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.
Mechanical behaviour of partially stabilized zirconia crystals (PSZC) with terbia and ceria additives was investigated under bending and indentation conditions. Test specimens were oriented along the [010] direction and along the axis of crystal growth. The PSZC bending strength (σb) was dependent on the crystallographic orientation of the specimens. The specimen volume subjected to stress influenced the PSZC strength. The highest mechanical characteristics were measured for ceria-doped crystals (σb = 1.9 GPa, Klc = 11.4 MPa m1/2, Ed = 366 GPa). The failure process was studied on the Vickers indentation, with special emphasis put on the development and propagation of lateral cracks. Anisotropy of lateral cracks in the (100) plane associated with that of the elastic moduli was revealed. At the same time anisotropy of radial cracks and hardness was not found. A new version of the equation to evaluate the fracture toughness (Kcv) on the Vickers indentation was derived. The Kcv values calculated by this equation correspond to those (Klc) obtained by an SENB method.  相似文献   

2.
Methyl methacrylate was polymerized in aqueous medium initiated by a copper(II)-ascorbic acidoxygen system at 40°C and a kinetic study of the reaction is presented. The rate of polymerization, Rp showed an increase, constancy and then a decrease with increase in the [Cu2+]. The order with respect to [Cu2+] was 0.5 in the rate increase region. The order in monomer concentration changed gradually from 1.0 to 2.0 with increase in [Cu2+]. Rp became independent of ascorbic acid (AA) concentration and oxygen concentration at high concentrations. These results indicate that termination by mutual interaction of chain radicals predominates at low [Cu2+] while termination was exclusively by metal ions at high [Cu2+]. Rp was also observed to increase with temperature and ionic strength and to Kp/Kt1/2 value was calculated and compared with literature values. Chain lengths were determined by viscometry for the polymers obtained under various experimental conditions.  相似文献   

3.
A Monte Carlo computer simulation model for the electrophoretic deposition of polymer chains on a discrete lattice is used to study the polymer density profile, interface growth, and its dependence on field, temperature, and molecular weight. The interface width (W) decreases WE−1/2 on increasing the field (E). Width (W) depends non-monotonically on the temperature (T): a power-law decay is followed by a power-law increase on raising the temperature. Monotonic decay of the interface width with the molecular weight is possibly a stretched exponential. Conformation and dynamics of a tracer chain is used to probe its characteristics in interface to bulk region. The root mean square (rms) displacement of the center of mass of the tracer chain shows an ultra-slow motion, Rtν (ν0.1–0.01 at E=0.1–1.0) as the driven chain moves deeper from interface to bulk. Longitudinal compression of the radius of gyration (Rg) of the chain increases with the field; transverse components (Rgx, Rgy) are larger than the longitudinal component (Rgz). The transverse component (Rgx(y)) becomes oscillatory due to periodic squeezing at high fields as the field competes with the polymer barriers.  相似文献   

4.
Edmund A. Di Marzio 《Polymer》1990,31(12):2294-2298
The entropy theory of glasses is used to derive the glass temperature, Tg, of a binary polymer blend in terms of the glass temperatures of the two substituents. The formula is Tg = B1Tg1 + B2Tg2, where Bi is the fraction of flexible bonds of substituent i. A bond is flexible if rotation about it changes the shape of the molecule. Bonds in side groups as well as in the backbone are to be counted. This formula assumes that the free volume, taken here to be the volume fraction of empty lattice sites, is the same for each of the three materials. It has no parameters. The above equation expressed in weight fractions, Wi, is (TgTg1)W111) + (TgTg2)W222) = 0, where ωi is the weight of a monomer unit and ggi is the number of flexible bonds per monomer unit. A more general treatment is given. One variation of the more general treatment which expresses the properties of the blend in purely additive terms gives Tg = B1Tg1 + B2Tg2 + KB1B2(Tg1Tg2)(V01V02), where V0i are the free volume fractions of the homopolymers at their glass temperatures and K is a constant. The added term is usually small. The most general form of the equation requires the energy of interaction between the two unlike molecules, which can be estimated by volume measurements on the blend.  相似文献   

