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 500 毫秒
1.
Hexakis(N—allylthiourea)tetracopper(I) Tetratrifluoromethanesulfonate, [Cu4{CH2=CHCH2NHC(S)NH2}6](CF3SO3)4 (sp.gr.P21/n, a = 13.5463(8), b = 24.129(2), c = 19.128(1)Å, β = 108.053(6)°, Z = 4, R = 0.0440 for 13548 unique reflections) was obtained by reduction of Cu(CF3SO3)2 with excess of N—allylthiocarbamide in benzene medium. Four crystallographical independent Cu atoms possess trigonal environment of three S atoms of CH2=CHCH2NHC(S)NH2 moiety and form Cu4S64+ adamantane—like fragments. The latteres are connected with CF3SO3 anions via (C)—H···F hydrogen bonds.  相似文献   

2.
Enantiomerically pure triflones R1CH(R2)SO2CF3 have been synthesized starting from the corresponding chiral alcohols via thiols and trifluoromethylsulfanes. Key steps of the syntheses of the sulfanes are the photochemical trifluoromethylation of the thiols with CF3Hal (Hal=halide) or substitution of alkoxyphosphinediamines with CF3SSCF3. The deprotonation of RCH(Me)SO2CF3 (R=CH2Ph, iHex) with nBuLi with the formation of salts [RC(Me)? SO2CF3]Li and their electrophilic capture both occurred with high enantioselectivities. Displacement of the SO2CF3 group of (S)‐MeOCH2C(Me)(CH2Ph)SO2CF3 (95 % ee) by an ethyl group through the reaction with AlEt3 gave alkane MeOCH2C(Me)(CH2Ph)Et of 96 % ee. Racemization of salts [R1C(R2)SO2CF3]Li follows first‐order kinetics and is mainly an enthalpic process with small negative activation entropy as revealed by polarimetry and dynamic NMR (DNMR) spectroscopy. This is in accordance with a Cα? S bond rotation as the rate‐determining step. Lithium α‐(S)‐trifluoromethyl‐ and α‐(S)‐nonafluorobutylsulfonyl carbanion salts have a much higher racemization barrier than the corresponding α‐(S)‐tert‐butylsulfonyl carbanion salts. Whereas [PhCH2C(Me)SO2tBu]Li/DMPU (DMPU = dimethylpropylurea) has a half‐life of racemization at ?105 °C of 2.4 h, that of [PhCH2C(Me)SO2CF3]Li at ?78 °C is 30 d. DNMR spectroscopy of amides (PhCH2)2NSO2CF3 and (PhCH2)N(Ph)SO2CF3 gave N? S rotational barriers that seem to be distinctly higher than those of nonfluorinated sulfonamides. NMR spectroscopy of [PhCH2C(Ph)SO2R]M (M=Li, K, NBu4; R=CF3, tBu) shows for both salts a confinement of the negative charge mainly to the Cα atom and a significant benzylic stabilization that is weaker in the trifluoromethylsulfonyl carbanion. According to crystal structure analyses, the carbanions of salts {[PhCH2C(Ph)SO2CF3]Li? L }2 ( L =2 THF, tetramethylethylenediamine (TMEDA)) and [PhCH2C(Ph)SO2CF3]NBu4 have the typical chiral Cα? S conformation of α‐sulfonyl carbanions, planar Cα atoms, and short Cα? S bonds. Ab initio calculations of [MeC(Ph)SO2tBu]? and [MeC(Ph)SO2CF3]? showed for the fluorinated carbanion stronger nC→σ* and nO→σ* interactions and a weaker benzylic stabilization. According to natural bond orbital (NBO) calculations of [R1C(R2)SO2R]? (R=tBu, CF3) the nC→σ*S? R interaction is much stronger for R=CF3. Ab initio calculations gave for [MeC(Ph)SO2tBu]Li ? 2 Me2O an O,Li,Cα contact ion pair (CIP) and for [MeC(Ph)SO2CF3]Li ? 2 Me2O an O,Li,O CIP. According to cryoscopy, [PhCH2C(Ph)SO2CF3]Li, [iHexC(Me)SO2CF3]Li, and [PhCH2C(Ph)SO2CF3]NBu4 predominantly form monomers in tetrahydrofuran (THF) at ?108 °C. The NMR spectroscopic data of salts [R1(R2)SO2R3]Li (R3=tBu, CF3) indicate that the dominating monomeric CIPs are devoid of Cα? Li bonds.  相似文献   

