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 26 毫秒
1.
Average values for dispersion γsd and polar γsd contributions of the solid surface tension γs γsd + γsp for poly(methylene oxide) (PMO) and Na-treated polytetrafluoroethylene (PTFE) are determined by a new computational analysis of wettability data. PMO displays γsd equals; 21.8 ± 0.9 and γsp = 11.5 ± 1.5 dyn/cm while Na-treated PTFE displays γsd = 36.1 ± 3.0 and γsp = 14.5 ± 2.9 dyne/cm. These surfaces present the highest fractional surface polarity ps = γsps = 0.29-0.35 yet encountered for organic polymers or oriented monolayers. These unusual surface tension properties are correlated with surface chemistry and adhesion phenomena.  相似文献   

2.
Polymers with fluoroalkyl side-groups on three of every four carbon atoms along the polymer main chain were made by treating 1 : 1 copolymers of perfluoroallyl (or methallyl)ethers and maleic anhydride with SF4, and then esterifying the resulting acid fluoride groups with 1H, 1H-pentadecafluorooctanol. Surface wettability of these polymers by polar and non-polar liquids was studied. The critical surface tension (γc) was found by Zisman's method. The dispersion force component of polymer surface energy (γsD) and the polar component of surface energy (γsp) were calculated by the respective methods of Fowkes and Owens. The γc values for several of the highly fluorinated polymers were lower than previously reported for any fluoropolymer but not so low as has been observed for the surface of an oriented perfluoroacid monolayer. In the range 11.4–18.5 dyne/cm, γc approximates γsD for the fluorinated surfaces; however at lower γc values, considerable difference between γc and γsD was noted.  相似文献   

3.
A series of bis(trifluoromethyl)carbinyl acrylate monomers [Y-C(CF3)2 O? CO? CH?CH2] in which Y is CH3, CH3CH2, CH3CH2CH2, CH3CH2CH2CH2, C6H5, H, F, CF3, N3, CN, and CH3OCH2CH2O, was prepared. Polymers were easily prepared from all of these monomers except where Y = CN, wherein a variety of initiation methods failed to produce high molecular polymer. Wettabilities of the polymer films were examined by means of contact angle measurements by using n-alkane test liquids and water. Values of the dispersion force contribution (γsd) and the polar force contribution (γsp) to the solid surface energy were calculated by employing both geometric and harmonic mean approximations. Values of γsd calculated by either method agreed well with γc (critical surface tension) values determined graphically from contact angle data employing n-alkane test liquids, confirming the suggestion that γc is an approximate measure of the dispersion force contribution to solid surface energy. Values of γsd ranged from 15 dyne/cm (Y = F or CF3) to 25 dyne/cm (Y = C6H5). Values of the polar force contribution to solid surface energy (γsp) varied from 0.6 dyne/cm (Y = CH3CH2CH2CH2) to 3.4 dyne/cm (Y = CH3OCH2CH2O) when calculated by the geometric mean equation. The values of γsp obtained from the harmonic mean equation followed the same trend upon varying substituents, but were larger in value, ranging from 2.9 dyne/cm (Y = CH3CH2CH2CH2) to 7.5 dyne/cm (Y γ CH3OCH2CH2O).  相似文献   

4.
The adhesive polydecapeptide poly(Lys‐Pro‐Thr‐Gln‐Tyr‐Ser‐Asp‐Glu‐Tyr‐Lys) (average repeating number, n = 5), which is the consensus sequence of the Asian freshwater mussel Limnoperna fortunei adhesive protein (Lffp), has been synthesized by the polycondensation of the active esters. The surface chemical experiments revealed the following characteristics of the freshwater adhesive protein: (i) wettability of the Lffp solution is affected by the polar component value (γsp) of the surface free energy of the substrate, and a substrate having a γsp less than 10 mJ·m–2 exhibits a reduced wettability of the Lffp solution; (ii) the comparison of wettability of native Lffp with synthetic Lffp suggests that the decapeptide sequence, ‐Lys‐Pro‐Thr‐Gln‐Tyr‐Ser‐Asp‐Glu‐Tyr‐Lys‐, contributes to the interaction with the underwater surface; (iii) the Lffp tends to adsorb on nonpolar surfaces that have a low γsp value; and (iv) the adsorption ability of the freshwater adhesive protein is less than that of the marine adhesive protein because of the higher hydrophilicity of the freshwater adhesive protein. An antifouling examination indicated that a γsp value of the substrate surface of less than 10 mJ·m–2 should achieve a higher antifouling effect towards the L. fortunei attachment. These results are the first findings for the development of a freshwater antifouling strategy based on the molecular mechanism underlying the attachment of L. fortunei.  相似文献   

