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.
Eight new hard‐sphere equations of state (EOS's) were obtained from molecular simulation data for the pair correlation function gHS(σ) vs. packing fraction η and combined with three theoretical schemes to obtain 21 new cubic EOS's for athermal hard‐sphere chains (AHSC's). The eight new hard‐sphere EOS models successfully reproduced isotropic fluid compressibility factor ZHS and gHS(σ) vs. η simulation data and predicted metastable liquid ZHS vs. η and virial coefficients up through B10. Moreover, calculated Z vs. η and reduced second‐virial coefficient vs. chain length m were compared with molecular simulation data for chains up to m = 201 for a set of representative (eight of twenty‐one) chain equations. Z vs. η for three AHSC binary mixtures was also successfully predicted. The results indicate that the new cubic EOS's give a satisfactory representation of simulation data for chain fluids and can be used to develop theoretically based cubic EOS's for “real” fluids including attractive effects. © 2015 American Institute of Chemical Engineers AIChE J, 61: 1677–1690, 2015  相似文献   

2.
The precise control on concentration profile of dispersion in functionally graded material (FGM) is essential for obtaining a desired material. A suitable simulation of parameters and an appropriate model that describes the motion of particles in the fluid can predict various aspects those are needed to produce FGM, by gravity sedimentation or centrifugation technique. Simulation was conducted to observe the changes in concentration profile, while using the following equations applicable to polymerizing fluid, and to determine the terminal velocities (Vm) of particles; Vm = {D2s ? ρl)g*(1 ? ?s)4.65}/(18μ0e) for gravity sedimentation and Vm = {D2s ? ρl)rω2(1 ? ?s)4.65}/(18μ0 e) for centrifugation, where D is the diameter of the spherical particle, ρs the density of solid particles, ρl the density of fluid, μ the viscosity of fluid, g* the acceleration due to gravity, ?s is the volume fraction of particles, and tc is the elapsed time of curing of thermosetting resin. b is a constant, r is the radius, and ω is the angular velocity. This simulation demonstrates that the time of centrifugation/sedimentation, particle size, distribution of particle size, and centrifugal/gravitational forces can be effectively utilized to attain a desired concentration profile in graded materials. Simulation also revealed that there exist the possibility of two graded profiles, namely low concentration profile and high concentration profile, in one sample of graded material, made either by centrifugation or sedimentation. Low concentration profile is more sensitive to particle size distribution as compared to high concentration profile. The present simulation method is also sensitive to concentration‐measuring methods. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci, 2006  相似文献   

3.
An equation which relates the volume term (V=M/(ρL-ρg)) of unassociated liquids to pressure P and temperature T has been obtained by the combination of (a) 03V(?I/?V)P→0=Tx-T for the effect of temperature on V at low (atmospheric) pressure and (b) - V(?P/?V)T = Px Vx/6/V6 p→0+9(P-p) for the effect of pressure on volume at constant temperature. In the equations, p is the vapour pressure; pL the density of the liquid and pg the vapour density. Often pg can be neglected compared with pL and p is small compared with the large pressures required to affect the densities of liquids appreciably. There are three constants: Tx, Px, which equals 4.455 × 109 N m2, and Vx which can be calculated by the addition of atomic values for all the atoms in the molecule and subtraction of a value (6.56 × 106 m3 mol1) for each bond. When V approximates to ML, the molar volume, the equation can be integrated to give the work and heat of isothermal compression. The viscosity of a liquid is related to the work of compression and solubilities in a liquid to the work required to bring the solute to the compressibility of the liquid. Many relationships can be derived and can be used to estimate properties of unassociated liquids.  相似文献   

4.
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   

5.
Glassmelting efficiency largely depends on heat transfer to reacting glass batch (melter feed), which in turn is influenced by the bulk density (ρb) and porosity (?) of the reacting feed as functions of temperature (T). Neither ρb(T) nor ?(T) functions are readily accessible from direct measurements. For the determination of ρb, we monitored the profile area of heated feed pellets and calculated the pellet volume using numerical integration. For the determination of ?, we measured the material density of feeds quenched at various stages of conversion via pycnometry and then computed the feed density at heat‐treatment temperature using thermal expansion values of basic feed constituents.  相似文献   