3.
CF3SO2N?SCl2 reagiert mit (CH3)2S[NSi(CH3)3]2, (C4H8)S[NSi(CH3)3]2 oder (C5H10)S[NSi(CH3)3]2 unter Trimethylchlorsilanabspaltung zu den achtgliedrigen S4N4-Derivaten S4N4(NSO2CF3)2(CH3)4 3 , S4N4(NSO2CF3)2(C4H8)2 4a und S4N4(NSO2CF3)2(C5H1 0)2 4b . In den achtgliedrigen SN-Ringen haben die Schwefelatome die Koordinationszahl 3 und 4. Die Röntgenstrukturanalyse von 4a ergab eine Sessel-Konformation. 4a kristallisiert orthorhombisch in der Raumgruppe Pna21 mit a = 17,641(4), b = 6,406(2), c = 19,130(4) Å, dx = 1,815 g cm?3 und Z = 4. Die mittleren S? N-Abstände betragen an den vierfach koordinierten Schwefelatomen 1,597 Å und an den Schwefelatomen mit der Koordinationszahl 3 1,650 Å. CF3SO2N? SCl2 reagiert mit trimethylzinnhaltigen S? N-Verbindungen zum bekannten CF3SO2N[Sn(CH3)3]S(CH3)NSO2CF3 und Dimethylzinndichlorid. Synthesis and X-Ray Structure Analysis of S4N4-Derivatives with Threefold and Fourfold Coordinated Sulfur Atoms CF3SO2N?SCl2 reacts with (CH3)2S[NSi(CH3)3]2, (C4H8)S[NSi(CH3)3]2 or (C5H10S[NSi(CH3)2]2 under elimination of (CH3)3SiCl to yield the eight-membered S4N4 derivatives S4N4?NSO2CF3)2(CH3)4, 3 , S4N4(NSO2CF3)2(C4H8)2 4a und S4N4(NSO2CF3)2(C5H1 0)2 4b . In the eight-membered SN-rings the sulfur atoms have the coordination number 3 and 4. The X-ray structure analysis of 4a revealed a chair conformation. 4a crystallizes in the orthorhombic space group Pna21 with a = 17.641(4), b = 6.406(2), c = 19.130(4) Å, dx = 1.815 g cm?3, and Z = 4. The average S? N distance was found to be 1.597 Å at fourfold coordinated sulfur atoms and 1.650 Å at sulfur with coordination number 3. CF3SO2N=SCl2 reacts with trimethyl tin-containing S? N compounds to the known CF3SO2N[Sn(CH3)3]S(CH3)NSO2CF3 and dimethyl tin dichloride.  相似文献   

4.
Trifluoromethyl trifluoromethanesulfonate has proved to be an excellent reservoir of difluorophosgene and a promising click ligation for amines in the preparation of urea derivatives, heterocycles, and carbamoyl fluorides under metal- and additive-free conditions. The reactions are rapid, efficient, selective, and versatile, and can be performed in benign solvents, giving products in excellent yields with minimal efforts for purification. The characteristics of the reactions meet the requirements of a click reaction. The use of trifluoromethyl trifluoromethanesulfonate as a click reagent is advantageous over other “CO” sources (e.g., TsOCF3, PhCO2CF3, CsOCF3, AgOCF3, and triphosgene) because this reagent is readily accessible; easy to scale up; and highly reactive, even under metal- and additive-free conditions. It is anticipated that CF3SO3CF3 will be increasingly as important as SO2F2 as a click agent in future drug design and development.  相似文献   

