Phosphinoarenesulfonamide-mediated Staudinger reduction (blue); triphenylphosphine-mediated Staudinger reduction (red).
Figure 1. DFT calculated reaction pathways: ortho-Phosphinoarenesulfonamide-mediated Staudinger reduction (blue); triphenylphosphine-mediated Staudinger reduction (red).
Molecules 27 05707 g001
Table 1. Staudinger reduction of methyl 4-azidobenzoate (1a) a.
Table 1. Staudinger reduction of methyl 4-azidobenzoate (1a) a.
Molecules 27 05707 i001
EntryReagentSolventTime(h)3 b (%)
1 c2aTHF30
22bTHF394
32cTHF399
4 d2bTHF10 min58
5 d2cTHF10 min81
62cMeCN393
72cDMSO395
82cDMF396
92cCH2Cl23>99
102cToluene397
11 d,e2cTHF/H2O (1:1)397
a Reactions were run on a 0.10 mmol scale in 1 mL of anhydrous THF at 30 °C under a N2 atmosphere for 3 h. 1.10 equiv of 2 was used. b Yields were determined by 1H NMR using 1,3,5-trimethoxybenzene as an internal standard. c 10.0 equiv of H2O was added. d 0.01 M. e 1.50 equiv of 2c was used.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, X.; Wang, Z.; Luo, W.; Wang, Z.; Yin, K.; Li, L. Ortho-Phosphinoarenesulfonamide-Mediated Staudinger Reduction of Aryl and Alkyl Azides. Molecules 2022, 27, 5707. https://doi.org/10.3390/molecules27175707

AMA Style

Li X, Wang Z, Luo W, Wang Z, Yin K, Li L. Ortho-Phosphinoarenesulfonamide-Mediated Staudinger Reduction of Aryl and Alkyl Azides. Molecules. 2022; 27(17):5707. https://doi.org/10.3390/molecules27175707

Chicago/Turabian Style

Li, Xingzhuo, Zhenguo Wang, Wenjun Luo, Zixu Wang, Keshu Yin, and Le Li. 2022. "Ortho-Phosphinoarenesulfonamide-Mediated Staudinger Reduction of Aryl and Alkyl Azides" Molecules 27, no. 17: 5707. https://doi.org/10.3390/molecules27175707

Article Metrics

Back to TopTop