5.
Lanthanide trihalide molecules LnX3 (X = F, Cl, Br, I) were quantum chemically investigated, in particular detail for Ln = Lu (lutetium). We applied density functional theory (DFT) at the nonrelativistic and scalar and SO‐coupled relativistic levels, and also the ab initio coupled cluster approach. The chemically active electron shells of the lanthanide atoms comprise the 5d and 6s (and 6p) valence atomic orbitals (AO) and also the filled inner 4f semivalence and outer 5p semicore shells. Four different frozen‐core approximations for Lu were compared: the (1s2–4d10) [Pd] medium core, the [Pd+5s25p6 = Xe] and [Pd+4f14] large cores, and the [Pd+4f14+5s25p6] very large core. The errors of Lu? X bonding are more serious on freezing the 5p6 shell than the 4f14 shell, more serious upon core‐freezing than on the effective‐core‐potential approximation. The Ln? X distances correlate linearly with the AO radii of the ionic outer shells, Ln3+‐5p6 and X?np6, characteristic for dominantly ionic Ln3+‐X? binding. The heavier halogen atoms also bind covalently with the Ln‐5d shell. Scalar relativistic effects contract and destabilize the Lu? X bonds, spin orbit coupling hardly affects the geometries but the bond energies, owing to SO effects in the free atoms. The relativistic changes of bond energy BE, bond length Re, bond force k, and bond stretching frequency vs do not follow the simple rules of Badger and Gordy (Re~BE~kvs). The so‐called degeneracy‐driven covalence, meaning strong mixing of accidentally near‐degenerate, nearly nonoverlapping AOs without BE contribution is critically discussed. © 2015 Wiley Periodicals, Inc.  相似文献   

6.
In this study, polyethylene glycol (PEG) of different molecular weight was added into poly(vinylidene fluoride) (PVDF) through melt blending. The hydrophilicity, processability, mechanical properties of the blends were investigated. Moreover, to understand fully, the crystallization and melting behavior, crystalline structure, molecular interactions of the blends were also studied. Results revealed that the addition of PEG evidently enhances the ductile and reduces the rheological torque of PVDF, what’s more, the contact angle of the blends with water was greatly reduced. The solid surface tension γs, chromatic dispersion part γ s d and polar part γ s p suggested the surface tension of the blend gradually increased from 38.31 to 55.65 mJ/m2 with the increase of PEG-20000 content.Wherein the dispersion part of the tension changed little, the polar part of the tension increased significantly from 5.74 to 20.41 mJ/m2, indicating that the addition of PEG greatly enhanced the polar part of surface tension of the blend, and therefore evidently enlarged the hydrophilic property of the blends. Besides, to elucidate the related effecting mechanism, the crystallization and molecular interaction of the blends were also studied.  相似文献   

7.
A theory of the fracture of polymers with network microstructure was developed that was based on the vector, or rigidity percolation (RP) model of Kantor and Webman, in which the modulus, E, is related to the lattice bond fraction p, via E ~ [p ? pc]τ. The Hamiltonian for the lattice was replaced by the strain energy density function of the bulk polymer, U = σ2/2E, where σ is the applied stress and p was expressed in terms of the lattice perfection via the bond density ν, with the entanglement molecular weight, ν = ρ/Me and appropriate measures of crosslink density for rubber, thermosets, and carbon nanotubes. The stored mechanical energy, U, was released by the random fracture of νDo[p ? pc] over stressed hot bonds of energy Do ≈ 330 kJ/mol. The polymer fractured critically when p approached the percolation threshold pc, and the net solution was obtained as σ = (2EνDo [p ? pc])1/2 with a fracture energy, G1c ~ [p ? pc]. The fracture strength of amorphous and semicrystalline polymers in the bulk was well described by, σ = [EDoρ/16 Me]1/2, or σ ≈ 4.6 GPa/Me1/2. Fracture by disentanglement was found to occur in a finite molecular weight range, Mc < M < M*, where M*/Mc ≈ 8, such that the critical draw ratio, λc = (M/Mc)1/2, gave the molecular weight dependence of the fracture as G1c ~ [(M/Mc)1/2 ? 1]2. The critical entanglement molecular weight, Mc, is related to the percolation threshold, pc, via Mc = Me/(1 ? pc). Fracture by bond rupture was in accord with Flory's suggestion, G/G* = [1 ? Mc/M], where G* is the maximum fracture energy. Fracture of an ideal rubber with p = 1 was determined not to occur without strain hardening at λ > 4, such that the maximum stress, σ = E (λ ? 1/λ) = 3.75E. The fracture properties of rubber were found to behave as σ ~ ν, σ ~ E, and G1c ~ ν. For highly crosslinked thermosets, it was predicted that σ ~ (Eν)1/2, σ ~ (X ? Xc)1/2, and G1c ~ ν?1/2, where X is the degree of reaction of the crosslinking groups and Xc defines the gelation point. When applied to carbon nanotubes (SWNT and MWNT) of diameter d and hexagonal bond density ν = j/b2, the nominal stress as a function of diameter is σ(d) = [16 EDo(p ? pc) j/b]1/2/d ≈ 211/d (GPa.nm) and the critical force, Fc(d) ≈ 166 d (nN/nm), in which j = 1.15, b = 0.142 nm, E ≈ 1 Tpa, and Do = 518 kJ/mol. For polymer interfaces with Σ chains per unit area of length L and width XL1/2, G1c is then ~ [p ? pc], where p ~ ΣL/X. The results predicted by the RP fracture model were in good agreement with a considerable body of fracture data for linear polymers, rubbers, thermosets, and carbon nanotubes. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 43: 168–183, 2005  相似文献   

