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.
The microstructural evolution of pure copper during friction-stir welding was found to be principally influenced by welding temperature. At temperatures below ~0.5 Tm (where Tm is melting point), the microstructure was shown to be essentially determined by continuous recrystallization, leading to significant grain refinement and related material strengthening in the stir zone. In contrast, grain structure development at temperatures above ~0.5 Tm was dominated by discontinuous recrystallization producing a relatively coarse grain structure in the stir zone and giving rise to material softening.  相似文献   

2.
The melting and recrystallization of nylon-6,10 thin films immersed in an aqueous solution of calcium chloride were investigated by DSC measurements. The crystal length, ζ, was determined as a function of the melting peak temperature, T m . The end surface free energy of nylon-6,10 crystals used for the ζT m conversion was derived thermodynamically. For films of 0.01 mm thickness, the original length of ζ (=7.6 structural units) at T m decreased step by step with increasing immersion time by the length near the structural unit (2.24 nm) per step. However, the suppression of the recrystallization after melting of the original crystals formed during the first cooling by the adsorbed calcium ions did not occur completely, even for films immersed for 30~60 min at 50°C.  相似文献   

3.
Abstract

Thermal cross‐linking of poly(vinyl methyl ether) (PVME) in the absence of cross‐linking agent, was detected rheologically. The linear viscoelastic properties of PVME were found to be greatly changed by the onset of the cross‐linking process. The viscoelastic material functions, such as dynamic shear moduli, G′ and G″, complex shear viscosity, η*, and loss tangent, tan δ, were found to be sensitive to the structure changes during the cross‐linking process and the formation of a three‐dimensional polymer network. At the onset temperature of the cross‐linking process, an abrupt increase in G′, G″, and η* (several orders of magnitude) during dynamic temperature ramps (2°C/min heating rate) was observed with some frequency dependence. The temperature dependence of tan δ was found to be frequency independent at the gel‐point, T gel, that is, the crossover in tan δ regardless of the value of frequency can be taken as an accurate method for determination of T gel. The coincidence of G′ and G″ at the gel‐point cannot be considered a general method for evaluation of T gel due to its high frequency dependence, that is, T gel determined from the crossover of G′ and G″ in the dynamic temperature ramp at 1 rad/sec is about 20°C less than at 100 rad/sec. Furthermore, a dramatic increase in η0 above the minimum (“v” shape) was observed at T = T gel in agreement with the value obtained from tan δ vs. T (190°C). The time–temperature‐superposition principle was found to be valid only for temperatures lower than the T gel (190°C); the principle failed at T ≥ 190°C. This was clearly seen in the low‐frequency region as a deviation from the terminal slope in the G′ curve. Similar behavior was observed in the modified Cole–Cole analyses (G″ vs. G′) that is, the curves start to deviate at 190°C.  相似文献   

4.
M. Kaya  Y. Elerman  I. Dincer 《哲学杂志》2018,98(21):1919-1932
The effect of heat treatment on the structural, magnetic and magnetocaloric properties of Ni43Mn46In11 melt-spun ribbons was systematically investigated using X-ray powder diffraction (XRD), scanning electron microscope (SEM), atomic force microscope (AFM), magnetic force microscope (MFM) and magnetic measurements. From the XRD studies, tetragonal and cubic phases were detected at room temperature for as-spun, quenched and slow-cooled ribbons. Furthermore, it was observed, upon annealing martensite transition temperatures increased when compared to the as-spun ribbon. To avoid magnetic hysteresis losses in the vicinity of the structural transition region, the magnetic entropy changes-ΔS m of the investigated ribbons were evaluated from temperature-dependent magnetisation-M(T) curves on cooling for different applied magnetic fields. The maximum ΔS m value was found to be 6.79 J kg?1 K?1 for the quenched ribbon in the vicinity of structural transition region for a magnetic field change of 50 kOe.  相似文献   

5.
ABSTRACT

Dielectric properties of 0.5Ba(Ti0.8Zr0.2)O3–0.5(Ba0.7Ca0.3)TiO3 ceramics were probed in the frequency range from 10 Hz to 100 THz in a broad temperature range (10–900 K). Polar soft phonon observed in infrared spectra softens with cooling; however, below 500 K, its frequency becomes temperature independent. Simultaneously, a central mode activates in terahertz and microwave spectra; and it actually drives the ferroelectric phase transitions. Consequently, the phase transitions strongly resemble a crossover between the displacive and order–disorder type. The central mode vanishes below 200 K. The dielectric relaxation in the radiofrequency and microwave range anomalously broadens on cooling below TC1, resulting in the nearly frequency independent dielectric loss below 200 K. This broadening comes from a broad frequency distribution of ferroelectric domain wall vibrations. Raman spectra reveal new phonons below 400 K, i.e. already 15 K above TC1. Several weak modes are detected in the paraelectric phase up to 500 K in Raman spectra. Activation of these modes is ascribed to the presence of polar nanoclusters in the material.  相似文献   

