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 93 毫秒
1.
The paper presents average flow visualizations and measurements, obtained with the Particle Image Velocimetry (PIV) technique, of a submerged rectangular free jet of air in the range of Reynolds numbers from Re = 35,300 to Re = 2200, where the Reynolds number is defined according to the hydraulic diameter of a rectangular slot of height H. According to the literature, just after the exit of the jet there is a zone of flow, called zone of flow establishment, containing the region of mixing fluid, at the border with the stagnant fluid, and the potential core, where velocity on the centerline maintains a value almost equal to the exit one. After this zone is present the zone of established flow or fully developed region. The goal of the paper is to show, with average PIV visualizations and measurements, that, before the zone of flow establishment is present a region of flow, never mentioned by the literature and called undisturbed region of flow, with a length, LU, which decreases with the increase of the Reynolds number. The main characteristics of the undisturbed region is the fact that the velocity profile maintains almost equal to the exit one, and can also be identified by a constant height of the average PIV visualizations, with length, LCH, or by a constant turbulence on the centerline, with length LCT. The average PIV velocity and turbulence measurements are compared to those performed with the Hot Film Anemometry (HFA) technique. The average PIV visualizations show that the region of constant height has a length LCH which increases from LCH = H at Re = 35,300 to LCH = 45H at Re = 2200. The PIV measurements on the centerline of the jet show that turbulence remains constant at the level of the exit for a length, LCT, which increases from LCT = H at Re = 35,300 to LCT = 45H at Re = 2200. The PIV measurements show that velocity remains constant at the exit level for a length, LU, which increases from LU = H at Re = 35,300 to LU = 6H at Re = 2200 and is called undisturbed region of flow. In turbulent flow the length LU is almost equal to the lengths of the regions of constant height, LCH, and constant turbulence, LCT. In laminar flow, Re = 2200, the length of the undisturbed region of flow, LU, is greater than the lengths of the regions of constant height and turbulence, LCT = LCH = 45H. The average PIV and HFA velocity measurements confirm that the length of potential core, LP, increases from LP = 45H at Re = 35,300 to LP = 78H at Re = 2200, and are compared to the previous experimental and theoretical results of the literature in the zone of mixing fluid and in the fully developed region with a good agreement.  相似文献   

2.
For the traction boundary value problem in finite elasticity, a bound is obtained for the total strain energy in terms of the L2 integral norms of the surface tractions and body forces, under the assumptions that the unstressed body occupies a convex domain and the displacement gradients are sufficiently small.This is an extension of known results in linear (infinitesimal) elasticity into finite elasticity.  相似文献   

3.
For the boundary value problem in finite elasticity in which nonzero tractions are given on a connected subdomain of the boundary, the rest of the boundary is stress-free, and there are no body forces, a bound is obtained for the strain energy in terms of the L 2 integral norm of the surface tractions with the constant involved depending only upon and the material constants.The result is obtained in the context of finite elasticity under the assumptions that the unstressed body occupies a convex domain and the displacement gradients are sufficiently small. In the context of the linear theory, the same result is obtained without these assumptions.
Zusammenfassung Wir betrachten ein Randwertproblem in der nichtlinearen Elastizität in dem nur ein zusammenhängendes Teilgebiet der Randfläche belastet ist, sonst aber keine Randbelastung oder Körperkräfte vorhanden sind. Eine Schranke für die Verzerrungsenergie wird mittels der L 2 Integralnorme der Randbelastung hergeleitet, wobei die auftretende Konstante nur von dem Teilgebiet und von den Eigenschaften des Materials abhängig ist.Das Ergebnis gilt für die nichtlinearen Elastizität, unter den Vorraussetzungen dass das unbelastete Material ein konvexes Gebiet besetzt und dass die Verschiebungsgradiente hinreichend klein sind. Das gleiche Ergebnis gilt in der linearen Elastizität ohne diese Vorraussetzungen.


