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 109 毫秒
1.
Reaction of 2, 4, 6‐tri‐tert‐butylphenol ( 1 ) with di‐n‐butylmagnesium in the molar ratio 1:1 allows the synthesis of {(nBu)Mg(μ‐OR)2Mg(nBu)} ( 2 ) (R = 2, 4, 6‐tBu3C6H2), which reacts with excess 1 to give the homoleptic alcoholate complex {(RO)Mg(μ‐OR)2Mg(OR)} ( 3 ) (R = 2, 4, 6‐tBu3C6H2). The structures of 2 and 3 were determined by X‐ray crystallography.  相似文献   

2.
Aqueous‐phase dissociation constants (Ka) for the conjugate acids of a series of 2‐azidoethanamine bases: R1N(R2)CH2CH2N3 ( 1 , R1 = CH3, R2 = H; 2 , R1 = CH3, R2 = CH3; 3 , R1 = CH2CH3, R2 = CH2CH3; 4 , R1/R2 =  CH2CH2CH2CH2 ; 5 , R1/R2 =  CH2CH2OCH2CH2 ; 6 , R1 = CH2CH3, R2 = CH2CH2N3) were measured and found to fall between those for analogous unfunctionalized and cyano‐functionalized ethanamines. To explore the possibility of a relationship existing between the constants and molecular geometry, a theoretically based study was conducted. In it, the Gibbs free energies of aqueous‐phase (equilibrium) conformers of the bases and their conjugate acids were determined via a density functional theory/polarizable continuum model method. The results indicate that an attractive interaction between the amine and azide groups that underlies the lowest‐energy gas‐phase conformer of 2 is negated in an aqueous environment by solvent–solute interactions. The magnitudes of the free energy changes of solvation and −TS (entropic) energies of the conformers of the 2‐azidoethanamines and their conjugate acids are observed to correlate with the magnitude of the separation between the conformers' amine and azide groups. However, those correlations are not by themselves sufficient to predict the relative free energies of a molecule's conformers in an aqueous environment. That insufficiency is due to the influence of the correlations being mitigated by three other parameters that arise within the thermodynamic framework employed to compute the observable. The nature of those parameters is discussed. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

3.
A `missing' member of the inverse crown ether family, namely μ4‐oxo‐tetrakis(μ‐2,2,6,6‐tetra­methyl­piperidinido)­di­mag­nes­ium­(II)­disodium(I), [Na2Mg2O(C9H18N)4], has been synthesized by blocking the alternative aromatic metallation route via the use of sterically hindered 1,3,5‐mesityl­ene as a solvent. [Na2Mg2O(NR2)4] (NR2 is 2,2,6,6‐tetra­methyl­piperidinide) is shown to form a cationic planar eight‐membered ring with alternating metal and N atoms, which captures at its core an oxide guest that lies on an inversion centre [principal dimensions: Na—O = 2.2405 (11) Å, Na—N = 2.445 (3) and 2.572 (3) Å, Mg—O = 1.8673 (9) Å, and Mg—N = 2.032 (2) and 2.063 (2) Å].  相似文献   

4.
The reaction of 6-bromo-1-hexene with Mg was studied in order to obtain information on the role of the solvent during the formation of Grignard reagents. The 5-hexenyl radical (Rnc.) is known to cyclize rapidly and irreversibly to the cyclopentylmethyl radical (Rc.). Changes in yields of the cyclized and non-cyclized Grignard compounds have been found on varying the solvent. Information on the radical pairs involved is obtained from the yields of the three possible coupling products (RncRnc, RncRc and RcRc). Results are correlated to the intensity of the CIDNP spectra of the Grignard compounds. It is found that basicity and viscosity of the solvents influence the reactions at the site of single electron transfer. Formation of Grignard compounds via radical pairs increases: (a) with decreasing basicity of the solvent, (b) with decreasing viscosity of the solvent, and (c) on dilution of THF with benzene. It is proposed that interaction between the radical and the π-electron rich solvent benzene plays a role in reactions run in Bz/THF mixtures.  相似文献   