6.
We consider massive photon decay reactions via intermediate states of electron-electron-holes and proton-proton-holes into neutrino-antineutrino pairs in the course of neutron star cooling. These reactions may become operative in hot neutron stars in the region of proton pairing where the photon due to the Higgs-Meissner effect acquires an effective mass m γ that is small compared to the corresponding plasma frequency. The contribution of these reactions to neutrino emissivity is calculated; it varies with the temperature and the photon mass as T 3/2 m γ 7/2 exp(−m γ /T) for T<m γ . Estimates show that these processes appear as extra efficient cooling channels of neutron stars at temperatures T≅109–1010 K. Zh. éksp. Teor. Fiz. 114, 385–397 (August 1998) Published in English in the original Russian journal. Reproduced here with stylistic changes by the Translation Editor.  相似文献   

7.
The melting temperature, T m, of copper has been determined from ambient pressure to 16 GPa using multi-anvil techniques. The melting curve obtained (T m=1355(5)+44.5(31)P?0.61(21)P 2, with T m in Kelvin and P in GPa) is in good agreement with both the previous experimental studies and with recent ab initio calculations.  相似文献   

8.
Abstract

Healing of symmetric interfaces of amorphous anionically polymerized high‐ and ultrahigh‐molecular weight (HMW and UHMW, respectively) polystyrene (PS) in a range of the weight‐average molecular weight M w from 102.5 (M w/M n = 1.05) to 1110 kg/mol (M w/M n = 1.15) was followed at a constant healing temperature, T h, well below the glass transition temperature of the polymer bulk [T g‐bulk = 105–106°C as measured by differential scanning calorimeter (DSC)]. The bonded interfaces were shear fractured in tension on an Instron tester at ambient temperature. Autoadhesion at symmetric HMW PS–HMW PS and UHMW PS–UHMW PS interfaces was detected mechanically after healing at T h = 38°C for 107 hr, and even at 24°C (for longer healing times). The occurrence of autoadhesion between the surfaces of the UHMW PS with M w = 1110 kg/mol at 24°C implies that the glass transition temperature at the interface, T g‐interface, of this polymer was a least lower: by 82°C than its DSC T g‐bulk, by 30–40°C than the Vogel temperature, T —the lowest theoretical value of a kinetic T g‐bulk at infinite long time—and by 20°C than T 2 (a “true” thermodynamic T g‐bulk corresponding to a second‐order phase transition temperature). To our knowledge, this is the first observation of such nature, which gives further evidence of the lowering of the T g at polymeric surfaces and the persistence of this effect at early stages of healing of polymer–polymer interfaces.  相似文献   

9.
Samples from sheets of the polymeric material Makrofol DE 7-2 have been exposed to 1 MeV protons of fluences in the range 2.5×1013–5×1015 p/cm2. The resultant effect of proton irradiation on the thermal properties of Makrofol has been investigated using thermogravimetric analysis and differential thermal analysis (DTA). The onset temperature of decomposition T o and the activation energy of thermal decomposition E a were calculated, and the results indicated that the Makrofol detector decomposes in one weight loss stage. Also, the proton irradiation in the fluence range 7.5×1013–5×1015 p/cm2 led to a more compact structure of Makrofol polymer, which resulted in an improvement in its thermal stability with an increase in the activation energy of thermal decomposition. The variation of transition temperatures with proton fluence has been determined using DTA. The Makrofol thermograms were characterized by the appearance of an endothermic peak due to the melting of the crystalline phase. The melting temperature of the polymer, T m, was investigated to probe the crystalline domains of the polymer. At a fluence range of 7.5×1013–5×1015 p/cm2, the defect generated destroys the crystalline structure, thus reducing the melting temperature. In addition, the VI characteristics of the polymer samples were investigated. The electrical conductivity was decreased with the increasing proton fluence up to 5×1015 p/cm2. Further, the refractive index, transmission of the samples and any color changes were studied. The color intensity Δ E was greatly increased with the increasing proton fluence and was accompanied by a significant increase in the red and yellow color components.  相似文献   

