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.
Network formation (branching) theory was used to examine the relationships between network structure and concomitant sound and vibration damping. For a series of model polyether-based polyurethane networks with varying stoichiometry and composition, the glass transition temperature Tg, was found to increase with increasing concentration of elastically active network chains, EANCs, as well as the ratio of branch OH group concentration to the total OH group concentration ρ. The values of (tan δ)max, the peak height of tan δ at Tg, linearly decrease with increasing concentration of EANCs, regardless of the ρ values. However, the loss area (LA), equal to the integral of the linear loss modulus-temperature curve, is independent of the concentration of EANCs and/or ρ. Utilizing group contribution analysis techniques, the value of the main chain -O- group contribution, LA-O-, is 19.1 GPa·K/g, a rather large value. This finding gives insight into why polyether urethanes are preferred for many damping applications.  相似文献   

2.
Using the theory of branching processes, structural parameters such as the molecular weights of elastically active network chains (EANCs), including dangling chains, backbone EANCs and dangling chains of networks built up by the alternating polyaddition of a bi- and trifunctional monomer, are characterized. The theory is compared with viscoelastic data on polyurethane networks prepared from poly(oxypropylene)triols and diisocyanate at various initial ratios of functional groups; in the calculation, the distribution of functionalities of the triols used and the possible incompleteness of the reaction is taken into account. The comparison reveals that both the length of the backbone EANC and the length of dangling chains contribute to the total width of the retardation spectrum.  相似文献   

3.
In the hydrogen‐bonding networks of 8‐hydroxy‐5‐hydroxy­methyl‐3,6‐dioxatricyclo­[6.3.1.01.5]dodecan‐2‐one and 5,7‐bis(hydroxy­methyl)‐3,6‐dioxatricyclo­[5.3.1.01.5]undecan‐2‐one, both C11H16O5, layers and double strands, respectively, lead to the formation of chains connected by hydroxy‐to‐hydroxy contacts, where the hydroxy­methyl group, present in both structures, acts as a donor. The secondary structures differ in the hydrogen bonding of these chains via the second hydroxy group, which is involved in hydroxy‐to‐carbonyl and hydroxy‐to‐hydroxy bonds, respectively.  相似文献   

4.
The effect of the initial mole ratio of reactive components on the shape and position of dynamic mechanical functions in the main transition and rubbery region was investigated for two series of networks made from poly(oxypropylene)diamine (D-400)-diglycidyl ether of Bisphenol A (DGEBA) and poly(oxypropylene)-triamine (T-403)-DGEBA. The networks were prepared with an excess of amine groups up to the highest conversion of epoxy groups; the ratio rH = 2 [ NH2 ]0/ [E]0 ranged from unity to 2,1 for networks from D-400 and from unity to 3,5 for networks from T-403. By using the theory of branching processes, structural parameters of these networks were calculated, in particular, the molecular weights of elastically active network chains (EANC's) including dangling chains, of backbone EANC's and of dangling chains. A comparison between theory and experiment led to the following conclusions: (a) the mechanical behaviour in the rubberlike region can be described either by using an affine deformation model (front factor A = 1), or by means of a phantom model (A = (fe-2)/fe, fe being functionality of the crosslink) with the contribution of permanent interchain interactions; (b) the temperature and frequency position of viscoelastic functions in the main transition region is conclusively affected by the concentration of EANC's; (c) the shape of visco-elastic functions, especially of retardation spectra in the main transition and rubbery region, depends on the detailed structure of EANC's, but it cannot be decided from the result which structural parameter has the strongest effect on the shape of the functions.  相似文献   

5.
Dynamic mechanical, differential scanning calorimetry and X-ray scattering behavior of ordered polyurethane systems, based on a diol with rigid (mesogenic) group in side chain (D), 2(4)-methyl-1,3-phenylene diisocyanate (DI) and two triols (T)--stiff trimethylolpropane (TMP) or flexible poly(oxypropylene)triol (PPT), was investigated during crosslinking and on the networks. The networks were prepared at various stoichiometric initial molar ratios of the reactive groups, [OH]T/[NCO]DI/[OH]D ranging from 1/2/1 to 1/20/19. From our measurements it follows that: (a) Power-law parameters, which are characteristic of the structure at the gel point (the gel strength S and the relaxation exponent n), are dependent on the initial ratio of the reactants. With increasing content of mesogenic diol in the system (increasing length of elastically active network chains, EANCs), the gel strength S increases and the relaxation exponent n decreases; higher S and lower n are found for stiffer TMP networks in comparison with more flexible PPT ones. (b) Introduction of crosslinks reduces the flexibility of the network chains in fully cured samples and inhibits conformational rearrangements required for ordering. A more complex thermal behavior was found for networks based on TMP in comparison with those based on PPT. (c) Strong physical interactions between the mesogens promote cyclization in the course of crosslinking; the fraction of bonds lost in intramolecular cycles is ∼15% for fully cured networks.  相似文献   

