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 31 毫秒
1.
The addition of carboxylic acids to dimethylamino-propinal ( 1a ) and 4-dimethyl-amino-but-3-in-2-on ( 1b ) gives, after rearrangement of the very instable primary adducts ( 2 ), Z-3-acetoxy-N,N-dimethylacrylamides and -crotonamides 3 to 8 in excellent yields and in a stereospecific manner. Similarly, the adducts of HCl and HBr to the alkynes 1a and 1b may be rearranged at low temperature by traces of acid to cis/trans equilibria of 3-halo-acrylamides and -crotonamides 9 and 10 . - On the other hand, treatment of 3-alkoxy-3-dimethylaminoacrolein with traces of acid yields alkylesters of E-3-dimethylaminoacrylic acid ( 12 , X = OR). - The preparative aspects of the rearrangement are discussed, and a brief outline of the spectroscopic properties of the compounds 3 to 8 is given.  相似文献   

2.
Kai Wu  Qing‐Yun Chen 《中国化学》2001,19(12):1273-1279
In an open glassware, heating a gas HCFC‐133a (CF3CH2C1) or HFC‐134a (CF3CH2F), KOH and a phenol (or an alcohol) in DMSO at 80°C gave ethers ROCF2CH2X and (E/Z)‐ROCF = CHX (X = Cl, F) in moderate yields.  相似文献   

3.
Sodium perfluoroalkanesulfinates [Cl (CF2)n SO2 Na (1), a , n = 4; b , n = 6; c , n = 8] with the reduction potentials about 0.95—1.00V could be oxidized readily with various oxidizing agents such as Mn (OAc)3 2H2O, Ce (SO4)2, HgSO4 and Co2O3 to generate perfluoroalkyl radicals which added to the olefins RCH ? CHR' to give two kinds of adducts, namely RCH (Rf) CHXR' (3, X ? H; 4, X ? OAc), with good yields depending upon the solvent system used. Different oxidizing agents showed slight variation on the yields of the adducts. The reaction time could be greatly shortened at higher temperature. Thus, this reaction provides a new way for introducing a perfluoroalkyl group into olefinic compounds.  相似文献   

4.
Per- and polyfluoroalkyl iodides [RFI, RF=Cl(CF2)4, 1a ; Cl(CF2)6, 1b ; Cl(CF2)8, 1c ; n-C6F13, 1d ; n-C8F17, 1e ] reacted with cyclic enol ethers such as 2,3-dihydrofuran (2) and 3,4-dihydro-2H-pyran (3) in aqueous acetonitrile in the presence of sodium dithionite and sodium bicarbonate at room temperature (10–15°C) to give the corresponding 2-(F-alkyl) hemiacetals in high yields. The adducts were oxidized with Ce(NH4)2(NO3)6 in acetonitrile or reduced with LiAlH4 in ether to form the corresponding 2-(F-alkyl)lactones or diols respectively in good yields. In the presence of p-toluenesulfonic acid, the adducts were refluxed in benzene and CH3CN to produce the corresponding 2,3-dihydro-4-(F-alkyl) furan and 3,4-dihydro-5-(F-alkyl)-2H-pyran. This is a new and effective method for preparing these useful organofluorine compounds.  相似文献   