6.
Densities (ρ, kg m?3), and viscosities (η, 0.1 kg m?1 s?1) of Bovine Serum Albumin (BSA), Egg Albumin, and Lysozyme in aqueous iodide salts of lithium, sodium, and potassium, along with cationic surfactant‐cetyltrimethyl ammonium bromide (CTAB) were measured at a temperature of 303.15 K. The 0.0010–0.0018 g %, w/v of each protein at an interval of 0.0002 mol L?1 in 0.2, 0.4, and 0.8 millimol L?1 of salt and CTAB are studied. Data are used for apparent molar volumes (V?, 10?6 m3 mol?1) and intrinsic viscosities ([η], dL kg?1), respectively. Data are regressed and extrapolated to zero concentrations for ρ0, η0, and limiting values and Sd, Sη and SV corresponding slopes for protein–salt structural interactions. With size of cations, the densities decrease as CTAB > LiI > NaI > KI and increase with salts concentrations, with salts the densities are as Lysozyme > BSA > Egg Albumin, viscosities and V? as BSA > Egg–Albumin > Lysozyme. The ρ and η values with CTAB higher and [η] are lower and converse at around 0.4 mmol L?1 salt and is effective for greater stability of proteins. The [η] in CTAB are higher than other salts and decreases with size of cations with stronger intermolecular forces. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

7.
Here, the limitations of characteristic length-based (Lchb) and grain size-based (Gb) criteria with two or three parameters were pointed out employing the apparent toughness tests of 12 different ceramics at a large span range of U-notch root radius (ρ) values. After comprehensively considering the potential influencing factors of stress intensity factor (Kc), ρ divided by critical notch tip radius (ρc) was proposed as the independent variable, and the data of 21 materials (covering ceramics, plastics, resins, rocks, and metals) was summarized and discussed to establish a simple and more applicable Kc prediction model. Results indicated that Kc/KIc was a power function of ρ/ρc with a power exponent n of 0.5 for ideal materials and less than 0.5 for actual materials. It was also found that ρc can be calculated simply by KIc2/(πσ02), where σ0 represented the inherent strength. This semiempirical criterion succeeded in unifying the Lchb and Gb criteria without introducing more parameters to increase the prediction accuracy of the Kc at the U-notch root for brittle materials like ceramics.  相似文献   

8.
Styrene was copolymerized in bulk with a number of esters of benzylidenecyanoacetic acid. The kinetic scheme of all pairs fitted the improved scheme of copolymerization, taking into account the effect of the penultimate unit. The Alfrey-Price Q and e values were calculated. Using the modified Taft equation, log (1/r1) = ρ*σ* + δEs, it was found that the relative reactivities of the ester monomers toward the polystyryl radical were correlated by the polar substituent constants σ* of the ester alkyl groups (ρ* = 0.14) and not by their steric substituent constants Es (δ = 0.008).  相似文献   

9.
Bed‐to‐wall heat transfer properties of a vertical heat tube in a fluidized bed of fine fluid catalytic cracking (FCC) particles are measured systematically using a specially designed heat tube. Two important surface hydrodynamic parameters, i.e. the packet fraction (δpa) and mean packet residence time (τpa) based on the packet renewal theory, are determined by an optical fiber probe and a data processing method. The experimental results successfully reveal the axial and radial profiles of heat‐transfer coefficient, the effects of superficial gas velocity, and static bed height on heat‐transfer coefficient, most of which can be explained successfully by the measured τpa, an indicator of packet renewal frequency. τpa is found to play a more dominant role than δpa on bed‐to‐wall heat transfer. With a fitted correction factor, the modified Mickley and Fairbanks model is able to predict the heat‐transfer coefficients with enough accuracy based on the determined packet parameters. © 2014 American Institute of Chemical Engineers AIChE J, 61: 68–83, 2015  相似文献   