6.
The transition states for fragmentation of dihydroxycarbene [C(OH)2] to H2 and CO2 and for the rearrangement of this carbene to formic acid were located by ab initio calculations. The relative energies of the transition states were determined at several levels of theory and the basis set dependence of the energies is discussed. At the best level of theory; using a basis set of double-zeta quality augmented by polarization functions and with the inclusion of extensive CI, we found that the transition state for fragmentation was considerably higher in energy than that for rearrangement. This finding is at variance with the predictions of the Woodward--Hoffmann rules because fragmentation represents an “allowed” reaction, whereas rearrangement is “forbidden.” In conformity with the Woodward–Hoffman rules, the transition state for rearrangement was found to be close in energy to H· + ·CO2H. The even higher energy of the transition state for concerted fragmentation to H2 and CO2 is attributed to the need for the latter fragment to remain substantially bent in order to permit H2 formation while maintaining a modicum of OH bonding. Difficulties in locating the transition state for concerted fragmentation are discussed and a new method for finding transition states is proposed.  相似文献   

7.
The room temperature reaction of Na4Sn2S6 · 5H2O with CoCl2 · 6H2O and 2-(aminomethyl)pyridine (2-AMP) or trans-1,2-diamino-cyclohexane (DACH) leads to crystallization of two compounds with the compositions [Co(2-(aminomethyl)pyridine)3]2 Sn2S6 · 10H2O ( 1 ) and [Co(trans-1,2-diaminocyclohexane)3]2Sn2S6 · 8H2O ( 2 ). In both compounds [Sn2S6]4– anions are present that are charge balanced each by two Co2+ centered complexes. Each of the two CoII cations are sixfold coordinated by six N atoms of three 2-AMP or DACH ligands within slightly distorted octahedra. In compound 1 , the two complexes are linked by one [Sn2S6]4– anion via strong N–H ··· S hydrogen bonds into centrosymmetric charge neutral trimeric units, that are further linked by weak C–H ··· S and N–H ··· S hydrogen bonds into chains that are directed along the a axis. These chains are further joined by N–H ··· O and O–H ··· O hydrogen bonds into a 3D network, with the H2O molecules forming chains along the b axis. The crystal structure of 2 is similar to that of 1 featuring trimeric units which are also linked into chains. Between the chains water molecules are embedded that link the chains into a 3D network. Upon heating 2 in a thermobalance the water and ligand molecules are removed in discrete steps, indicating that compounds with more condensed thiostannate networks will form.  相似文献   

8.
Temperature dependence of the shape and linewidth of the broad-line NMR spectra of commercially available high-density polyethylene (PE/HD), low-density polyethylene (PE/LD with ~3 per cent CH3), block-copolythene (PE/AA with ~3 per cent acrylic acid) and polyethylene single crystal (PE/SC) were investigated to obtain information on the effect of branching and structural changes on the glass-transition temperature (Tg) and activation energy of molecular motion (E). The relatively lower value of Tg ~- ?100° for PE/HD compared to Tg ~- ?85° and ?60° for PE/CH3 and PE/AA, along with the estimated lower value of E ~- 2·1 kcal/mole for PE/HD compared to 2·6 and 5·1 kcal/mole for PE/CH3 and PE/AA, respectively, were interpreted by a molecular reorientational process connected mainly with the sidegroups of the main polymeric chains.  相似文献   