5.
Investigations on Aromatic Amino-Claisen Rearrangements The thermal and acid catalysed rearrangement of p-substituted N-(1′,1′-dimethylallyl)anilines (p-substituent=H (5) , CH3 (6) , iso-C3H7 (7) , Cl (8) , OCH3 (9) , CN (10) ), of N-(1′,1′-dimethylallyl)-2,6-dimethylaniline (11) , of o-substituted N-(1′-methylallyl)anilines (o-substituent=H (12) , CH3 (13) , t-C4H9 (14) , of (E)- and (Z)-N-(2′-butenyl)aniline ((E)- and (Z)- 16 ), of N-(3′-methyl-2′-butenylaniline (17) and of N-allyl- (1) and N-allyl-N-methylaniline (15) was investigated (cf. Scheme 3). The thermal transformations were normally conducted in 3-methyl-2-butanol (MBO), the acid catalysed rearrangements in 2N -0,1N sulfuric acid. - Thermal rearrangements. The N-(1′,1′-dimethylallyl)anilines rearrange in MBO at 200-260° with the exception of the p-cyano compound 10 in a clean reaction to give the corresponding 2-(3′-methyl-2′-butenyl)anilines 22–26 (Table 2 and 3). The amount of splitting into the anilines is <4% ( 10 gives ? 40% splitting). The secondary kinetic deuterium isotope effect (SKIDI) of the rearrangement of 5 and its 2′,3′,3′-d3-isomer 5 amounts to 0.89±0.09 at 260° (Table 4). This indicates that the partial formation of the new s?-bond C(2), C(3′) occurs already in the transition state, as is known from other established [3,3]-sigmatropic rearrangements. The rearrangement of the N-(1′-methylallyl)anilines 12–14 in MBO takes place at 290–310° to give (E)/(Z)-mixtures of the corresponding 2-(2′-Butenyl)anilines ((E)- and (Z)- 30,-31 , and -32 ) besides the parent anilines (5–23%). Since a dependence is observed between the (E)/(Z)-ratio and the bulkiness of the o-substituent (H: (E)- 30 /(Z)- 30 =4,9; t-C4H9: (E)- 32 /(Z)- 32 =35.5; cf. Table 6), it can be concluded, that the thermal amino-Claisen rearrangement occurs preferentially via a chair-like transition state (Scheme 22). Methyl substitution at C(3′) in the allyl chain hinders the thermal amino-Claisen-rearrangement almost completely, since heating of (E)-and (Z)- 16 , in MBO at 335° leads to the formation of the expected 2-(1′-methyl-allyl) aniline (33) to an extent of only 12 and 5%, respectively (Scheme 9). The main reaction (?60%) represents the splitting into aniline. This is the only observable reaction in the case of 17 . The inversion of the allyl chain in 16 - (E)- and (Z)- 30 cannot be detected - indicated that 33 is also formed in a [3, 3]-sigmatropic process. This is also true for the thermal transformation of N-allyl- (1) and N-allyl-N-methylaniline (15) into 2 and 34 , respectively, since the thermal rearrangement of 2′, 3′, 3′-d3- 1 yields 1′, 1′, 2′-d3- 2 exclusively (Table 8). These reaction are accompanied to an appreciable extent by homolysis of the N, C (1′) bond: compound 1 yields up to 40% of aniline and 15 even 60% of N-methylaniline ((Scheme 10 and 11). The activation parameters were determined for the thermal rearrangements of 1, 5, 12 and 15 in MBO (Table 22). All rearrangements show little solvent dependence (Table 5, 7 and 9). The observed ΔH values are in the range of 34-40 kcal/mol and the ΔS values very between -13 to -19 e.u. These values are only compatible with a cyclic six-membered transition state of little polarity. - Acid catalysed rearrangements. - The rearrangement of the N-(1′, 1′-dimethylallyl) anilines 5-10 occurs in 2N sulfuric acid already at 50-70° to give te 2-(3′-methyl-2′-butenyl)anilines 22-27 accompanied by their hydrated forms, i.e. the 2-(3′-hydroxy-3′-methylbutyl) anilines 35-40 (Tables 10 and 11). The latter are no more present when the rearrangement is conducted in 0.1 N sulfuric acid, whilst the rate of rearrangement is practically the same as in 2 N sulfuric acid (Table 12). The acid catalysed rearrangements take place with almost no splitting. The SKIDI of the rearrangement of 5 and 2′, 3′, 3′-d3- 5 is 0.84±0.08 (2 N H2SO4, 67, 5°, cf. Table 13) and thus in accordance with a [3,3]-sigmatropic process which occurs in the corresponding anilinium ions. Consequently, the rearrangement of a 1:1 mixture of 2′, 3′, 3′-d3- 5 and 3, 5-d2- 5 in 2 N sulfuric acid at 67, 5° occurs without the formation of cross-products (Scheme 13). In the acid catalysed rearrangement of the N-1′-methylallyl) anilines 12-14 at 105-125° in 2 N sulfuric acid the corresponding (E)- and (Z)-anilines are the only products formed (Table 14 and 15). Again no splitting is observed. Furthermore, a dependence of the observed (E)/(Z) ratio and the bulkiness of the o-substituent ( H : (E)/(Z)- 30 = 6.5; t- C 4 H 9: (E)- 32 /(Z)- 32 = 90; cf. Table 15) indicates that also in the ammonium-Claisen rearrangement a chair-like transition state is preferentially adopted. In contrast to the thermal rearrangement the acid catalysed transformation in 2 N-O, 1 N sulfuric acid (150-170°) of (E)- and (Z)- 16 as well as of 1 and 15 , occurs very cleanly to yield the corresponding 2-allylated anilines 33, 2 and 34 (Scheme 15 and 18). The amounts of the anilines formed by splitting are <2%. During longer reaction periods hydration of the allyl chain of the products occurs, and in the case of the rearrangement of (E)- and )Z)- 16 the indoline 45 is formed (Scheme 15 and 18). All transformations occur with inversion of the allyl chain. This holds also for the rearrangement of 1 , since 3′, 3′-d2- 1 gives only 1′, 1′-d2- 2 (Scheme 17). The activation parameters were determined for the acid catalysed rearrangement of 1, 5, 12 and 15 in 2 N sulfuric acid (Table 22). The ΔH values of 27-30 kcal-mol and the ΔS values of +9 to -12 e.u. are in agreement with a [3, 3]-sigmatropic process in the corresponding anilinium ions. The acceleration factors (kH+/kΔ) calculated from the activation parameters of the acid catalysed and thermal rearrangements of the anilines are in the order of 105 - 107. They demonstrate that the essential driving force of the ammonium-Claisen rearrangement is the ‘delocalisation of the positive charge’ in the transition state of these rearrangements (cf. Table 23). Solvation effects in the anilinium ions, which can be influenced sterically, also seem to play a role. This is impressively demonstrated by N-(1′, 1′-dimethylallyl)-2, 6-dimethylaniline (11) : its rearrangement into 4-(1′, 1′-dimethylallyl)-2, 6-dimethylaniline (43) cannot be achieved thermally, but occurs readily at 30° in 2 N sulfuric acid. From a preparative standpoint the acid catalysed rearrangement in 2 N-0, 1 N sulfuric acid of N-allylanilines into 2-allylanilines, or if the o-positions are occupied into 4-allylanilines, is without doubt a useful synthetic method (cf. also [17]).  相似文献   

