首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 539 毫秒
1.
Thionyl tetrafluoride (SOF4) is a valuable connective gas for sulfur fluoride exchange (SuFEx) click chemistry that enables multidimensional linkages to be created via sulfur–oxygen and sulfur–nitrogen bonds. Herein, we expand the available SuFEx chemistry of SOF4 to include organolithium nucleophiles, and demonstrate, for the first time, the controlled projection of sulfur–carbon links at the sulfur center of SOF4‐derived iminosulfur oxydifluorides (R1?N=SOF2). This method provides rapid and modular access to sulfonimidoyl fluorides (R1?N=SOFR2), another array of versatile SuFEx connectors with readily tunable reactivity of the S?F handle. Divergent connections derived from these valuable sulfonimidoyl fluoride units are also demonstrated, including the synthesis of sulfoximines, sulfonimidamides, and sulfonimidates.  相似文献   

2.
We report here the development of a suite of biocompatible SuFEx transformations from the SOF4‐derived iminosulfur oxydifluoride hub in aqueous buffer conditions. These biocompatible SuFEx reactions of iminosulfur oxydifluorides (R‐N=SOF2) with primary amines give sulfamides (8 examples, up to 98 %), while the reaction with secondary amines furnish sulfuramidimidoyl fluoride products (8 examples, up to 97 %). Likewise, under mild buffered conditions, phenols react with the iminosulfur oxydifluorides (Ar‐N=SOF2) to produce sulfurofluoridoimidates (13 examples, up to 99 %), which can themselves be further modified by nucleophiles. These transformations open the potential for asymmetric and trisubstituted linkages projecting from the sulfur(VI) center, including versatile S?N and S?O connectivity (9 examples, up to 94 %). Finally, the SuFEx bioconjugation of iminosulfur oxydifluorides to amine‐tagged single‐stranded DNA and to BSA protein demonstrate the potential of SOF4‐derived SuFEx click chemistry in biological applications.  相似文献   

3.
Sulfur(VI) fluoride exchange (SuFEx) is a new family of click chemistry based transformations that enable the synthesis of covalently linked modules via SVI hubs. Here we report thionyl tetrafluoride (SOF4) as the first multidimensional SuFEx connector. SOF4 sits between the commercially mass‐produced gases SF6 and SO2F2, and like them, is readily synthesized on scale. Under SuFEx catalysis conditions, SOF4 reliably seeks out primary amino groups [R‐ NH2 ] and becomes permanently anchored via a tetrahedral iminosulfur(VI) link: R−N=(O=)S(F)2. The pendant, prochiral difluoride groups R−N=(O=) SF2 , in turn, offer two further SuFExable handles, which can be sequentially exchanged to create 3‐dimensional covalent departure vectors from the tetrahedral sulfur(VI) hub.  相似文献   

4.
Sulfur(VI)-fluoride exchange (SuFEx) chemistry, an all-encompassing term for substitution events that replace fluoride at an electrophilic sulfur(VI), enables the rapid and flexible assembly of linkages around a SVI core. Although a myriad of nucleophiles and applications works very well with the SuFEx concept, the electrophile design has remained largely SO2-based. Here, we introduce S≡N-based fluorosulfur(VI) reagents to the realm of SuFEx chemistry. Thiazyl trifluoride (NSF3) gas is shown to serve as an excellent parent compound and SuFEx hub to efficiently synthesize mono- and disubstituted fluorothiazynes in an ex situ generation workflow. Gaseous NSF3 was evolved from commercial reagents in a nearly quantitative fashion at ambient conditions. Moreover, the mono-substituted thiazynes could be extended further as SuFEx handles and be engaged in the synthesis of unsymmetrically disubstituted thiazynes. These results provide valuable insights into the versatility of these understudied sulfur functionalities paving the way for future applications.  相似文献   