5.
The title compounds were synthesized by reacting the elements in sealed tantalum tubes in a high‐frequency furnace. They crystallize with the Mo2FeB2 structure, a ternary ordered variant of the U3Si2 type, space group P4/mbm. All compounds were characterized through Guinier powder patterns and the lattice parameters were obtained from least‐squares fits. Four structures were refined from single crystal X‐ray data: a = 740.5(1) pm, c = 372.5(1) pm, wR2 = 0.0430, 247 F values, 13 variables for Y2Ni1.90Mg, a = 764.5(1) pm, c = 394.39(9) pm, wR2 = 0.0371, 310 F values, 12 variables for La2Ni2Mg, a = 754.4(1) pm, c = 385.20(9) pm, wR2 = 0.0460, 295 F values, 12 variables for Pr2Ni2Mg, and a = 752.53(8) pm, c = 382.33(5) pm, wR2 = 0.0183, 291 F values, and 12 variables for Nd2Ni2Mg. A refinement of the occupancy parameters indicated small defects on the nickel site of the yttrium compound, resulting in the composition Y2Ni1.90Mg for the investigated single crystal. The compounds with cerium, samarium, and gadolinium to thulium as rare earth component were characterized through their Guinier powder patterns. The cell colume of Ce2Ni2Mg is smaller than that of Pr2Ni2Mg, indicating intermediate‐valent cerium. The structures can be considered as an intergrowth of distored AlB2 and CsCl related slabs of compositions LnNi2 and LnMg. Chemical bonding in La2Ni2Mg and isotypic La2Ni2In is compared on the basis of extended Hückel calculations.  相似文献   

6.
A remarkable solvent‐controlled enantiodivergence is seen in the hydroquinidine 1,4‐phthalazinediyl diether ((DHQD)2PHAL)‐catalyzed chlorocyclization of unsaturated carbamates. Eyring plot analyses of this previously unreported reaction are used to probe and compare the R‐ and S‐selective pathways. In the CHCl3/hexanes solvent system, the pro‐R process shows a surprising increase in selectivity with increasing temperature. These studies point to a strongly solvent‐dependent entropy–enthalpy balance between the pro‐R and pro‐S pathways.  相似文献   

7.
The structure of the title compound, {[Mg(C4H7O2)2(H2O)3]·H2O}n, features one‐dimensional ...(μ2‐ib)Mg(μ2‐ib)Mg... zigzag chains (ib is isobutyrate) parallel to the c axis. The octahedral Mg environment is completed by three fac‐oriented terminal water ligands, as well as one further monodentate end‐on coordinated ib ligand. In the crystal structure, the hydrophobic ib groups are all oriented within one half of the coordination perimeter of each chain, whereas the water ligands, together with hydrogen‐bonded noncoordinated solvent water molecules, define the other half. Along the a axis, neighbouring strands are oriented so that both the hydrophilic and hydrophobic sides are adjacent to each other. This results in an extensive hydrogen‐bonding network within the hydrophilic areas, also involving an additional solvent water molecule per formula unit. There are van der Waals contacts between the aliphatic isopropyl groups of the hydrophobic areas.  相似文献   

8.
The standard enthalpy and entropy changes for the stepwise and overall complex formation homogeneous equilibria between Ag(I) and thiocarbonyl ligands or for the single reactions (series of reactions) in H2O, CH3OH (MeOH), C2H5OH (EtOH), C3H7OH (n-PrOH), CH3CHOHCH3 (Is-PrOH), and CH3COCH3 (Ac) are linearly correlated (corr. coeff. R > 0.9). The intercepts of the ?H????S regression straight lines ?H?=??Hres???τ ?S depend on both solvent and number of coordinated ligands (hereafter coordination level) in the coordination sphere of the substrate AgLn. The slopes τ?=?d?H/d?S (hereafter susceptivities/sensitivities) depend only on solvent. Robust and sound hard solvent parameter scales, founded on solute–solute soft–soft interactions, can be set up on ?Hres and/or τ bases. At 25 °C, the transfer standard free energies for the overall complex formation reactions at the single coordination levels n = 1, 2, or 3 between the various solvents are uncorrelated. When considered as unique set with n = 1, 2, and 3, the results linearly correlated (0.85?≤?R?≤?0.97) with slope?≈?1 and intercepts depending on the actual couple of solvent media. The enthalpy(entropy) changes are uncorrelated.  相似文献   