5.
Products from the reaction of 11-dihomodriman-8α-ol-12-one with several reagents such as MeSO3SiMe3, CF3SO3SiMe3, Sc(CF3SO3)3, conc. H2SO4 in EtOH (30% solution), and Amberlist-15 ion-exchange resin were studied. 11-Dihomodrim-8(9)-en-12-one and its oxime were synthesized. The reaction of its oxime with H3PO4 (86%) or CF3CO2H produced (1S,2S,4aS,8aS)-2,5,5,8a-tetramethyldecahydro-1H-naphtho [1,2-e]-3-methyl-4,5-dihydro-[1,2,6]-oxazine; with p-TsCl in Py, (1S,2S,4aS,8aS)-2,5,5,8a-tetramethyldecahydro-1H-naphtho[1,2-d]-2-methylpyrroline-N-oxide; and with PCl5 in Et2O, 11-acetylaminodrim-8(9)-ene and 11-methylaminooxodrim-8(9)-ene.  相似文献   

6.
The reaction of (NH4)2PbCl6 and fuming sulfuric acid (65 % SO3) in a sealed glass tube at 250 °C led to colorless single crystals of Pb[S3O10] (orthorhombic, Pbcn, Z = 4, a = 10.9908(4), b = 8.5549(3), c = 8.0130(3) Å, V = 753.42(5) Å3). The compound shows a three‐dimensional linkage of the tenfold oxygen coordinated Pb2+ ions and exhibits the unusual trisulfate anion, [S3O10]2–, that consists of three vertex connected [SO4] tetrahedra. The distances S–O within the S–O–S bridges of the anion are remarkable asymmetric with distances of 155 and 169 pm, respectively. This structural feature is well reproduced by calculations on a PBE0/cc‐pVTZ and a MP2/cc‐pVTZ level of theory. Similar calculations allow also for an inspection of the yet unknown corresponding acid, H2S3O10. Also for this acid non‐symmetric S–O–S bridges are predicted. The thermal behavior of Pb[S3O10] is characterized by the loss of two equivalents of SO3 at low temperature and the decomposition of intermediate Pb[SO4] at higher temperature.  相似文献   

7.
《Tetrahedron: Asymmetry》2000,11(13):2765-2779
The ligands 6-[(diphenylphosphanyl)methoxy]-4,8-di-tert-butyl-2,10-dimethoxy-5,7-dioxa-6-phosphadibenzo[a,c]cycloheptene, 1, (S)-4-[(diphenylphosphanyl)methoxy]-3,5-dioxa-4-phosphacyclohepta[2,1-a;3,4a′]dinaphthalene, (S)-2, and (S)-4-[(diphenylphosphanyl)methoxy]-2,6-bis-trimethylsilanyl-3,5-dioxa-4-phosphacyclohepta[2,1-a;3,4-a′]dinaphthalene, (S)-3, (S)-2-(3,5-dioxa-4-phosphacyclohepta[2,1-a;3,4-a′]dinaphthalen-4-yloxymethyl)pyridine, (S)-4, and (S)-2-(3,5-dioxa-4-phosphacyclohepta[2,1-a;3,4-a′]dinaphthalen-4-yloxy)pyridine, (S)-5, have been easily prepared.The cationic complexes [Pd(η3-C3H5)(L-L′)]CF3SO3 (L–L′=1–(S)-5) and [Pd(η3-PhCHCHCHPh)(L–L′)]CF3SO3 (L–L′=(S)-2–(S)-4) were synthesized by conventional methods starting from the complexes [Pd(η3-C3H5)Cl]2 and [Pd(η3-PhCHCHCHPh)Cl]2, respectively. The behavior in solution of all the π-allyl- and π-phenylallyl-(L–L′)palladium derivatives 614 was studied by 1H, 31P{1H}, 13C{1H} NMR and 2D-NOESY spectroscopy. As concerns the ligands (S)-4 and (S)-5, a satisfactory analysis of the structures in solution was possible only for palladium–allyl complexes [Pd(η3-C3H5)((S)-4)]CF3SO3, 11, and [Pd(η3-C3H5)((S)-5)]CF3SO3, 12, since the corresponding species [Pd(η3-PhCHCHCHPh)((S)-4)]CF3SO3, 13, and [Pd(η3-PhCHCHCHPh)((S)-5)]CF3SO3, 14, revealed low stability in solution for a long time. The new ligands (S)-2–(S)-5 were tested in the palladium-catalyzed enantioselective substitution of (1,3-diphenyl-1,2-propenyl)acetate by dimethylmalonate. The precatalyst [Pd(η3-C3H5)((S)-2)]CF3SO3 afforded the allyl substituted product in good yield (95%) and acceptable enantioselectivities (71% e.e. in the S form). A similar result was achieved with the precatalyst [Pd(η3-C3H5)((S)-3)]CF3SO3. The nucleophilic attack of the malonate occurred preferentially at allylic carbon far from the binaphthalene moiety, namely trans to the phosphite group. When the complexes containing ligands (S)-4 and (S)-5 were used as precatalysts, the product was obtained as a racemic mixture in high yield. The number of the configurational isomers of the Pd-allyl intermediates present in solution in the allylic alkylation and the relative concentrations are considered a determining factor for the enantioselectivity of the process.  相似文献   

