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 890 毫秒
1.
Start up from rest and relaxation from steady shear flow experiments have been performed on monodisperse polystyrene solutions with molecular weight ranging from 1.3 × 105 to 1.6 × 106 and concentration c ranging from 5% to 40%. A method of reduced variables based on the use of a characteristic time τw is proposed. τw is defined as the product of zero shear viscosity with the steady state elastic compliance.Reduced steady and transient viscometric functions so obtained depend on the ratio M/Me (where Me is the entanglement molecular weight). Limiting forms are obtained when M/Me ? 18. In steady flow, a simple correlation is found between shear and normal stresses.In stress relaxation experiments, independent of shear rate, the long-time behaviour can be characterised by a single relaxation time τ1, which is identical for shear and normal stresses. τ1 can be simply related to the zero shear rate viscosity and the limiting elastic compliance.  相似文献   

2.
Flow visualization experiments have been carried out on these melts flowing from a reservoir into a capillary die. The existence and magnitude of vortices at the die entrance have been determined over a range of extrusion conditions. The vortex size is interpreted in terms of the theory of viscoelastic fluid mechanics. It is found that the second-order fluid-perturbation solution cannot represent the observed experimental results. The data are correlated with (i) a Weissenberg number τchVL\?gt(γ?w)γ?w ≡ Ψ1γ?w/2η  (N1)w/ 2(σ12)w measured at the die wall and (ii) with the deformation-rate dependence of relaxation time. Interpretation of vortex formation and size in terms of elongational viscosity is offered.Several polystyrene and polyethylene melts have been rheologically characterized as part of this study with measurements of viscosity η and principal normal stress difference N1. The zero shear viscosity η0 of the polystyrenes varies with the 3.5 power of the weight-average molecular weight Mw while the principal normal stress difference coefficient Ψ1 varies with the sixth power of Mw when evaluated at a shear rate of 1 sec?1.  相似文献   

3.
The observation made in Part 2 that squeezing flow with a superimposed rotation results in an equilibrium situation with the applied load just balancing the normal stresses generated in the test fluid is used to develop a new technique (the Torsional-Balance Rheometer) for measuring the viscometric functions of elastic liquids.The Rheometer utilizes conventional torsional flow and its novel feature is that the applied load is fixed and the associated shear rate at the rim determined, in contrast to the usual situation where the shear rate is fixed and the total normal force measured.It is argued that the Torsional Balance has significant advantages over other rheometers in the very high shear-rate range, since the normal stresses being measured themselves supply a mechanism for keeping the top plate (which is free to float on the test fluid) at a constant separation from the rotating bottom plate, hence allowing very small gaps to be considered. Consistent data are shown to be possible for shear rates in excess of 105 s?1.  相似文献   

4.
A novel pressure sensor plate (normal stress sensor (NSS) from RheoSense, Inc.) was adapted to an Advanced Rheometrics Expansion System rheometer in order to measure the radial pressure profile for a standard viscoelastic fluid, a poly(isobutylene) solution, during cone–plate and parallel-plate shearing flows at room temperature. We observed in our previous experimental work that use of the NSS in cone-and-plate shearing flow is suitable for determining the first and second normal stress differences N 1 and N 2 of various complex fluids. This is true, in part, because the uniformity of the shear rate at small cone angles ensures the existence of a simple linear relationship between the pressure [i.e., the vertical diagonal component of the total stress tensor (Π22)] and the logarithm of the radial position r (Christiansen and coworkers, Magda et al.). However, both normal stress differences can also be calculated from the radial pressure distribution measured in parallel-plate torsional flows. This approach has rarely been attempted, perhaps because of the additional complication that the shear rate value increases linearly with radial position. In this work, three different methods are used to investigate N 1 and N 2 as a function of shear rate in steady shear flow. These methods are: (1) pressure distribution cone–plate (PDCP) method, (2) pressure distribution parallel-plate (PDPP) method, and (3) total force cone–plate parallel-plate (TFCPPP) method. Good agreement was obtained between N 1 and N 2 values obtained from the PDCP and PDPP methods. However, the measured N 1 values were 10–15% below the certified values for the standard poly(isobutylene) solution at higher shear rates. The TFCPPP method yielded N 1 values that were in better agreement with the certified values but gave positive N 2 values at most shear rates, in striking disagreement with published results for the standard poly(isobutylene) solution.
J. J. MagdaEmail:
  相似文献   

