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 78 毫秒
1.
A combined study of molecular dynamics (MD) simulation, experimental, and linear regression analysis method is presented for hindered phenol of 3,9‐bis[1,1‐dimethyl‐2‐{b‐(3‐tertbutyl‐4‐hydroxy‐5‐methylphenyl)propionyloxy}ethyl]‐2,4,8,10‐tetraoxaspiro‐[5,5]‐undecane (AO‐80)/nitrile‐butadiene rubber/linear phenolic resin (AO‐80/NBR/PR) composites with different AO‐80 contents to quantitatively establish the relations between microstructure and damping performance. The number of hydrogen bonds (NHBs), the fractional free volume (FFV), and the binding energy (Ebinding) of AO‐80/NBR/PR composites with different AO‐80 content are calculated by MD simulation from the microscopic scale. Damping parameters, including the loss factor peak (tan δmax) and the loss peak area (TA) (tan δ > 0.3), are obtained by dynamic mechanical analysis from macroscopic scale. The quantitative relationships between microstructure parameters (NHBs, Ebinding, and FFV) and macroscopic damping properties (tan δmax and TA) are obtained by linear regression analysis. This research is expected to provide a theoretical guidance for improving the damping performance of rubber‐based organic hybrid composites.  相似文献   

2.
Φ80 mm‐diameter, highly <110>‐oriented β‐SiC wafers were ultra‐fast fabricated via halide chemical vapor deposition (CVD) using tetrachlorosilane (SiCl4) and methane (CH4) as precursors. The effects of deposition temperature (Tdep) and total pressure (Ptot) on the orientations, microstructures, and deposition rate (Rdep) were investigated. Rdep dramatically increased with increasing Tdep where maximum Rdep was 930 μm/h at Tdep = 1823 K and Ptot = 4 kPa, leading to a maximum of 1.9 mm in thickness in 2 h deposition. The <110>‐oriented β‐SiC was obtained at Tdep > 1773 K and Ptot = 1–4 kPa. Growth mechanism of <110>‐oriented β‐SiC has also been discussed under consideration of crystallographic planes, surface energy, and surface morphology.  相似文献   

3.
The dynamic hysteresis scaling behaviors of Nb‐doped Pb(Zr0.52Ti0.48)O3 ceramics have been investigated as a function of electric field amplitude (E0) and frequency (f) at different temperatures (T). The loop area <A> of saturated loops is found to follow various power laws as <A> ∝ E00.3065 at fixed f and <A> ∝ f 0.0120 at fixed E0. Furthermore, the linear scaling relation <A> k3 fαE0β + b3 is estimated under various temperatures. The exponents α (=0.01) and β (=0.10) are T‐independent, whereas the slopes k3 and y‐intercepts b3 are T‐dependent because the increasing temperature in the same phase range only decreases the threshold field of the reversal rather than change the dynamic reversal process.  相似文献   

4.
Highly oriented <111> and <110> β‐SiC films were prepared on Si(100) single crystal substrates by laser chemical vapor deposition using a diode laser (wavelength = 808 nm) and HMDS (Si(CH3)3–Si(CH3)3) as a precursor. The effects of laser power (PL), total pressure (Ptot), and deposition temperature (Tdep) on the orientation, microstructure, and deposition rate (Rdep) were investigated. The orientation of the β‐SiC films changed from <111> to random to <110> with increasing PL and Ptot. The <111>‐, randomly, and <110>‐oriented β‐SiC films exhibited dense, cauliflower‐like, and cone‐like microstructures, respectively. Stacking faults were observed in the <111>‐ and <110>‐oriented films, and aligned parallel to the (111) plane in the <111>‐oriented film, whereas they were perpendicular to the (110) plane in the <110>‐oriented film. The highest Rdep of the <111>‐oriented β‐SiC film was 200 μm/h at Ptot = 200 Pa and Tdep = 1420 K, whereas that of the <110>‐oriented film was 3600 μm/h at Ptot = 600 Pa and Tdep = 1605 K.  相似文献   