10.
This effort reports on novel fluorinated polyamide (FPA) and polyamide 1010 (PA1010)-based blends and graphene reinforced nanocomposite. PA1010/FPA (80:20) blend was opted as matrix material on the basis of molecular weight, thermal, and shear stress performance. Graphene was obtained through in situ chemical method of graphene oxide reduction. PA1010/FPA/Graphene nanocomposites was developed using various graphene loadings (up to 5 wt.%). Thin film coatings were prepared on glass substrate. Consequently, the PA1010/FPA/Graphene attained regular spongy morphological pattern. PA1010/FPA/Graphene 3 also showed improved T0 and Tmax of 534 and 591 °C relative to the neat blend (T10 423 °C; Tmax 551 °C). Limiting oxygen index measurement indicated better non-flammability of PA1010/FPA/Graphene 1–3 nanocomposite series (57–60%) relative to the blend series (28–31%). UL94 tests also showed V-0 rating for nanocomposites. Furthermore, PA1010/FPA/Graphene 3 nanocomposite revealed significantly high tensile strength (62 MPa), flexural modulus (1690 MPa), and adhesive properties to be utilized as coating materials. The nanocomposite coatings also displayed outstanding barrier properties against O2 and H2O compared with neat blends.  相似文献   

11.
A study of electrophysical and thermodynamic properties of C60 single crystals under step shock loading has been carried out. The increase and the following reduction in specific electroconductivity of C60 fullerite single crystals at step shock compression up to pressure 30 GPa have been measured. The equations of state for face centred cubic (fcc) C60 fullerite as well as for two-dimensional polymer C60 and for three-dimensional polymer C60 (3D-C60) were constructed. The pressure–temperature states of C60 fullerite were calculated at step shock compression up to pressure 30 GPa and temperature 550 K. The X-ray diffraction studies of shock-recovered samples reveal a mixture of fcc C60 and a X-ray amorphous component of fullerite C60. The start of the formation of the X-ray amorphous component occurs at a pressure P m≈ 19.8 GPa and a temperature T m≈ 520 K. At pressures exceeding P m and temperatures exceeding T m, the shock compressed fullerite consist of a two-phase mixture of fcc C60 fullerite and an X-ray amorphous component presumably consisting of the nucleators of polymer 3D-C60 fullerite. The decrease in electroconductivity of fullerite can be explained by the percolation effect caused by the change of pressure, size and number of polymeric phase nuclei.  相似文献   

12.
Magnetite is the oldest magnet and the first material where the concept of a strong correlations driven metal–insulator transition was suggested and found at TV = 124 K in the so-called Verwey phase transformation. Recently, the structure below TV was solved revealing subtle electronic structure in the form of trimeron lattice that, according to yet another recent communication, may be switched within femtosecond range. In this review article, we argue that the same change of trimeron lattice can be achieved by a magnetic field, in the phenomenon called the easy axis switching. The results of many of our experiments show that although this process is best viewed by magnetization studies, it is also reflected in magnetostriction, causes some changes in electronic transport and can be observed microscopically by NMR that proved electronic order alteration. All those facts suggest that the axis switching process observed and studied by us is intimately linked with the fast change of electronic trimeron order mentioned above.  相似文献   

13.
Amorphous Fe78Si9B13 alloy ribbons were heat treated between 296 and 763 K, using heating rates between 1 and 4.5 K/min. Whereas one ribbon partially crystallized at T x = 722 K, the other one partially crystallized at T x = 763 K. The partially crystallized ribbon at 722 K, heat treated using a triangular form for the heating and cooling rates, was substantially less fragile than the partially crystallized at 763 K where a tooth saw form for the heating and cooling rates was used. Vickers microhardness and hyperfine magnetic field values behaved almost concomitantly between 296 and 673 K. The Mössbauer spectral line widths of the heat-treated ribbons decreased continuously from 296 to 500 K, suggesting stress relief in this temperature range where the Vickers microhardness did not increase. At 523 K the line width decreased further but the microhardness increased substantially. After 523 K the line width behave in an oscillating form as well as the microhardness, indicating other structural changes in addition to the stress relief. Finally, positron lifetime data showed that both inner part and surface of Fe78Si9B13 alloy ribbons were affected distinctly. Variations on the surface may be the cause of some of the high Vickers microhardness values measured in the amorphous state.  相似文献   

14.
Abstract

Films of high‐molecular‐weight amorphous polystyrene (PS, M w = 225 kg/mol, M w/M n = 3, T g‐bulk = 97°C, where T g‐bulk is the glass transition temperature of the bulk sample) and poly(methyl methacrylate) (PMMA, M w = 87 kg/mol, M w/M n = 2, T g‐bulk = 109°C) were brought into contact in a lap‐shear joint geometry at a constant healing temperature T h, between 44°C and 114°C, for 1 or 24 hr and submitted to tensile loading on an Instron tester at ambient temperature. The development of the lap‐shear strength σ at an incompatible PS–PMMA interface has been followed in regard to those at compatible PS–PS and PMMA–PMMA interfaces. The values of strength for the incompatible PS–PMMA and compatible PMMA–PMMA interfaces were found to be close, both being smaller by a factor of 2 to 3 than the values of σ for the PS–PS interface developed after healing at the same conditions. This observation suggests that the development of the interfacial structure at the PS–PMMA interface is controlled by the slow component, i.e., PMMA. Bonding at the three interfaces investigated was mechanically detected after healing for 24 hr at T h = 44°C, i.e., well below T g‐bulks of PS and PMMA, with the observation of very close values of the lap‐shear strength for the three interfaces considered, 0.11–0.13 MPa. This result indicates that the incompatibility between the chain segments of PS and PMMA plays a negligible negative role in the interfacial bonding well below T g‐bulk.  相似文献   

