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 15 毫秒
1.
2.
In Z? CH? CH?CH? Y compounds (Z or Y being an alkyl group) the ethylenic part of the spectra is often very complex and the 3J(H? C?C? H) coupling constant which is a good tool for determining the configuration, is not easily determined. We have studied such allylic derivatives and many configurations have been assigned through stereospecific synthesis. Except a very few cases, δ CH(Z) of the cis isomer is larger than δ CHZ of the trans isomer. In alcohols RCH?CH? CHOHR′ the stereoisomers behave differently in solutions with europium, praseodymium, holmium and dysprosium complexes. The spectra of the trans isomers remain strongly coupled but 3J(H? C?C? H) becomes easy to measure in the cis compounds.  相似文献   

3.
1H NMR spectra of several aliphatic and phenyl substituted alkenediynes have been obtained. Chemical shifts and coupling constants of these compounds are discussed in conjunction with some compounds described in the literature. Chemical shifts of the protons from isomeric alkenediynes R? C?C? C?C? CH?CH2, R? CH?CH? C?C? C?CH and R? CH?CH? C?C? C?C? CH3 (R = H, alkyl, C6H5, C6H4OCH3-p) are well correlated with cis/trans-isomerism and electronic effects of substituents at the C?C bond. The coupling constants were found to be only slightly dependent on the substitution at the double bond. We could resolve couplings over a maximum of eight bonds in the alkenediyne system.  相似文献   

4.
Summary The preparation of the series ofcis- andtrans-[Co(NH3)4(RNH2)Cl]2+ complexes (withcis, R = Me orn-Pr andtrans, R = Me, Et,n-Pr,n-Bu ori-Bu) is described. The u.v-visible spectra indicate a decrease of the ligand field on increasing chain length. Infrared spectra show an enhanced Co-Cl bond strength compared to the pentaammine. Partial molar volumes of the complex cations do not reveal steric compression. From proton exchange studies in D2O it follows that [Co(NH3)5Cl]2+ and thecis- andtrans-[Co(NH3)4-(CH3NH2)C1]2+ complexes exchange the amine protons on the grouptrans to the chloro faster than those on thecis. A coordinated methylamine group exchanges its amine protons slower than a corresponding NH3 group in the parent pentaammine, but the methyl introduction accelerates the exchange of the other NH3 groups. The aquation of thetrans-alkylamine complexes (studied at 52° C) is acceleratedca. 10 times compared to the parent pentaammine, irrespective of the nature of the alkyl group. Thecis complexes do not show this acceleration of aquation. In base hydrolysis (studied at 25° C) thecis complexes are the most reactive (a factor 20 over the parent ion). Thecis/trans product ratio in base hydrolysis and the competition ratio in the presence of azide ions were calculated from the 500 MHz1H n.m.r. spectra, which display distinctly different alkyl resonances for each individual complex. Thecis ions react under stereochemical retention of configuration; thetrans compounds give 10±1%trans tocis rearrangement. The ionic strength (4 mol dm–3) and the pH do not affect this result. The same product ratio is obtained in methanol-water and DMSO-water mixtures. Ammoniation in liquid ammonia gives the same ratios as in base hydrolysis, base-catalyzed solvolysis in neat methylamine gives stereochemical retention for both thecis- andtrans-methylamine ion. The product competition ratio (Co-N3)/(Co-OH2) for thecis compounds and the bulkier amines (R =n- andi-Bu), 15–25% at 1 mol dm–3N 3 , isca. twice that of thetrans compounds and the pentaammine. The results are interpreted in the classical conjugate base mechanism, and discussed in the context of current ideas about stereochemistry of base hydrolysis.Prof. C. R. Píriz Mac-Coll from Uruguay is a guest at the Free University of Amsterdam.  相似文献   

5.
In the 1H-NMR spectrum of polychloroprene dissolved in C6D6, the ?CH proton signal was separated into two triplet peaks. These triplet signals were assigned to the ?CH proton in the trans-1,4 and cis-1,4 isomers by measurement of 1H-NMR spectra of 3-chloro-1-butene and a mixture of trans- and cis-2-chloro-2-butene as model compounds for the 1,2, trans-1,4 and cis-1,4 isomers. In 1H-NMR spectra (220 Mcps) of polychloroprene dissolved in C6D6, two triplet signals were separated completely from which the relative concentrations of trans-1,4 and cis-1,4 isomers could be obtained quantitatively.  相似文献   