5.
The excess molar volumes (VE) and the deviations in molar refraction (ΔR) at 298.15 K were determined for the binary systems {diisopropyl ether (DIPE) + 1‐propanol}, {Tert‐amyl methyl ether (TAME) + methanol}, {DIPE + trihexyltetradecylphosphonium bis(2,4,4‐trimethylpentyl)phosphinate ([P666,14][TMPP])}, {TAME + [P666,14][TMPP]}, {methanol + [P666,14][TMPP]} and {1‐propanol + [P666,14][TMPP]} using a digital vibrating‐tube densimeter and a precision digital refractometer. The VE and ΔR were correlated with the Redlich–Kister equation for binary systems. In addition, the ternary VE and ΔR data at 298.15 K were predicted for the ternary systems {DIPE + 1‐propanol + [P666,14][TMPP]} and {TAME + methanol + [P666,14][TMPP]} by using the binary contribution model of Radojkovi? with correlated sub‐binary Redlich–Kister parameters. © 2011 Canadian Society for Chemical Engineering  相似文献   

6.
We have prepared polycrystalline R2NiMnO6 (R = Nd, Eu, Gd, Dy, and Y) samples by conventional solid‐state reaction and all the samples have shown monoclinic structure with P21/n space group. With the decrease in rare‐earth ionic size (<rR>), the <Ni–O–Mn> bond angle decreases, correspondingly a decrease in ferromagnetic (FM) Curie temperature is noticed. In the dielectric measurement, the dielectric anomaly shifts to high temperature with the decrease in the <rR> and shows no correlation with the FM Curie temperature and hints the absence of apparent magnetodielectric (MD) coupling. Appearance of multiple relaxations in the dielectric study suggests the electrical heterogeneity of the system. The dielectric/impedance analysis has revealed a close correlation between dc conductivity and the dielectrics; in fact, both dc resistivity and the grain relaxations follow the variable range hopping mechanism. The thermal activation of charge carriers at the grain boundary led to Maxwell–Wagner interfacial polarization. Finally, dielectric study under magnetic field showed no effect, it implies that not only the intrinsic MD is absent, but also the extrinsic MD due to the lack of magnetoresistance.  相似文献   

7.
Structural, electrical, and up‐conversion (UC) properties of Na0.5Bi4.5‐xErxTi4O15 (NBT–xEr3+) (0.00 < x < 0.40) ceramics have been studied. All the ceramic samples possessed a single‐phase orthorhombic structure. The unit cell volume, the lattice parameters a, b, and c, and orthorhombic distortion analyzed on the basis of Rietveld refinement were observed to decrease with increasing Er3+ contents (x). The average values of grain size were found to slightly decrease with increasing x. Raman spectroscopy revealed that (Bi2O2)2+ layers remained unaffected in the modified compositions, and Er3+ substitution for Bi3+ occurred predominantly at the A‐site in the perovskite blocks causing the cationic disorder and a slight decrease in the TiO6 octahedral distortion. NBT–xEr3+ ceramic with x = 0.20 achieved the optimized photoluminescence. The relative intensity of green and red UC emissions could be tuned by changing Er3+ ions concentrations. Ferroelectric measurements revealed that Er3+ doping led to reduction in 2Pr and 2Ec. However, dielectric measurements showed that the incorporation of Er3+ ions increased Tc with simultaneously lowered tanδ at high temperature, implying that this ceramics can be suitable for high‐temperature sensor applications.  相似文献   

8.
Dynamic hysteresis of Nb‐doped Pb(Zr1?xTix)O3 (PZT, 0.20 ≤ ≤ 0.60) ceramics were studied systemically at different field (E) and frequency (f). The hysteresis loops were strongly dependent on E and f. The measured coercive fields (Ec0) were far lower than the calculated values based on the Landau–Ginzburg theory, and increased dramatically from the rhombohedral phase to the tetragonal phase while had less variation with composition in the same phase. With increasing E or f for each composition, three types of loops were observed: linear, minor, and saturated loops. The cross fields (E0), cross polarizations (P0), and hysteresis areas (<A>) showed different variation regularities with f. Similar varying curves were observed for all PZT ceramic samples by normalizing E0, P0, and <A> with E/Ec0, which indicated the same polarization switching for different domain structures. Further analyses revealed that the switching processes can be divided into three stages: space charge polarization, domain switching, and steady state. The first and second stage occurred at ~0.5 and ~1.5 E/Ec0, respectively. These results would be very helpful to further understand the polarization switching of ferroelectric ceramics.  相似文献   

