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.
Piezoelectric Lead Zirconate Titanate Ceramic Fiber/Polymer Composites   总被引:1,自引:0,他引:1  
Piezoelectric lead zirconate titanate (PZT) ceramic fiber/polymer composites were fabricated by a novel technique referred to as "relic" processing. Basically, this involved impregnating a woven carbon-fiber template material with PZT precursor by soaking the template in a PZT stock solution. Careful heat treatment pyrolized the carbon, resulting in a PZT ceramic relic that retained the fibrous template form. After sintering, the densified relic was backfilled with polymer to form a composite. Optimized relic processing consisted of soaking activated carbon-fiber fabric twice in an intermediate concentration (405-mg PZT/(1-g solution)) alkoxide PZT solution and sintering at 1285°C for 2 h. A series of piezoelectric composites encompassing a wide range of dielectric and piezoelectric properties was prepared by varying the PZT-fiber orientation and polymer-matrix material. In PZT/Eccogel polymer composites with PZT fibers orientated parallel to the electrodes, K = 75, d 33= 145 pC/N, d h= 45 ± 5 pC/N, and d hgh= 3150 × 10−15 m2/N were measured. Furthermore, in composites with a number of PZT fibers arranged perpendicular to the electroded surfaces, K = 190, d 33= 250 pC/N, d h= 65 ± 2 pC/N, and d h gh= 2600 × 10−15 m2/N.  相似文献   

2.
A multilayer piezoelectric ceramic/polymer composite with 2–2 connectivity was fabricated by thermoplastic green machining after co-extrusion. The multilayer ceramic body was composed of piezoelectrically active lead zirconate titanate (PZN)–lead zinc niobate (PZN)-lead zirconate titanate (PZT) layers and electrically conducting PZN–PZT/Ag layers. After co-extruding the thermoplastic body, which consisted of five piezoelectric layers interspersed with four conducting layers, it was computer numeric-controlled machined to create periodic channels within it. Following binder burnout and sintering, an 18 vol% array of 190 μm thin PZT slabs with a channel size of 880 μm was fabricated. The channels were filled with epoxy in order to fabricate a PZN–PZT/epoxy composite with 2–2 connectivity. The piezoelectric coefficient (effective d 33) and hydrostatic figure of merit ( d h× g h) of the PZN–PZT/epoxy composite were 1200 pC/N and 20 130 × 10−15 m2/N, respectively. These excellent piezoelectric characteristics as well as the relatively simple fabrication procedure will contribute in widening the application range of the piezoelectric transducers.  相似文献   

3.
The woven replication process was used to fabricate lead zirconate titanate (PZT)/polymer composites with 1–3, 2–3, and 3–3 connectivities by starting with novoloid-derived carbon fiber, woven fabric, and nonwoven felt templates, respectively. Activated carbon-fiber template material was impregnated with PZT by soaking it in a solution containing stoichiometric amounts of dissolved lead, zirconium, titanium, and niobium ions. Heat treatment burned out the carbon, leaving a PZT replica with the same form as the template material. Replicas were sintered in a controlled atmosphere and backfilled with an epoxy polymer to form final composites. This method, which is believed to be adaptable for mass production, is capable of producing composites with extremely fine microstructures. Woven composite samples have fiber tow diameters of 200 to 250 μm and spacings between tows of about 150 to 250 μm. Average d 33= 90 pC/N, g 33= 211 mV · m/N, and dhgh hydrophone figure of merit of 2100 × 10−15 m2/N values are reported for woven PZT/polymer composites.  相似文献   

4.
Ceramic-polymer composites with 0–3 connectivity were prepared from (Pb0.5, Bi0.5) (Ti0.5, (Fe1- x Mn x )0.5)O3 ( x = 0.00 to 0.02) ceramic powder and epoxy polymer. Epoxy was dissolved in a methanol solution and the ceramic powder was added to form a suspension. Powder-polymer coacervates were precipitated from the suspension by the addition of water. Colloidal filtration followed by cold-pressing consolidated the coacervates to form composites with uniform microstructure. Composites fabricated in this manner could be poled with higher electric fields than could be sustained by the conventionally prepared composites. The values of piezoelectric coefficients, d 33, d h, and g h were as high as 65, 41 pC/N, and 145 mV · m/N, respectively.  相似文献   