6.
Enamino -thial and -thiones R1C(S)CH?CHNR (R1 = H or alkyl; R2 = Me or Et) have been shown by NMR spectra to exist in two rotational forms, s-cis and s-trans, the populations of the latter being approximately the same as in the case of the parent oxa analogues. An increase of the order of 2 to 4 Kcal/mole in the heights of C? C and C? N rotation barriers (ΔG*) was found on comparing the title compounds with their oxa analogues. IR spectra failed as a tool to establish the rotational equilibrium. IR absorption bands of the νC? C, νC? H (in the NMe2 group) and γHC?CH vibrations have been found, but the νC?S band could not be assigned unambiguously. Anomalies concerning the frequency and intensity of the νC?C band are discussed.  相似文献   

7.
We have applied various theoretical methods to gain detailed insights into the isomers as well as the transition states (TSs) along the corresponding reaction pathways for RSNO (R=H, C n H2n+1 n ≤ 4). On the basis of G2 and G2MP2 results, the relative order of stability for R=H is estimated to be trans-HSNO > cis-HSNO > HNSO > cis-HONS trans-HONS, while it is cis-CH3SNO trans-CH3SNO > CH3NSO > trans-CH3ONS > cis-CH3ONS for R=CH3. A similar trend is also obtained from the B3P86 method with considerably less computing effort if the nearly isoenergetic isomers cis-HONS and trans-HONS are ignored. Based on the results of B3P86, cis-RSNO is more stable than trans-RSNO when R=H is replaced by alkyl groups except for R=t-Bu. Natural bond orbital analyses allow us to explore whether the high reactivity of S-nitrosothiols is due to the strong negative hyperconjugation (). The mesomeric effect of S-nitrosothiols, although non-negligible, does not cause the breakage of N–O bond due to the compensation of columbic attraction between N and O.Electronic Supplementary Material Supplementary material is available for this article at and is accessible for authorized users.  相似文献   

8.
The deuterioformylation of (Z)- or (E)-2-butene catalyzed by [DIOP]Pt(SnCl3)-Cl
  • 1 DIOP=2,3-O-isopropylidene-2,3-dihydroxy-1,4-bis(diphenylphosphino)butane.
  • gives predominantly erythro- or threo-1,3-[2H]2-2-methylbutanal respectively. Hence, hydroformylation by this catalytic system must take place with cis-stereochemistry.  相似文献   

    9.
    Methods for the calculation of the torsional angle of the C? C linkage
  • 1 Voir Réf. 1.
  • are applied to some 4-germa-1,3-dioxanes. It is thereby shown that 4,4-diethyl-2trichloromethyl-4-germa-1,3-dioxane in CCl 4 and C 6 D 6 adopts the chair conformation, with the equatorial C? Cl 3 group distorted by the presence of the germanium: the torsional angle of the Ge-CH 2 -CH 2 -O-fragment is 45°. The most stable 6-alkyl derivatives ( cis isomers) have the same conformation; the less stable trans - tert -butyl isomer prefers a skew-boat form compatible with a torsional angle of about 60°.  相似文献   

    10.
    A study of the α-cleavage processes of aliphatic amines (using RLCH2N(R)CH2Rs and RLCH(NR)Rs
  • 1 RL represents the larger Rs the smaller alkyl group throughout this paper and both, unless indicated to the contrary, have similar degrees of lsubstitution on the α-carbon atom.
  • as typical substrates) at several ionizing voltages indicates that the loss ratio of large and small alkyl radicals, [M ? RL]/[M ? Rs], decreases with ionizing voltage. This ratio, greater than 1·0 at 70 eV, may become less than unity at low voltage (15 eV, 10 eV) in some cases. Comparison of similarly structured compounds with nitrogen or oxygen as the heteroatom suggests that the effects of ionizing voltage are lessened when the fragment is of greater stability (e.g. amine vs. ether or imine vs. ketone). The effects of lowering the ionizing voltage became much less pronounced as the alkyl groups become larger or similar in size.  相似文献   

