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.
New lead‐free piezoelectric ceramics of 0.9BaTiO3–(0.1?x)(Bi0.5Na0.5)TiO3xBiMO3, M=Al and Ga, where x=0.00‐0.10, were fabricated by the solid‐state reaction technique. The effect of BiMO3 contents on the perovskite structure, phase transition, and dielectric, ferroelectric, and piezoelectric properties was investigated. X‐ray diffraction patterns showed that the ceramics exhibit a monophasic perovskite phase up to x=0.06, suggesting stabilized perovskite structures with B‐site aliovalent substitutions. Compositional‐dependent phase transitions were observed from tetragonal to pseudo‐cubic phase with increasing BiMO3 amounts. Al3+ ions were found to stabilize the transition temperature of the ceramics, while significantly decreasing transition temperature, and a change in the dielectric peak were found with an increasing amount of Ga3+. Regarding Al3+ substitution, the remanent polarization (Pr) values were found to decrease slightly with the Al3+ amount. With regard to Ga3+ substitution, Pr values decreased with the Ga3+ amount up to 0.06 and then increased slightly. The ceramics became softer with a higher degree of substitution according to the lower coercive field (Ec), when compared with 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics. Ceramics with a lower degree of substitution and tetragonal phase showed butterfly strain loops that correlated with normal ferroelectric behavior.  相似文献   

2.
Based on kinetic considerations, the following equation, connecting the zero‐shear viscosity of polymeric solutions with temperature and the molecular weight and concentration of the polymer was derived: RTln ηR = KBφMn /(1 + BφMn), where ηR is relative viscosity (i.e., the ratio of the solution viscosity to the solvent viscosity); K represents a change in enthalpy of viscous flow from a pure solvent to a pure polymer at the same temperature or from a polymer of low molecular weight (M) to one of higher molecular weight, and has the dimensions of energy (e.g., J/mol) because the ratio BφMn/(1 + BφMn) is dimensionless; φ is the volume or molar fraction of a polymer in solution (concentration units can be used in dilute solutions); B is a constant related to the stiffness of the chains of the polymer in a given solvent; and at BφMn >> 1, ln ηR = K/RT. The equation describes published data on the zero‐shear viscosity of four polar and nonpolar polymers in nine solvents with R2 > 0.98. This approach allows the use of solutions of moderate concentrations for the characterization of polymers and opens a way for a single‐point degree of polymerization (DP) determination of polymers at moderate concentrations if constants K, B, and n of the equation are known. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 85: 2064–2073, 2002  相似文献   

3.
Creep experiments have been applied to probe the zero‐shear viscosity, η0, of polyethylene chains directly and precisely in a constant‐stress rheometer at 190°C. Such experiments, when combined with precise measurements of the weight‐average molecular weight, Mw, calibrated relative to linear chains of high‐density polyethylene, are shown to provide a very sensitive approach to detect low levels (0.005 branches per 1000 carbons) of long‐chain branching (LCB). This detection limit is shown to be insensitive to whether the molecular weight distribution (MWD) breadth, Mw/Mn, rises from about two to ten. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

4.
Hydrolyzed cellulose–polyacrylonitrile graft copolymer is a polyelectrolyte gel suspension with a high viscosity in water. It is a closely packed swollen gel particle suspension in the appropriate concentration range and has similar rheological properties to other thickeners of this type. Viscosities η in either water or salt solution are reduced to a single master curve by use of the reduced viscosity function η/cQ, where c is weight fraction of polymer and Q is swelling volume in excess solvent of the same ionic strength. The effective molecular weight between crosslinks, Mc, determined from shear modulus, corresponds to Mc values for other closely packed gel thickners of similar η/cQ. Among all examples of this class of thickener, the plateau values of η/cQ, which occur at cQ > 2, are approximately inversely proportional to Mc.  相似文献   