8.
Reactions of Li+ [(η5-C5H5)Re(NO)(PPh3)] with 2- and 4-chloroquinoline or 1-chloroisoquinoline give the corresponding σ quinolinyl and isoquinolinyl complexes 3 , 6 , and 8 . With 3 and 8 there is further protonation to yield HCl adducts, but additions of KH give the free bases. Treatment of 3 with HBF4⋅OEt2 or H(OEt2)2+ BArf gives the quinolinium salts [(η5-C5H5)Re(NO)(PPh3)(C(NH)C(CH)4C (CH)(CH))]+ X ( 3-H + X; X=BF4/BArf, 94–98 %). Addition of CF3SO3CH3 to 3 , 6 , or 8 affords the corresponding N-methyl quinolinium salts. In the case of [(η5-C5H5)Re(NO)(PPh3)(C(NCH3)C(CH)4C (CH)(CH))]+ CF3SO3 ( 3-CH3 + CF3SO3), addition of CH3Li gives the dihydroquinolinium complex (SReRC,RReSC)-[(η5-C5H5)Re(NO)(PPh3)(C(NCH3)C(CH)4C (CHCH3)(CH2))]+ CF3SO3 ((SReRC,RReSC)- 5 + CF3SO3, 76 %) in diastereomerically pure form. Crystal structures of 3-H + BArf, 3-CH3 + CF3SO3, (SReRC, RReSC)- 5 + Cl, and 6-CH3 + CF3SO3 show that the quinolinium ligands adopt Re⋅⋅⋅ C conformations that maximize overlap of their acceptor orbitals with the rhenium fragment HOMO, minimize steric interactions with the bulky PPh3 ligand, and promote various π interactions. NMR experiments establish the Brønsted basicity order 3 > 8 > 6 , with Ka(BH+) values >10 orders of magnitude greater than the parent heterocycles, although they remain less active nucleophilic catalysts in the reactions tested. DFT calculations provide additional insights regarding Re⋅⋅⋅ C bonding and conformations, basicities, and the stereochemistry of CH3Li addition.  相似文献   

9.
The reactions of py‐hz ligands ( L1–L5 ) with Pb(CF3SO3)2?H2O resulted in some rare examples of discrete single‐stranded helical PbII complexes. L1 and L2 formed non‐helical mononuclear complexes [Pb L1 (CF3SO3)2]?CHCl3 and Pb L2 (CF3SO3)2][Pb L2 CF3SO3]CF3SO3?CH3CN, which reflected the high coordination number and effective saturation of PbII by the ligands. The reaction of L3 with PbII resulted in a dinuclear meso‐helicate [Pb2 L3 (CF3SO3)2Br]CF3SO3?CH3CN with a stereochemically‐active lone pair on PbII. L4 directed single‐stranded helicates with PbII, including [Pb2 L4 (CF3SO3)3]CF3SO3?CH3CN and [Pb2 L4 CF3SO3(CH3OH)2](CF3SO3)3?2 CH3OH?2 H2O. The acryloyl‐modified py‐hz ligand L5 formed helical and non‐helical complexes with PbII, including a trinuclear PbII complex [Pb3 L5 (CF3SO3)5]CF3SO3?3CH3CN?Et2O. The high denticity of the long‐stranded py‐hz ligands L4 and L5 was essential to the formation of single‐stranded helicates with PbII.  相似文献   