6.
Ruthenium(II) Phthalocyaninates(2–): Synthesis and Properties of (Acido)(carbonyl)phthalocyaninato(2–)ruthenate(II), [Ru(X)(CO)Pc2?]? (X = Cl, Br, I, NCO, NCS, N3) (nBu4N)[Ru(OH)2Pc2?] is reduced in acetone with carbonmonoxid to blue-violet [Ru(H2O)(CO)Pc2?], which yields in tetrahydrofurane with excess (nBu4N)X acido(carbonyl)phthalocyaninato(2–)ruthenate(II), [Ru(X)(CO)Pc2?]? (X = Cl, Br, I, NCO, NCS, N3) isolated as red-violet, diamagnetic (nBu4N) complex salt. The UV-Vis spectra are dominated by the typical π-π* transitions of the Pc2? ligand at approximately 15100 (B), 28300 (Q1) und 33500 cm?1 (Q2), only fairly dependent of the axial ligands. v(C? O) is observed at 1927 (X = I), 1930 (Cl, Br), 1936 (N3, NCO) 1948 cm?1 (NCS), v(C? N) at 2208 cm?1 (NCO), 2093 cm?1 (NCS) and v(N? N) at 2030 cm?1 only in the MIR spectrum. v(Ru? C) coincides in the FIR spectrum with a deformation vibration of the Pc ligand, but is detected in the resonance Raman(RR) spectrum at 516 (X = Cl), 512 (Br), 510 (N3), 504 (I), 499 (NCO), 498 cm?1 (NCS). v(Ru? X) is observed in the FIR spectrum at 257 (X = Cl), 191 (Br), 166 (I), 349 (N3), 336 (NCO) and 224 cm?1 (NCS). Only v(Ru? I) is RR-enhanced.  相似文献   

7.
The ligand exchange MX5·L + *L?MX5·*L + L for the octahedral adducts MX5·L, in an inert solvent (CH2Cl2 or CHCl3) with neutral ligands, proceeds via a dissociative D mechanism when M = Nb, X = Cl and L = phosphoryl compound. A dissociative interchange Id mechanism is suggested when M = Nb or Ta, and X = F. A first order rate law and positive values for ΔS* (+4 to +14 cal K?1 mol?1) are observed for the exchanges on the pentachloride adducts. However, a second order rate law and large negative values for ΔS* (-15 to -24 cal K?1 mol?1) are found for the intermolecular neutral ligand exchange (measured by 1H-NMR.) and for the intramolecular fluorine exchange (measured by 19F-NMR.) reactions on the pentafluoride adducts. The fluorine exchange is 2 to 5 times faster than the ligand exchange. The exchanges, on the pentachloride and on the pentafluoride adducts, are slowed down with increasing donor strength of the phosphoryl compound.  相似文献   