9.
The dilute solution properties of linear, 18-arm, and 270-arm star polybutadienes have been studied in a theta solvent and in a good solvent. Values of the radius of gyration RG, the second virial coefficient A2, the intrinsic viscosity [η], and the diffusion coefficient D0 have been measured for each polymer. The ratios RT/RG, RV/RG, and RH/RG for each type of polymer are used to compare the four dilute solution properties. RT is termed the “thermodynamic radius.” It is the radius of the hard sphere with the same excluded volume as the polymer coil. RT is calculated from A2 by RT = (3A2M2/16ηNA)1/3. RV and RH are equivalent hard spheres defined for the intrinsic viscosity and translational diffusion coefficient, respectively. RT/RG, RV/RG, and RH/RG increase from about 0.7 for linear polymer coils as the number of arms in the star increases. Values of the ratios for the 18-arm stars are less than the value for the hard-sphere, but the values of the ratios of the 270-arm stars are equal to the hard-sphere limit within experimental error.  相似文献   

10.
Magnesium dicyanamide tetrahydrate Mg[N(CN)2]2 · 4 H2O was synthesized by aqueous ion exchange starting from Na[N(CN)2] and Mg(NO3)2 · 6 H2O. The crystal structure was solved and refined on the basis of powder X‐ray diffraction data (P21/c, Z = 2, a = 737.50(2), b = 732.17(1), c = 971.67(2) pm, β = 98.074(1)°, wRp = 0.059, Rp = 0.046, RF = 0.075). In the crystal there are neutral complexes [Mg[N(CN)2]2(H2O)4] which are only connected via hydrogen bonds. Above 40 °C the tetrahydrate decomposes into anhydrous Mg[N(CN)2]2.  相似文献   

11.
In this work we study the geometric properties of the triblock copolymer micelles with solubilized low-mass molecules in a selective solvent using the Monte Carlo technique on a cubic lattice. The triblock copolymers are of the ABA type, with the two insoluble blocks at the ends. The size of the micelles is characterized by the squared radius of gyration of the micellar core, Rg2, while the shape is treated by the asphericity b and the acylindricity c, which are defined in terms of the principal moments of the radius of gyration tensor. The parameters varied are the amount of solubilizate molecules, the polymer concentration, the interaction parameters between A and B, A and solvent, solute and solvent, solute and B block, and the A and B block length. The micelle size, characterized by Rg2, grows with increasing concentration of the solubilizates and/or the polymer, and stronger interactions between the incompatible species. The A block length is found not to modify Rg2 monotonously, while an increase in B block length results in a decrease in Rg2 at high concentrations. As the size expands, the micellar shape becomes less spherical but retains its cylindricity. In addition to an increase in the averaged Rg2, the distribution of Rg2 becomes broader and the system less homogeneous.  相似文献   

12.
Two alkaline earth–tetrazole compounds, namely catena‐poly[[[triaquamagnesium(II)]‐μ‐5,5′‐(azanediyl)ditetrazolato‐κ3N1,N1′:N5] hemi{bis[μ‐5,5′‐(azanediyl)ditetrazolato‐κ3N1,N1′:N2]bis[triaquamagnesium(II)]} monohydrate], {[Mg(C2HN9)(H2O)3][Mg2(C2HN9)2(H2O)6]0.5·H2O}n, (I), and bis[5‐(pyrazin‐2‐yl)tetrazolate] hexaaquamagnesium(II), (C5H3N6)[Mg(H2O)6], (II), have been prepared under hydrothermal conditions. Compound (I) is a mixed dimer–polymer based on magnesium ion centres and can be regarded as the first example of a magnesium–tetrazolate polymer in the crystalline form. The structure shows a complex three‐dimensional hydrogen‐bonded network that involves magnesium–tetrazolate dimers, solvent water molecules and one‐dimensional magnesium–tetrazolate polymeric chains. The intrinsic cohesion in the polymer chains is ensured by N—H...N hydrogen bonds, which form R22(7) rings, thus reinforcing the propagation of the polymer chain along the a axis. The crystal structure of magnesium tetrazole salt (II) reveals a mixed ribbon of hydrogen‐bonded rings, of types R22(7), R22(9) and R24(10), running along the c axis, which are linked by R24(16) rings, generating a 4,8‐c flu net.  相似文献   