15.
Measurements of the hysteresis loop and pyroelectric current density have been carried out. It has been shown that the function describing the remanent polarization decay over time generated by a prolonged transverse electric field is for TGS qualitatively the same as for other uniaxial ferroelectrics (TGSe, Rochelle salt), regardless of the fact that different electrode–sample systems were used. A prolonged application of an electric potential Vs at temperature T = TA < TC (TC is the critical temperature of the paraelectric–ferroelectric phase transition) to a side ring electrode of a round plate sample changes pyroelectric properties of TGS and leads to the memory effect. For T < TA, the polarization P values obtained by time integration of electric current density measured after Vs disconnection differ from those measured before Vs application by a constant value, and therefore, the first derivative ?P/?T remains unchanged provided that the temperature TA is not exceeded.  相似文献   

16.
Novel polyurethane (PU) adhesive was prepared and coated on poly(methyl methacrylate) (PMMA) and poly(methyl methacrylate)/fullerene (PMMA/Full-C60) composite. Dip-coating technique was employed as facile and cost-effective procedure to coat polyurethane on film substrate. The properties of PU/PMMA and PU/PMMA/Full-C60 composite were studied using Fourier transform infrared spectroscopy, Field Emission Scanning Electron Microscopy, tensile, adhesion, thermal and flammability measurement. Testing polyurethane-coated PMMA exhibited crumpled surface while fullerene addition formed unique pattern of dispersed spherical structures. Fullerene nanofiller loading improved the adhesion and mechanical properties of composite films due to polymer–carbon interaction. In PU/PMMA/Full-C60 0.5 composite with 0.5 wt.% nanofiller, tensile strength (71.4 MPa) was increased by 18.6% while tensile modulus was increased by 143.85% compared with PU/PMMA. In PU/PMMA/Full-C60 0.5, T0 of 473 °C and Tmax of 655 °C were observed. Increasing the fullerene content up to 0.5 wt.% decreased the peak heat release rate to 131 kW/m2. Novel polyurethane-coated PMMA/Full-C60 composite have potential applications as adhesive coatings in electronic and automotive appliances.  相似文献   

17.
We report on the electrical resistivity and far-infrared reflectance measurements of LaO1?xFxFeGe samples. Furthermore, we introduce a new method to probe the energy gap and determine its value. The onset transition temperature was 22.8 K for x = 0.13, and a clear anomaly was observed at 90 K in the ρ(T) curve for x = 0.11 with Tc = 20.6 K. We clearly observed the phonon-suppressed feature in reflectance spectra where F-doping caused a strong suppression of a peak at 200 cm?1. The energy gap above Tc, 2Δ = 2.10 meV, was determined from the measured spectra based on the changes in reflectivity by F-doping.  相似文献   

18.
19.
We have investigated by electron tomography, in a transmission electronic microscope, the interactions between dislocations in olivine single crystals and polycrystals deformed in axial compression at T < 1000 °C (T < 0.5Tm). Dislocations are mostly of the [0?0?1] type, except in the polycrystal where [1?0?0] and [0?0?1] dislocations have been activated. A few 〈1?0?1〉 junctions have been found and characterized. Many collinear interactions have been identified either involving direct interactions between crossing dislocations of opposite Burgers vectors or indirect interactions between dislocations gliding in parallel planes and sessile dislocation loops. We suggest that collinear interaction, already identified as the primary source of strain hardening in FCC metals, is the main dislocation interaction mechanism in olivine deformed at temperatures below 1000 °C.  相似文献   

20.
Abstract

In this work, we have studied on double-layered perovskite (Ruddlesden–Popper) manganite structure in Pr1.75Sr1.25Mn2O7 synthesised by sol–gel method. The crystal structure of the double-layered perovskite is found as tetragonal from the X-ray diffraction analysis with I4/mmm space group. A high Curie temperature, TC = 305 K is observed from the temperature dependence of magnetisation measurement. The isothermal magnetisation curves showed that magnetic phase transition is second order due to the positive slope of the Arrott plots. Maximum magnetic entropy change (ΔSM) and adiabatic temperature change (ΔTad) values are calculated as 3.99 J kg?1 K?1 and 2.1 K under external magnetic field of 70 kOe, respectively. Since our double-layered perovskite manganite sample has desired TC value and relatively high ΔSM, it can be a potential candidate as a magnetocaloric material for room temperature magnetic cooling systems.  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