8.
The population of the conformations obtained by rotation around the C(2)? N and the N? C(O) bonds of AllNAc, GlcNAc, and GlcNMeAc derivatives was investigated by 1H-NMR spectroscopy. The AllNAc-derived α-D -and β-D -pyranosides 4–7 , the AllNAc diazirine 16 , and the GlcNAc-derived axial anomers α-D - 8–10 prefer the (Z)-anti-conformation. A significant population of the (Z)-syn-conformer in the (Z)-syn/(Z)-anti-equilibrium for the equatorial anomers β-D - 8–10 and the GlcNAc diazirine 17 was evidenced by an upfield shift of H? C(2), downfield shifts of H? C(1) and H? C(3), and by NOE measurements. The population of the (Z)-syn-conformation depends on the substituent at C(1) and is highest for the hexafluoroisopropyl glycoside. The population of the (Z)-syn-conformation of β-D - 14 decreases with increasing polarity of the solvent, but a substantial population is still observed for solutions in D2O. Whereas the α-D -anomers of the hemiacetal 22 and the methyl glycoside 21 prefer the (Z)-anti-conformation in D2O solution, the corresponding β-D -anomers are mixtures of the (Z)-anti-and (Z)-syn-conformers. The diazirine 17 self-associates in CD2Cl2 solution at concentrations above 0.005M at low temperatures. The axial anomers of the GlcNMeAc derivatives α-D - 26–28 are 2:1 to 3:1 mixtures of (Z)-anti-and (E)-anti-conformers, whereas the corresponding β-D -glycosides are ca. 1:3:6 mixtures of (Z)-syn-, (Z)-anti-, and (E)-anti-conformers.  相似文献   

9.
Replacement and Oxidation Reactions of N-Dichlorophosphanyl Triphenylphosphazene, Ph3P?N? PCl2 The title compound ( 1 ) reacts with MeOH, EtOH, PhOH, EtSH, and water forming N-phosphanyl or N-phosphinoyl phosphazenes, resp., Ph3P?N? PX2 (X ? OPh( 8 ), SEt( 9 )) or Ph3P?N? PH(O)X (X ? Cl( 3 ), OH( 4 ), OMe( 5 ), OEt( 7 )). The reaction of 1 with P(NEt2)3 yields Ph3P?N? P(NEt2)2 ( 10 ). Ph3P?N? PF2( 11 ) and Ph3P?N? PH(O)F ( 12 ) are obtained by chlorine-fluorine exchange. The N-phosphanyl compounds 1 , 8 , 9 and 11 are oxidized by NO2 yielding the corresponding N-phosphoryl derivatives, Ph3P?N? P(O)X2 (X ? Cl( 2 ), OPh( 13 ), SEt ( 14 ), F( 15 )). The thiophosphoryl compounds, (Ph3P?N? P(S)X2 (X ? Cl( 16 ), OPh( 17 ), F( 18 )) are obtained by oxidizing 1 , 8 , and 11 with sulfur.  相似文献   

10.
Six bis(silyl)acetylenes (XMe2Si? C?C? SiMe2X) with the following varied silicon substituents X were prepared: 1 (Me, Me); 2 (H, H); 3 (C1, H); 4 (CI, CI); 5 (MeO, H); 6 (MeO, MeO). While 1 and 2 may be prepared by the reaction of dilithio- or bis(bromomagnesium)-acetylide with the appropriate chlorosilane, similar reactions designed to give 3–6 yielded oligomers, XMe2Si? (? C?C? SiMe2)n? X, 7, X=CI, MeO, as the major products, indicating that the acetylenic functionality on silicon activates the chlorosilane towards nucleophilic substitution. Compounds 3 and 4 were prepared by free radical chlorination of 2. Methanolysis of 3 and 4 gave quantitative yields of 5 and 6 respectively. Compounds 1–6 undergo a Diels–Alder reaction with α-pyrone to produce, after loss of carbon dioxide, bis(silyl)-substituted benzene derivatives. The order of reactivity has been determined to be: 4=6>3=5>1>2, indicating that chloro or alkoxy substituents favor the cycloaddition with 2- pyrone. The adducts formed by compounds 3–6 undergo an unusually facile hydrolysis or elimination to give 1,1,3,3-tetramethyl-1,3-disila-2-oxaindane.  相似文献   