10.
Lü Jian 《中国化学》2011,29(2):283-287
The effect of metallic ions on the nitrolysis of DAPT [3,7‐diacetyl‐1,3,5,7‐tetraazabicyclo(3.3.1)nonane] and HA (hexamine) was investigated by experimental and theoretical approaches. The combinatorial reagent, M(NO?3)n/Ac2/NH4NO3 (M=Mg2+, Cu2+, Pb2+, Bi3+, Fe3+ and Zr4+), was found to be efficient in the experiment of the nitrolysis of DAPT. A key intermediate during the nitrolysis of DAPT was detected by 1H NMR. The formation mechanism of the intermediate was proposed and analyzed. Some discrepant results for the nitrolysis of DAPT and HA catalyzed by different metallic nitrates were explained based on hard‐soft and acid‐base principle and stabilized energy of ion‐complex. From the latter point of view, some cations with high polarizable ligands, e.g., OSO2CF3?, (CF3SO2)2N?, and (C4F9SO2)2N?, can increase the yields. Two newly designed catalysts, Cu[(CF3SO2)2N]2 and Cu[(C4F9SO2)2N]2, were tested to be highly efficient.  相似文献   

11.
Synthetic Application of Epoxynitrones I. Nitrone, a New α-Methylidene-γ-lactone Annelating Reagent The N-(2, 3-epoxypropyliden)-cyclohexylamine-N-oxide/CF3SO3SiR3 reagent descried in this communication opens a new and interesting entry to the versatile N-substituted N-propenylnitrosonium ions of type b (Scheme 6). One of the uses of this reagent is shown to be the synthesis of α-methylidene-γ-lactones from olefins. This new method shows similar features as the method based on 2, 3-dichloropropylidenamine-oxide/AgBF4 originally developed for the same purpose by Petrzilka, Felix and Eschenmoser. Epoxynitrone 18 can be transformed to the positively charged heterodiene of type b (Scheme 5) using the highly electrophilic reagents CF3SO3SiMe3 ( 23 ) and CF3SO3Si (t-Bu)Me2 ( 24 ), respectively. Low temperature 1H- and 13C-NMR. spectroscopy at ?78° showed the sole formation of the nitrone-O-silyl-ethers a (Scheme 5). Epoxid opening leading to the diene b and subsequent reactions are observed only at about ?30°. The diene b prepared in situ, adds to isolated double bonds by way of an inverse Diels-Alder reaction to afford cycloadducts of type 27 (Scheme 7). Their stable cyanoderivatives, e.g. 28 (Scheme 7), can be isolated and transformed via 31 , 44 and 54 into cis annelated α-methylidene-γ-lactones of type 55 (Scheme 11). Using trisubstituted olefins, substitution at the lower substituted olefinic C-atom competes efficiently with the cycloaddition (e.g. 34 , Scheme 8).  相似文献   

12.
In order to synthesize poly-(fluorinated alkanesulfonamides) a series of model experiments were carried out: (1) reactions of fluorinated alkanesulfonyl fluorides with amines, (2) reactions of fluorinated alkanesulfonyl chloride with amines and (3) reactions of sodium salts of fluorinated alkanesulfonamides with alkyl iodides of fluorinated alkanesulfonic acid esters. Seventeen new fluorinated alkanesulfonamides were prepared in good yields, namely: RFO(CF2)2SO2NR1R2 (1a-h), R1R2NSO2RFSO2NR1R2 (2a-h) and [Cl (CF2)4O(CF2)2SO2NH(CH2)3]2 (3). Reaction of RFSO2NH2 with equivalent amount of NaOCH3 and methyl iodide was shown to give both the N-mono- and N,N-di-substituted amides. Consequently the N-monosubstituted alkanesulfonamides were chosen as monomers for syntheses of the poly-(fluorinated alkanesulfonamides) and two new polymers were synthesized. The effect of the condition of the polycondensation on M?n of the polymers were discussed and elemental composition, 19F NMR, IR, M?n, Tg, tensile strength, thermal and chemical stabilities of the polymers were measured. Several new perfluoroalkanesulfonyl chlorides CISO2RFSO2Cl (4a-c) and fluorinated alkanesulfonic acid esters (6a-d) were synthesized. However, reaction of CFCl2CF2O(CF2)2SO2F with AlCl3 was found to give Cl3CCF2O(CF2)2SO2F (5) instead of the expected sulfonyl chloride.  相似文献   