    11.
    Studies on the Electronic Influence of Organoligands. XIII. Synthesis and Characterization of 2-Functionalized Vinyl Rhodoximes 2-Functionalized vinyl rhodoximes [Rh(dmgH)2 (PPh3)cis/trans-CH = CHZ] ([Rh]? CH = CHZ) ) ( 1 ) can be prepared with a wide variation of the substituent Z (cis: OEt ( 1 a ), OPh ( 1 b ), Cl ( 1 c ), Me ( 1 j ), Ph ( 1 k ), SMe ( 1 l ), SPh ( 1 m ); trans: SPh ( 1 d ), Me ( 1 e ), Ph ( 1 f ), CMe3 ( 1 g ), SiMe3 ( 1 h )) by oxidative addition of XCH = CHZ and/or by nucleophilic addition of HC?CZ and Me3SiC?CZ, respectively, to [Rh]?. 1 a is converted to [Rh]? CH2CHO ( 2 ) already in a weakly acid medium. 1 l is isomerized to trans-[Rh]? CH = CHSMe ( 1 n ) in the presence of acids. The complexes 1 are characterized by microanalysis and by 1H, 13C and 31P NMR spectroscopy. The magnitude of the coupling constants 1J(103Rh, 31P) reveals only a small effect of Z on the (NMR) trans influence of the vinyl ligands CH = CHZ. The molecular structures of cis-[Rh]? CH = CHSPh ( 1 m ) and trans-[Rh]? CH = CHSPh ( 1 d ) show a distorted octahedral coordination of Rh with a mutual trans position of triphenyl-phosphine and the 2-phenylmercaptovinyl ligands. Van der Waals interactions exist between the sulfur and the equatorial dimethylglyoximato ligands in the cis complex 1 m .  相似文献   

    12.
    The cationic polymerization of cis- and trans-ethyl propenyl ethers (EPE, CH3? CH?CH? O? C2H5), initiated by a mixture of hydrogen iodide and iodine (HI/I2 initiator) at ?40°C in nonpolar media (toluene and n-hexane), led to living polymers of controlled molecular weights and a narrow molecular weight distribution (MWD) (M?w/M?n = 1.2–1.3). The geometrical isomerism of the monomer did not affect the living character of the polymerization. 13C NMR stereochemical analysis of the polymers showed that the living propagating end is sterically less crowded than nonliving counterparts generated by conventional Lewis acids (e.g., BF3OEt2). New block copolymers between EPE (cis or trans) and isobutyl vinyl ether were also prepared by sequential living polymerization of the two monomers.  相似文献   

    13.
    A series of ruthenium hydride compounds containing substituted bidentate pyrrole‐imine ligands were synthesized and characterized. Reacting RuHCl(CO)(PPh3)3 with one equivalent of [C4H3NH(2‐CH=NR)] in ethanol in the presence of KOH gave compounds {RuH(CO)(PPh3)2[C4H3N(2‐CH=NR)]} where trans‐Py‐Ru‐H 1, R = CH2CH2C6H9; cis‐Py‐Ru‐H 2, R = Ph‐2‐Me; and cis‐Py‐Ru‐H 3, R = C6H11. Heating trans‐Py‐Ru‐H 1 in toluene at 70°C for 12 hr resulted a thermal conversion of the trans‐Py‐Ru‐H 1 into its cis form, {RuH(CO)(PPh3)2[C4H3N(2‐CH=NCH2CH2C6H9)]} (cis‐Py‐Ru‐H 1) in very high yield. The 1H NMR spectra of trans‐Py‐Ru‐H 1, cis‐Py‐Ru‐H 2, cis‐Py‐Ru‐H 3, and cis‐Py‐Ru‐H 1 all show a typical triplet at ca. δ–11 for the hydride. The trans and cis form indicate the relative positions of pyrrole ring and hydride. The geometries of trans‐Py‐Ru‐H 1, cis‐Py‐Ru‐H 1, and cis‐Py‐Ru‐H 3 are relatively similar showing typical octahedral geometries with two PPh3 fragments arranged in trans positions.  相似文献   