5.
Higashitani and Pritchard (H-P) carried out an analysis of the hole pressure (PH) for viscoelastic fluids which leads to expressions relating PH to the shear stress (σ), the wall shear stress (σw), and the normal stress differneces (N1 and N2). Although very good agreement their theory and experimental results has been obtained for several polymer solutions and three polymer melts, it is known that at least two of the key assumptions in the theory are violated. In this study flow birefingence has been used to determine the stress field (i.e. σ and the normal stress difference, σ11 — σ22) in the region of a slot placed perpendicular to the flow direction for a polystyrene melt. Values of σ and σ11 — σ22 were then used to evaluate the integrand in the expression relating Pe1 where Pe1 is the hole pressure measured at the base of the rectangular slot, σ and N1. Values of Pe1 evaluated using flow birefringence data agreed well with those obtained using the same integral expression and cone-and-plate values of N1 and σ, and with directly measured values of Pe1. This agreement occurred even though the stress field was found to be asymmetric around the centerline of the slot and with secondary flow in the slot. A detailed evaluation of the values of N1/2σ, which constitute the integral in the H-P theory, along the centerline of the slot revealed most of the contributions to the integral canceled in integrating from the base of the slot where σ is zero at the centerline of the slit-die. The main contributions to the integral occurred from the integration taken from the centerline of the slit-die to the upperdie  相似文献   

6.
Relaxation of the second normal stress difference (N 2) following step strain of a concentrated monodisperse polystyrene solution has been studied using mechanical and optical rheometry. Measurements of normal thrust in a parallel plate geometry are corrected for strain inhomogeneity and combined with independent measurements of the first normal stress difference (N 1) to determine N 2. Optical experiments were performed using a novel configuration where flow birefringence data collected using multiple light paths within the shear plane are combined with the stress-optical law to determine all three independent stress components for shearing deformations. This technique eliminates end effects, and provides an opportunity to oversample the stress tensor and develop consistency checks of experimental data. N 2 is found to be nonzero at all accessible times, and relaxes in roughly constant proportion to N 1. This reflects nonaffine distribution of chain segments, even well within the regime of chain retraction at short times. Data collected with the two techniques are reasonably consistent with each other, and with results of previous studies, generally lying between the predictions of the Doi-Edwards model with and without the independent alignment approximation. The normal stress ratio –N 2/N 1 shows pronounced strain thinning in the nonlinear regime.  相似文献   

7.
We investigate the behavior of flow variables, thermodynamic variables and their interaction in rapidly sheared (S) homogeneous compressible turbulence using rapid distortion theory (RDT). We subject an initially isotropic and incompressible flow field to homogeneous shear-rate of various strengths quantified by a gradient Mach number (M g ) based on characteristic wavenumber. Our objective is to characterize the behavior of flow/thermodynamic fluctuations and their linear interactions during the course of turbulence evolution. Even though the mean shear-rate is held constant, the gradient Mach number progressively diminishes with time as the relevant wavenumber increases due to the mean deformation. The evolution exhibits three distinct phases which we categorize based on the character of pressure as: (i) Pressure-released (PR) stage which is observed when ${St < \sqrt{M_{g0}}}$ and pressure effects are negligible; (ii) Wave-character (WC) stage wherein ${\sqrt{M_{g0}} < St < M_{g0}}$ and the wave character of pressure is in evidence; and (iii) Low-Mach number (LM) stage when St > M g0, where M g0 is the initial gradient Mach number. In the PR regime we find that the thermodynamic fluctuations evolve from their initial state but velocity fluctuations grow unhindered by pressure fluctuations. In the WC regime, the pressure fluctuations become significant and flow-thermodynamic interaction commences. This interaction brings about equipartition of dilatational kinetic energy and thermodynamic potential energy. The interaction also results in stabilization of turbulence, and the total kinetic energy growth comes to a near standstill. Ultimately in the LM stage, kinetic energy starts increasing again with the growth rate being very similar to that in incompressible RDT. However, the thermodynamic fluctuations continue to grow despite the gradient Mach number being substantially smaller than unity. Overall, the study yields valuable insight into the linear processes in high Mach number shear flows and identifies important closure modeling issues.  相似文献   