5.
The relic processing technique was used to fabricate fine-scale piezoelectric lead zirconate titanate (PZT) ceramic fiber/polymer shell composites. In this technique sacrificial activated carbon fabrics were soaked in a PZT precursor solution, dried, and heat treated to form piezoceramic relics. Relics were embedded with polymer, which was allowed to cure, and the resulting composites were polished, electroded, and poled. Different facets of the composite- forming process were examined: structural modifications, soaking, firing, and polymer impregnation. The physical and electromechanical properties of the unique resulting composite were evaluated. Optimized PZT shell composites with 39 vol% ceramic exhibited the following property values: K ε200, tan δε5.5%, d 33ε290 pC/N, d hε100 pC/N, d h g hε6000 ± 10-15 m2N, k pε0.19, and k tε0.28.  相似文献   

6.
Fine-scale, large-area lead zirconate titanate (PZT) ceramic fiber/polymer composites were developed using the relic method. Carbon fabrics were used as a starting template material. These fabrics were soaked in PZT stock solution, stacked, and fired, removing the carbon and resulting in a relic structure identical to that of the original template. The relics were then sintered, backfilled with polymer, polished, and poled, resulting in a piezoelectric ceramic/polymer composite. The processing of the larger-area composites involved scaling up the procedure utilized in the fabrication of the smaller composites. The application of an optimum uniaxial pressure of 580 Pa to the stacks during firing was found to improve the piezoelectric properties and facilitate the increase in sample area. Scale-up to large area was achieved using two approaches: (1) the use of a larger template material, and (2) tiling smaller area relics together in an array configuration. The dielectric and piezoelectric properties of relic composites of 2.5 × 2.5 cm2 area with ∼37 vol% PZT were K = 150 ± 8, d33= 180 ± 11 pC/N, dh= 85 ± 7 pC/N, and d hgh= 5525 × 10−15m2/N. The properties of large-area composites were comparable with those of small area.  相似文献   

7.
Thermal expansion of the low-temperature form of BaB2O4 (β-BaB2O4) crystal has been measured along the principal crystallographic directions over a temperature range of 9° to 874°C by means of high-temperature X-ray powder diffraction. This crystal belongs to the trigonal system and exhibits strongly anisotropic thermal expansions. The expansion along the c axis is from 12.720 to 13.214 Å (1.2720 to 1.3214 nm), whereas it is from 12.531 to 12.578 Å (1.2531 to 1.2578 nm) along the a axis. The expansions are nonlinear. The coefficients A, B , and C in the expansion formula L t = L 0(1 + At + Bt 2+ Ct 3) are given as follows: a axis, A = 1.535 × 10−7, B = 6.047 × 10−9, C = -1.261 × 10−12; c axis, A = 3.256 × 10−5, B = 1.341 × 10−8, C = -1.954 × 10−12; and cell volume V, A = 3.107 × 10−5, B = 3.406 × 10−8, C = -1.197 × 10−11. Based on α t = (d L t /d t )/ L 0, the thermal expansion coefficients are also given as a function of temperature for the crystallographic axes a , c , and cell volume V.  相似文献   

8.
Crystals of β-Ca2SiO4 (space group P 121/ n 1) were examined by high-temperature powder X-ray diffractometry to determine the change in unit-cell dimensions with temperature up to 645°C. The temperature dependence of the principal expansion coefficients (αi) found from the matrix algebra analysis was as follows: α1= 20.492 × 10−6+ 16.490 × 10−9 ( T - 25)°C−1, α2= 7.494 × 10−6+ 5.168 × 10−9( T - 25)°C−1, α3=−0.842 × 10−6− 1.497 × 10−9( T - 25)°C−1. The expansion coefficient α1, nearly along [302] was approximately 3 times α2 along the b -axis. Very small contraction (α3) occurred nearly along [     01]. The volume changes upon martensitic transformations of β↔αL' were very small, and the strain accommodation would be almost complete. This is consistent with the thermoelasticity.  相似文献   