10.
Tomography, an efficient nonintrusive technique, was employed to visualize the flow in continuous‐flow mixing and to measure the cavern volume (Vc) in batch mixing. This study has demonstrated an efficient method for flow visualization in the continuous‐flow mixing of opaque fluids using two‐dimensional (2‐D) and 3‐D tomograms. The main objective of this study was to explore the effects of four inlet‐outlet configurations, fluid rheology (0.5–1.5% xanthan gum concentration), high‐velocity jet (0.317–1.660 m s?1), and feed flow rate (5.3 × 10?5?2.36 × 10?4 m3 s?1) on the deformation of the cavern. Dynamic tests were also performed to estimate the fully mixed volume (Vfully mixed) for the RT, A310, and 3AM impellers in a continuous‐flow mixing system, and it was found that Vfully mixed was greater than Vc. Incorporating the findings of this study into the design criteria will minimize the extent of nonideal flows in the continuous‐flow mixing of complex fluids and eventually improve the quality of end‐products. © 2013 American Institute of Chemical Engineers AIChE J, 60: 315–331, 2014  相似文献   

11.
The electrochemical behaviour of the series of ten [Rh(RCOCHCOR′)(P(OPh)3)2] complexes with R, R′ = CF3, CF3 (1), CF3, CH3 (2), CF3, Ph (C6H5) (3), CF3, Fc (ferrocenyl = (C5H5)Fe(C5H4)) (4), CH3, Ph (5), CH3, CH3 (6), Ph, Ph (7), Fc, CH3 (8), Fc, Ph (9) and Fc, Fc (10) were studied in acetonitrile containing 0.100 mol dm−3 tetra-n-butylammonium hexafluorophosphate as supporting electrolyte utilizing a glassy carbon working electrode. Results are consistent with Rh(I) being first oxidized in an electrochemically irreversible two-electron transfer process at peak anodic potentials ranging Epa(Rh) = 0.124–0.881 V vs. Fc/Fc+. For the ferrocene-containing complexes (4), and (8)–(10) the rhodium oxidation was followed by the electrochemically reversible oxidation of the ferrocenyl group in a one-electron transfer process at a slightly more positive potential. Relationships were established between the electrochemical quantity Epa(Rh) and kinetic parameter log k2 as well the sum of experimental group electronegativities (Gordy Scale) of the R and R′ groups (χR + χR′), the Hammett σ values (σR + σR′) and the Lever ligand parameter EL for the [Rh(RCOCHCOR′)(P(OPh)3)2] complexes: Epa(Rh) (vs. Fc/Fc+/V) = 0.31 (χR + χR′)–1.09 = 0.56 (σR + σR′) + 0.28 = SMEL) + (IM − 0.66 V) = −0.23 log k2 − 0.03 (k2 = second order rate constant for the oxidative addition of methyl iodide to rhodium). A profound shift of Epa(Rh) to a more positive potential was observed for Rh(I) substrates containing β-diketonato ligands with increasing electronegative substituents R and R′. An exponential dependence of Epa(Rh) on the pKa of the β-diketone was obtained.  相似文献   

12.
Two types of permanent antistatic agents, polyethylene wax grafted with sodium acrylate (PEW‐g‐AAS) and polypropylene (PP) wax grafted with sodium acrylate (PPW‐g‐AAS), were prepared using a solution grafting method and applied to PP for enhancing antistatic properties. The grafting degree was determined using back titration method and structures were confirmed by Fourier transform infrared spectroscopy. The antistatic properties of PEW‐g‐AAS/PP blends and PPW‐g‐AAS/PP blends were characterized by surface resistivities (ρs) and volume resistivities (ρv), and a combination of contact angle measurements, scanning electron microscope, permittivity, and dielectric loss were used to investigate the surface and inner structures of the blends. Results showed ρs and ρv of PEW‐g‐AAS/PP blends dropped significantly (4–7 magnitudes) above a critical addition at 10%, where a electrostatic dissipative network formed; PPW‐g‐AAS revealed an inferior antistatic performance than PEW‐g‐AAS due to its better compatibility and smaller dispersed phase in the matrix. Further, the antistatic blends treated in 80°C water, 80°C air, and room temperature were investigated, and the results were interpreted from surface energy. Moreover, the addition of antistatic agent had little impact on tensile strength of the PP matrix. © 2012 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