8.
The use of a sliding plate rheometer (SPR) to determine the first normal stress difference of molten polymers and elastomers at high shear rates is demonstrated. The simple shear flow in this instrument is not subject to the flow instabilities that limit the use of rotational rheometers to shear rates often below 1 s−1. However, issues of secondary flow and wall slip must be addressed to obtain reliable data using an SPR. A highly entangled, monodisperse polybutadiene and a commercial polystyrene were the polymers studied. The inclusion of the polystyrene made it possible to compare data with those obtained by Lodge using a stressmeter, which is an instrument based on the measurement of the hole pressure. The data from the two instruments are in good agreement and are also close to the predictions of an empirical equation of Laun based on the storage and loss moduli.  相似文献   

9.
The MTR 25 is a multitask rheometer (for shear and squeeze flow) with 25 kg of normal force and a partitioned plate. Torque and normal force are measured at both, the inner disk and the outer ring of the plate. The first and second normal stress differences can be determined from a single test. The axial stiffness is high (107 N/m) by using rigid springs and strain gauges for the load cell. Monodisperse polystyrene (M w = 206 kg/mol, 180°C) has been sheared in the range from 0.05 to 47 s − 1. The viscosity and first normal stress difference are highly reproducible. The second normal stress difference scatters and mirrors the instability at the rim. A critical comparison is made between the MTR 25 method and the single transducer evaluation method (RMS 800 method, Schweizer, Rheol Acta 41:337–344, 2002): Both yield excellent and coinciding viscosity and first normal stress difference data. The RMS 800 method gives more stable second normal stress difference data, since the normal force from the outer ring, which is influenced by edge fracture, is not used. Data for the RMS 800 method can be acquired on the MTR 25. The high normal force capacity permits larger samples and higher shear rates than on the RMS 800.
Thomas SchweizerEmail:
  相似文献   

10.
Rheological characterizations were carried out for two polystyrenes. One was a linear polymer with M w =222,000 g/mol and M w /M n =2, while the other was a randomly branched polystyrene with M w =678,000 g/mol and a broad molecular weight distribution. Experiments performed included oscillatory shear to determine the storage and loss moduli as functions of frequency and temperature, viscosity as a function of shear rate and pressure, and multi-angle light scattering to determine the radius of gyration as a function of molecular weight. The presence of branching in one sample was clearly revealed by the radius of gyration and the low-frequency portion of the complex viscosity curve. Data are also shown for three polyethylene copolymers, one (LLDPE) made using a Ziegler catalyst and two made using metallocene catalysts, one (BmPE) with and one (LmPE) without long-chain branching (LCB). While the distribution of comonomer is known to be much more uniform in LmPE than in LLDPE, the pressure shift factors were the same for these two polymers. The pressure and temperature shift factors of the two polystyrenes were identical, but, in the case of polyethylene, the presence of a small amount of LCB in the BmPE had a definite effect on the shift factors. These observations are discussed in terms of the relative roles of free volume and thermal activation in the effects of temperature and pressure.  相似文献   