5.
Polymer brushes present a unique architecture for tailoring surface functionalities due to their distinctive physicochemical properties. However, the polymerization chemistries used to grow brushes place limitations on the monomers that can be grown directly from the surface. Several forms of click chemistry have previously been used to modify polymer brushes by postpolymerization modification with high efficiency, however, it is usually difficult to include the unprotected moieties in the original monomer. We present the use of a new form of click chemistry known as SuFEx (sulfur(VI) fluoride exchange), which allows a silyl ether to be rapidly and quantitatively clicked to a polymer brush grown by free‐radical polymerization containing native ‐SO2F groups with rapid pseudo‐first‐order rates as high as 0.04 s?1. Furthermore, we demonstrate the use of SuFEx to facilely add a variety of other chemical functional groups to brush substrates that have highly useful and orthogonal reactivity, including alkynes, thiols, and dienes.  相似文献   

6.
This review summarizes literature data and the authors’ own research results from the past 14–15 years relating to practical, valuable organosilicon carbofunctional sulfur‐containing compounds of general formula R4−n Si(Sx R1)n , where R is alkyl, arylalkoxyl, aroxyl or even sylatranyl fragment, R1 is hydrogen, alkyl, aryl, alkenyl, etc., n = 1–3, x = 1–10, having sulfur functional groups such as thiol, sulfide, di‐ and polysulfide, as well as sulfur heteroatomic groups such as thiocarbamide, dioxothiocarbamide, dithiourethane, thiuramdisulfide, etc. The compounds reviewed have been found to be effective, for example, as ingredients for rubber compositions for non‐flammable, water‐ and wear‐proof tires or as ion‐exchanging and complexing sorbents of heavy and noble metals.  相似文献   

7.
The rate constants of the reactions of ethoxy (C2H5O), i‐propoxy (i‐C3H7O) and n‐propoxy (n‐C3H7O) radicals with O2 and NO have been measured as a function of temperature. Radicals have been generated by laser photolysis from the appropriate alkyl nitrite and have been detected by laser‐induced fluorescence. The following Arrhenius expressions have been determined: (R1) C2H5O + O2 → products k1 = (2.4 ± 0.9) × 10−14 exp(−2.7 ± 1.0 kJmol−1/RT) cm3 s−1 295K < T < 354K p = 100 Torr (R2) i‐C3H7O + O2 → products k2 = (1.6 ± 0.2) × 10−14 exp(−2.2 ± 0.2 kJmol−1/RT) cm3 s−1 288K < T < 364K p = 50–200 Torr (R3) n‐C3H7O + O2 → products k3 = (2.5 ± 0.5) × 10−14 exp(−2.0 ± 0.5 kJmol−1/RT) cm3 s−1 289K < T < 381K p = 30–100 Torr (R4) C2H5O + NO → products k4 = (2.0 ± 0.7) × 10−11 exp(0.6 ± 0.4 kJmol−1/RT) cm3 s−1 286K < T < 388K p = 30–500 Torr (R5) i‐C3H7O + NO → products k5 = (8.9 ± 0.2) × 10−12 exp(3.3 ± 0.5 kJmol−1/RT) cm3 s−1 286K < T < 389K p = 30–500 Torr (R6) n‐C3H7O + NO → products k6 = (1.2 ± 0.2) × 10−11 exp(2.9 ± 0.4 kJmol−1/RT) cm3s−1 289K < T < 380K p = 30–100 Torr All reactions have been found independent of total pressure between 30 and 500 Torr within the experimental error. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 860–866, 1999  相似文献   

8.
Photolysis of t-BuHgCl/KI with PhC(R2)C(R1)NO2 forms PhC(R2)C(R1)Bu-t when R1 = R2 = H or in low yield when R1 = H, R2 = Ph. When R1 ≠ H, or when R2 = Ph, reactions with t-BuHgI/KI/hv proceed mainly via PhC(R2)C(R1)NO2·-, PhC(R2)C(R1)N(OBu-t)OHgX+, PhC(R2)C(R1)NO and PhC(R2)C(R1)N(OBu-t)HgX to form a variety of novel products including the dimeric bisnitronic esters ( 6 ) with R1 = Me or Ph and R2 = H; PhCH(R2)C(R1) = NOBu-t with R1 = Me or Ph and R2 = H or R1 = H and R2 = Ph; PhC(R2)(OBu-t)C(R1)NOH with R1 = H or Me and R2 = Ph; and 3-phenyl-2-R1-indoles with R1 = H, Me, Ph, PhS or t-BuS and R2 = Ph. Nitrosoaromatics react with t-BuHgX in the dark to form ArN(OBu-t)(OBu-t)HgX+ which condenses with ArNO to form the azoxy compound. tert-Butyl radicals will add to RNO2 [R = Ph, Ph2CCH, Ph2CC(Ph)] in the presence of t-BuHgI2 to form products derived from RN(OBu-t)OHgI+.  相似文献   