8.
Localized molecular orbitals (LMOs) for several octahedral complexes are presented. Wavefunctions are calculated within the PRDDO approximations and localized by the Boys criterion. Complexes of general formula (NH3)x(CO)6-xM, M = Cr0 or Mn+ and x = 1, 2, or 3 illustrate the general trends for carbonyl complexes. Weak to moderate π-bonding results in three equivalent inner shell LMOs dominantly of metal 3s, 3p and 3d character but highly delocalized to the carbonyls. These three LMOs flank the M-CO bond axis. Other π back-bonding situations result in metal-ligand double bonds which are nonequivalent and have σ-π separability [(NH3)5(py)Mn+] and also equivalent double bonds [(NH3)5(NO)Cr+].  相似文献   

9.
New scaling laws for chain networks are derived to describe the fundamental relationships between the viscosity exponent (k), viscoelastic exponent (m), stretched exponent (β), spatial dimension (d). fractal dimension (df), and a universal constant (γ). The scaling of the total number of monomers and the radius of gyration is defined by df. We have discovered γ = m/β to be a universal constant which relates the shear modulus of a polymer gel melt to the shear modulus near the glass transition. Analyzing the size-dependent shear viscosity, we have determined γ = 3dfcd/(7d−5dfc) = 2.647 for d = 3 where dfc is the fractal dimension of critical clusters at the gel point. By using γ, the present theory extends previous work pertaining to systems near the sol-gel transition, and shows how properties far from the critical point can be explained. The theoretical prediction is in good agreement with viscoelastic measurements.  相似文献   

10.
The shear stress σ, two components of birefringence, and extinction angle were measured for a concentrated polystyrene solution in step-shear deformation of magnitude of shear 0.3 ≤ γ ≤ 4.0. The stress-optical coefficient did not depend on either γ or time. The first and the second normal-stress differences v1 and v2 were evaluated with the use of the stress-optical law. Over a certain range of long times, σ could be factored as σ = γh(γ)G(t) and the quantity h(γ) agreed with the prediction of the Doi–Edwards theory based on the de Gennes tube model of entangled polymer chains. At short times the effect of γ on σ/γ was smaller than at long times. The relaxation spectrum became approximately independent of γ at the short-time end of the rubbery plateau region. The ratios v1/σ and v2/v1 were independent of time and were in quantitative agreement with those predicted by the Doi–Edwards theory: v1/σ was equal to γ, v2/v1 was negative, and |v2/v1| decreased with increasing γ.  相似文献   