9.
Linear polyurethanes (PUs) were prepared in solution in two steps. In the first step, toluene‐diisocyanate (TDI) 80:20 (T) was reacted with a short‐chain poly(oxypropylene) diol (V) to obtain prepolymers characterized by various ratios of R = [T]°/[V]°. In the second step, the prepolymers were reacted with the extender 1,4‐cyclohexanedimethanol (E) with various extension ratios RE = (R ? 1)[V]°/[E]°. The PU properties were analyzed by various techniques. In particular, the molecular weights determined by gel permeation chromatography (GPC), the thermal properties such as glass transition temperature Tg and specific heat variation ΔCp measured by differential scanning calorimetry (DSC), the chemical composition of the PUs and the various types of hydrogen bonds present in the polymers evidenced by Fourier‐transform infrared (FTIR) spectroscopy, the monophasic structure evidenced by small‐angle X‐ray scattering (SAXS), and the existence of only one Tg confirmed that these PUs were linear, amorphous and monophasic. Positron annihilation lifetime spectroscopy (PALS) analysis showed that the mean volume of the nanoholes increased with increasing temperature, but was not dependent on the composition, as expected for a monophasic system. A stoichiometric semi‐empirical model was proposed that relates the PU blocks' micro‐composition to the R and RE macro‐parameters, chosen for the synthesis. The polymer assumes various expressions of the general formula X? [(TV)k? (TE)n]m? X for different values of the R and RE ratios. The micro‐parameters k and m have a direct connection with the experimental mean molecular weights of the prepolymer and the polymer, respectively: n depends only on R. The model could foresee the density of hydrogen bonds and distinguished the bonds connected to either V or E, which could be shown by FTIR analysis. This paper shows that, when using stereo‐irregular diols and blends of 2,4‐ and 2,6‐TDI, non‐stereoregular PUs are obtained. If low‐molecular‐weight diols are used and R < 3.3, it is quite improbable that the PU blocks separate into macrophases and therefore monophasic amorphous PUs are obtained. Monophasic PUs can be useful for applications such as in the field of membrane gas and vapour separation. Copyright © 2005 Society of Chemical Industry  相似文献   

10.
A relatively simple procedure is described to calculate the coordinates, temperature (Taz), pressure (Paz), and composition (x1az), of an azeotrope (for either an isothermal or isobaric binary system at low pressure and with a single or double homogeneous azeotrope) from an expression for the liquid‐phase excess molar Gibbs function (gE). General results are based on (?gE/?x1)az = RTaz ln[p2*(Taz)/p1*(Taz)], and Paz = (γ1p1*)az = (γ2p2*)az, where R is the gas constant, p* is saturation vapour pressure, and γ is liquid‐phase activity coefficient. Specific results are given for the Redlich‐Kister, van Laar, Wilson, and NRTL equations. Numerical examples are provided for both an isothermal and an isobaric system. The procedure provides a means to obtain azeotrope coordinates that are consistent with a gE expression obtained either from experimental data or from a model. It is applicable to polyazeotropy, whereas the criteria given previously are generally not applicable to polyazeotropy.  相似文献   