5.
In this study, the viscoelastic behavior of hydrophobically modified polyelectrolytes obtained from the hydrolysis of cationic acid salts (CAS's) as a function of their zwitterion fraction (x) and anion fraction (z) was studied. The dynamic viscosity (η′) dependence on frequency of polymer solutions of polybetaine/anionic polyelectrolyte (APE) with various compositions of x and z in 0.1N NaCl showed typical shear thinning behavior. η′ of a solution of CAS 4 (M2‐4 (4 mol % hydrophobe)) attained a maximum value in the presence of 1.67 equiv of NaOH (corresponding to an x : z ratio of 33 : 67) and decreased with any further addition of NaOH. We suggest this maximum to be a result of a combined effect of coil expansion and hydrophobic association. The influence of the temperature and concentration on η′ of CAS 4 (M2‐4) treated with 1.67 equiv of NaOH was also investigated. The rheology of CAS 4 (M2‐4) samples treated with 1.67, 1.81, and 2.0 equiv of NaOH suggested a reversible network. However, for APE 7 (M2‐5 (5 mol % hydrophobe)), elastic behavior was dominant, and the formation of highly interconnected three‐dimensional networks was suggested. At lower x : z ratios, the effect of coil expansion due to a higher APE fraction was more than counterbalanced by the lower degree of intermolecular hydrophobic associations, whereas at higher x : z ratios, coil contraction became the predominant effect. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

6.
Lead‐free perovskite (1‐x)(K0.48Na0.48Li0.04)Nb0.95Sb0.05O3x(Bi0.5Na0.5)HfO3 piezoelectric ceramics were prepared by a traditional ceramic fabrication method. An investigation was conducted to assess the effects of (Bi0.5Na0.5)HfO3 content on the crystal structure, microstructure, phase‐transition temperatures, and piezoelectric properties of the ceramics. The X‐ray diffraction results, combined with the temperature dependence of dielectric properties, revealed that the ceramics experienced a structural transition from an orthorhombic phase to a tetragonal phase with the addition of (Bi0.5Na0.5)HfO3, and a coexistence of orthorhombic and tetragonal phases was identified in the composition range of 0.005≤x≤0.015. An obviously improved piezoelectric activity was obtained for the ceramics with compositions near the orthorhombic‐tetragonal phase boundary, among which the composition x=0.005 exhibited the maximum values of piezoelectric constant d33, and planar and thickness electromechanical coupling coefficients (kp and kt) of 246 pC/N, 0.435, and 0.554, respectively. Furthermore, the Curie temperature of the ceramics was found decreasing with the increase in (Bi0.5Na0.5)HfO3 content, but still maintaining above 300°C for the phase boundary compositions. These results indicate that the ceramics are promising lead‐free candidate materials for piezoelectric applications.  相似文献   

7.
Ternary solid solutions of (1 ? x)(0.8Bi0.5Na0.5TiO3–0.2Bi0.5K0.5TiO3)– xNaNbO3 (BNKT–xNN) lead‐free piezoceramics were fabricated using a conventional solid‐state reaction method. Pure BNKT composition exhibited an electric‐field‐induced irreversible structural transition from pseudocubic to ferroelectric rhombohedral phase at room temperature. Accompanied with the ferroelectric‐to‐relaxor temperature TF‐R shifted down below room temperature as the substitution of NN, a compositionally induced nonergodic‐to‐ergodic relaxor transition was presented, which featured the pinched‐shape polarization and sprout‐shape strain hysteresis loops. A strain value of ~0.445% (under a driving field of 55 kV/cm) with large normalized strain of ~810 pm/V was obtained for the composition of BNKT–0.04NN, and the large strain was attributed to the reversible electric‐field‐induced transition between ergodic relaxor and ferroelectric phase.  相似文献   