9.
Summary The subject of this work is to determine the dynamic and static elastic properties of polymer networks and to compare the experimental results with theoretical predictions.It is found that poly(dimethylsiloxane) (PDMS) networks on the one hand and polyisoprene (IR) and natural rubber (NR) networks on the other hand show different behaviour in frequency dependence of the storage modulusG is observed in the range 10–3 – 1 Hz, contrary to IR or NR. The reason for this is the absence of entanglements in PDMS.Comparing the measured static moduli with those calculated by rubber elasticity theory, we found that the front factorA· <r 2>/<r 2>0 is near to 0.5 in case of PDMS, but near to I for IR/NR networks. Both parts of the front factor may possibly cause this difference. As the PDMS chains possess a relatively high mobility we can assume that fluctuations of the junction points are less restricted than in IR/NR. This causes a structure factor A smaller than 1 in agreement with Flory's recent theory.According to James and Guth, the network chains tend to contract during the crosslinking process. This will be more likely in the networks of entanglement-free, highly mobile PDMS chains than in IR or NR rubbers. Hence the memory term <r 2>/<r 2>0 is smaller for PDMS than for IR/NR.Both alternative explanations are based on the different mobility of the chains considered. It may be assumed that the front factor is influenced by both effects and not only by the fluctuations of crosslinks.Dedicated to Professor Dr. K. Ueberreiter on his 70th birthday  相似文献   

10.
Two polymeric silver(I) complexes, [Ag4(TAA)2(TA)2] · 2[Ag(TA)2] (I) and [Ag2(TA)(BA)] n (II), where TAA is tris(2-aminoethyl)amine, TA is terephthalic acid, BA is butane-1,4-diamine, were synthesized and structurally characterized by elemental analyses and X-ray diffraction. Complex I consists of mononuclear [Ag(TA)2] units and TA-bridged tetranuclear [Ag4(TAA)2(TA)2] units, which are further linked via Ag…Ag interactions to form chains. In complex II, the silver atoms are linked by the TA and BA groups to form chains, which are further linked via Ag…Ag interactions to form layers. In both complexes, molecules are finally linked through intermolecular hydrogen bonds to form three-dimensional networks.  相似文献   

11.
The synthesis of trans-[Ni(dbm)2(RN=C(Me)NHR)2] (dbm?=?1,3-diphenylpropanedionate; R?=?phenyl, p-tolyl, 3,5-dimethylphenyl, 3,5-dichlorophenyl) has been achieved by reaction of [Ni(dbm)2(H2O)2] with two equivalents of the amidine ligands, RN=C(Me)NHR. X-ray crystallographic studies reveal that trans-[Ni(dbm)2(RN=C(Me)NHR)2] (R?=?3,5-dimethylphenyl) exhibits an intramolecular N?CH···O hydrogen bond, which along with the large steric bulk of the amidine ligands may enforce the trans geometry. Electrochemical studies show a strong dependence of the oxidation potential on the substituent groups despite their remote position. DFT calculations indicate that the HOMO consists of Ni-ligand ??* orbitals with a significant contribution from the amidine ligands.  相似文献   

12.
In this paper, we have used the hydrogen-bonding interactions, combining the designed diamine ligands and anionic metal chlorides, into the construction of a series of new pillar-layered supramolecular complexes. The flexible molecule N,N,N′,N′-tetra-p-methoxybenzyl-ethylenediamine (L1) bearing doubly protonated H-bond donors, has been synthesized and reacted with the metal chlorides (such as [PdCl4]2?, [FeCl4]? and [CoCl4]2?) via weak C–H···Cl interactions, yielding crystal products [H2 L1]2+·Cl?·[FeCl4]? (1), 0.5H2O ? [H2 L1]2+·Cl?·0.5[PdCl4]2? (2) and [2-hydroxy naphthyl]1.5 ? 2[H2 L1]2+·2Cl?·[CoCl4]2? (3). The 3-D networks are organic double layers formed by the self-assembly of the ligands through extensive hydrogen-bonding interactions (C–H···O or C–H···π interactions) and further interconnected by [PdCl4]2?/[FeCl4]?/[CoCl4]2? in a pillar fashion, constructing into pillar-layered networks with channels accessible to various guest molecules. The inclusion property of [H2 L1][CoCl4] was studied, varieties of guest molecules, such as 2-hydroxy naphthyl, phenanthrene and hydroquinone, can be included in the framework.  相似文献   