5.
Mass transfer coefficients at cylindrical, H2 evolving electrodes, were measured by determining the reduction rate of K3Fe(CN)6. The variables studied were: gas discharge rate V, diameter of the cylinder D, height and position of the cylinder. The diameters ranged from 0·2–2·5 cm, the cd from 25-380 mA/cm2. For horizontal cylinders, the following correlation was found: log K = a + 2·17 log(V0·11/D0·08). The application of gas evolving cylindrical electrodes in industrial electrolysis is discussed in comparison with rotating electrodes.  相似文献   

6.
The divalent selective electrode together with high precision solid state, digital pH -mv -meter makes broader application of potentiometry in physical and inorganic chemistry a certainty. The above set-up is used to determine the stoichiometric constants, K, for Ca and Mg ions association with formates, acetates, propionates and butyrates at 25°, 35° and 45°C in aqueous media. The K-values were converted to infinite dilution KA values were found to be 8.4 LM−1, 10.4 LM−1, 19.1 LM−1 and 19.3 LM−1 for calcium salts of formate, acetate, propionate and butyrate respectively. Also KA values for Mg salts of formate, acetate, propionate and butyrate were found to be 7.8 LM−1, 9.5 LM−1, 13.1 LM−1 and 13.1 LM−1 respectively. Other thermodynamic parameters such as ΔG°, ΔH° and ΔS° are also obtained from the variation of KA with temperature for each salt. The data are interpreted relative to each other on basis of pKa of the corresponding organic acid. Their temperature behaviour is similar to those salts derived from strong acids such as sulphates, rather than weak acids.  相似文献   

7.
Microwave dielectric properties of (Zn1/3Nb2/3)0.40(Ti1−xSnx)0.60O2 ceramics were investigated as a function of SnO2 content (0.15 ≤ x ≤ 0.30). A single phase with tetragonal rutile structure was obtained through the entire composition. The unit-cell volume of the specimens was increased with SnO2 content, due to the larger ionic radius of Sn4+ (0.69 Å) than that of Ti4+ (0.605 Å) for octahedral site. Dielectric constant (K) of the sintered specimens was affected by the dielectric polarizability. Quality factor (Qf) was dependent on the degree of reduction of Ti4+ ion. With an increase of SnO2 content, the temperature coefficient of resonant frequency (TCF) of the specimens decreased due to the decrease of the octahedral distortion of rutile structure.  相似文献   

8.
Anastasios Dondos 《Polymer》1977,18(12):1250-1252
The relation proposed between the intrinsic viscosity and the molecular weight of a polymer1,2 (1/[η]versus 1/M1/2), has been applied in this paper in cases when the polymer is dissolved in a binary solvent mixture. We have then shown that even in this case we can calculate in a very accurate way the molecular weight of the polymer from its intrinsic viscosity, especially in the molecular weight range where the Mark-Houwink-Sakurada equation is the least applicable (Mw < 150 000).  相似文献   

9.
The particle collision behavior and heat transfer performance are investigated to reveal the heat transfer enhancement and fouling prevention mechanism in a Na2SO4 circulating fluidized bed evaporator. The particle collision signals are analyzed with standard deviation by varying the amount of added particles ε (1%–3%), circulation flow velocity u (0.37–1.78 m·s-1), and heat flux q (7.29–12.14 kW·m-2). The results show that the enhancement factor reach up to 14.6% by adding polytetrafluoroethylene particles at ε = 3%, u = 1.78 m·s-1, and q = 7.29 kW·m-2. Both the standard deviation of the particle collision signal and enhancement factor increase with the increase in the amount of added particles. The standard deviation increases with the increase in circulation flow velocity; however, the enhancement factor initially decreases and then increases. The standard deviation slightly decreases with the increase in heat flux at low circulation flow velocity, but initially increases and then decreases at high circulation flow velocity. The enhancement factor decreases with the increase in heat flux. The enhancement factor in Na2SO4 solution is superior to that in water at high amount of added particles. The empirical correlation for heat transfer is established, and the model results agree well with the experimental data.  相似文献   