    14.
    Intramolecular Diels–Alder (IMDA) transition structures (TSs) and energies have been computed at the B3LYP/6‐31+G(d) and CBS‐QB3 levels of theory for a series of 1,3,8‐nonatrienes, H2C?CH? CH?CH? CH2? X? Z? CH?CH2 [? X? Z? =? CH2? CH2? ( 1 ); ? O? C(?O)? ( 2 ); ? CH2? C(?O)? ( 3 ); ? O? CH2? ( 4 ); ? NH? C(?O)? ( 5 ); ? S? C(?O)? ( 6 ); ? O? C(?S)? ( 7 ); ? NH? C(?S)? ( 8 ); ? S? C(?S)? ( 9 )]. For each system studied ( 1 – 9 ), cis‐ and trans‐TS isomers, corresponding, respectively, to endo‐ and exo‐positioning of the ? C? X? Z? tether with respect to the diene, have been located and their relative energies (ErelTS) employed to predict the cis/trans IMDA product ratio. Although the ErelTS values are modest (typically <3 kJ mol?1), they follow a clear and systematic trend. Specifically, as the electronegativity of the tether group X is reduced (X?O→NH or S), the IMDA cis stereoselectivity diminishes. The predicted stereochemical reaction preferences are explained in terms of two opposing effects operating in the cis‐TS, namely (1) unfavorable torsional (eclipsing) strain about the C4? C5 bond, that is caused by the ? C? X? C(?Y)? group’s strong tendency to maintain local planarity; and (2) attractive electrostatic and secondary orbital interactions between the endo‐(thio)carbonyl group, C?Y, and the diene. The former interaction predominates when X is weakly electronegative (X?N, S), while the latter is dominant when X is more strongly electronegative (X?O), or a methylene group (X?CH2) which increases tether flexibility. These predictions hold up to experimental scrutiny, with synthetic IMDA reactions of 1 , 2 , 3 , and 4 (published work) and 5 , 6 , and 8 (this work) delivering ratios close to those calculated. The reactions of thiolacrylate 5 and thioamide 8 represent the first examples of IMDA reactions with tethers of these types. Our results point to strategies for designing tethers, which lead to improved cis/trans‐selectivities in IMDAs that are normally only weakly selective. Experimental verification of the validity of this claim comes in the form of fumaramide 14 , which undergoes a more trans‐selective IMDA reaction than the corresponding ester tethered precursor 13 .  相似文献   

    15.
    Abstract

    The purpose of this study was to synthesize trans-l and determine the equilibriurr constant with cis-1. Oniy the synthesis1 and x-ray structure2 of the cis isomer have bcen reported. Four prior synthetic routes to make the vans isomer3 gave only cis product. For example, intrarmolecular ring closure of the cis or trains isomers of 4 (R= (CH2)3OH) with LiH or thermal closure of the cis or trans 4 (R= (CH2)2) gave only cis-1. Since both iosmers of 1,8-dioxabicyclo[4.4.0] decane are known and readily equilibrate (57% cis and 43% trans), the apparent inaccessibility of trans-1 attracted our attention. Thc preparation of trans-1 was achieved by treatment of cis-1 with Lawesson's reagent (LR) to provide cis-2. followed by oxidation with m-chloroperbenzoic acid/trifluoroacetic acid to give a 5:1 mixture of cis:trans 1, respectively. An unexpected formation of the sulfur analogue of 1 was observed on treatment of cis-1 with P2S5/pyridine at reflux temperatures to give a 1.6:1 mixture of cis:trans 3, respectively. Thermal quilibration of 1 at 204°C provided an equilibrium ratio of 99.5% cis and 0.5% of the trans isomer. However, equilibration of 3 at 250°C led to 82.2:17.8 ratio in favor of the cis isomr. These results are consistent with semiemperical MO calculations. The stereochemical outcome on treatment of 4 with LR was also investigated. X-ray structures for six compounds: trans-1, cis-2, cis and trans-3; cis-4 (R=Ph), and cis-5, (R = Ph) wen determined.  相似文献   

    16.
    On the Synthesis of 1-Aryl- and-1-Alkyl-2, 3-dimethyl-quinoxalinium Perchlorates. 2nd Communication
  • 1 1. Mitt.: [1].
  • . Synthesis and 1 H-NMR. Spectra of 2, 3-Dimethyl-1-phenyl-6-X-quinoxlinium Perchlorates The synthesis of the title compounds ( 5 ) which have been useful as precursors for a lot of conventional and new-type dyes [2-8] has yet been limited to examples with X?H [2] [3] [11] [15] and with electron-donating [4] [12] or at best slightly electron-accepting [1] [6] substituents X and R. We now describe a method suitable even for compounds 5 with strongly electron-accepting substituents: N-monosubstituted o-phenylendiamines 4 , were condensed with 2, 3-butanedione and perchloric acid in mixed solvents containing an excess of diethyl ether. The products - mostly substituted at position 6 of the quinoxalinium ring - are chracterized by 1H-NMR. spectra, elemental analyses and in most cases by isolation of the corresponding bases 6 . Correlations of chemical shifts with Hammett's σp [18] are given by equations (1)-(5).  相似文献   