13.
Three types of metal complexes containing coordinated zwitterionic 8-Quinolinol(oxine) are isolated from the reaction ofMOx 2 (M=divalent Ni, Mn, or Mg; HO x =oxine) and haloacetic acidsRCO2H (R=CF3, CCl3, CHCl2, or CH2Cl) in benzene. These types are:M(O2CR)Ox·HOx forM=Ni,R=CCl3, CHCl2, and CH2Cl and forM=Mn,R=CHCl2.MOx(HOx) (RCO2)MOx·nH2O forM=Ni, Mn, or Mg,R=CF3 andn=1,1, and 4, respectively.MO x (HOx) (RCO2)2 MOx forM=Mn andR=CCl3. These types are compared with the simple mixed chelateMn(O2CCH2Cl)Ox. Interrelated reactions are suggested to explain the formation of these metal complexes and the contributing factors are discussed. The coordination of the zwitterion to the metal ion through its phenolate oxygen and the presence of the triatomic system+N–H...O in the three types of metal complexes are evidenced by typical infrared bands. Analytical and spectral data are in accordance with the suggested formulations.
Koordination von zwitterionischem 8-Chinolinol (Oxin) an gemischten Oxinat-Carboxylat-Komplexen des divalenten Nickel, Mangan und Magnesium
Zusammenfassung Drei Typen von Metallkomplexen mit koordiniertem zwitterionischem 8-Chinolinol (Oxin) wurden aus der Reaktion vonMOx 2 [M=Ni(II), Mn(II), Mg(II); HOx=Oxin] mit Halogen-essigsäurenRCOOH (R=CF3, CCl3, CHCl2, CH2Cl) in Benzol isoliert. Es werden Reaktionswege zur Bildung der Komplexe diskutiert. Die Koordination des Zwitterions über den phenolischen Sauerstoff und die Präsenz der Gruppierung+N–H...O in allen Typen der untersuchten Metallkomplexe wird auf Grund typischer IR-Banden nachgewiesen.
  相似文献   

14.
Phase diagrams were studied for (R4N)2[Nd(NO3)5]-C n H2n + 2-chloroform liquid ternaries (where R4N+ is trialkylbenzylammonium and n = 10, 12, 14, or 15) at T = 298.15−333.15 K. (R4N)2[Nd(NO3)5]-C n H2n + 2 binaries are two-phase liquid systems at all temperatures in this range. One phase (phase I) is an almost pure hydrocarbon solvent. The other (phase II) is enriched in (R4N)2[Nd(NO3)5]. The C n H2n + 2 solubility in phase II decreases with increasing the alkyl chain length of the hydrocarbon solvent and increases with increasing temperature. The title liquid ternaries are characterized by a homogeneous solution field and a two-phase liquid solution field. One phase is enriched in (R4N)2[Nd(NO3)5] and chloroform; the other is enriched in the hydrocarbon solvent. Liquid-liquid phase separation fields enlarge with increasing C n H2n + 2 alkyl chain length and slightly narrow with increasing temperature. Critical solution points at fixed temperatures depend on C n H2n + 2. Original Russian Text ? A.K. Pyartman, V.A. Keskinov, P.V. Zaitsev, 2009, published in Zhurnal Neorganicheskoi Khimii, 2009, Vol. 54, No. 3, pp. 531–534.  相似文献   