8.
The viscoelastic properties of narrowly distributed linear poly(ethylene-co-styrene) copolymers with different mole fractions of styrene (xS = 0–20.5 mol %) and molecular weights (Mw = 64–214 kg/mol) were analyzed in the molten state at different temperatures by means of oscillatory rheometry. Analyzing the thermorheological properties of the polymers, we found that the time temperature superposition principle is fulfilled. The corresponding shift factors follow up to 16.5 mol % of styrene units the Arrhenius behavior of neat polyethylene. For a styrene content of about 20 mol %, the polymers no longer crystallize and a transition from Arrhenius to WLF behavior of pure polystyrene was observed. The zero shear viscosity, η0, of the polymers was derived from the mastercurves. The determination of the plateau modulus by the well-known tan δ-min criterion is not possible due to the beginning crystallization in the corresponding temperature range. An approximate calculation of this value is based on the characteristic relaxation time λx = 1/ωx, corresponding to the crossover of G′ and G′. Indeed, the characteristic modulus Gpx calculated as η0x is a good approximation for the plateau modulus Gp. The viscosity–molecular weight and relaxation time–molecular weight scaling relations were established for three copolymers with different molecular weights and nearly the same styrene content. For both material parameters, the scaling exponent is around 3.4, confirming the linear architecture of the investigated polymers. The mixing rules describing the change of such material parameters like zero shear viscosity or plateau modulus independent of styrene content are of logarithmic linear character using the weight fraction of styrene units instead of the mole fraction. The relations found allow the prediction of melt state properties for polymers with arbitrary styrene content. In the future, when catalysts with sufficient activity for the synthesis of high styrene content copolymers are available, these predictions will have to be checked. © 1997 John Wiley & Sons, Inc. J Appl Polym Sci 65:209–215, 1997  相似文献   

9.
A study of the recent literature emphasized the importance of blending polymers with oils for improving the performance characteristics (like flexibility, corrosion resistance, etc.) of adhesives, coatings and laminates. Investigation of the available data revealed that several properties of such oil‐polymer blends could be correlated by molar refraction (RM), with reasonable accuracies. The properties of the linseed oil‐polystyrene (PS) and linseed oil‐polymethyl methacrylate (PMMA) blends studied are iodine value (IV), hydroxyl value (HV), inherent viscosity (η), melting temperature (TM), and glass transition temperature (Tg). In the case of safflower, palm and peanut oil‐sucrose polyester formulations, the viscosity at 40 °C (η40) and the melting point TM have been correlated by RM with the average absolute deviations (ē) of 17.8% and 3.0%, respectively. Using the orientation polarization PO (to represent polarity) in addition to RM, η40 and TM of oil‐polyester formulations could be calculated with ē values of 8.9% and 1.7%, compared to 17.8% and 3.0% using RM alone. The results indicated the importance of PO in improving the accuracy of predictions for properties.  相似文献   

10.
The rates of change of polymer properties (glass transition temperature, weight fraction sorbed water, and polymer molecular weight) were determined in poly (DL ‐lactide‐co‐glycolide) films under accelerated storage conditions. Films were stored at 70°C and 95%, 75%, 60%, 45%, or 28% relative humidity. Weight fraction sorbed water was determined by thermogravimetric analysis, the glass transition temperature (Tgmix) of the polymer/water mixture by modulated temperature differential scanning calorimetry, and PLGA number‐average molecular weight (Mn) by size exclusion chromatography (SEC). Rates of moisture increase and Tgmix decrease were related to the decrease in PLGA Mn through a modification of the Gordon‐Taylor equation. Understanding the relative rates of polymer physical and chemical degradation will allow for improved design of PLGA formulations that control rates of drug delivery and preserve drug stability. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

11.
A series of lead‐free perovskite solid solutions of (1 ? x) Na0.5Bi0.5TiO3(NBT)—x BaSnO3(BSN), for 0.0 ≤ x ≤ 0.15 have been synthesized using a high‐temperature solid‐state reaction route. The phase transition behaviors are studied using dielectric and Raman spectroscopic techniques. The ferroelectric to relaxor phase transition temperature (TFR) and the temperature corresponding to maximum dielectric permittivity (Tm) are estimated from the temperature‐dependent dielectric data. Dielectric studies show diffuse phase transition around ~335°C in pure NBT and this transition temperature decreases with increase in x. The disappearance of x‐dependence of A1 mode frequency at ~134 cm?1 for x ≥ 0.1 is consistent with rhombohedral‐orthorhombic transition. In situ temperature dependence Raman spectroscopic studies show disappearance and discontinuous changes in the phonon mode frequencies across rhombohedral (x < 0.1)/orthorhombic (x ≥ 0.1) to tetragonal transition.  相似文献   