10.
Rate constants or exchange current densities of electrode surface processes involving adatom arrays are conveniently evaluated by determining that sweep rate, s0 (the reversibility parameter), in a linear potential sweep (LPS) experiment, below which the process just remains kinetically reversible, ie its overpotential is sensibly zero. Transition to irreversibility is characterized by peak potentials, Ep, becoming linear in the log of the sweep rate, s, following a region of independence of s for < s0. A suitable extrapolation procedure enables s0 to be evaluated. However, if s0 is large and/or the resistivity of the solution is appreciable, the IRu drop associated with uncompensated resistance in the measurement system can be comparable with the increase of Ep with log s, when s > s0, rendering evaluation of s0 inaccurate. While compensation or empirical correction for this may be made, it is desirable that the nature of the IRu effect of the LPS I vs E profiles be understood in a more fundamental way. It is the purpose of this communication to provide such a treatment of this effect through evaluation of the actual time-dependent potential that becomes applied to the electrode, and to propose criteria based on the product of s0, Ru and the reaction pseudocapacitance, Cφ for indicating the anticipated extent of the IRu effect in the evaluation of s0. While the transition in the Ep vs log s plot may be due to IRu effects as well as to kinetic irreversibility, when the former are appreciable, it is shown that the corresponding transition in the value of Cφ with increasing log s can usually allow spurious IRu and significant irreversibility effects to be distinguished.  相似文献   

11.
The hydrodynamic characteristics in aqueous solution at ionic strength I=0.2  of carboxymethylchitins of different degrees of chemical substitution have been determined. Experimental values varied over the following ranges: the translational diffusion coefficient (at 25.0°C), 1.1<107×D<2.9 cm2 s−1; the sedimentation coefficient, 2.4<s<5.0 S; the Gralen coefficient (sedimentation concentration-dependence parameter), 130<ks<680 mL g−1; the intrinsic viscosity, 130<[η]<550 mL g−1. Combination of s with D using the Svedberg equation yielded ‘sedimentation–diffusion' molecular weights in the range 40 000<M<240 000 g mol−1. The corresponding Mark–Houwink–Kuhn–Sakurada (MHKS) relationships between the molecular weight and s, D and [η] were: [η]=5.58×10−3 M0.94; D=1.87×10−4 M−0.60; s=4.10×10−15 M0.39. The equilibrium rigidity and hydrodynamic diameter of the carboxymethylchitin polymer chain is also investigated on the basis of wormlike coil theory without excluded volume effects. The significance of the Gralen ks values for these substances is discussed.  相似文献   

12.
The miscibility of poly(methyl methacrylate) (PMMA) and styrene-acrylonitrile random copolymers (SAN) blends was investigated on the basis of the Flory—Orwoll—Vrij equation of state theory. To obtain the equation of state parameters (P*, V*sp, T*: characteristic parameters), the pressure—volume—temperature (PVT) behaviour was measured for PMMA and a series of SANs with various acrylonitrile contents. The exchange energy parameter Xij was also calculated by fitting the theory to some phase diagrams of PMMA/SAN blends. The Flory—Huggins interaction parameter χ was separated into two contributions based on the equation of state theory for mixtures: the exchange energy term χinter and the free volume term χfree. Both the temperature and copolymer composition dependences of χinter and χfree were estimated by calculations using the equation of state parameters. There exists a region in which χinter is negative, leading to a miscibility window in PMMA/SAN blends. However, the immiscibility at high temperatures in the blends cannot be explained only by χinter; it is caused by the free volume contribution, χfree. The miscibility window behaviour in PMMA/SAN blends may be explained within the framework of the equation of state theory.  相似文献   

13.
Forward recoil spectrometry (FRES) was used to measure the tracer diffusion coefficients D*PS and D*PXE of deuterated polystyrene (d-PS) and deuterated poly(xylenyl ether) (d-PXE) chains in high molecular weight protonated blends of these polymers. The D*s were shown to be independent of matrix molecular weights and to decrease as M−2, where M is the tracer molecular weight, suggesting that the tracer diffusion of both species occurs by reptation. These D*s were used to determine the monomeric friction coefficients ζ0,PS and ζ0,PXE of the individual PS and PXE macromolecules as a function of ф, the volume fraction of PS in the PS:PXE blend. Since ζ0,PSζ0,PXE at each ф, the rate at which a PS molecule reptates is much greater than that of a PXE molecule, even though both chains are diffusing in identical surroundings. Part of this difference may be due to the difficulty of backbone bond rotation of the PXE molecule. However, even when measured at a constant temperature increment above the glass transition temperature, ζ0,PS and ζ0,PXE were observed to be markedly composition dependent. In addition the ratio ζ0,PS0,PXE varied from a maximum of 4 × 10−2 near ф=0.85 to a minimum of 5 × 10−5 for ф=0.0. These results show that intramolecular barriers do not solely determine the ζ0s of the components in this blend. Clearly, the interactions between the diffusing chains and the matrix chains also influence ζ0.  相似文献   