15.
16.
Tetrabenzyltitanium (B4Ti), tribenzyltitanium chloride (B3TiCl), tetra(p-methylbenzyl)titanium (R4Ti) and tri(p-methylbenzyl)titanium chloride (R3TiCl) have been used as catalysts for ethylene and propylene polymerization activated by AlEt2Cl. B4Ti-AIEt2Cl in solution polymerizes ethylene readily but its activity decays rapidly. B4Ti was also supported on Cab-O-Sil, Alon C, and Mg(OH)Cl. The last support was found to give catalyst with longest lifetime with a rate of polymerization, Rp = 7.0 g/hr-mmole Ti-atm ethylene. 14CO counting techniques gave 1.13 × 10?3 mole of propagating center per mole of B4Ti; the rate constant of propagation, kp = 540 l./mole-sec. None of the tetravalent titanium compounds polymerize propylene in solution. However, when supported on Mg(OH)Cl, Cab-O-Sil, Alon C, Cab-O-Ti, and charcoal, they all polymerize propylene. In this work the supports were characterized by various techniques, including the paramagnetic probe method, to determine the concentration and nature of surface hydroxyls. Those factors controlling the rate and stereospecificity of propylene polymerization were investigated. The system B3TiCl–Mg(OH)Cl–AlEt2Cl is the most active with Rp = 2.89 g/hr-mmole Ti-atm propylene. The concentration of propagation center is 0.9 × 10?3 mole per mole of B3TiCl; kp = 32 l./mole-sec. This catalyst gave only about 70% stereoregular polymer. Diethyl ether is found to raise stereospecificity to 100%, but there is a concommittent tenfold decrease of activity. Other interesting catalyst systems are: (π-C5H5)TiMe3–Mg(OH)Cl–AlEt2Cl (1.56, 89.5); (π-C5H5)TiMe2–Mg(OH)Cl–AlEt2Cl (0.075, 94.5); and (π-C5H5)TiMe3–Alon C–Al-Et2Cl (0.08,97.2), where the first number in the parenthesis is Rp in g/mmole Ti-hr-atm and the second entry corresponds to percentage yield of stereoregular polypropylene. Hafnocene and titanocene supported on Mg(OH)Cl produce only oligomers of propylene.  相似文献   

17.
The nitrogen protonation energies of the imino bases HN?CHR, where R is H, CH3, NH2, OH, and F, have been evaluated to determine the dependence of absolute and relative protonation energies on geometry, basis set, and correlation effects. Reliable absolute protonation energies require a basis set larger than a split-valence plus polarization basis, the inclusion of correlation, and optimized geometries of at least Hartree–Fock 4-31G quality. Consistent relative protonation energies can be obtained at the Hartree–Fock level with smaller basis sets. Extending the split-valence basis set by the addition of polarization functions on all atoms decreases the computed absolute Hartree–Fock nitrogen protonation energies of the imino bases HN?CHR except when R is F, but increases the oxygen protonation energies of the carbonyl bases O?CHR.  相似文献   

18.
Data are presented to show that two correlations of viscosity–concentration data are useful representations for data over wide ranges of molecular weight and up to at least moderately high concentrations for both good and fair solvents. Low molecular weight polymer solutions (below the critical entanglement molecular weight Mc) generally have higher viscosities than predicted by the correlations. One correlation is ηsp/c[η] versus k′[η], where ηsp is specific viscosity, c is polymer concentration, [η] is intrinsic viscosity, and k′ is the Huggins constant. A standard curve for good solvent systems has been defined up to k′[η]c ≈? 3. It can also be used for fair solvents up to k′[η]c ≈? 1.25· low estimates are obtained at higher values. A simpler and more useful correlation is ηR versus c[η], where ηR is relative viscosity. Fair solvent viscosities can be predicted from the good solvent curve up to c[η] ≈? 3, above which estimates are low. Poor solvent data can also be correlated as ηR versus c[η] for molecular weights below 1 to 2 × 105.  相似文献   

19.
Catalytic cyclometallation of bicyclo[2.2.1]hept-2-ene, spirocyclo[2.2.1]hept-2-ene-7.1'-cyclopropane andendo-tricyclo[5.2.1.02,6]deca-3,8-diene (dicyclopentadiene) with R2Mg (R =n-Pr,n-Bu) leading to the formation of polycyclic magnesacyclopentanes (MC) has been studied. The yield of MC and the selectivity of the reaction have been shown to depend on the ratio of starting reagents, the solvent, and temperature. The most probable scheme of the transformation studied is proposed.Translated fromIzvestiya Akademii Nauk, Seriya Khimicheskaya, No. 1, pp. 165–169, January, 1993.  相似文献   

20.
Summary Relationships between R M values and pH of aqueous stationary phase were determined for some heterocyclic bases and their N-oxides chromatographed in partition systems of the type: organic solvent (benzene, chloroform or mixture of chloroform with cyclohexane)—aqueous buffer solution. In view of large differences in pK A values of the bases and their N-oxides, the R Mr o (NNO) values (change of R M during suppressed ionization, due to reaction) could only be estimated by graphical extrapolation of the chromatographic data. The R Mr o values thus determined were in most cases in the range 2.2–2.7 R M units.  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