11.
Spectroscopic and physical properties of Nd3+-doped alkali lead borate glasses of type 20R 2O · 30PbO · 49.5B2O3 · 0.5Nd2O3 (R = Li and K) and alkaline-earth lead borate glasses 20RO · 30PbO · 49.5B2O3 · 0.5Nd2O3 (R = Ca, Ba, and Pb) have been investigated. Optical absorption spectra have been used to determine the Slater-Condon (F2, F4, and F6), spin orbit ξ4f, and Racah parameters (E1, E2, and E3). The oscillator strengths and the intensity parameters Ω2, Ω4, and Ω6 have been determined by the Judd-Ofelt theory, which, in turn, provide the radiative transition probability (A), total transition probability (A T ), radiative lifetime (τ R ), and branching ratio (β R , %) for the fluorescent levels. The lasing efficiency of the prepared glasses has been characterized by the spectroscopic quality factor (Ω46), the value of which is in the range 0.2–1.5, typical of Nd3+ in different laser hosts. A red shift of the peak wavelength is observed upon addition of alkali or alkaline-earth oxides to the lead borate glass. A higher value of the W2 parameter for potassium-doped glass indicates a higher covalency for this glass matrix. The relative intensity of the peaks 4I9/24F7/2, 4S3/2 has also been studied. The text was submitted by the authors in English.  相似文献   

12.
《Ceramics International》2020,46(7):8675-8681
The dielectric properties and bipolar polarization-electric field (P-E) and strain-electric field (S-E) dynamic hysteresis of a relaxor [001]c 0.73Pb(Mg1/3Nb2/3)O3-0.27PbTiO3 (PMN-0.27PT) single crystal were investigated to reveal more details of the temperature-induced phase transitions. Different linear scaling relations for ferroelectric hysteresis area <A>, coercive field Ec, saturation polarization Ps and remnant polarization Pr versus temperature τ were measured in different temperature regions. For each measurement frequency f, all hysteresis parameters were found to decrease linearly with temperature in the temperature range of the single rhombohedral (R) phase or tetragonal (T) phase, and the rate of decrease in the T phase was observed to be much larger than the corresponding rate in the R phase. In the temperature range near the R-T phase transition, the exponent α in the power law <A>∝f α for the R phase was found to be smaller than that for the T phase, and the magnitude of α depended strongly on temperature when the crystal was in the R-T coexisting phase state. Our experimental and theoretical results indicate that the difference in the activation energy and dipole moment in the R and T phases may lead to the observed discrepancy for the P-E and S-E hysteresis behaviour in different temperature regions.  相似文献   

13.
Polymer matrix composites based on brominated epoxy as the matrix and aluminum nitride (AlN) particles as the filler were prepared. The influences of the size, content, and size distribution of AlN on the thermomechanical properties, including the glass‐transition temperature (Tg), coefficient of thermal expansion (CTE), dynamic storage modulus (E′), dynamic loss modulus (E″), and loss factor (tan δ), of the composites were investigated by thermomechanical analysis and dynamic mechanical analysis. There was a total change trend for Tg; that is, Tg of the composites containing nano‐aluminum nitride (nano‐AlN; 50 nm) was lower than that of the micro‐aluminum nitride (micro‐AlN; 2.3 μm) filled composites, especially at high nano‐AlN contents. The Tg depression of the composites containing nano‐AlN was related to the aggregation of nano‐AlN and voids in the composites. On the other hand, the crosslink density of the epoxy matrix decreased for nano‐AlN‐filled composites, which also resulted in a Tg depression. The results also show that E′ and E″ increased, whereas tan δ and CTE of the composites decreased, with increasing the AlN content or increasing nano‐AlN fraction at the same AlN content. These results indicate that increasing the interfacial areas between AlN and the epoxy matrix effectively enhanced the dynamic modulus and decreased CTE. In addition, at a fixed AlN content of 10 wt %, a low E′ of pre‐Tg (before Tg temperature) and high Tg were observed at the smaller weight ratio of nano‐AlN when combinations of nano‐AlN plus micro‐AlN were used as the filler. This may have been related to the best packing efficiency at that weight ratio when the bimodal filler was used. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