9.
Reactions of triaminophosphane (R2N)2P–NH2, (R = 1Pr) 1a, with aminodihalogenophosphanes 1Pr2N–PX2, 2a–c [X = CL (a), Br (b), I(c)], in the presence of a base yielded the P-hydrogeno-iminophosphoranyl-halogenophosphanes (R2N)2PH = N–PX–N(1Pr)2 4a–c [X = Cl (a), Br (b), I(c)]. Analogous reactions between 1a and 1b (b: R = c-hexyl) and chloroiminophosphane (Cl–P = N–Mes*, (Mes* = 2,4,6-tBu3C6H2) 6 , gave the P-hydrogeno(iminophosphoranyl)-σ23-iminophosphanes, (R2N)2PH = N–P = N–Mes* 8a and 8b. In solution 8a, 8b eliminated amine, yielding σ2, λ3-iminophosphanyl-substituted 1,3,2,4-diazadiphosphetidines [(R2N)PN(P = N–Mes*)]2, 10a, 10b , and 11 ( 10a and 10b : cis; 11: trans). The X-ray structure analyses of compounds 4a, 4b, 8a, and 11 are discussed. © 1996 John Wiley & Sons, Inc.  相似文献   

10.
A Heck–Matsuda process for the synthesis of the otherwise difficult to access compounds, β‐arylethenesulfonyl fluorides, is described. Ethenesulfonyl fluoride (i.e., vinylsulfonyl fluoride, or ESF) undergoes β‐arylation with stable and readily prepared arenediazonium tetrafluoroborates in the presence of the catalyst palladium(II) acetate to afford the E‐isomer sulfonyl analogues of cinnamoyl fluoride in 43–97 % yield. The β‐arylethenesulfonyl fluorides are found to be selectively addressable bis‐electrophiles for sulfur(VI) fluoride exchange (SuFEx) click chemistry, in which either the alkenyl moiety or the sulfonyl fluoride group can be the exclusive site of nucleophilic attack under defined conditions, making these rather simple cores attractive for covalent drug discovery.  相似文献   

11.
The reaction of elemental Se with 1,3-dimethylimidazolium iodide in methanolic K2CO3 yields 1,3-dimethyl-2(3H)-imidazoleselone for which three addition products, two with bromine and one with iodomethane, have been synthesized and for which X-ray crystallographic analysis shows the structure to consist of a selenium-substituted planar heterocyclic ring with bond distances and angles significantly different from those noted for the previously reported sulfur analog [1,3-dimethyl-2(3H)-imidazolethione, dmit]. Crystal data: C5H8N2Se, space group C mcm, M = 175.03, a = 8.625(3), b = 11.447(6), c = 6.900(4) Å, V = 681.24 Å3, Z = 4, Dc = 1.707 g cm−3, D0 = 1.68 g cm−3, λ = 0.71073 Å (Mo-Kα), μ = 5.35 mm−1, R = 0.034, and Rw = 0.031.  相似文献   