9.
Piezoelectric ceramic/polymer composites with 1-3 connectivity were made by weaving sized lead zirconate titanate (PZT) fiber bundles through a honeycomb support. Bundles comprised of fine-scale, 20-50 μm green fibers, made using the viscous suspension spinning process, were sized to increase their manageability. The sizing step comprised of soaking the green PZT fiber bundles in an aqueous solution of poly(vinyl alcohol), then pulling the wet fibers through a steel sizing die. Sizing resulted in dense and flexible fiber bundles, which facilitated composite construction and led to composites with increased volume fractons. Sintering, polymer embedding, and machining produced a composite exhibiting 1-3 connectivity. Composites with 10 vol% PZT yilded d 33 values of 230 pC/N and a dielectric constant of 130.  相似文献   

10.
High-density lead zirconate titanate (PZT) ceramics were fabricated for the first time at a temperature as low as 800°C via the hot isostatic pressing (HIP) of a PZT powder with a modified composition of 0.92Pb(Zr0.53Ti0.47)O3—0.05BiFeO3—0.03Ba(Cu0.5W0.5)O3 that contained 0.5 mass% MnO2. The resultant PZT ceramics exhibited a microstructure that was denser and finer than that of PZT sintered at 935°C, which is the lowest temperature for the densification of the same composition via normal sintering. The relevant dielectric and piezoelectric properties of the HIPed PZT ceramics were as follows: coefficient of electromechanical coupling ( K 31), 31.8%; mechanical quality factor ( Q m), 1364; piezoelectric constant ( d 31), −73.7 × 10−12 C/N; relative dielectric constant (ɛ33T0), 633; dielectric loss factor (tan δ), 0.5%; Curie temperature ( T c), 285°C; and density (ρ), 8.06 g/cm3. In addition to these reasonably good piezoelectric properties, the HIPed PZT exhibited better mechanical properties—particularly, higher fracture strength—than the normally sintered PZT.  相似文献   

11.
The effect of a transverse tensile stress on the electric-field-induced 90°-domain reorientation in tetragonal lead zirconate titanate (PZT) near the morphotropic phase boundary was investigated in situ using X-ray diffraction (XRD). The XRD intensity ratio, I (002)/ I (200), which represents the ratio of the volume of the c -domains to that of the a -domains on the PZT surface, was examined as a function of the electric field at various stress levels. It was found that a transverse tensile stress changes the electric-field dependence of I (002)/ I (200), especially at higher electric fields. Without a transverse tensile stress, I (002)/ I (200) began to saturate at E ≈ 800 kV/m. With a transverse tensile stress of 75 MPa, I (002)/ I (200) increased with an upward curvature with the electric field, indicating that the transverse tensile stress enhanced the field-induced 90°-domain reorientation, and increased the effective piezoelectric coefficients at larger electric fields. At E = 900 kV/m, the estimated d 31,domain changed from −200 × 10−12 V/m at zero stress, to −350 × 10−12 V/m at 75 MPa.  相似文献   