12.
Bismuth vanadate compounds including BiVO4, Bi25VO40, and the composite Bi25VO40/Bi7V11O9 with photocatalytic properties were firstly synthesized via a facile electrochemical strategy. The amount of NaOH in the electrolyte and electric currents were discovered to be crucial variables to the composition, morphology, and crystallinity of the products. Monoclinic phase BiVO4 and Bi25VO40 with sillenite structure can be obtained at 2 A with NaOH concentration of 0–10?4 and 0.1–0.4M, respectively. The acquirement of Bi25VO40/Bi7V11O9 can be achieved with electric current 3–5 A in an electrolyte containing 0.4M NaOH. All the electrosynthesized samples exhibit good visible‐light‐driven photocatalytic activity in the degradation of rhodamine B, much higher than that of the commercial TiO2 (P25).  相似文献   

13.
Lattice dynamics and phase transitions of Aurivillius ferroelectric Bi3Ti1?xWxNbO9+δ (BTWN100x, 0 ≤ x ≤ 15%) ceramics have been investigated by temperature‐dependent Raman spectra (80–800 K) and far‐infrared (FIR) reflectance spectra (6–300 K). The frequency, intensity, and line width of phonon modes as well as complex dielectric functions of BTWN100x ceramics have been extracted by fitting Raman and FIR reflectance spectra with the multi‐Lorentzian oscillator model. It was found that the dielectric constants at a certain frequency region become lower with increasing W compositions at room temperature due to the compressive stress on W–O bands. Moreover, the temperature‐dependent behavior of optical phonon modes in BTWN100x ceramics indicates that there are two intermediate phases during the phase transition from paraelectric I4/mmm to ferroelectric A21am phase on cooling. The phase diagram of BTWN100x ceramics as a function of W composition has been improved.  相似文献   

14.
The field-controlled phase transition is a promising concept for the design of novel multiferroic materials. Rare-earth samarium-modified bismuth ferrite (Bi1−xSmxFeO3) possesses a morphotropic phase boundary (MPB) that has similar free energies between the polar and nonpolar phases, making it an exceptional candidate. In this study, we investigated the electric field cycling-dependent behavior of ferroelectricity in Bi1−xSmxFeO3 ceramics near MPB. During electric field cycling, a significantly enhanced remanent polarization was observed. Cycled Bi0.86Sm0.14FeO3 and Bi0.84Sm0.16FeO3 exhibited enhanced ferroelectric (remanent polarization >30 μC/cm2) and magnetic (remanent magnetization >0.20 emu/g) properties at room temperature. Through a systematic study of dynamic hysteresis measurements and a structural analysis, these results were attributed to a field cycling-induced nonpolar-to-polar phase transition. In situ high temperature measurements showed a previously unreported sharp anomaly of the piezoelectric coefficient (d33) near the magnetic transition point (TN). These results indicated a strong magnetoelectric coupling in rare earth-modified bismuth ferrite materials, suggesting the possibility of magnetically modulated piezoelectricity.  相似文献   

15.
Cptt2ZrCl2 (Cptt = η5?tBu2C5H3) was synthesized by the reaction of LiCptt with ZrCl4 and characterized by X‐ray crystallographic studies. It was used as catalyst for ethylene polymerization. Structural analysis was carried out on the polyethylene (PE) catalyzed by Cptt2ZrCl2 via wide‐angle X‐ray diffraction (WAXD) and small‐angle X‐ray scattering (SAXS). The degree of crystallinity (Wc,x) was calculated by WAXD. The semiaxises of the particles (a, a, b) of PE were determined by SAXS and it could be found that the crystalline particles of PE are mainly rod shaped determined by the characteristic function v0 (r). The radius of gyration Rg, crystalline thickness Lc, the thickness of noncrystalline region La, long period L, electron‐density difference between the crystalline and noncrystalline regions ηc ? ηa, and the invariant Q are determined by SAXS. The results also indicate that a transition zone exists between the traditional “two phases” with a clear dimension of 1.3 nm. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 96: 169–175, 2005  相似文献   