14.
Liquid phase mixing time (θmix) was measured in mechanically agitated contactors of internal diameter 0.57 m, 1.0 m and 1.5 m. Tap water was used as the liquid phase. The impeller speed was varied in the range of 0.4-9.0 r/s. Three types of impellers, namely disc turbine (DT), pitched blade downflow turbine (PTD) and pitched-blade upflow turbine (PTU) were employed. The ratio of impeller diameter to vessel diameter (D/T) and the ratio of impeller blade width to impeller diameter (W/D) were varied over a wide range. The effects of impeller clearance from the tank bottom (C), the blade angle (φ), the number of blades (nb), the blade thickness (k) and the total liquid height (H/T) were studied in detail. Mixing time was measured using the conductivity method.

Mixing time was found to have a strong dependance on the flow pattern generated by the impeller. Mixing time was found to decrease by decreasing the impeller clearance in the case of DT and PTU. However in the case of PTD it increases with a decrease in the impeller clearance. Similar trend of the effect of impeller clearance on θmix, was observed for all the other PTD impellers with different diameter, number of blades and blade angle (except 60° and 90°). All the impeller designs were compared on the basis of power consumption and on this basis optimum design recommendations have been made. For PTD impellers, a correlation has been developed for the dimensionless mixing time.  相似文献   

15.
Poly-p-phenylene (PPP) was synthesized from benzene according to the Kovacic method. Electrodes were made from this electronic insulator by cold- or hot-pressing of the loose, brown powder, under the addition of 7.5 wt. % soot (Corax L®, Degussa AG). The electrochemical insertion and removal of anions HSO4, ClO4 in this material in aqueous solutions of the corresponding acids was investigated by slow cyclic voltammetry.

Initially, only a surface layer of about 0.1 mm thickness takes part in the electrochemical processes, which are reversible. A maximum concentration of anions in the solid of [(−C6H4−)+7 A] is attainable. The maximum degree of insertion is equal to 0.14. The insertion potential UI shifts strongly into the negative direction with increasing concentration c of the acid. A linear UI/c relationship is observed as in the case of graphite, where the intercalation potential is more positive by 20–200 mV for the same electrolyte. The round trip current efficiency for the insertion/removal cycle increases with increasing acid concentration attaining 100% in 14 M H2SO4 or 11.3 M CHlO4. For a given concentration, increases in the same order as with graphite (H2SO4 < HClO4 < HBF4), being somewhat lower for a given electrolyte composition. From anodic current limitation (jlim = 5–10 mA cm−2), a diffusion coefficient of about D = 2 × 10−7 cm2 s−1 is derived for the transport of anions in the bulk of PPP. The striking similarity of our results to former findings with graphite is thoroughly discussed. Some general conclusions are derived thereof.  相似文献   


16.
Correlation between the equation of state and the temperature dependence of the self-diffusion coefficient D for polymers such as polystyrene (PS) and polydimethyl siloxane (PDMS) and simple liquids such as argon, methane and benzene and the pressure dependence of D for oligomers such as dimethyl siloxane (DMS) and simple liquids such as cyclohexane and methanol has been examined based on the equation of state derived previously. The experimental data used were published by Antonietti et al. and McCall et al. for polymers, by McCall for linear dimethylsiloxanes and by Jonas et al. and Woolf et al. for simple liquids. The expression for D in this work is given by

where A1(M) is a function of molecular weight Mw, C1(T) and P1(T) are functions of temperature and B1, n1 and m1 are constants determined experimentally. For simple liquids, the values of n1 obtained range from 0.3 to 1.2, with an average , and m1 is in the range 0.5–1.2, with . For polymers, values of n1 are in the range 2.5–7.0 for PS and 0.5–1.3 for PDMS and m1 for DMS is in the range 0.8–1.0. The relation Dη/T = f(M) is found to be useful for simple liquids over a wide range of temperature including the critical region and for pressures up to ≈5 kbar