11.
Reaction of the acyclic (diamino)carbene (ADC) :C(NiPr2)2 (1) with different dihaloboranes of the type RBX2 (R = Mes, Dur; X = Cl, Br) smoothly afforded a novel class of ADC-stabilized borane adducts. For MesBBr2 however, the reaction did not stop at the adduct level, but an uncommon rearrangement process occurred, which eventually resulted in the formation of a 5-membered boracycle after elimination of mesitylene. Chemical reduction of the ADC borane adducts by KC8 selectively yielded air stable 1,2-azaboretidines. Detailed DFT studies suggest a reduction mechanism involving a highly reactive borylene intermediate, which is converted into the boracycles via a rearrangement/C–H activation sequence.  相似文献   

12.
When a mixture of (E)- and (Z)-1-propenylnaphth-2-yl-allylether ((E/Z)- 5 ) is heated to 182° only the (E)-isomer rearranges to give the ‘out-of-ring’ product (E/Z)- 16 , (Z)- 5 remains unchanged. At higher temperature (Z)- 5 yields 2-methyl-naphtho[2,1-b]furane ( 15 ) as the main product. The mixture of β-chloro-allyl derivatives (E/Z)- 6 behaves in a similar way. These findings led us to suspect that the ‘out-of-ring’ products 16 and 18 are formed by direct [1, 5s] allyl migration from the starting ethers (E)- 5 and (E)- 6 . Kinetic' measurements made on (E)- and (Z)- 5 and the independently synthesized (E)- and (Z)-1-allyl-1-propenyl-1 H-naphthalen-2-ones ((E)- and (Z)- 17 ) show however, that the ethers (E)- 5 and (E)- 6 undergo a double [3s, 3s] rearrangement (i.e. Claisen followed by Cope rearrangement) and hydrogen migration to yield the ‘out-of-ring’ products (E/Z)- 16 and (E/Z)- 18 (Scheme 9). In the (Z)-series steric factors prevent the intermediate naphthalenones (Z)- 17 and (Z)-19 from undergoing the Cope rearrangement and instead, at higher temperature, cleavage of the allyl group occurs (Scheme 11). The isopropenyl derivative 7 behaves in a similar way (Scheme 5). Rearrangement of (E/Z)-1-propenylnaphth-2-yl benzyl ether ( 8 ) requires a higher temperature (214°). The nature of the products obtained (Scheme 4) makes the occurrence of a direct sigmatropic [1,5s] shift of the benzyl group very unprobable. In the case of (E/Z)-2-propenylnaphth-1-yl allyl ether ( 10 ) both isomers rearrange to yield the ‘out-of-ring’ product 30 and the para-Claisen product 32 (Scheme 7). This experiment also provides evidence against a sigmatropic [1,5s] shift of the allyl group. The same conclusion can be drawn from the thermal behaviour of (E/Z)-2-propenylphenyl allyl ether (11) and 6-t-butyl-2-propenylphenyl allyl ether ( 12 ) where only 11 yields traces of the ‘out-of-ring’ product 35 (Scheme 8). Up to this date there is no evidence whatsoever for the existence of a sigmatropic [1,5s] migration of an allyl group from oxygen to carbon. Thermal rearrangement of (E/Z)-1-propenylnaphth-2-yl propargyl ether ( 9 ) yields only (E/Z)-1-propenyl-benz[e]indan-2-one ( 27 ) (and its secondary product 28 ). The mechanism for this reaction is given in Scheme 12. Treatment of a mixture of (E/Z)- 18 with base yields the (Z)-cyclisation product 2,4-dimethylnaphth[2,1-b]oxepine ( 43 ) (Scheme 13).  相似文献   