The second author was a visitor at Georgia Institute of Technology, School of Mathematics, at the time that the revised version was prepared.  相似文献   

4.
For the displacement boundary value problem in nonlinear elastostatics with zero body force, an integral bound for the strain energy is obtained in terms of theL 2-norms of the given boundary displacements and their tangential derivatives (assumed sufficiently small). The constants involved depend upon the strain energy density function and upon the geometry of the domain.  相似文献   

5.
A closed-form model for the computation of heat transfer rates through the rectangular-partitioned enclosures is investigated. The rectangular-partitioned enclosures may contain solid or gas with or without a constant and uniformly distributed heat generation. The conduction in the enclosures is considered as two-dimensional, whereas one-dimensional heat transfer through the fin-type partition is assumed. Dimensionless heat flux plots are parametrically studied by varying the aspect ratio (L/H) of the enclosure, the ratio of thermal conductivities of the enclosure to the fin-type partition (k a /k f ), the Biot number (β a =h a L/k a ), and the reduced partition thickness (t */L). It is demonstrated through an example problem that there is a large error in using one-dimensional analysis, particularly at lower values of k a /k f , and β a .  相似文献   

6.
An analysis is made of the unsteady lift exerted on a stationary rigid body immersed in an incompressible, plane-wall turbulent boundary layer. The lift is expressed as a surface integral over the body involving theupwash velocity induced by the “free” vorticity Ω (found by taking explicit account of the interaction of the body with the flow and excluding the bound vorticity) and a harmonic function X2that depends only on the shape of the body. The upwash velocity is the free-field velocity given in terms of Ω by the Biot–Savart formula, augmented by the velocity field of a conventional distribution of image vortices in the wall. The function X2can be interpreted as the velocity potential of flow past the body, produced by motion of the wall at unit speed towards the body. Detailed predictions are made of the lift on a slender airfoil placed in the outer region of the boundary-layer. When the airfoil chord is large compared to the boundary-layer thickness, vortex shedding into the wake causes the magnitude of the net upwash velocity near the trailing edge to be small. The main contributions to the surface integral are then from the nose region, where the upwash velocity may be estimated independently of the fluctuations near the trailing edge. Analytical results for a thin plate airfoil of chord 2a at distance h from the wall show that the lift increases as a/h increases; it is ultimately independent of a and scales with the ratio of h to the hydrodynamic wavelength. Application is made to determine the sound generated by the airfoil in a weakly compressible boundary layer flow over a finite elastic plate.  相似文献   

7.
The current state-of-the-art for estimating the maximum rise time of a strain gage which is subjected to an axially sweeping strain pulse ist rg≤0.8L/c+0.5μsec wheret rg is the 10/90 rise time of the strain gage,* L is the gage length andc is the strain-pulse velocity. This paper shows that the effect of the 0.8L/c term can be significantly reduced by utilizing an analytical compensation technique. In addition, it is shown that the 0.5 μsec additive constant can be reduced to approximately 0.1 μsec by reinterpreting data published by a previous author.  相似文献   

8.
This paper is concerned with the prediction of micro-bubble dissolution characteristics in water and seawater when microbubbles are generated by a Sadatomi-type micro-bubble generator (2003) with a spherical body in a flowing liquid tube. In the experiments, in order to know the effects of the salinity on the characteristics, tap water and an artificial seawater with different salt concentrations of 1 and 3 wt% were used as the test liquids. Parameters measured were the Sauter mean diameter of bubbles, dBS, the void fraction, α, the rising velocity of bubbles, uG, the interfacial area concentration, a, the volumetric mass transfer coefficient, KLa, and the liquid-side mass transfer coefficient, KL. In the analysis, for predicting α, KLa and KL, some correlations in the literatures were tested against the present data. Furthermore, in order to improve the predictability, new correlations were developed based on the present data. The prediction of KLa with the new correlation agreed well with Nishino et al.’s [T. Nishino, K. Terasaka, M. Ishida, Application for several micro-bubble generators for gas absorber, in: Proceedings of the Annual Meeting of the Japanese Society for Multiphase Flow, 2006, pp. 276–277 (in Japanese)] and Li and Tsuge’s [P. Li, H. Tsuge, Water treatment by induced air flotation using microbubbles, Journal of Chemical Engineering of Japan 39 (2006) 896–903; P. Li, H. Tsuge, Ozone transfer in a new gas-induced contactor with microbubbles, Journal of Chemical Engineering of Japan 39 (2006) 1213–1220] data for different aeration systems using several different micro-bubble generators.  相似文献   