11.
Pressure distribution measurements for a polyisobutylene/decalin solution D1 in the Truncated Cone-and-Plate (TCP) apparatus are combined with elastic hole pressures obtained for the same solution on the Lodge Stressmeter® in order to provide two independent estimates of the second normal-stress difference (N 2). The values ofN 2 from the TCP apparatus, obtained by numerical differentiation of a function of the center-hole pressure and the pressure gradient, are in good agreement with measurements made on the same sample by Tanner et al. with a direct method, namely the Tilted Trough Experiment, and by Christiansen et al. with a method that requires an extrapolation to the pressure at the free surface of coneand-plate rheogoniometer data obtained with flush-mounted pressure transducers. The viscosities from the modified Stressmeter for low shear rates extend over five decades of shear rate, including a zero-shear-rate region, and agree with the data of Christiansen on a torque-driven flow. The Higashitani-Pritchard-Baird-Lodge (HPBL) equation relatingN 1N 2 to the hole pressure gives good agreement with the data over a certain range of shear stress. The Newtonian hole pressures for several liquids at 20 and 46 °C compare well with a finite-element calculation for a two-dimensional Poiseuille flow. When the elastic hole pressures from the Stressmeter are combined with the extrapolated rim pressures from the TCP Apparatus in order to extract the value ofN 2, an agreement betweenN 2 from the center-hole pressure andN 2 from the rim pressure can only be obtained up to a shear rate of about 40 s–1, beyond which the value of –N 2 from the rim pressure diverges abruptly to negative values. It is possible that this constitutes the first quantitative estimate of an edge effect in cone-and-plate rheometry. Alternatively, the elastic hole pressure in cone-and-plate flow is not equivalent to the elastic hole pressure in Poiseuille flow, at least at high shear rates. The data of Christiansen et al. with flush-mounted pressure transducers appear to confirm this second possibility. Finally, a single set of shift factors obeying the Williams, Landel and Ferry equation superposes the viscosity, the first and the second normal-stress difference within experimental scatter, which can be less than 1% for a certain combination of normal-stress differences. The data were recorded at 3, 20, 30, and 46 °C in the shear rate range 1–260 s–1.  相似文献   

12.
Some correlations involving the shear viscosity of polystyrene melts   总被引:1,自引:0,他引:1  
Based upon a compilation of steady-shear and dynamic-shear viscosity data from the literature for polystyrene melts, an assessment has been made concerning the relative merits of the Cross and Carreau models in describing the shear-rate dependence of such viscosities. It is shown that the Cross model is decidedly more appropriate for PS of BMWD. Based upon master plots, it is demonstrated that the Cox-Merz relation applies to PS of both BMWD and NMWD. It is also shown that the Cox-Merz relation applies even into the second-Newtonian regime, with being independent ofM w and MWD. In addition, the applicability of the Prest-Porter-O'Reilly relationship between shear viscosity and recoverable shear compliance is corroborated in the case of PS of NMWD.  相似文献   

13.
Shear-banding phenomenon in the entangled polymer systems was investigated in a planar Couette cell with the diffusive Rolie-Poly (ROuse LInear Entangled POLYmers) model, a single-mode constitutive model derived from a tube-based molecular theory. The steady-state shear stress ?? s was constant in the shear gradient direction while the local shear rate changed abruptly, i.e., split into the bands. We focused on the molecular conformation (also calculated from the Rolie-Poly model) around the band boundary. A band was found also for the conformation, but its boundary was much broader than that for the shear rate. Correspondingly, the first normal stress difference (N 1) gradually changed in this diffuse boundary of the conformational bands (this change of N 1 was compensated by a change of the local pressure). For both shear rate and conformation, the boundary widths were quite insensitive to the macroscopic shear rate but changed with various parameters such as the diffusion constant and the relaxation times (the reptation and the Rouse times). The broadness of the conformational banding, associated by the gradual change of N 1, was attributed to competition between the molecular diffusion (in the shear gradient direction) and the conformational relaxation under a constraint of constant ?? s.  相似文献   