13.
I. Tabushi  K. Yamamura 《Tetrahedron》1975,31(16):1827-1831
Charge-transfer transition energies, association constants, and molar extinction coefficients of complexes between tetracyclo[3.2.0.02,7.04,6]heptane(quadricyclane) and olefine type or quinone type acceptors were measured in methylene chloride at 20°. An excellent linear relationship (r = 0·9938) with a slope of 0·96 was observed in plots of νmax(CT) for a series of complexes of quadricyclane against νmax(CT) for the corresponding complexes of N,N-dimethyl-aniline, indicating that quadricyclane forms electron donor-acceptor complexes of weak interactions. The first ionization potential, estimated from the hyperbolic relationship between the charge-transfer transition energy (ECT) and the ionization potential, was an exceptionally low value (8·28–8·32 eV) for a saturated hydrocarbon and was indeed in the same order of magnitude with that of norbornadiene, which was well reproduced by EHM and MINDO/1 calculations applying Koopmans' theorem to the calculated HOMO energy.  相似文献   

14.
By means of a galvanic cell, emf values were measured for the solid-state reactionsnKCl+ MCl2 = KnMCln+2 for all existing compounds in the pseudobinary systems withM = Mg and Mn. ThusΔGr values could be calculated and, from their linear temperature dependence in the range 550–730 K, reaction entropies could be determined. EnthalpiesΔHr were calculated using the Gibbs-Helmholtz relation; they are compared with values found by solution calorimetry at room temperature. The magnitude of the entropy term for the free enthalpy of the formation reactions is discussed for the different compounds. For the modifications ofKMCl3 the lattice parameters for the cubic, tetragonal, and one of the orthorhombic phases were determined by X-ray photographs at varying temperatures. By DSC measurements the transition enthalpy for the tetragonal to cubic transition of KMnCl3 at 659 K was found to be 0.20–0.4 kJ · mole?1, compared to 4.6 kJ · mole for the transition of the stable room-temperature modification with the NH4CdCl3 structure to the metastable GdFeO3 structure.  相似文献   

15.
Two cobalt(II) coordination polymers, {[Co(μ‐4,4′‐bipy)(4,4′‐bipy)2(H2O)2]·(OH)3·(Me4N)·4,4′‐bipy·4H2O}n ( 1 ) and {[Co(μ‐4,4′‐bipy)(H2O)4]·suc·4H2O}n ( 2 ) (4,4′‐bipy = 4,4′‐bipyridine, suc = succinate dianions), were hydrothermally synthesized and structurally characterized by X‐ray diffraction analysis, UV‐Vis‐NIR, and ICP. The main structure feature common to the both polymers is presence of the infinite linear chains, [Co(μ‐4,4′‐bipy)(4,4′‐bipy)2(H2O)2]n ( 1 ) and [Co(μ‐4,4′‐bipy)(H2O)4]n ( 2 ), respectively. In 1 , the chains are further linked by the hydrogen‐bond and π‐π stacking interaction, producing extended layer structure. The 4,4′‐bipy molecules in 1 play three different roles. In 2 , the chains are linked into three‐dimensional network structure via complicated hydrogen bonding system. The variable temperature (2.0~300 K) magnetic susceptibility of 1 indicates a tendency of spin‐transition in the temperature range of 110 K to 22 K, which attributes to the transition of high‐spin to low‐spin from Co2+(d7) ion. Also, the result of surface photovoltage spectroscopy (SPS) reveals that the polymer 1 has significant photoelectric conversion property in the region of 300‐800 nm.  相似文献   

16.
A new compound consisting of Keggin-type polyoxometalate and transition metal complex, [Cr(apca)2(H2O)2]4[SiW12O40]·12H2O(1), (apca??=?3-aminopyrazine-2-carboxylate anion), was synthesized with conventional method and characterized by single crystal X-ray diffraction, elemental analysis, IR spectroscopy, TG analyses, PXRD. Fluorescent property was explored. The [SiW12O40]4? anions and complex cations [Cr(apca)2(H2O)2]+ in compound 1 are connected by hydrogen bonds and π–π interaction, forming the 3D structure. Especially, [Cr(apca)2(H2O)2]+ cations are linked into left-hand helical chain extending along c-axis through hydrogen bonds, forming a channel A. The chains connect other four chains from four directions also through hydrogen bonds, then the four chains enclose a channel B whose inside dimension is a little bigger than that of A, polyoxoanions fill in the channel B, and free waters in the channel A. Compound 1 shows an asymmetric wide emission band at λmax?=?435 nm, which may belongs to the intraligand π*?→?π transition.  相似文献   