1 kbar = 100 MPa There is a close correlation between ln(D/T) and p and βT through ln(D/T)ln Dc−1p−β−1T, where Dc is D at the critical temperature and p and βT are the thermal expansion coefficient and compressibility, respectively. The molecular weight dependence of D for polymers and simple liquids is discussed based on the experimental data and recent theory of Doi and Edwards. A new model for the mechanism of self-diffusion in the liquid state is proposed.  相似文献   


17.
18.
In this work, the computational fluid dynamics method is used to study the liquid hydrodynamics behavior in the microchannel without central insert (MC1) and the central insert microchannel (MC2), respectively. The maximum deviation between simulation and experiment is 24%. The formations of flow patterns are explained based on contours and force analysis where the flow pattern maps are established by two-phase flow rate. The effects of aqueous phase viscosity and two-phase flow rate on the characteristic sizes of each flow pattern are also explored. Specifically, four unconventional flow patterns are found in MC2, namely the unique droplet flow, the unique slug flow, the unique coarse annular flow and the unique film annular flow. Though the insert occupies part of the channel, the pressure difference in the channel is significantly reduced compared with MC1. Moreover, the insert significantly changes the formation velocity range of each flow pattern, greatly broadens the formation range of annular flow and also has an important influence on the characteristic size of the flow pattern. The organic-phase dimensionless axial size (Lo/W) and the dimensionless radial size (Do/W) of the droplet (slug) are negatively related to the aqueous-phase viscosity (μa) and flow rate (ua). The Do/W of the annular is negatively correlated with μa and positively correlated with organic-phase flow rate (uo). This study provides direct numerical evidence that the insert is key to the formation of bicontinuous phase flow pattern, as well as further strengthens our understanding of the flow characteristics and optimization design of insert microchannels.  相似文献   

19.
The solute-solute and solute-solvent interactions of drug semicarbazide hydrochloride with carbohydrates (D-glucose/D-sucrose) are investigated by using volumetric, viscometric and acoustic properties. The measurements of the densities ρ, ultrasonic speeds u, and viscosities η. of semicarbazide hydrochloride in 5% and 10% D-glucose/D-sucrose+water (w/w) solutions were carried out at temperatures (293.15-318.15) K and at pressure, p=101 kPa. The apparent molar volumes, Vϕ, limiting apparent molar volumes,Vϕ°, apparent molar compressibilities, Ks, ϕ, limiting apparent molar compressibilities, Ks,ϕ°, partial molar expansibilities, Eϕ°, transfer volumes, Vϕ,tr° and transfer compressibilities, Ks,ϕ,tr° have been calculated from the experimental data. The viscosity data were examined by using the Jones-Dole equation and the viscosity A and B coefficients were evaluated. The results are discussed in terms of solute-solute and solute-solvent interactions in these solutions. The structure making/breaking ability of semicarbazide hydrochloride is examined using the sign of temperature derivative of B-coefficient, dB/dT.  相似文献   

20.
Small-angle neutron scattering experiments in the range of q2 from 0.01 to 25 nm−2 have been carried out on branched epoxide resins based on bisphenol-A at the Institute Laue—Langevin (I.L.L) in Grenoble (q=(4π/λ) sin(θ/2)). Measurements were made with six samples in the range of MW from 1500 to 19 000 and four concentrations between 1.3 and 10% (w/w) in deuterated diglyme. The results are as follows: (i) The mean square radius of gyration follows a relationship S2z=4.69×10−4M1.20W (nm2). (ii) In all cases fairly large second virial coefficients A2 are obtained which, however, decrease strongly with molecular weight. Above MW=2500, the virial coefficient follows the relationship A2=1.6M−0.85W (mol cm3g−2). (ii) The reciprocal particle scattering factor as a function of q2 exhibits only a slight upturn and otherwise shows the behaviour of a randomly branched polycondensate. The slight upturn is discussed as being caused by the finite volume of the monomeric unit. Possible reasons for the high exponent in the S2z versus MW dependence are briefly discussed.  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