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.
Equilibrium swelling measurement was employed to correlate the Hildebrand (one-dimensional) solubility parameter (δ) of styrene butadiene rubber (SBR) with using different kinds of solvents. The δ was estimated to be a range from 17 MPa1/2 to 19 MPa1/2. It was attempted to calculate the three-dimensional solubility parameter (HSP) of SBR by inputting the swelling data into a professional software program, and the output results were δd = 18.0 MPa1/2, δp = 2.9 MPa1/2 and δH = 2.3 MPa1/2. A new Flory-Huggins interaction parameter (χHSP) between SBR and solvents was calculated by HSP values for the purpose of better predicting rubber swelling responses. The results turned out that the swelling ratio (q) increased with a decrease of χHSP value, showing a roughly linear relationship. Furthermore, a correlation index relating to χHSP value was introduced and mathematical fitted as a functional equation in this work. By using this equation, it is easy to characterize the relative interactions between SBR and each ingredient involved in the formulation and make a pre-screening of the candidate additives for the formula design. Taking advantage of this method one can response better to the challenges in selecting the most suitable additives and replacing the outdated ones by the new emerging ones, bio-sourcing with environmentally friendly ones in particular.  相似文献   

2.
In order to understand the swelling behavior of hydrogenated nitrile rubber (HNBR) more fully, the total solubility parameters (δt) of HNBR (Therban 2568) were determined by equilibrium swelling tests. Then, the swelling responses were analysed by a computer program to determine the Hansen three-dimensional solubility parameter (HSP). The HSP values – determined from lightly cured rubber samples – were estimated as δd = 18.4 (J/cm3)1/2, δp = 6.0 (J/cm3)1/2, δh = 4.5 (J/cm3)1/2 and δt = 19.9 (J/cm3)1/2. The energy difference (Ra) between HNBR and solvents or solvent mixtures have been calculated by their HSP values and proven to be useful for predicting the swelling behaviour of HNBR. The swelling volume decreases with increasing Ra values. Using blended solvents, a clear correlation between Ra values and the rubber swelling response was established. Thus, it may be possible to use the Hansen three-dimensional solubility parameters to predict swelling phenomena of cured rubber articles in mixed fluids such as bio-fuels or lubricants. Also, the HSP values may be used to predict the response of rubber seals or gaskets when replacing toxic or expensive fluids with more favorable environmentally friendly or inexpensive ones.  相似文献   

3.
4.
14/15N N.M.R. and 11B N.M.R. Data of Trifluoromethylthioamino-boranes with Natural Isotope Abundance (Part 2) 14/15N as well as 11B-NMR data for trifluoromethylthioabminoboranes of the types XnB[N(SCF3)2], with X = F, Cl, Br, N3, or NHSCF3, n = 0, 1 or 2, and Cl3?nB(NHSCF3)n with n = 1, 2 or 3, as well as for the amine-borne Me3NBCl2N(SCF3)2 and the cyclic borazene (CF3SNHBNH)3 are reported. These data are used, together with a qualitative analysis of the bonding situation based on observed rotational barriers and known structures, to analyse for B ← N back donation in these compounds. Relatively small variations in δ14/15N compared to those observed for alkylaminoboranes as well as large variations in δ11B are suggestive of small contributions only from back bonding. In addition the ?halogene like”? nature of the (CF3S)2N group is confirmed. For the series X2BN(SCF3)2 (X = F, Cl, Br on N3), XB[N(SCF3)2]2 (X = Cl, Br, N3 or N(SCF3)2) and Cln?3B(NHSCF3)n (n = 1, 2 or 3) a linear relationship for δ11B and δ14/15N is observed. It is furtheron demonstrated that hitherto known δ14/15N/11B correlations are valid only in case of strong B ← N back donation.  相似文献   