14.
Physical properties of mica-reinforced linear (LPB) and star-branched polybutadiene (SPB) composites have been studied with special reference to the effect of silane coupling agent, degree of crosslinking, and degree of mica loading, as well as the molecular architecture of polybutadienes. Tensile properties increase and swelling decreases with addition of mica for linear polybutadiene. Star-branched polybutadiene shows a reverse behavior, especially beyond 5% of mica. The improvements in mechanical properties are more pronounced in the case of silane-treated mica compositions of both type of polybutadienes. Dynamic mechanical spectra were obtained for linear and star-branched polybutadienes. Effects of mica loading and silane treatment on dynamic moduli are discussed. Dynamic mechanical moduli (E′, E′) of composites increase with increasing mica content for linear polybutadiene but decrease for star-branched polybutadiene beyond certain mica loadings. Effective damping regions were determined in terms of frequency and temperature. The glass transition temperature (Tg) increased slightly, and the damping peak (tan δ) broadened due to the rubber—filler interaction, especially after silane treatment for both polymers. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 66: 1943–1952, 1997  相似文献   

15.
Er‐doped 0.94Bi0.5Na0.5TiO3‐0.06BaTiO3 (BNT‐6BT: xEr, x is the molar ratio of Er3+ doping) lead‐free piezoceramics with = 0–0.02 were prepared and their multifunctional properties have been comprehensively investigated. Our results show that Er‐doping has significant effects on morphology of grain, photoluminescence, dielectric, and ferroelectric properties of the ceramics. At room temperature, the green (550 nm) and red (670 nm) emissions are enhanced by Er‐doping, reaching the strongest emission intensity when = 0.0075. The complex and composition‐dependent effects of electric poling on photoluminescence also have been measured. As for electrical properties, on the one hand, Er‐doping tends to flatten the dielectric constant‐temperature (εrT) curves, leading to temperature‐insensitive dielectric constant in a wide temperature range (50°C–300°C). On the other hand, Er‐doping significantly decreases the ferroelectric‐relaxor transition temperature (TF–R) and depolarization temperature (Td), with the TF–R decreasing from 76°C to 42°C for x = 0–0.02. As a result, significant composition‐dependent electrical features were found in ferroelectric and piezoelectric properties at room temperature. In general, piezoelectric and ferroelectric properties tend to become weaker, as confirmed by the composition‐dependent piezoelectric coefficient (d33), planar coupling factor (kp), and the shape of polarization‐electric field (PE), current‐electric field (J–E), bipolar/unipolar strain‐electric field (S–E) curves. Furthermore, to understand the relationship between the TF–R/Td and the electrical properties, the composition of = 0.0075 has been intensively studied. Our results indicate that the BNT‐6BT: xEr with appropriate Er‐doping may be a promising multifunctional material with integrated photoluminescence and electrical properties for practical applications.  相似文献   

16.
Experimental densities (ρ), ultrasonic speeds (u), and refractive indices (nD) of binary mixtures of dichloromethane (DCM) with acetone (ACT) and dimethylsulfoxide (DMSO) were measured over the whole composition range at T?=?298.15, 303.15, and 308.15?K. From the experimental data, excess molar volume (VE), deviations in isentropic compressibility (Δks), deviations in intermolecular free length (ΔLf), deviations in refractive index (ΔnD), and deviations in ultrasonic speed (Δu) were calculated. Moreover, the Benson–Kiyohara theory was applied to the binary mixtures to obtain the theoretical Δks values. The COSMO calculations depending on density functional theory were utilized to estimate the σ-profiles for the DCM, ACT, and DMSO. The interpreted σ-profile trends were found supportive with the experimental findings. Applicability of different empirical and semi-empirical relations of refractive index data were tested against the measured results, and good agreement has been obtained. The possible results of intermolecular molecular interactions among mixture components were interpreted.  相似文献   

17.
Summary Four narrow-distribution samples of sodium hyaluronate with weight-average molecular weights M w of 3.8 × 103 to 1.1 × 104 in 0.02 and 0.1 M aqueous NaCl at 25°C have been studied by small-angle X-ray scattering. Data for their z-average radii of gyration <S2>z 1/2 and particle scattering functions P(θ), together with previous <S2>z data from light scattering for high molecular weights, are compared with relevant theories for the wormlike chain with or without excluded volume, using the parameters estimated previously from intrinsic viscosity ([η]) data. It is shown that for M w lower than 1.1 × 104, <S2>z, P(θ), and [η] of the polysaccharide in the aqueous salts are all consistently explained by this model without excluded volume. Received: 8 December 1997/Accepted: 26 December 1997  相似文献   