13.
On triplet excitation (E)- 2 isomerizes to (Z)- 2 and reacts by cleavage of the C(γ), O-bond to isomeric δ-ketoester compounds ( 3 and 4 ) and 2,5-dihydrofuran compounds ( 5 and 19 , s. Scheme 1). - On singulet excitation (E)- 2 gives mainly isomers formed by cleavage of the C(γ), C(δ)-bond ( 6–14 , s. Scheme 1). However, the products 3–5 of the triplet induced cleavage of the C(γ), O-bond are obtained in small amounts, too. The conversion of (E)- 2 to an intermediate ketonium-ylide b (s. Scheme 5) is proven by the isolation of its cyclization product 13 and of the acetals 16 and 17 , the products of solvent addition to b . - Excitation (λ = 254 nm) of the enol ether (E/Z)- 6 yields the isomeric α, β-unsaturated ε-ketoesters (E/Z)- 8 and 9 , which undergo photodeconjugation to give the isomeric γ, δ-unsaturated ε-ketoesters (E/Z)- 10 . - On treatment with BF3O(C2H5)2 (E)- 2 isomerizes by cleavage of the C(δ), O-bond to the γ-ketoester (E)- 20 (s. Scheme 2). Conversion of (Z)- 2 with FeCl3 gives the isomeric furan compound 21 exclusively.  相似文献   

14.
Reactions of Pt(PPh3)4 with the sulfines, XYCSO, (X, Y = aryl, S-aryl, S-alkyl, Cl) yield coordination compounds of the type Pt(PPh3)2(XYCSO). Infrared, 31P and 1H NMR spectra reveal that in all cases the sulfine ligand is coordinated side-on via the CS π-bond (Pt—η2-CS). Reactions of Pt(PPh3)4 with either the E- or Z-isomer of (p-CH3C6H4)(CH3S)CSO yields the corresponding E- or Z-coordination compound, Pt(PPh3)2[E-(p-CH3C6H4)(CH3S)CSO] or Pt(PPh3)2[Z-(p-CH3C6H4)(CH3S)CSO], indicating that the configuration of the sulfine ligand is retained upon coordination to the Pt(PPh3)2 unit. The compounds Pt(PPh3)2(XYCSO), containing reactive CX and/or CY bonds (X, Y = S-aryl, S-alkyl, Cl), undergo a rearrangement in solution to give complexes of the type PtX(PPh3)2(YCSO).  相似文献   

15.
The dimethyl acetals 4 of (E)-2,4-pentadienal and of (E,E)- and (E,Z)-2,4-hexadienals undergo regio- and stereospecific cycloaddition reactions with in-situ-generated acylnitroso dienophiles 5a and 5b , leading thereby to the corresponding dihydrooxazines 7a–d and 8c–d . cis-Glycolization of these latter adducts stereospecifically gave the dihydro derivatives 9a–d and 10d which, after sequential hydrogenolysis, deacetalization, and instant cyclization, led to the aminodeoxyribose derivatives 17a, 17f , and 18 , and to the amino-dideoxyallose compounds 17c and 17h . These piperidino-deoxysugar derivatives exhibit a strong anomeric effect, i.e. OH? C(1) is always axial, which is explained in terms of a nN(π)-σ*(C? OH) orbital compression, as compared to the less pronounced one in the more classical pyranose series.  相似文献   

16.
Preparation of Halogeno Pyridine Rhenates(III), [ReX6?n(Py)n](3?n)? (X = Br, Cl; n = 1?3) Crystal Structures of trans-[(C4H9)4N][ReBr4(Py)2], mer-[ReCl3(Py)3], and mer- [ReBr3(Py)3] The mixed halogeno-pyridine-rhenates(III), [ReX6?n(Py)n](3?n)? (X = Br, Cl), n = 1?3, have been prepared for the first time by reaction of the tetrabutylammoniumsalts (TBA)2[ReX6] (X = Br, Cl) in pyridine with (TBA)BH4 and separation by chromatography on Al2O3. Apart from the monopyridine complexes only the trans and mer isomers are formed from the bis-and tris-pyridine compounds. The X-ray structure determinations of the isotypic neutral complexes mer- [ReX3(Py)3] (monoclinic, space group P 21/n, Z = 4; for X = Cl: a = 9,1120(8), b = 12,5156(14), c = 15,6100(13) Å, β = 91,385(7)°; for X = Br: a = 9,152(5), b = 12,852(13), c = 15,669(2) Å, β = 90,43(2)°) reveal, due to the stronger trans influence of pyridine compared with Cl and Br, that the Re? X distances in asymmetric Py? Re? X3 axes with ReCl3 = 2,397 Å and ReBr3 = 2,534 Å are elongated by 1,3 and 1% in comparison with symmetric X1? Re? X2 axes with ReCl1 = ReCl2 = 2,367 Å and ReBr1 = 2,513 and ReBr2 = 2,506 Å, respectively. The Re? N bond lengths are roughly equal with 2,12 Å. Trans-(TBA)[ReBr4(Py)2] crystallizes triclinic, space group P1 , a = 9,2048(12), b = 12,0792(11), c = 15,525(2) Å, α = 95,239(10), β = 94,193(11), γ = 106,153(9)°, Z = 2. The unit cell contains two independent but very similar complex anions with approximate D2h(mmm) point symmetry.  相似文献   