5.
The effect of the 2′-(3-trifluoromethyl-2,2,3-trifluorocyclobutyl)ethyl radical in fluorosiloxane on the thermodynamics of dissolution of various low-molecular-mass compounds in this polymer is studied via inverse gas chromatography. The activity coefficients at infinite dilution, Ω 1 , in the temperature range 20–100°C are determined for C6–C8 n-alkanes, cyclohexane, benzene, toluene, tetrachloromethane, trichloromethane, 2-butanone, and ethyl acetate. Flory-Huggins interaction parameters χ 12 are calculated. It is shown that hydrocarbons are poor solvents (Ω 1 > 6, χ 12 > 0.5) and that the studied fluorosiloxane is not inferior to commercial polymethyl(3,3,3-trifluoropropyl)siloxane with respect to stability in nonpolar liquid media in the range 20–70°C. Trichloromethane (above 50°C) and compounds containing a carbonyl group are found to be good solvents. The dispersion component of the solubility of the polymer, δ 2, is determined to be 14.0 (J/cm3)1/2 at 20°C and slightly lower, 13.5 (J/cm3)1/2, at 100°C. Possible causes of these low values of δ2 are discussed.  相似文献   

6.
Surface tension data can be used for estimating the solubility of polymers in liquids. By determining the apolar and the polar components of the surface tension of polymers and of solvents, the attractive free energy, δG 121, of a polymer (1) in a given solvent (2) can be established. By also taking into account the contactable surface area of two polymer molecules, immersed in a liquid, δG 121 can be expressed in units of kT. Solubility then is favored when -1.5 kT < δG 121 < 0 for apolar systems, and when -1.5 kT < δG 121 for polar systems. In polar solvents, hydrogen bonding can often increase δG 121 from <-1.5 kT to > + 1.5 kT. Positive values are frequently attained and this strongly shifts the behavior from insolubility to solubility. A number of proteins exemplify this behavior.  相似文献   

7.
Nitrogen-rich compounds involving the cyclo-pentazole anion (cyclo-N5) have attracted extensive attention due to higher energy release and environmental friendliness than traditional high energy density materials (HEDMs). However, the synthesis of stable HEDMs with cyclo-N5 is still a challenge. In this study, the effect of nine solvents on the geometrical and electronic structures and solvation energies of Zn(N5)2, one of the recently synthesized nitrogen-rich compounds, was studied using the density functional theory and the polarized continuum model. The results indicated an increase in the stability of Zn(N5)2 in the solution phase compared to the vacuum phase, and the stability of Zn(N5)2 increases with increasing dielectric constants. The energy gap of frontier molecular orbitals and the absolute value of total energy in water are the largest, revealing that Zn(N5)2 is more stable in water than in other solvents. To understand the stabilization mechanism of Zn(N5)2 by water, further studies were performed with the natural bond orbital (NBO) analysis and the quantum theory of atoms in molecules (QTAIM) analysis using the explicit solvent model. The charge transfer and the hydrogen bonds are observed between Zn(N5)2 and water, which are beneficial to improvement of the stability of Zn(N5)2. This may indicate the solvents that have strong interactions with the cyclo-N5 candidate may improve the possibility of success of synthesis.  相似文献   

8.
9.
The electrochemical performances of the α-, γ-, and δ-MnO2 with different crystallographic structures were systematically investigated in 0.5 mol/L Li2SO4, 0.5 mol/L Na2SO4, 1 mol/L Ca(NO3)2, and 1 mol/L?Mg(NO3)2 electrolytes. The results showed that the electrochemical performances of the manganese dioxides depended strongly on the crystallographic structures of MnO2 as well as the cation in the electrolytes. Because the δ-MnO2 consists with layers of structure and the interlayer separation is 7 Å, which is suitable for insertion/extraction of some alkaline and alkaline–earth cations, the δ-MnO2 electrode showed the higher specific capacitance than that of α-MnO2 and γ-MnO2. We also found that the α-, γ-, and δ-MnO2 electrodes in the Mg(NO3)2 electrolyte showed a higher specific capacitance, while all the α-, γ-, and δ-MnO2 electrodes in the Li2SO4 electrolyte exhibited a better cycle life. The reason for the different behavior of Li+ and Mg2+ during the charge/discharge process can be ascribed to the charge effect of the cations in the electrolytes. The ex situ X-ray diffraction (XRD) and long-time cyclic voltammogram measurements were used to systematically study the energy storage mechanism of MnO2-based electrodes. A progressive crystallinity loss of the materials is also observed upon potential cycling at the oxidized states. A reasonable charge/discharge mechanism is proposed in this work.  相似文献   