    17.
    Two iridium(I) complexes, [IrCl(COD)(PEt3) n ], n = 1 or 2, have been prepared and structurally characterised. Although [IrCl(COD)(PEt3)] is a known compound the spectroscopic data on both compounds is presented and discussed. In addition, the X-ray crystal structure of the previously described orthometallated isomer of Vaska's compound, [IrHCl(CO)(PPh3) 2-PPh2(C6H4)], is reported to show the hydride ligand trans- to the carbonyl ligand.  相似文献   

    18.
    Contributions to the Chemistry of Phosphorus. 117. Synthesis and Properties of the Hexaorganyl-octaphosphanes(6) P8R6, R = Me, Et, Pri The Hexaorganyl-octaphosphanes(6) P8Me6 ( 1 ), P8Et6 ( 2 ), and P8Pr ( 3 ) have been obtained by reacting mixtures of the corresponding organyldichlorophosphanes and phosphorus(III) chloride with magnesium. In the case of 1 and 2 the organyl-cyclophosphanes (PR)n can also be used in the reaction with phosphorus(III) chloride and magnesium. Besides, mainly P7R5 as well as other polycyclic organylphosphanes are formed. 2 and 3 have been isolated as pure substances, whereas 1 was concentrated to ?50 mol-% in the product mixture. According to their 31P-NMR spectra the three compounds possess a pentalane-analogous P8-skeleton with the substituents within each five-membered ring in trans position and the substituents of different five-membered rings next to the zero bridge in cis position; the organyl groups in the 3, 7 position are trans oriented with respect to the free electron pairs of the bridgehead atoms. Therefore, the structures of 1 – 3 differ from the known tert-butyl compound P8Bu, whereas the corresponding phosphorus hydride P8H6 has the same pentalane-analogous P8-skeleton, thus being a bicyclo[3.3.0]octaphosphane.  相似文献   

    19.
    Alkene elimination, alkyl loss, styryl and alkyl migrations to oxygen are the major fragmentations of (Z)- and (E)-alkane-sulfinyl-2-phenylethenes (cis- and trans-styryl alkyl sulfoxides) under electron impact. Alkene elimination has been found to go through a McLafferty-type hydrogen rearrangement to yield the benzylic carbon and the sulfinyl oxygen, respectively. This is the most facile and interesting process for compounds with longer alkyl chains (R?C3H7), and is a unique feature for styryl sulfoxides. Product ion analyses show the migrating aptitude to carbon and to oxygen to be qualitatively similar.  相似文献   

    20.
    The salt elimination reaction of the transition carbonyl metal-lates [L(CO)nM](Na/K) (M = Cr, Mo, W, Mn, Re, Fe, Co, Ni; L= CO, n5-C5R5, PR3; n= 1-4; R= alkyl, aryl) with the base-stabilized galliumhalides ClaGaR3 -a(Do) (R = H, alkyl, halide; Do = THF, N(CH3)3, NC7H13) or ClaGa[(CH2)3N-R2](R)2 - a yielded almost quantitatively the transition metal-substituted, gallanes [L(CO)nM]aGaR3 - a(Do) and [L(CO)n-M]aGa[(CH2)3NR2](R)2 - a, respectively. Residual halide functionalities in these complexes were selectively replaced by various other groups. The new compounds were characterized by means of elemental analysis, 1H-, 13C-, 31P-NMR, MS, and lR v(CO) data. The single-crystal X-ray structure analysis of trans-(Ph3P)(CO)3Co-Ga[(CH2) 3N(C2H5)2](R)( 6s : R = Cl, 6t : R= CH3) showed s̀(Co-Ga) lengths of 237.78(4) and 249.5(1) pm, respectively. A short s̀(Fe-Ga) contact of 236.18(3) pm was found for (n5-C5H5)(CO)2Fe-Ga-Cl2[N(CH 3)3] ( 5a ). Low-pressure MOCVD experiments were performed to give thin films of analytically pure CoGa alloy.  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