13.
A new method for solving the integral equation based on the Ornstein‐Zernike equation for binary mixture is proposed. Then the radial distribution function obtained for both the Percus‐Yevick and the hypernetted chain closure equations are used to calculate the residual chemical potential at infinite‐dilution and at reduced temperatures T* = 2, 1.5 (supercritical isotherms) over a varying range of reduced densities ρ* = 0.1 to 0.6 for various types of the Lennard‐Jones mixture in terms of size ratios D and energy ratios C. To examine the ability of the integral equation approach for the residual chemical potential calculations, the results are compared with the Monte‐Carlo simulation data and the van der Waals I results (Shing et al., 1988). It is seen that at ρ* = 0.1 to 0.5, the deviation of the integral equation results from the MC simulation data is less than the reported statistical fluctuation.  相似文献   

14.
More realistic dynamic bed‐expansion experiments using a three‐phase anaerobic fluidized bed reactor (AFBR) with and without internal biogas production were conducted for the establishment of correlation equations for the mean volume ratio of wakes to bubbles (k). A predictive model was also developed for the expansion characteristics of the three‐phase AFBR with internal biogas production. The predicted bed‐expansion heights (HGLS) deviated by only ±10% from the experimental measurements for the three‐phase AFBR. According to the modeling results, if a three‐phase AFBR is loaded into a carrier with low specific gravity (dry density of carrier, ρmd = 1.37 g cm?3; wet density of carrier, ρmw = 1.57 g cm?3) and operated at a high superficial liquid velocity (ul = 4.0 cm s?1), the ratio of HGLS to HLS at a high superficial gas velocity (ug = 1.5 cm s?1) can reach as high as 271%. A higher fluidized‐bed height has a greater effect on the bed‐expansion behavior because of the decrease in liquid pressure (surrounding gas bubbles) along the fluidized‐bed height. From parametric sensitivity analyses, HGLS is most sensitive to the parameter reactor width (X), especially within a small ΔX/X0 range of ±10%; sensitive to ρmw, diameter of the carrier, ρmd and total mass of carrier and least sensitive to ul, biofilm thickness and ug. Copyright © 2005 Society of Chemical Industry  相似文献   

15.
Glass formation behavior of the TeO2–WO3–Na2O system was studied by using conventional melt‐quenching technique. A wide glass formation range was determined for the first time in the literature and thermal, physical, and structural characterization of sodium‐tungsten‐tellurite glasses were realized using differential scanning calorimetry (DSC) and Fourier transform infrared (FTIR) spectroscopy techniques. Glass transition (Tg) and crystallization (Tc/Tp) temperatures, glass stability (?T), density (ρ), molar volume (VM), oxygen molar volume (VO), and oxygen packing density (OPD) values and structural transformations in the glass network were investigated according to the equimolar substitution of TeO2 by Na2O+WO3 and changing Na2O or WO3 at constant TeO2.  相似文献   

16.
The influence of thermocycling process at VO2 phase transition temperature (Tt≈68°C) on the resistivity temperature dependences and dielectric constant (ε) of ceramics on a basis of VO2 and V2O5–P2O5 glasses has been studied. The differential thermal analysis and dilatometric measurements for tested ceramics has been made. Vanadium dioxide based ceramics exhibits a variation of electrical properties with thermocycling. The resistivity increases, ε and electrical resistivity jump ρsm in the vicinity of Tt decrease with a growth of a number of thermocycles. The reason of such behavior is microcracks formation due to a sharp variation of thermal expansion in the vicinity of Tt. After some repetition of thermocycling process the stabilization of ceramics electrical properties has been observed, but the resistivity jump ρsm has been disappeared practically. Possible reasons of the stabilization have been discussed. The stabilization of electrical resistivity jump ρsmsm∼102–103) after thermocycling for thin ceramic samples (a thickness ∼0·4 mm or less) has been shown. ©  相似文献   