10.
Preparation and Spectroscopic Characterization of the Pure Bondisomers [OsCl5(NCS)]2? and [OsCl5(SCN)]2? The oxidation of [OsCl5I]2? with (SCN)2 in CH2Cl2 yields the bondisomers [OsCl5(NCS)]2? and [OsCl5(SCN)]2?, which are isolated as pure compounds by ion exchange chromatography on DEAE-Cellulose. Only the salts of the N-isomer show significant shifts in the vibrational and electronic spectra caused by polarization of the terminal S depending on the size of the cations and the polarity of the solvents. In the IR and Raman spectra νCN(S), νCS(N) and δNCS are found at higher wave numbers than νCN(N), νCS(S) and δSCN. In the optical spectrum of the red [OsCl5(SCN)]2? the charge-transfer S→Os is nearly constant at 538 nm, but the N→Os transition of the yellow to violet coloured N-isomer shifts from 480 nm in organic solvents or in presence of large alkylammonium cations to 516 nm in aqueous solution and to 544 nm in the solid Cs-salt. The optical electronegativities are calculated to χopt(–SCN) = 2.6 and χopt(–NCS) = 2.6–2.8. According to spinorbit coupling and to lowered symmetry (C4v) the splitted intraconfigurational transitions are observed at 10 K as weak peaks in the regions 600, 1000 and 2000 nm. The O? O transitions are calculated from the vibrational fine structure. The lowest level of both isomers is confirmed by peaks in the electronic raman spectra. With the parameters ζ(OsIV) = 3200 cm?1 and B(? SCN) = 316 cm?1 or B(? NCS) = 288 cm?1 there is a good fit of calculated and experimental data, resulting in the nephelauxetic series: F? > CI? > SCN? > Br? > NCS? > I?.  相似文献   

11.
The 29Si-NMR chemical shifts δ(29Si) of (CH3)4?nSiXn compounds and some 13C-NMR chemical shifts δ(13C) of analogous carbon compounds are discussed by means of relative paramagnetic screening constants σ*, calculated by a simplified model. In this model only the Si(3P)- and C(2P)-orbitals are considered; for the calculations, the electronegativities of Si, C and the X-substituents and a single empirical parameter are necessary. The calculated values of σ* are in good agreement with the change of the chemical shifts which are observed for the (CH3)4?nMXn compounds with different X and n. These results clearly show that δ(29Si) and δ(13C) depend primarily on the σ-charge of the Si- and C-atom, and that (P? d)π-interactions on the Si-atom are of minor importance.  相似文献   

12.
We measured the 15N-, 1H-, and 13C-NMR chemical shifts for a series of aromatic diamines and aromatic tetracarboxylic dianhydrides dissolved in DMSO-d6, and discuss the relationships between these chemical shifts and the rate constants of acylation (k) as well as such electronic-property-related parameters such as ionization potential (IP), electronic affinity (EA), and the energy ε of the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO). The 15N chemical shifts of the amino group of diamines (δN) depend monotonically on the logarithm of k (log k) and on IP. We inferred the reactivities of diamines whose acylation rates have not been measured from their δN, and we propose an arrangement of diamines in the order of their reactivity. The 1H chemical shift of amino hydrogens (δH) and the 13C chemical shift of carbons bonded to nitrogen (δC) are roughly proportional to δN, but these shifts are not as closely correlated with log k and IP. Although the 13C chemical shifts of the carbonyl carbon of dianhydrides (δC,) varies much less than the δC and δN of diamines, δC, can be an index of acylation reactivity for dianhydrides because it is closely correlated with εLUMO. These facts indicate that the chemical shifts of diamines and dianhydrides are displaced according to their electron-donor and electron-acceptor properties, and that these chemical shifts can be used as indices of the electronic properties of monomers. Changes in reactivity caused by the introduction of trifluoromethyl groups into diamines and dianhydrides are inferred from the displacements of δN and δC © 1992 John Wiley & Sons, Inc.  相似文献   