17.
The reaction of of 4‐amino‐5‐ethyl‐2H‐1,2,4‐triazole‐3(4H)‐thione (AETT, L ) with furfural in methanol led to the corresponding Schiff‐Base ( L1 ). The reaction of L1 with [Cu(PPh3)2]Cl in methanol gave to the neutral compound [( L1 )Cu(PPh3)2Cl] ( 1 ). By recrystallization of 1 from CH3CN the complex [( L1 )Cu(PPh3)2Cl]·CH3CN ( 1a ) was obtained. All compounds were characterized by infrared spectroscopy, elemental analyses as well as by X‐ray diffraction studies. Crystal data for L1 at ?80 °C: space group with a = 788.4(1), b = 830.3(2), c = 928.8(2) pm, α = 84.53(1)°, β = 65.93(1)°, γ = 72.02(1)°, Z = 2, R1 = 0.0323; for 1 at ?100 °C: space group with a = 1166.3(1), b = 1423.8(2), c = 1489.1(2) pm, α = 62.15(1)°, β = 72.04(1)°, γ = 88.82(1)°, Z = 2, R1 = 0.0338 and for 1a at ?100 °C: space group P21/c with a = 1294.1(1), b = 1019.8(2), c = 3316.9(4) pm, β = 94.73(1)°, Z = 4, R1 = 0.0435.  相似文献   

18.
Syntheses of the title compounds, viz. N(CH2CH2O)3GeY ( 2 Y?Fluorenyl; 4 Y?PhC?C) by the reaction of X3GeY ( 1 Y?Fluorenyl, X?Br; 5 Y?PhC?C, X?Cl) with N(CH2CH2OSnR3)3 ( 3 R?Et; 6 R?Bu) are reported including the preparation of the new compound 1 . Identity and structures were established by elemental analyses, 1H and 13C NMR spectroscopy. 2 and 4 were characterized by mass spectrometry. Single crystal structures of 1 , 2 and 4 were determined by X-ray diffraction methods.  相似文献   

19.
Synthesis and characterisation of (Z)-1-[2(triphenylstannyl)vinyl]-l-cyclododecanol, c-(CH2)11C(OH)CH=CHSnPh3, are reported, together with its halogenation by I2, Br2 and ICI to yield derivatives of the types c-(CH2)11C(OH)CH=CHSnPhs?nXn (n=1, 2; X=I, Br, Cl, respectively). The molecular structures of two compounds have been determined by X-ray diffraction analysis. The tin atom exhibits a distorted tetrahedral geometry in the crystal of (Z)-l-[2–(triphenylstannyI)vinyl]-l-cyclododecanol, but a trigonal bipyramidal geometry in the monoiodo-derivative (Z)-l-[2]-(diphenyliodo-stannyl)vinyl)-1-cyclododecanol and other derivatives, in which significant intramolecular coordinative interaction HO→Sn ia observed. And the formation of a five-membered tin containing ring is significant for their antitumour activities.  相似文献   

20.
Chlorination of a mixture of C86 isomers no. 16 (Cs) and no. 17 (C2) with VCl4 or a (TiCl4+Br2) mixture afforded crystalline chlorides with 16 to 22 Cl atoms per fullerene cage. Single crystal X‐ray diffraction with the use of synchrotron radiation enabled us to determine the chlorination patterns of C86(16)Cl16, C86(17)Cl18, C86(17)Cl20, and C86(17)Cl22. At these degrees of chlorination, addition patterns of C86(16) and C86(17) chlorides have some features in common, owing to the close similarity in the cage structures of both isomers. The average energy of C?Cl bonds decreases with increasing number of attached Cl atoms.  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