16.
Dense 0.6(Bi0.9La0.1)(Ga0.05Fe0.95)O3-0.4(Pb1-xBax)TiO3 (BLGF-PBT, x = 0, 0.1, 0.2, 0.3, and 0.4) ceramics with different Ba contents were prepared by the solid-state reaction method, and effects of Ba contents on the structure and multiferroic properties were investigated. X-ray diffraction results indicate that the Ba-modified BLGF-PBT ceramics exhibit single perovskite structure without detectable secondary phases, and a transition from tetragonal phase to rhombohedral one takes place with the increase of Ba content. The piezoelectric constant d33, remanent polarization Pr, and remanent magnetization Mr are improved by the Ba substitution simultaneously. Values of Pr, Mr, and d33 of BLGF-PBT ceramics for the composition of x = 0.3 with the coexistence of tetragonal and rhombohedral phases are 20 μC/cm2, 0.05 emu/g, and 256 pC/N, respectively, much higher than those without Ba modification. The significantly improved d33 and Mr indicate that BLGF-PBT ceramics with Ba modification provide alternative materials for multifunctional devices.  相似文献   

17.
While the reddish‐orange emitting phosphors M2Si5N8:Eu2+(M = Ca, Sr) have been intensively investigated as potential materials for white‐light‐emitting diodes, in this study, optical energy storage properties of (Ca1?xSrx)2Si5N8: Eu2+, Tm3+ (x = 0–1) solid solutions were tuned by cation substitution, which was commonly used to tune color point for improving w‐LEDs. Partial substitution of either Ca by Sr or Sr by Ca resulted in a redshifted Eu2+ emission which had a demarcation point at x = 0.5. Furthermore, the (Ca1?xSrx)2Si5N8: Eu2+, Tm3+ materials exhibited similar persistent‐ and photostimulated luminescence behaviors with a maximum intensity at about x = 0.2. Such optical energy storage characters of the samples were attributed to the more appropriate trap depths (322–333 K) and higher density of energy level traps indicated by the thermoluminescence analysis.  相似文献   

18.
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   

19.
By conventional ceramics sintering technique, the lead‐free 0.85Bi0.5Na0.5(1?x)Li0.5xTiO3‐0.11Bi0.5K0.5TiO3‐0.04BaTiO3 (x =0–0.15) piezoelectric ceramics were obtained and the effects of Li dopant on the piezoelectric, dielectric, and ferroelectric properties were studied. With increasing Li addition, the temperature‐dependent permittivity exhibited the normal ferroelectric‐to‐ergodic relaxor (FE‐to‐ER) transition temperature (TFEER, abbreviated as TF‐R) decreasing down to room temperature. The increasing Li content also enhanced the diffuseness of the FE‐to‐ER transition behavior. For composition with x = 0.15, a large unipolar strain of 0.37% ( = Smax/Emax = 570 pm/V) was achieved under 6.5 kV/mm applied electric field at room temperature. Both unipolar and bipolar strain curves related to the temperature closely, and when the temperature reached the TF‐R, the normalized strain achieved a maximum value (e.g., for x = 0.10, = 755 pm/V) owing to the electric‐field‐induced ER‐to‐FE state transition.  相似文献   

20.
Influence of K/Na ratio in (KxNa1?x)NbO3 on the ferroelectric stability and consequent changes in the electrical properties of 0.99(Bi0.5Na0.4K0.1)TiO3–0.01(KxNa1?x)NbO3 (BNKT–KxNN) ceramics were investigated. Results showed that change of K/Na ratio in KNN induces a phase transition from ferroelectric to ergodic relaxor phase with a significant disruption of the long‐range ferroelectric order, and correspondingly adjusts the ferroelectric–relaxor transition point TF?R to room temperature. Accordingly, giant strain of ~0.46% (corresponding to a large signal d33* of ~575 pm/V) which is comparable to that of Pb‐based antiferroelectrics is obtained at a K/Na ratio of ~1, and the emergence of large strain response induced by the change of K/Na ratio of KNN can be well explained by the correlation between the position of ferroelectric–ergodic relaxor phase boundary in the BNKT–KxNN system and the tolerance factor t of the end number (KxNN). In situ high‐energy X‐ray scattering experiments with external field reveals that the large strain response in the studied system is likely related to the electric field‐induced distortion from the pseudocubic structure.  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