13.
The decay processes of the lowest excited singlet and triplet states of five methylated angelicins (4,6,4′-trimethyl-angelicin, MA, and four methylated thioangelicins, MTA; see Scheme 1) were investigated in live solvents by stationary and pulsed fluorometric and flash photolytic techniques. In particular, the solvent effects on absorption, fluorescence, quantum yields of fluorescence (φF) and triplet formation (φT), lifetimes of fluorescence (τF) and the triplet state (τT) and the quantum yields of singlet oxygen production (φΔ) were investigated. Semiempirical (ZINDO/S-CI) calculations were carried out to obtain information (transition probabilities and nature) on the lowest excited singlet and triplet states. The quantum mechanical calculations and the solvent effect on the photophysical properties showed that the lowest excited singlet state (S1) is a partially allowed π,π* state, while the close-lying S2 state is n,π* in nature. The efficiencies of fluorescence, S1→T1 intersystem crossing (ISC) and S1→ S0 internal conversion (IC) strongly depend on the energy gap between S1, and S2 and are explained in terms of the so-called proximity effect. In fact, for MA in cyclohexane, only the S1→ S0 internal conversion is operative, while in acetonitrile and ethanol, where the n.π* state is shifted to higher energy, the efficiencies of fluorescence and ISC increase significantly. The energy gap between S1 and S2 increases in MTA, where the furanic oxygen is replaced by a sulfur atom. Consequently, the solvent effect on the photophysical parameters of MTA is less marked than for MA; e.g. fluorescence and triplet-triplet absorption are also detectable in the nonpolar cyclohexane. The lowest excited singlet state of molecular oxygen O2(1Dg) was produced efficiently in polar solvents by energy transfer from the T1 state of MA and MTA.  相似文献   

14.
Preparation and characterization of bondisomeric bromorhodanorhenates(IV) The new compounds [ReBr5(SCN)]2?, [ReBr5(NCS)]2?, cis/tr.-[ReBr4(NCS)(SCN)]2?, cis-[ReBr4(NCS)2]2?, mer-[ReBr3(NCS)3]2? are prepared from [ReBr6]2? by ligand exchange with NaSCN, KSCN, or (SCN)2 in organic solvents and isolated by ion exchange chromatography on DEAE cellulose. The bondisomers are significantly distinguished by the frequencies of inner ligand vibrations: vCN(S) > vCN(N), vCS(N) > vCS(S), δNCS δSCN. The electronic absorption spectra measured at 10 K exhibit in the region 5700 to 15300 cm?1 all intraconfigurational transitions (t2g3) splitted into 8 Kramers doublets by lowered symmetry (C4v, C2v, Cs) and spin orbit coupling. The O–O-transitions are deduced form vibrational fine structure and confirmed by electronic Raman bands in some cases. The magnetic moments are in the range of 3.0 to 3.9 B.M.  相似文献   

15.
Correlations among the three components, δd2 (dispersion), δh2 (hydrogen‐bonding), and δp2 (polar) that make up the Hansen solubility parameter equation, δo2 = δd2 + δh2 + δp2, have been analyzed for a large number of organic solvents. A relationship is found that enables δh and δp to be estimated if δo and δd are known. This relationship is applied to a variety of common polymers and remarkably good agreement is obtained with tabulated values for δh and δp. Additional correlations are found that can be expressed in approximate functional form. The analysis also reveals relationships, expressed as inequalities, among the parameters that limit their range of possible values. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 4337–4343, 2004  相似文献   

16.
In this article, the structural, electronic, and spectroscopic properties of osmabenzyne Os{≡CC(SiH3)=C(CH3)C(SiH3)=CH}Cl2(PH3)2 are explored in the gas phase and five solvents. The effects of solvents on the structural parameters, frontier orbital energies, and spectroscopic parameters of the complex are elucidated using the polarizable continuum model. The wavenumbers of selected IR-active vibrations in different solvents are obtained and correlated with the Kirkwood–Bauer–Magat equation. In addition, thermodynamic parameters of solvation are calculated for the complex. 1H and 13C NMR chemical shifts are estimated using the gauge-invariant atomic orbital method.  相似文献   