9.
A hydraulic jump is characterized by a highly turbulent flow with macro-scale vortices, some kinetic energy dissipation and a bubbly two-phase flow structure. New air–water flow measurements were performed in a large-size facility using two types of phase-detection intrusive probes: i.e. single-tip and double-tip conductivity probes. These were complemented by some measurements of free-surface fluctuations using ultrasonic displacement meters. The void fraction measurements showed the presence of an advective diffusion shear layer in which the void fractions profiles matched closely an analytical solution of the advective diffusion equation for air bubbles. The free-surface fluctuations measurements showed large turbulent fluctuations that reflected the dynamic, unsteady structure of the hydraulic jumps. The measurements of interfacial velocity and turbulence level distributions provided new information on the turbulent velocity field in the highly-aerated shear region. The velocity profiles tended to follow a wall jet flow pattern. The air–water turbulent integral time and length scales were deduced from some auto- and cross-correlation analyses based upon the method of Chanson [H. Chanson, Bubbly flow structure in hydraulic jump, Eur. J. Mech. B/Fluids 26 (3) (2007) 367–384], providing the turbulent scales of the eddy structures advecting the air bubbles in the developing shear layer. The length scale Lxz is an integral air–water turbulence length scale which characterized the transverse size of the large vortical structures advecting the air bubbles. The experimental data showed that the dimensionless integral turbulent length scale Lxz/d1 was closely related to the inflow depth: i.e. Lxz/d1 = 0.2–0.8, with Lxz increasing towards the free-surface.  相似文献   

10.
This paper is concerned with constitutive equations in which the stress at an instant of time,t, say, is a functional of the deformation gradient history up to and includingt. It is assumed that the deformation gradient histories are of bounded variation and are continuous, or piece-wise continuous with a countable number of salti. It is shown that if the stress functional isn times Fréchet differentiable in the sense of the supremum norm, it can be represented by a series of multiple integrals in the deformation gradient history, with an error or ordern + 1 in the displacement gradients. The relation of this type of integral representation to that of Green and Rivlin is discussed.  相似文献   

11.
The direct identification of the cohesive law in pure mode I of Pinus pinaster is addressed in this work. The approach couples the double cantilever beam (DCB) test with digital image correlation (DIC). Wooden beam specimens loaded in the radial-longitudinal (RL) fracture propagation system are used. The strain energy release rate in mode I (G I) is uniquely determined from the load–displacement curve by means of the compliance-based beam method (CBBM). This method relies on the concept of equivalent elastic crack length (a eq) and therefore does not require the monitoring of crack propagation during test. DIC measurements are processed with two different purposes. Firstly, the physical evidence of a eq is discussed with regard to actual estimation of the crack length based on post-processing full-field displacement measurements. Secondly, the crack tip opening displacement in mode I (w I) is determined from the displacements near the initial crack tip. The cohesive law in mode I (σ I???w I) is then identified by numerical differentiation of the G I???w I relationship. The methodology and accuracy on this reconstruction are addressed. Moreover, the proposed procedure is validated by finite element analyses including cohesive zone modelling. It is concluded that the proposed data reduction scheme is adequate for assessing the cohesive law in pure mode I of P. pinaster.  相似文献   

12.
A theory for the prediction of the size dependence of torsional rigidities of nanosized structural elements is developed. It is shown that, to a very good approximation, the torsional rigidity (D) of a nanosized bar differs from the prediction of standard continuum mechanics (Dc) as (DDc)/Dc=Ah0/a where A is a non-dimensional constant, a is the size scale of the cross-section of the bar and h0 is a material length equal to the ratio of the surface elastic constant to the bulk elastic constant. The theory developed is compared with direct atomistic calculations (“numerical experiment”) of the torsional rigidity bars made of several FCC metals modeled using the embedded atom method. Very good agreement is obtained between theory and simulation. The framework presented here can aid the development of design methodologies for nanoscale structural elements without the need for full scale atomistic simulations.  相似文献   

13.
In general Stokes problems, no boundary conditions exist for the pressure. But pressure is an L2(Ω) function and can uniquely be represented as the divergence of a precisely defined vector field. In the 2-D case, this vector field can in turn be represented as the sum of a gradient (of a pressure-potential) and the curl of a second scalar potential. The latter potential is entirely determined by the first one. A variational equation is obtained for such pressure potential class, which exists and is uniquely characterized. This variational problem is well-posed. Finite element approximations can easily be realized and ensure high convergence rates for the L2(Ω) norm of the pressure.  相似文献   