13.
In contrast to RFSO3CH2R(1)(R=hydrogen, alkyl and perfluoroalkyl) and RFSO3CF2RF′ (2), the reactions of difluoromethyl perfluoroalkanesulfonates RFSO3CF2H (3) With nucleophiles are more complicated. Halide inos, X? (X = F, Cl, I) and ethanol only attack the alkoxyl carbon atom, cleaving the C? O bond to give HCF2X (4) and HCF2OEt (5) respectively. Other reagents such as RCO2? (R=CH3, CF3), C6H5S? etc. can either attack the carbon or sulfur atom of 3 to give the corresponding products of C? O and S? O bond cleavages. More basic nucleophiles RO? (R = C6H5, Et) mainly abstract the proton of the HCF2 moiety to produce difluorocarbene. Ether and benzene, which can be alkylated by methyl perfluoroalkanesulfonate, do not react with 3 under similar conditions. The reaction rate of 3 with KF is much slower than that of 1 (R = H). All these data seem to indicate that the shielding effect caused by the two fluorine atoms on the methyl carbon in 3 prevents to some extent the nucleophilic attack on this carbon, but not so completely as in 2 due to the presence of a hydrogen atom.  相似文献   

14.
郭丽  虞忠衡  朱士正  陈庆云 《化学学报》2005,63(10):897-902
用密度泛函理论研究了CF3SO3CF2CF3+F的碳氧键断裂反应的机理. 首先, 用DFT方法优化了反应物、中间体、过渡态、产物的平衡构型, 分析了碳氧键断裂反应的势能面变化. 发现在SN2反应机理中, 除了S—O断裂SN2反应外, 引起C—O键断裂的同面进攻也是一个可能的反应途径. 理论计算表明, 最终反应的产物是受热力学控制的, S—O键的断裂绝对地优于C—O的断裂. 因此, C—O断裂的同面机理虽然是可能的, 但却难以被实验观察到. 本文还讨论了端基 —F3在同面SN2反应中的邻位效应, 以及基组对这个效应的影响.  相似文献   

15.
Using the concept of the additivity of partial molar heat pulses, a series of mixtures containing all or some of the following anions (S2?, SO2?4, SO2?3, S2O2?3) and S have been determined by the use of selected but non-selective reactions. The accuracy at the mM level is within 1.5%.  相似文献   

16.
The reactions of hydrogen atoms produced by the mercury-photosensitized decomposition of H2 with bis(trifluoromethyl)disulfide has been studied. The rate coefficient for the primary reaction, H + CF3SSCF3 → CF3SH + CF3S, was determined in competition with the reaction H + C2H4S → SH + C2H4 to have the value k = (3.0 ± 0.18) × 1014 exp[-(4560 ± 140)/RT] cm3 mol?1 S?1. The high A factor can be partially accounted for by assuming free rotation for the two CF3 groups and the SCF3 groups about the S—S bond in the transition state. The relatively high activation energy is attributed to inductive and orbital overlap effects. CH3SH, H2S, and CF3SH all react with CF3SSCF3 to yield solid complexes which were not explored further.  相似文献   

17.
On Sn[OCH(CF3)2]2 and Sn(OCH2CF3)2 (n = 1, 2) The sulfoxylates S[OCH(R)CF3]2, 1 and 2 and the disulfides S2[OCH(R)CF3]2, 5 and 6 (R = CF3, H) are obtained by reacting SCl2 or S2Cl2, respectively, and the lithium alcoxides LiOCH(R)CF3. Chlorine and compound 2 give ClS(O)OCH2F3 and CF3CH2Cl, whereas the sulfur-sulfur bound is cleaved in 5 and 6 furnishing SCI2, 1 and 2 , respectively. The 19F n.m.r. spectrum of 5 and the 1H n.m.r. spectrum of 6 are interpreted in terms of hindered rotation about the sulfur-sulfur axis.  相似文献   