14.
We propose a simple, robust method to measure both the first and second normal stress differences of polymers, hence obtaining the full set of viscometric material functions in nonlinear shear flow. The method is based on the use of a modular cone-partitioned plate (CPP) setup with two different diameters of the inner plate, mounted on a rotational strain-controlled rheometer. The use of CPP allows extending the measured range of shear rates without edge fracture problems. The main advantage of such a protocol is that it overcomes limitations of previous approaches based on CPP (moderate temperatures not exceeding 120 °C, multiple measurements of samples with different volume) and yields data over a wide temperature range by performing a two-step measurement on two different samples with the same volume. The method was tested with two entangled polystyrene solutions at elevated temperatures, and the results were favorably compared with both the limited literature data on the second normal stress difference and the predictions obtained with a recent tube-based model of entangled polymers accounting for shear flow-induced molecular tumbling. Limitations and possible improvements of the proposed simple experimental protocol are also discussed.
Graphical abstract The effects of edge fracture in start-up shear experiments can be circumvented with the use of a cone-partitioned plate (CPP) geometry. Such a device consists of an inner measuring plate surrounded by an outer nonmeasuring corona. The radius of the sample exceeds that of the measuring plate so that the measured volume is not affected by edge instability. However, the measured first normal stress difference is an apparent one (Napp,1), owing to the contribution of the nonmeasured part of the sample. The figure depicts a schematic design of a modular CPP geometry. Such a fixture is built in a way that the inner tool and the outer partition can be easily replaced, in order to have different measuring diameters (i.e., 6 and 10 mm). From the corresponding signals of the Napp,1, the effective first and second normal stress differences can be calculated.
  相似文献   

15.
Nonlinear viscoelasticity of PP/PS/SEBS blends   总被引:1,自引:0,他引:1  
The nonlinear viscoelastic behavior of polypropylene/polystyrene (PP/PS) blends compatibilized or not with the linear triblock copolymer (styrene-ethylene-/butylene-styrene, SEBS) was investigated. Start-up of steady-shear at rates from 0.1 to 10 s–1 was carried out using a controlled strain rotational rheometer and a sliding plate rheometer for strain histories involving one or several shear rates. The shear stress and first normal shear stress difference were measured as functions of time, and the morphologies of the samples before and after shearing were determined. For each strain history except that involving a single shear rate of 0.1 s–1 the blends showed typical non-linear viscoelastic behavior: a shear stress overshoot/undershoot, depending on the history, followed by a steady state for each step. The first normal stress difference increased monotonically to a steady-state value. The values of the stresses increased with the addition of SEBS. The shear stress overshoot and undershoot and the times at which they occurred depended strongly on the strain history, decreasing for a subsequent shear rate step performed in the same direction as the former, and the time at which stress undershoot occurred increased for a subsequent shear rate step performed in the opposite direction, irrespective of the magnitude of the shear rate. This behavior was observed for all the blends studied. The time of overshoot in a single-step shear rate experiment is inversely proportional to the shear rate, and the steady-state value of N1 scaled linearly with shear rate, whereas the steady-state shear stress did not. The average diameter of the dispersed phase decreased for all strain histories when the blend was not compatibilized. When the blend was compatibilized, the average diameter of the dispersed phase changed only during the stronger flows. Experimental data were compared with the predictions of a model formulated using ideas of Doi and Ohta (1991), Lacroix et al. (1998) and Bousmina et al. (2001). The model correctly predicted the behavior of the uncompatibilized blends for single-step shear rates but not that of the compatibilized blends, nor did it predict morphologies after shearing.  相似文献   

16.
The shear viscosity of clay-based coating colors containing latex and carboxymethyl cellulose (CMC) has been measured over a relatively large shearrate region. In the shear-rate range of 50–1500 s–1 the measurements were performed using a rotational viscometer and, at higher shear rates extending into the region 105 – 106 s–1, a high pressure capillary viscometer was employed. The viscosity of the clay colors increased with increasing CMC-concentration, but the influence of the CMC-content was less pronounced at higher shear rates. The apparent shear-thinning behavior of the investigated colors could, in part, be attributed to the shear-thinning of the corresponding polymer (CMC) solution constituting the liquid phase of the color, but the influence of another factor was also indicated. At low shear rates, the interaction between the color components can produce relatively high viscosity levels, but in the high shear rate region these interactions appear to be less important for the viscosity level. It is also of interest to note that the viscosity dependence on the solids content in the high shear-rate region could be described with reasonable accuracy using an empirical equation neglecting interactions between the color components.  相似文献   