17.
The solubility behaviors of poly(sulfonyldiphenylene phenylphosphonate) (PSPPP), a very efficient flame retardant for poly(ethylene terephthalate) (PET), in more than 50 solvents were examined. Its solubility parameters (δ) were determined by the intrinsic viscosity and turbidic titration methods. The two methods obtained consistent results, δ = 21.0–21.6 J1/2/cm3/2 and δ = 21.0 J1/2/cm3/2, and the three‐dimensional solubility parameters were δd = 18.9 J1/2/cm3/2, δp = 8.8 J1/2/cm3/2, and δh = 5.9 J1/2/cm3/2. The miscibility of PSPPP with PET was estimated by the calculation of the heats of mixing, which were related to the difference between the solubility parameters of PSPPP and PET. Fourier transform infrared was used to examine the interactions between PSPPP and PET macromolecules, which were the internal factors of polymer–polymer miscibility. The results showed that PSPPP and PET were miscible within a very wide composition range, especially with less than 15 wt % PSPPP, a composition of interest for the preparation of flame‐retardant PET. © 2003 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 41: 2296–2301, 2003  相似文献   

18.
The solubilities of sulfadiazine (SD), sulfamerazine (SMR) and sulfamethazine (SMT) in some 1-propanol + water co-solvent mixtures were measured at five temperatures from 293.15 to 313.15 K over the polarity range provided by the aqueous solvent mixtures. The mole fraction solubility of all these sulfonamides was maximal in the 0.80 mass fraction of 1-propanol solvent mixture (δ solv = 28.3 MPa1/2) and minimal in water (δ = 47.8 MPa1/2) at all temperatures studied. The apparent thermodynamic functions Gibbs energy, enthalpy, and entropy of solution were obtained from these solubility data by using the van’t Hoff and Gibbs equations. Apparent thermodynamic quantities of mixing were also calculated by using the ideal solubilities reported in the literature. Nonlinear enthalpy–entropy relationships were observed for these drugs in the plots of enthalpy versus Gibbs energy of mixing. The plot of ?mix H° versus ?mix G° shows different trends according to the slopes obtained when the mixture compositions change. Accordingly, the mechanism for the solution process of SD and SMT in water-rich mixtures is enthalpy driven, whereas it is entropy driven for SMR. In a different way, in 1-propanol-rich mixtures the mechanism is enthalpy driven for SD and SMR and entropy driven for SMT. Ultimately, in almost all of the intermediate compositions, the mechanism is enthalpy driven. Nevertheless, the molecular events involved in the solution processes remain unclear.  相似文献   

19.
The temperatuure dependence of the vicinal OCH2CH2 coupling constant in the six-membered ring δ-valerolactone suggests the presence of at least two conformers of unequal energy. Analysis of this temperature dependence by the Wood-Fickett-Kirkwood method in six solvents produces free energy differences in the narrow range 0.8–1.4 kcal mol?1, with a mean of 1.0 kcal mol?1 and no apparent dependence on solvent polarity. Thus the two conformers, probably the half-chair and the classic boat, have similar polarities. The analogous four- and five-membered lactones show little or no temperature dependence for their analogous coupling constants, in agreement with the presence of a single conformation for these rings.  相似文献   

20.
In order to examine 13C-SCS of substituted benzoic acids, chemical shifts of the acid form (I) and the dissociated form (II) have been obtained separately. Single substituent parameters, σ0 or σ+ are correlated with the shifts for the carboxyl (δco) or ipso carbons (δipso), respectively. Among the available five equations which are developed for the analysis with dual (or divided) substituent parameters (DSP), the Swain-Lupton equation (eqn 3) and the Taft-Swain-Lupton equation (eqn 4) give much better correlations, not only for δco and δipso but also for the results for ring carbons (C(2), C(5), C(6)), except for those attached to or neighboured by substituents. It is concluded that the SCS of aromatic compounds are best analyzed with substituent parameters derived from reactions or equilibria on the basis of linear free energy relationships.  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