17.
《应用陶瓷进展》2013,112(2):132-135
Abstract

The nanocrystalline BaTiO3 based PTC powders had been prepared by a simple sol–gel method starting from comparatively cheap raw materials. The powders and ceramics were characterised by thermogravimetry–differential scanning calorimetry, X-ray diffraction and scanning electron microscopy, while electronic properties of the ceramics were also studied. The good material electronic properties were obtained, including a Curie temperature (T C) of 100°C, a resistivity at room temperature (ρ 25°C) of 18 Ω cm, a high resistivity step ratio (ρ max/ρ min) of 1·2×106, a temperature coefficient of resistivity (α 30) of 17% °C?1 and a withstand voltage intensity (V b) of 196 V mm?1.  相似文献   

18.
We report in this article the results of nanosilica (SiO2)‐filled epoxy composites with different loadings and their electrical, thermal, mechanical, and free‐volume properties characterized with different techniques. The morphological features were studied by transmission electron microscopy, and differential scanning calorimetry was used to investigate the glass‐transition temperature (Tg) of the nanocomposites. The properties of the nanocomposites showed that the electrical resistivity (ρ), ultimate tensile strength, and hardness of the composites increased with SiO2 weight fraction up to 10 wt % and decreased thereafter; this suggested that the beneficial properties occurred up to this weight fraction. The temperature and seawater aging had a negative influence on ρ; that is, ρ decreased with increases in the temperature and aging. The free‐volume changes (microstructural) in the composite systems correlated with seawater aging but did not correlate so well with the mechanical properties. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011  相似文献   

19.
Electrically conductive nanocomposites of HCl‐doped polyaniline (PANI–HCl) nanocolloid particles with water‐soluble and film‐forming polymers such as poly(vinyl alcohol) (PVA) and methylcellulose (MC) were prepared by the redispersion of preformed MC‐coated submicrometric PANI–HCl particles in PVA and MC solutions under sonication for 1 h and the casting of the films from the dispersions followed by drying. The submicrometric polyaniline (PANI) particles were prepared by the oxidative dispersion polymerization of aniline in an acidic (1.25M HCl) aqueous ethanol (30 : 70) medium with MC as a steric stabilizer. The particles contained 4.7 wt % MC and had a conductivity of 7.4 S/cm. They had an oblong shape of 203 nm (length) and 137 nm (breadth). Sonication broke the oblong‐shaped particles to sizes of ~10 nm in the PVA matrix and ~60 nm in the MC matrix. The electrical conductivity of these films was measured, and the percolation threshold was determined. The composites had the characteristics of a low percolation threshold at a volume fraction of PANI of 2.5 × 10?2 in the PVA matrix and at a volume fraction of 3.7 × 10?2 in the MC matrix. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

20.
A novel process was developed to prepare electrically conducting maleic anhydride grafted polypropylene (gPP)/expanded graphite (EG) nanocomposites by solution intercalation. The conducting percolation threshold at room temperature (Φc) of the nanocomposites was 0.67 vol %, much lower than that of the conventional conducting composites prepared by melt mixing (Φc = 2.96 vol %). When the EG content was 3.90 vol %, the electrical conductivity (σ) of the former reached 2.49 × 10?3 S/cm, whereas the σ of the latter was only 6.85 × 10?9 S/cm. The TEM, SEM, and optical microscopy observations confirmed that the significant decrease of Φc and the striking increase of σ might be attributable to the formation of an EG/gPP conducting multiple network in the nanocomposites, involving the network composed of particles with a large surface‐to‐volume ratio and several hundred micrometers in size, and the networks composed of the boards or sheets of graphite with high width‐to‐thickness ratio and particles of fine microscale or nanoscale sizes. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 88: 1864–1869, 2003  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