18.
The kinetic equation describing the thermal decomposition reaction of NNHT obtained by TG‐DTG data, integral isoconversional non‐linear method and integral method of treating TG‐DTG curves is . The specific heat capacity (Cp) of NNHT was determined with the continuous Cp mode of the microcalorimeter. The equation of Cp (T) was obtained. The standard molar heat capacity of NNHT was 218.41 J mol−1 K−1 at 298.15 K. With the help of the onset temperature (Te) and maximum peak temperature (Tp) from the non‐isothermal DTG curves of NNHT at different heating rates (β), the apparent activation energy (EK and EO), and the pre‐exponential constant (AK) of the thermal decomposition reaction obtained by Kissinger’s method and Ozawa’s method, Cp obtained by microcalorimetry, density (ρ) and thermal conductivity (λ), the decomposition heat (Qd, taking half‐explosion heat), Zhang‐Hu‐Xie‐Li’s formula, Smith’s equation, Friedman’s formula, Bruckman‐Guillet’s formula, and Wang‐Du’s formulas, the values (Te0 and Tp0) of Te and Tp corresponding to β→0, thermal explosion temperature (Tbe and Tbp), adiabatic time‐to‐explosion (tTIad), 50 % drop height (H50) of impact sensitivity, critical temperature of hot‐spot initiation (Tcr), thermal sensitivity probability density function [S(T)] versus temperature (T) relation curves for spheroidic NNHT with radius of 1 m surrounded with ambient temperature of 300 K, peak temperature corresponding to the maximum value of S(T) versus T relation curve ( ), safety degree (SD), and critical ambient temperature(Tacr) of thermal explosion of NNHT are calculated. The following results of evaluating the thermal safety of NNHT are obtained: TSADT=Te0=453.34 K, TSADT=Tp0=454.86 K, Tbe=462.68 K, Tbp=467.22 K, tTIad=1.03 s, H50=17.69 cm, Tα=461.4 K. SD=72.74 %, PTE=27.26 %, and Tacr=321.96 K.  相似文献   

19.
Electropolymerization of O‐phenylenediamine (o‐PD) on Pt‐electrode from a deoxygenated aqueous acid medium was carried out using cyclic voltammetry technique. The kinetic parameters were calculated by means of electrochemical data. The experimentally obtained kinetic equation was RP,E = kE [monomer]1.19 [acid]1.23 [electrolyte]0.87 from the value of the anodic current density using cyclic voltammetry technique. The apparent activation energy (Ea) is found to be 28.34 kJ mol?1. The polymer films obtained have been characterized by X‐ray diffraction, elemental analysis, scanning electron microscopy, 1H‐NMR, 13C‐NMR, UV‐visible, and IR spectroscopy. The mechanism of the electrochemical polymerization reaction has been discussed. TGA is used to confirm the proposed structure and determination of the number of water molecules in the polymeric chain unit. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

20.
Several polymer binders based on 3,3‐bis(azidomethyl)oxetane (BAMO) were studied to explore the compatibility and interaction of the energetic binders with three common energetic oxidants. The compatibilities were studied by differential scanning calorimetry and ratings were obtained according to evaluated standards. The results showed that all the binders based on BAMO had good compatibility with cyclotrimethylenetrinitramine, cyclotetramethylenetetranitroamine and hexanitrohexazaiso‐wurtzitane. The work of adhesion (Wa) between binders and explosives was tested via measurement of contact angle and the results are in the following order: chain‐extended poly(3,3‐bis(azidomethyl)oxetane) (PBAMO) by isophorone diisocyanate (IPDI‐CE) with diethyl bis(hydroxymethyl) malonate (IPDI‐DBM‐CE) > chain‐extended PBAMO by IPDI‐CE > PBAMO. In addition, similar results were found in the binding energies reported by molecular dynamics, and the average values of Ebinding for the IPDI‐DBM‐CE system were larger than Ebinding for the other systems due to the formation of hydrogen bonds between –COOEt and –NO2, which improve the bonding abilities. © 2017 Society of Chemical Industry  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