17.
Schweizer et al. [J Rheol 48(6):1345–1363, 2004] showed nonlinear step shear rate data for a polystyrene melt (M w=200 kg/mol, M w/M n=1.06, T=175°C). For different rheometers, cone angles, and sample sizes, the delayed normal force rise observed therein relative to a compliance-free reference N 1 (from a thermodynamically consistent reptation model) is shown to depend on rheometer compliance characterized by the instrument stiffness K A. K A can be obtained from mapping N 1 on the measured N 1,meas. or directly from mechanical contact measurement with a mismatch of 20–30%. The ranking of the stiffnesses found is K A(RMS 800)>K A(MCR 300)>K A(ARES LR2). Once K A is known, N 1,meas.-data can be corrected by solving the ill-posed Volterra equation involved in it. The correction shown for experiments with the 0.15-rad cone angle gives very good results. The characteristic decay time of the normal force after cessation of flow scales linearly with the axial response time t a calculated from K A, cone angle, and sample radius. The torque decay time is practically independent of t a.Extended Version of a paper presented at the 2nd Annual European Rheology Conference in Grenoble, France, April 21–25, 2005.  相似文献   

18.
The rheological properties of seawater with the addition of surfactant additive (cetyltrimethyl ammonium chloride (CTAC)/sodium salicylate (NaSal)) are measured at different temperatures, including shear viscosity and first normal stress difference (N1). The effects of the temperature, the salts, and CTAC/NaSal concentration on the rheological properties of test solutions are investigated, and the corresponding influence mechanisms are analyzed. It shows that the addition of salt can decrease the shear viscosities of the solutions, and also decrease N1 and even eliminate the sharp augment of N1 above a certain shear rate. The growing elasticity can be characterized by the increase of the initial shear rate for shear-thickening inception. High temperature can also remove the sharp increase of N1 with salt. Nevertheless, the increase of CTAC/NaSal concentration can withstand the elimination of the sharp augment of N1.  相似文献   

19.
Linear high-density polyethylenes with molar masses M w between 240 and 1,000,000 g/mol, obtained by metallocene catalysts, were characterized in shear using oscillatory and creep tests. The polydispersities of the molar mass distributions (MMDs) lay between 1 and 16. The resulting zero shear-rate viscosities η0 covered a range from 2.5×10−3 to around 108 Pas. Above a critical molar mass of M c≈2,900 g/mol, the experimental results can be described by the relation η0M w3.6, independently of the MMD. The oscillatory data were fitted with a Carreau–Yasuda equation. The resulting parameters were correlated to molecular structure. The parameter a, being a quantity for the width of the transition between the Newtonian and the non-Newtonian regime, showed a dependence on the molar mass M w but not on M w/M n. The parameter λ of the Carreau-Yasuda equation was found to be the reciprocal crossover frequency for all samples with a log-Gaussian MMD. λ depends on the molar mass M w and also on M w/M n.
Helmut MünstedtEmail: Phone: +49-9131-8527604Fax: +49-9131-8528321
  相似文献   

20.
In this study, we studied the contact line motion of second-order fluids theoretically and experimentally. The theoretical study showed that the positive first normal stress difference (N 1) increases the contact line velocity while the second normal stress difference (N 2) does not affect the contact line motion. The increased contact line velocity is caused by the hoop stress acting on the curved stream lines near the contact line. The hoop stress increases the liquid pressure near the contact line, and the increased pressure changes the surface profile to have the smaller curvature and smaller dynamic contact angle. The contribution of N 1is 1 order of magnitude smaller than the contribution from the viscous component when the Deborah number remains O(1). For experiments, silicone oils of different kinematic viscosities (1,000–200,000 mm 2/s) were used while eliminating the drying problem and shear-thinning effect near the contact line. The silicone oils were well fitted to the second-order fluid model with the positive first normal stress difference. The spreading rate of a silicone oil drop on a solid surface was faster than the spreading rate predicted by the theory for Newtonian fluids. As the theory predicts that N 1increases the contact line velocity and the experimental result confirms the theoretical prediction, the effect of N 1is established.  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