12.
The metastable crystal structure of strontium- and magnesium-substituted LaGaO3 (LSGM) was studied at room and intermediate temperatures using powder X-ray diffractometry and Rietveld refinement analysis. With increased strontium and magnesium content, phase transitions were found to occur from orthorhombic (space group Pbnm ) to rhombohedral (space group R [Threemacr] c ) at the composition La0.825Sr0.175Ga0.825Mg0.175O2.825 and, eventually, to cubic (space group Pm [Threemacr] m ) at the composition La0.8Sr0.2Ga0.8Mg0.2O2.8. At 500°C in air and at constant strontium and magnesium content, a phase transformation from orthorhombic (space group Pbnm ) to cubic (space group Pm [Threemacr] m ) was observed. For the orthorhombic modification, thermal expansion coefficients were determined to be α a ,ortho = 10.81 × 10−6 K−1, α b ,ortho = 9.77 × 10−6 K−1, and α c ,ortho = 9.83 × 10−6 K−1 (25°–400°C), and for the cubic modification to be αcubic= 13.67 × 10−6 K−1 (500°–1000°C).  相似文献   

13.
Open-cell hydroxyapatite (HA) foams, produced through the novel technique of gelcasting foams with relative porosities ranging from 0.72 to 0.90, were characterized for pore-size distribution, surface area, permeability, compressive strength, elastic modulus, and microstructural features. The porous structure, which is composed of an array of spherical cells interconnected through windows, had a mode pore diameter in the range 17–122 μm, as demonstrated by mercury porosimetry. The BET specific surface area increased from 1.5 to 3.8 m2/g as the sample porosity increased. The compressive strength and elastic modulus were in the range 1.6–5.8 MPa and 3.6–21.0 GPa, respectively. The permeability constants, k 1 (Darcian) and k 2 (non-Darcian), were strongly dependent on porosity fraction and varied widely, from 1.22 × 1011 to 4.31 × 1010 m2 and from 1.75 × 106 to 8.06 × 105 m, respectively. This combination of properties make the HA foams suitable for a variety of potential applications in the biomedical field, preferentially nonloading, including materials for bone repair, carriers for controlled drug-delivery systems, and matrixes for tissue engineering.  相似文献   

14.
Low-Temperature Sintering of Lead-Based Piezoelectric Ceramics   总被引:3,自引:0,他引:3  
The low-temperature sintering of lead-based piezoelectric ceramics has been studied. The sintering temperature of lead zirconate titanate (PZT) ceramics could be reduced from ∼ 1250° to ∼960°C by the addition of a small amount of the lower-melting frit, B2O3–Bi2O3—CdO. It exhibited the following dielectric and piezoelectric properties: Kp= 0.52 to 0.58, Qm= 1000, εT330= 800 to 1000, tan δ= 50 × 10−4, ρ= 7.56 to 7.64 g/cm3. Ceramics with the aid of suitable dopants (CdO, SiO2, and excess PbO) in the Pb-(Ni1/3Nb2/3)O3—PZT family could be sintered at 860° to 900°C. For these materials, Kp= 0.56 to 0.61, Qm= 1000, εT330= 1500 to 2000, tan δ≤ 50 × 10−4, ρ= 7.80 to 8.03 g/cm3. The microstructure, sintering mechanism, and the effects of various impure additions have been analyzed by means of scanning electron microscopy, scanning transmission electron microscopy, electron probe microanalysis, and X-ray photoelectron spectroscopy.  相似文献   

15.
The characterization and properties of ceramic composites containing the phases Al2TiO5, ZrTiO4, and ZrO2 are described. The range of compositions investigated gives very low average thermal expansions (α24–1000°C as low as −2.0 × 10−6°C−1) and excellent high-temperature stability. The low thermal expansions are apparently due to a combination of microcracking by the titanate phases and a contractive phase transformation by the ZrO2. The crystal chemistry and microstructure of the product are processing dependent. Although the composites represent a complex microcracking system, the low thermal expansions and high-temperature stability make them potential candidates for commercial applications requiring thermal shock resistance.  相似文献   