14.
Many disperse systems show a typical non-Newtonian flow at relatively high concentrations of the disperse particle. However, two Newtonian viscosities η and η0 can be, respectively, determined at high and low rates of shear. Expect for very low particle content, ηs is proportional to exp(mϕ), where ηs is a medium viscosity, m a constant which might reflect the particle-particle interaction and ϕ the volume fraction. In considering this relationship, a new type of equation which describes the relation between the zero shear relative viscosity ηr0( ≡ η0s) of the disperse system and ϕ is proposed as follows. ln ηr0 = A(p)ϕ + am3ϕ2, where A(p), the Einstein-Simha constant, is a function of the axial ratio p of dispersing particles, and a is a constant (⋍ 0.03) which depends slightly on the particle shape.The equation has been compared with the experimental results obtained for several disperse systems. A number of disperse systems of spherical particles are described well by the choice A(p) = 2.5 and a = 0.027, and a system of rod-like particles with p = 50 by the choice A(p) = 215.6 and a = 0.033. m for rod-like particles is larger than that for spherical particles.  相似文献   

15.
Nanofibers used in current ceramic matrix composites (CMCs) are usually wavy and of finite length. Here, a shear-lag model for predicting the tensile strength and work of fracture of a composite containing a “single matrix crack”, as a function of all the material parameters including constant plus Coulomb interfacial friction, is presented for a CMC containing wavy, finite-length nanofibers having a statistical distribution of strengths. Literature results are recovered for straight infinite fibers, with strength and toughness depending on a characteristic strength σc and a characteristic length δc. For nanofibers of finite length L, radius of curvature R, and with Coulomb friction coefficient μ, the strength and toughness are found to depend only on σc, δc, and two new dimensionless parameters μδc/R and L/δc. Parametric results for a typical carbon nanotube CMC show an optimal region of morphology (L and R) that maximizes composite strength and toughness, exceeding the properties of composites with straight (R=∞) and/or long (L=∞) reinforcements. Therefore, while factors such as the nanofiber strength distribution and the nanofiber–matrix interface sliding resistance may not be easily controlled, the tuning, via processing, of purely geometrical properties of the nanofibers (L and R) alone can aid in maximizing composite properties. These results have important application in hybrid CMCs such as “fuzzy fiber” CMCs, where micron-scale fibers are covered with a forest of nanofibers such that the nanofiber array can bridge longitudinal cracks and thus improve delamination resistance.  相似文献   

16.
The strain gradient exists near a crack tip may significantly influence the near-tip stress field. In this paper, the strain gradient and the internal length scales are introduced into the basic equations of mode III crack by the modified gradient elasticity (MGE). By using a complex function approach, the analytical solution of stress fields for mode III crack problem is derived within MGE. When the internal length scales vanish, the stress fields can be simplified to the stress fields of classical linear elastic fracture mechanics. The results show that the singularity of the shear stress is made up of two parts, r−1/2 part and r−3/2 part, and the sign of the stress σyz changes. With the increase of lx, the peak value of σyz decrease and its location moves farther from the fracture vertex. The influence of strain gradient for mode III crack problem cannot be ignored.  相似文献   

17.
Crack propagation and coalescence processes are the fundamental mechanisms leading to progressive failure processes in rock masses, in which parallel non-persistent rock joints are commonly involved. The coalescence behavior of the latter, which are represented as pre-existing coplanar flaws (cracks), is numerically investigated in the present study. By using AUTODYN as the numerical tool, the present study systematically simulates the coalescence of two pre-existing coplanar flaws in rock under compression. The cumulative damage failure criterion is adopted in the numerical models to simulate the cumulative damage process in the crack initiation and propagation. The crack types (shear or tensile) are identified by analyzing the mechanics information associated with the crack initiation and propagation processes. The simulation results, which are generally in a good accordance with physical experimental results, indicate that the ligament length and the flaw inclination angle have a great influence on the coalescence pattern. The coalescence pattern is relatively simple for the flaw arrangements with a short ligament length, which becomes more complicated for those with a long ligament length. The coalescence trajectory is composed of shear cracks only when the flaw inclination angle is small (such as β ? 30°). When the pre-existing flaws are steep (such as β ? 75°), the coalescence trajectory is composed of tensile cracks as well as shear cracks. When the inclination angle is close to the failure angle of the corresponding intact rock material, and the ligament length is not long (such as L ? 2a), the direct shear coalescence is the more favorable coalescence pattern. In the special case that the two pre-existing flaws are vertical, the model will have a direct tensile coalescence pattern when the ligament length is short (L ? a), while the coalescence between the two inner flaw tips is not easy to achieve if the ligament length is long (L ? 2a).  相似文献   