12.
A series of chiral pentane‐2,4‐diyl‐based thioether‐amine ligands [ 4 and 5 ; (R,S)‐ and (S,S)‐R1SCH(CH3)CH2CH(CH3)NHR2, respectively, where 4a R1 = iPr, R2 = Ph; 4b R1 = tBu, R2 = Ph; 4c R1 = 1‐Ad, R2 = Ph; 5a R1 = iPr, R2 = Ph; 5b R1 = tBu, R2 = Ph; 5c R1 = 1‐Ad, R2 = Ph; 5d R1 = iPr, R2 = 4‐MeOC6H4; 5e R1 = iPr, R2 = 4‐MeC6H4; 5f R1 = iPr, R2 = 3,5‐Me2C6H3] with stereogenic S‐ and N‐donor atoms has been prepared starting from cyclic sulfates via optically pure γ‐aminoalcohol or 2,4‐dimethylazetidine intermediates. The synthesis of the novel diastereomerically related ligand sets 4 and 5 was accomplished starting from the same source of chirality. The modular ligand structure and the novel synthetic strategies developed for their synthesis allowed the easy modification of the ligands’ (i) S‐ and (ii) N‐substituents, as well as (iii) the relative stereochemistry within the ligand backbone. Six‐membered [Pd(N,S)Cl2]‐type chelate complexes of the diastereomerically related ligands 4a and 5a were synthesized and characterized by X‐ray crystallography in the solid phase, by density functional theory calculations and in solution by NMR spectroscopy. The coordination of 5a resulted in the formation of a single chair conformation by the stereospecific locking of both stereolabile (N and S) donor atoms. In contrast, compound 4a forms rapidly equilibrating palladium species due to the fast inversion of the sulfur donor. Ligands with stereochemically fixed donor atoms provided robust and efficient catalytic systems that can be effectively applied in alkylene carbonates as green reaction media. Remarkably, the phosphine‐free catalysts are air‐stable, and at room temperature in the presence of moisture gave excellent ee’s (up to 93%) in asymmetric allylation processes thanks to the double stereoselective coordination.  相似文献   

13.
《Polyhedron》1987,6(5):849-853
Addition of R4NX (R = Et, n-Pr, n-Bu) to solutions of SeX2 (X = Cl, Br) in acetonitrile results in the formation of tri- and tetra-haloselenate(II) complexes (SeX3 and SeX42−). Raman spectroscopic characterization of these solutions and solids, R4NSeX3 and (R4N)2SeX4, which may be crystallized from them, are reported. R4NSeCl3 compounds are shown to be relatively unstable; they are readily destroyed by laser light, 2R4NSeCl3 = (R4N)2SeCl6+Se(O) and especially sensitive to hydrolysis, 2SeCl3+H2O = SeOCl3+2HCl+Cl+ Se(O). The Raman spectra of solutions of SeX3 and SeX42− are consistent with T-shaped and square planar structures, respectively.  相似文献   

14.
The reactions of 4‐hydroxy‐2,2,6,6‐tetramethylpiperidinium N‐oxide, an oxammonium ion abbreviated R2NO+, have been studied. The previously unreported triflate salt was used in this study because the anions of the usual chloride and bromide salts can themselves be oxidized. Reactions between R2NO+ and alcohols produce ketones and aldehydes; the rate constant for PhCH2OH is 4.4 × 10−3 L mol−1 s−1 in acetonitrile at 298 K. The immediate product is the hydroxylamine, R2NOH, but its further comproportionation reaction with R2NO+ yields the stable piperidinyl oxyl radical, R2NO·. The rate constant of this reaction is 1.78 × 103 L mol−1 s−1 at 298 K. The possibility of using R2NO+ and MTO as co‐catalysts for the oxidation of alcohols was explored, but the competitive rates are such that the resultant is not particularly attractive. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 381–385, 1999  相似文献   

15.
Abstract

Compounds of the following structure

(R1O)2(X)P[sbnd]Y–P(X)(OR2)2

(X = O, Y = Sn (n = 1–4), R1 = R2 = Me, iPr;

X = S, Y = Sn (n = 1–4), R1, R2 = Me, Et, iPr, iBu;

X = S, Y = S-Se-S, S-Te-S, R1 = R2 = Me

were prepared and their NMR spectra were analysed. Depending on the number of sulfur atoms, bonded between the phosphorus atoms, typical ranges of the P-P coupling constants were found for the different sulfanes investigated: 2JPP from-10 to-20 Hz, 3JPP less than 3 Hz, 4JPP from +10 to +13 Hz and 5JPP less than 1 Hz. For the small vicinal coupling constants and the relatively large values of 4JPP different possibilities of their interpretation are given.  相似文献   