16.
Thermal conductivity ( k ), electrical resistivity ( p ), total hemispherical emittance (εt), and normal spectral emittance (ε0.65μ) of dense, arc-cast uranium monocarbide (5.3 wt % total carbon) were measured in the temperature range 1150° to 2050°K. The results were as follows: k , 0.057 cal/sec-cm-deg, 1200° < T < 2050°K, probable error ± 0.002; p, 20.4 × 10−6+ 114.8 T × 10−9 ohm-cm, 1175° < T < 2050°K, probable error ± 1.7 × 10−6; εt0.42, 1250° < T < 1980°K, probable error ± 0.02; ε0.65 0.539 – 0.02 T × 10−3 1150° < T < 1890°K, probable error ± 0.02. Experimental methods are discussed and error sources are analyzed. Uranium monocarbide exhibited typical metallic behavior in its thermophysical properties.  相似文献   

17.
The thermal and electrical properties of MoSi2 and/or SiC-containing ZrB2-based composites and the effects of MoSi2 and SiC contents were examined in hot-pressed ZrB2–MoSi2–SiC composites. The thermal conductivity and electrical conductivity of the ZrB2–MoSi2–SiC composites were measured at room temperature by a nanoflash technique and a current–voltage method, respectively. The results indicate that the thermal and electrical conductivities of ZrB2–MoSi2–SiC composites are dependent on the amount of MoSi2 and SiC. The thermal conductivities observed for all of the compositions were more than 75 W·(m·K)−1. A maximum conductivity of 97.55 W·(m·K)−1 was measured for the 20 vol% MoSi2-30 vol% SiC-containing ZrB2 composite. On the other hand, the electrical conductivities observed for all of the compositions were in the range from 4.07 × 10–8.11 × 10 Ω−1·cm−1.  相似文献   

18.
The thermal expansion of the hexagonal (6H) polytype of α-SiC was measured from 20° to 1000°C by the X-ray diffraction technique. The principal axial coefficients of thermal expansion were determined and can be expressed for that temperature range by second-order polynomials: α11= 3.27 × 10–6+ 3.25 × 10–9T – 1.36 × 10–12 T 2 (1/°C), and ş33= 3.18 × 10–6+ 2.48 × 10–9 T – 8.51 × 10–13 T 2 (1/°C). The σ11 is larger than α33 over the entire temperature range while the thermal expansion anisotropy, the δş value, increases continuously with increasing temperature from about 0.1 × 10–6/°C at room temperature to 0.4 × 10–6/°C at 1000°C. The thermal expansion and thermal expansion anisotropy are compared with previously published results for the (6H) polytype and are discussed relative to the structure.  相似文献   

19.
Ceria-stabilized tetragonal zirconia polycrystals were obtained by thermal treatment of amorphous powder prepared by the sol–gel method. Detailed XRD profile analysis was employed to study microstructural disorder and crystallite size and shape; in particular, no fluctuation of stoichiometry was found, the main cause of disorder being attributable to dislocations. Thermal expansion measurements were carried out by high-temperature XRD at 294, 473, 673, 873, and 1073 K using silicon as an internal standard. Thermal expansion coefficients are anisotropic and changes in the stabilizer content have little effect on them. A mean value, α a = 10.6 × 10−6 (K−1) and α c = 13.5 × 10−6 (K−1), can be assumed for Zr1− x Ce x O2 with x in the range 0.12–0.18.  相似文献   

20.
Porous lead zirconate titanate–lead zinc niobate (PZT–PZN) piezoelectric ceramics with a high degree of pore alignment were fabricated using directional freeze casting of a ceramic/camphene slurry. Well-aligned pores were formed as the replica of the camphene dendrites that grew in a preferential orientation, while a high porosity of 90% was achieved by employing a low initial solid loading of 5 vol%. As the orientation angle of the pores to the poling direction was decreased, the hydrostatic piezoelectric properties, such as hydrostatic piezoelectric strain coefficient ( d h), the hydrostatic piezoelectric voltage coefficient ( g h), and the hydrostatic figure of merit, increased significantly. The sample containing pores aligned parallel to the poling direction showed an extremely high HFOM of 161019 × 10−15 Pa−1, which was ∼1300 times higher than that (124 × 10−15 Pa−1) of the dense sample, owing to the presence of aligned pores.  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