17.
The compounds [C4H8ON(C3H5)2]+[Cu4Cl5] (I), [C4H8ONH(C3H5)]+[CuBr2] (II), and [C4H8ONH(C3H5)]+[CuBr1.41Cl0.59] (III) were prepared for the first time by ac electrochemical synthesis from mono- and di-N-allyl derivatives of morpholinium and copper(I) halides in ethanol solution and structurally characterized. In the structure of I π-complex, the centrosymmetric Cu8Cl10 fragments are associated into layers perpendicular to the b axis. The N,N"-diallylmorpholinium cation functions as a bridge, which coordinates two copper atom of the adjacent inorganic fragments by both allyl groups. The trigonal-pyramidal surrounding of the Cu(I) atom, as well as the distorted tetrahedral coordination sphere of Cu(2), involves three chlorine atoms and the C=C bond, whereas the planar trigonal surrounding of the Cu(3) atom and trigonal-pyramidal surrounding of the Cu(4) atom involve only chlorine atoms. In the isostructural II and III σ-complexes, the edge-shared CuX4 tetrahedra form the infinite copper-halide chains running along the a axis. The inorganic fragments and organic N-allylmorpholinium cations are united into the three-dimensional crystal structures by N–H···X and C–H···X (X = Cl, Br) hydrogen bonds.  相似文献   

18.
High, electrochemically controlled electrical conductivity (0.2–0.3 S cm?1) was measured for the first time for the metal–metal bonded polymeric compound [Ru(bpy)(CO)2]n (bpy = 2,2′-bipyridine). The conductivity depended on the degree of partial oxidation of the initially zero-valent Ru to RuI in the chains and was found to increase by four orders of magnitude after 5% oxidative doping. Linear conductance vs. doping level dependence was found for low levels and the conductivity of films could repeatedly be switched on and off. The conduction process is ascribed to charge delocalisation on the Ru chains in analogy to that of other similar chains formed by stacking of planar transition metal d8 complexes.  相似文献   

19.
In this study, we have deliberately utilized the second-sphere coordination approach into the construction of supramolecular inclusion solids Cl ? [H2 L1]·[InCl4] (Crystal I) and Br ? [H2 L1]·[TeBr6] (Crystal II). The chloride or bromine anions can be encapsulated inside the host assemblies formed by the diamine molecule (4,6-dimethyl-1,3-phenylene) bis(N,N-dibenzylmethane) (L1) and the metal complexes ([InCl4]? and [TeBr6]2?) via second-sphere interactions. The inclusion complexes have been structurally characterized by X-ray crystallography, indicating that weak C–H···Cl and C–H···Br hydrogen bonding synthons play a significant role in the construction of host framework. 2-D networks are formed in both complexes by the interconnection of 1-D networks through the multiple weak hydrogen bonding interactions with [InCl4]? or [TeBr6]2?. The guest Cl? or Br? anions are encapsulated inside the host cages through N–H···Cl hydrogen bonds. The inclusion selectively was studied for the two host assemblies.  相似文献   

20.
Crystals of MnNbOF5 · 4H2O were studied by X-ray diffraction, differential scanning calorimetry (DSC) and thermogravimetry. DSC showed a phase transition in the temperature range of 282 to 296.3 K. The compound is dehydrated in two stages in the temperature ranges of 65–131.1 and 131.1–190°C, two H2O molecules being removed in each stage. The structure of MnNbOF5 · 4H2O was determined at 153 K (α phase; space group P21/c) and at 297 K (β phase, space group C2/m). The structure of both phases is formed of the octahedral complexes [NbF4(O/F)2/2]0.5? and [Mn(H2O)4(O/F)2/2]0.5+ linked by bridging O and F atoms to infinite chains. The isolated niobium-manganese chains are connected by O-H…F hydrogen bonds. In the α phase, the Nb, O, and F atoms in the trans position relative to O are disordered with respect to the inversion center of the structure. Transition to the β phase is accompanied by splitting of all ligand positions at the Nb and Mn atoms into two equally probable positions.  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