18.
The salts [RuII(L–L)3](CF3SO3)2 (L–L = bpy or phen) have been prepared in high yields via reactions of [RuII(DMF)6](CF3SO3)2 (DMF = N,N-dimethylformamide), generated in situ by reduction of [RuIII(DMF)6]-(CF3SO3)3, with an excess of bpy or phen at room temperature in DMF solutions.  相似文献   

19.
A general class of C3‐symmetric Ag9 clusters, [Ag9S(tBuC6H4S)6(dpph)3(CF3SO3)] ( 1 ), [Ag9(tBuC6H4S)6(dpph)3(CF3SO3)2] ? CF3SO3 ( 2 ), [Ag9(tBuC6H4S)6(dpph)3(NO3)2] ? NO3 ( 3 ), and [Ag9(tBuC6H4S)7(dpph)3(Mo2O7)0.5]2 ? 2 CF3COO ( 4 ) (dpph=1,6‐bis(diphenylphosphino)hexane), with a twisted trigonal‐prism geometry was isolated by the reaction of polymeric {(HNEt3)2[Ag10(tBuC6H4S)12]}n, 1,6‐bis(diphenylphosphino)hexane, and various silver salts under solvothermal conditions. The structures consist of discrete clusters constructed from a girdling Ag9 twisted trigonal prism with the top and bottom trigonal faces capped by diverse anions (i.e., S2? and CF3SO3? for compound 1 , 2×CF3SO3? for compound 2 , 2×NO3? for compound 3 , and tBuC6H4S? and Mo2O72? for compound 4 ). This trigonal prism is bisected by another shrunken Ag3 trigon at its waist position. Interestingly, two inversion‐related Ag9 trigonal‐prismatic clusters are dimerized by the Mo2O72? ion in compound 4 . The twist is amplified by the bulkier thiolate, which also introduces high steric‐hindrance for the capping ligand, that is, the longer dpph ligand. Four more silver–sulfur clusters (namely, compounds 5 – 8 ) with their nuclearity ranging from 6–10 were solely characterized by single‐crystal X‐ray diffraction to verify the above‐described synergetic effect of mixed ligands in the construction of Ag9 twisted trigonal prisms. Surprisingly, only cluster 1 emits yellow luminescence at λ=584 nm at room temperature, which may be attributed to a charge transfer from the S 3p orbital to the Ag 5s orbital, or mixed with metal‐centered (MC) d10→d9s1 transitions. Upon cooling from 300 to 80 K, the emission intensity was enhanced along with a hypsochromic shift. The good linear relationship between the maximum emission intensity and the temperature for compound 1 in the range of 180–300 K indicates that this is a promising molecular luminescent thermometer. Furthermore, cyclic voltammetric studies indicated that the diffusion‐ and surface‐controlled redox processes were determined for compounds 1 and 3 as well as compound 4 , respectively.  相似文献   

20.
We have developed a novel and simple protocol for the direct incorporation of a difluoromethyl (CF2H) group into alkenes by visible‐light‐driven photoredox catalysis. The use of fac‐[Ir(ppy)3] (ppy=2‐pyridylphenyl) photocatalyst and shelf‐stable Hu's reagent, N‐tosyl‐S‐difluoromethyl‐S‐phenylsulfoximine, as a CF2H source is the key to success. The well‐designed photoredox system achieves synthesis of not only β‐CF2H‐substituted alcohols but also ethers and an ester from alkenes through solvolytic processes. The present method allows a single‐step and regioselective formation of C(sp3)–CF2H and C(sp3)?O bonds from C=C moiety in alkenes, such as hydroxydifluoromethylation, regardless of terminal or internal alkenes. Moreover, this methodology tolerates a variety of functional groups.  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