11.
The intermolecular interactions existing at three different sites between phenylacetylene and LiX (X = OH, NH2, F, Cl, Br, CN, NC) have been investigated by means of second‐order Møller?Plesset perturbation theory (MP2) calculations and quantum theory of “atoms in molecules” (QTAIM) studies. At each site, the lithium‐bonding interactions with electron‐withdrawing groups (? F, ? Cl, ? Br, ? CN, ? NC) were found to be stronger than those with electron‐donating groups (? OH and ? NH2). Molecular graphs of C6H5C?CH···LiF and πC6H5C?CH···LiF show the same connectional positions, and the electron densities at the lithium bond critical points (BCPs) of the πC6H5C?CH···LiF complexes are distinctly higher than those of the σC6H5C?CH···LiF complexes, indicating that the intermolecular interactions in the C6H5C?CH···LiX complexes can be mainly attributed to the π‐type interaction. QTAIM studies have shown that these lithium‐bond interactions display the characteristics of “closed‐shell” noncovalent interactions, and the molecular formation density difference indicates that electron transfer plays an important role in the formation of the lithium bond. For each site, linear relationships have been found between the topological properties at the BCP (the electron density ρb, its Laplacian ?2ρb, and the eigenvalue λ3 of the Hessian matrix) and the lithium bond length d(Li‐bond). The shorter the lithium bond length d(Li‐bond), the larger ρb, and the stronger the π···Li bond. The shorter d(Li‐bond), the larger ?2ρb, and the greater the electrostatic character of the π···Li bond. © 2012 Wiley Periodicals, Inc.  相似文献   

12.
The heat of fusion of virgin and melt-processed polytetrafluoroethylene (PTFE) was determined using the Clapeyron equation. Experimental data were obtained from PVT experiments and high-temperature x-ray diffraction measurements. For virgin, as-polymerized PTFE, the melting temperature is given by where, for Tm in degrees Celsius, A = 346.3±1.2, B = 0.095±0.003, and P is the pressure in kilograms per square centimeter. At the end of the atmospheric-pressure melting interval, the amorphous and crystalline specific volumes V1 and Vc are 0.6517 and 0.492 cm3/g, respectively. Thus the heat of fusion is 24.4 cal/g, or nearly twice the value reported previously. The increases in enthalpy and volume at the melting point both indicate a degree of crystallinity of about 75–80% although infrared, x-ray, and NMR data give much higher levels. Data from calorimetry, NMR, and dynamic mechanical measurements indicate that in virgin PTFE some of the crystals continue to experience torsional oscillations at temperatures below the room-temperature transitions. This indicates that there are at least two kinds of crystalline regions. For previously melted PTFE, Tm is determined by A = 328.5±0.7 and B = 0.095±0.002, the volumes are Vam = 0.6349 and Vcr = 0.4855 cm3/g, and the heat of fusion is 22.2 cal/g. The entropy of fusion for PTFE is much closer to that of polyethylene than was previously believed.  相似文献   

13.
Valence-bond calculations are reported for the isoelectronic series of molecules and ions: N2, CO, BF, NO+ and CN?. The most important structures are N?N, C?O, Bπ? F, N+?O and C?N. Hybridization of the 2s and 2p orbitals is important. Only two or three structures are required to obtain an energy lower than that obtained with the molecular orbital approximation. Structures in which the electronegative element loses a σ-orbital or gains a π-orbital are favored. π-bonds tend to be favored over σ-bonds. The bond in NO+ resembles that in CO, whereas that in CN? resembles the bonding in N2.  相似文献   

14.
Interactions between noble metals and rare gases have become an interesting topic over the last few years. In this work, a computational study of the open‐shell (d10s1) and closed‐shell (d10s and d10s2) noble metals (M = Cu, Ag, and Au) with three heaviest rare gas atoms (Rg = Kr, Xe, and Rn) has been performed. Potential energy curves based on ab initio [MP2, MP4, QCISD, and CCSD(T)] and DFT functionals (M06‐2X and CAM‐B3LYP) were obtained for ionic and neutral AuXe complexes. Dissociation energies indicate that neutral metals have the lowest and cationic metals have the highest affinities for interaction with rare gas atoms. For the same metals, there is a continuous increase in dissociation energies (De) from Kr to Rn. The nature of bonding and the trend of De and equilibrium bond lengths (Re) have been interpreted by means of quantum theory of atoms in molecules, natural bond orbital, and energy decomposition analysis. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
Various p-substituted benzyl p-hydroxyphenyl methyl sulfonium salts ( 2 ) were synthesized and their initiator activities were evaluated in bulk polymerization of glycidyl phenyl ether (PGE). The order of the activity was found to be 2b (X = CH3) > 2a (X = H) ≈ 2c (X = Cl) > 2d (X = NO2), indicating that the introduction of an electron-donating group enhanced the activity. In Hammett's plots, the logarithm of the ratio of the polymerization rates (log kx/kH) was correlated with σ+ρ better than with σp and a negative ρ+ value (-1.18) was obtained. Reaction of 2a with benzyl mercaptan mainly gave dibenzyl sulfide and p-hydroxyphenyl methyl sulfide. The obtained results seemed to demonstrate that the OH group of the aryl group yielded no proton as initiator for the polymerization, whereas the benzyl group caused the polymerization, which was initiated by the corresponding benzyl cation formed by C? S bond cleavage. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
Photochemistry of Conjugated γ,δ-Epoxyenones: The Influence of a Hydroxy Substituent in ?-Position On 1n, π*- or 1π,π*-excitation (λ ≥ 347 or λ=254 nm), the ?-hydroxy-γ;,δ-epoxyenone 8 undergoes fission of the C(γ)? O bond followed by the cleavage of the C(δ)-C(?) bond. This hitherto unknown sequence of reactions is evidenced by the structure determination of the new type products 10–17 and 25 , including a synthetic proof for 12 and the X-ray analysis of 11 (X-ray data: triclinic P1; a=7,386(2), b=8,904(4), c=9,684(5)Å; α=82,29(4)°, β=74,46(3)°, γ=82,29(3)°; Z=2). The selective 1π,π*-excitation also induces competitive C(γ)-C(δ) bond cleavage to yield the bicyclic acetal 18 and a ketonium-ylide intermediate a , which photochemically forms a carbene b giving the allene 19 and the cyclopropene 20 . On 1n,π*-excitation of the acetate 9 the initial C(γ)-O bond fission is, in contrast to the behaviour of the corresponding alcohol 8 , followed by a 1,2-methyl shift affording (E/Z)- 28 or by a cyclization-autoxidation process yielding the lactone 29 .  相似文献   