16.
Heteronuclear transition‐metal–main‐group‐element carbonyl complexes of AsFe(CO)3, SbFe(CO)3, and BiFe(CO)3 were produced by a laser vaporization supersonic ion source in the gas phase, and were studied by mass‐selected IR photodissociation spectroscopy and advanced quantum chemistry methods. These complexes have C3v structures with all of the carbonyl ligands bonded on the iron center, and feature covalent triple bonds between bare Group 15 elements and Fe(CO)3. Chemical bonding analyses on the whole series of AFe(CO)3 (A=N, P, As, Sb, Bi, Mc) complexes indicate that the valence orbitals involved in the triple bonds are hybridized 3d and 4p atomic orbitals of iron, leading to an unusual (dp–p) type of transition‐metal–main‐group‐element multiple bonding. The σ‐type three‐orbital interaction between Fe 3d/4p and Group 15 np valence orbitals plays an important role in the bonding and stability of the heavier AFe(CO)3 (A=As, Sb, Bi) complexes.  相似文献   

17.
The reaction of fac‐[MIIIF3(Me3tacn)]⋅x H2O with Gd(NO3)3⋅5H2O affords a series of fluoride‐bridged, trigonal bipyramidal {GdIII3MIII2} (M=Cr ( 1 ), Fe ( 2 ), Ga ( 3 )) complexes without signs of concomitant GdF3 formation, thereby demonstrating the applicability even of labile fluoride‐complexes as precursors for 3d–4f systems. Molecular geometry enforces weak exchange interactions, which is rationalized computationally. This, in conjunction with a lightweight ligand sphere, gives rise to large magnetic entropy changes of 38.3 J kg−1 K−1 ( 1 ) and 33.1 J kg−1 K−1 ( 2 ) for the field change 7 T→0 T. Interestingly, the entropy change, and the magnetocaloric effect, are smaller in 2 than in 1 despite the larger spin ground state of the former secured by intramolecular Fe–Gd ferromagnetic interactions. This observation underlines the necessity of controlling not only the ground state but also close‐lying excited states for successful design of molecular refrigerants.  相似文献   

18.
Treatment of N,N‐dichloroperfluoroalkanesulfonylamines with sulfur powder at room temperature gave the title products RfSO2N=SCl2 in good yields. They reacted readily with dimethyl sulfoxide, chloral, DMF, benzophenone, and similar compounds to form to corresponding imines RfSO2N=YR1R2 (Y: S, C). A reaction mechanism, one involving formation of a four‐membered intermediate, is proposed. © 1999 John Wiley & Sons, Inc. Heteroatom Chem 10: 41–48, 1999  相似文献   

19.
The title compound anilinium chloride–4‐bromo‐N‐phenyl­benzene­sulfonamide (1/1), C6H8N+·Cl·C12H10BrNO2S, displays a hydrogen‐bonded ladder motif with four independent N—H⋯Cl bonds in which both the NH group of the sulfonamide molecule and the NH3 group of the anilinium ion [N⋯Cl = 3.135 (3)–3.196 (2) Å and N—H⋯Cl = 151–167°] are involved. This hydrogen‐bonded chain contains two independent R42(8) rings and each chloride ion acts as an acceptor of four hydrogen bonds.  相似文献   

20.
A series of organotin(IV) compounds R3Sn(A) where R = Me or Ph and A is a chromogenic nitrophenolate ligand were prepared and studied as possible colorimetric sensors for anions (F, Cl, Br, AcO, H2PO4). Equilibrium constants for a complete set of reactions between R3Sn(A) with A = 2‐amino‐4‐nitrophenolate (ANP) or 4‐nitrophenolate and anions (X) involving formation of complexes R3Sn(A)(X) and substitution products R3Sn(X) and R3Sn(X)2 were determined by UV‐vis and 1H NMR titrations in MeCN and DMSO. The binding selectivity was AcO > F > H2PO4 > Cl ≫ Br in both solvents and both for R = Me and Ph with higher affinity for R = Ph. Compounds with A = ANP were found to have the optimum properties as anion sensors allowing optical detection of F, AcO and H2PO4 anions in the 5–100 µM range by appearance of an intense absorption band of free ANP resulting from its substitution with the analyte. Selectivity and affinity of anion interactions with R3Sn(ANP) are similar to those for thiourea receptors, but the organotin receptor produces a much larger naked eye detected optical signal, operates equally well in nonpolar and polar solvents and tolerates the presence of up to 20% vol. of water in DMSO. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号