17.
The structural, electronic, and magnetic properties of the stoichiometric (001) surface of double perovskite Sr2FeMoO6 have been studied by using a 10‐layer FeMoO4 and SrO terminated (001)‐oriented slab model and the first‐principles projector augmented wave potential within the generalized gradient approximation as well as taking into account the on‐site Coulomb repulsive (U = 2.0 eV for Fe and 1.0 eV for Mo). An outwards relaxation is observed for several layers near surface, and the accompanying layer rumpling has a decrease tend from surface layer to inner layer. Along Fe–O–Mo–O–Fe or Mo–O–Fe–O–Mo chains, the oxygen atom is closer to the adjacent Mo atom than to the adjacent Fe atom. In FeO6 or MoO6 octahedra, the two axial TM?O bonds are not equal, and especially, the surface dangling bond makes the remaining one axial TM?O bond slightly shorter than four equally equatorial TM?O bonds. The half‐metallic nature and a complete (100%) spin polarization character ensure the FeMoO4 and SrO terminated (001)‐oriented slab of double perovskite Sr2FeMoO6 a potential application in spintronics devices. The Fe+3 and Mo+5 ions are still in the (3d5, S = 5/2) and (4d1, S = 1/2) states with positive and negative magnetic moments respectively and thus antiferromagnetic coupling via oxygen between them. There is no direct interaction between two nearest Fe–Fe or Mo–Mo pairs, whereas the hybridizations between Fe 3d and 4s, O 2s and 2p, as well as Mo 4d, 5s and 5p orbitals are fairly significant. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

18.
Proton magnetic resonance studies of the pyridine and γ-picoline complexes of trimethyl-and triethylthioboranes provide strong evidence that the pπdπ bonding in the boron-sulphur bond is much weaker than the pπ? pπ bonding in the boron-oxygen bond.  相似文献   

19.
The kinetics of α-methylene-γ-butyrolactone (α-MBL) homopolymerization was investigated in N,N-dimethylformamide (DMF) with azobis(isobutyronitrile) as initiator. The rate of polymerization (Rp) was expresed by Rp = k[AIBN]0.54[α-MBL]1.1 and the overall activation energy was calculated as 76.1 kJ/mol. Kinetic constants for α-MBL polymerization were obtained as follows: kp/kt1/2 = 0.161 L1/2 mol?1/2·s?1/2; 2fkd = 2.18 × 10?5 s?1. The relative reactivity ratios of α-MBL(M2) copolymerization with styrene (r1 = 0.14, r2 = 0.87) were obtained. Applying the Qe scheme led to Q = 2.2 and e = 0.65. These Q and e values for α-MBL are higher than those for MMA  相似文献   

20.
The optical pumping method of alkali molecules by atom—molecule exchange collisions is applied to obtain the magnetic shielding difference σ(Na) — σ(Na2) = (29 ± 16) × 10?6 between Na atoms and Na2 molecules and the scalar nuclear spin—spin coupling constants ds = (306 ± 30)s?1 of 23Na39K.  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