18.
Transition from gradient to countergradient scalar transport in a statistically planar, one-dimensional, developing, premixed turbulent flame is studied both theoretically and numerically. A simple criterion of the transition referred to is derived from the balance equation for the combustion progress variable, with the criterion highlighting an important role played by flame development. A balance equation for the difference in velocities $\bar{u}_b$ and $\bar{u}_u$ conditioned on burned and unburned mixture, respectively, is numerically integrated. Both analytical and computed results show that; (1) The flux $\overline{\rho u'' c''}$ is gradient during an early stage of flame development followed by transition to countergradient scalar transport at certain instant t tr . (2) The transition time is increased when turbulence length scale L is increased or when the laminar flame speed S L and/or the density ratio are decreased. (3) The transition time normalized using the turbulence time scale is increased by u??. Moreover, the numerical simulations have shown that the transition time is increased by u?? if a ratio of u??/S L is not large. This dependence of t tr on u?? is substantially affected by (i) the mean pressure gradient induced within the flame due to heat release and (ii) by the damping effect of combustion on the growth rate of mean flame brush thickness. The reasonable qualitative agreement between the computed trends and available experimental and DNS data, as well as the agreement between the computed trends and the present theoretical results, lends further support to the conditioned balance equation used in the present work.  相似文献   

19.
Lift coefficients, CL, of single bubbles in linear shear flows are measured to investigate effects of the bubble shape, the liquid velocity gradient and the fluid property on CL. The range of the Morton number, M, tested is from logM = − 6.6 to − 3.2. The shapes of bubbles are spherical and ellipsoidal. A correlation of bubble aspect ratio for single bubbles in infinite stagnant liquids proposed in our previous study can give good evaluations for bubbles in the linear shear flows. The CL of spherical bubbles at low bubble Reynolds numbers, Re, depend on the dimensionless shear rate Sr and Re and decrease with increasing Re. These characteristics agree with the Legendre-Magnaudet correlation. The use of a single dimensionless group such as Re, the Eötvös number, the Weber number and the Capillary number cannot correlate CL of non-spherical bubbles. The trend of the critical Re for the reversal of the sign of CL is the same as that for the onset of oscillation of bubble motion, which supports the mechanism proposed by Adoua et al., at least within the range of −6.6 ≤ logM ≤ −3.2. An experimental database of CL is provided for validation of available CL models and CFD.  相似文献   

20.
The near wake of a wall-mounted finite-length square cylinder with an aspect ratio of 7 is investigated based on the proper orthogonal decomposition (POD) of the PIV data measured in three spanwise planes, i.e., z/d = 6, 3.5 and 1.0, near the cylinder free end, mid-span and fixed end (wall), respectively. The Reynolds number based on free-stream velocity (U ) and cylinder width (d) is 9,300. A two-dimensional (2D) square cylinder wake is also measured and analyzed at the same Reynolds number for the purpose of comparison. The structures of various POD modes show marked differences between the two flows. While the coefficients, a 1 and a 2, of the POD modes 1 and 2 occur within an annular area centered at a 1 = a 2 = 0 in the 2D wake, their counterparts are scattered all over the entire circular plane at z/d = 1.0 and 3.5 of the finite-length cylinder wake. Flow at z/d = 6 is dominated by POD mode 1, which corresponds to symmetrical vortex shedding and accounts for 54.0 % of the total turbulent kinetic energy (TKE). On the other hand, the POD modes 1 and 2, corresponding to anti-symmetrical vortex shedding, are predominant, accounting for about 45.0 % of the total TKE, at z/d = 3.5 and 1. It has been found that the flow structure may be qualitatively and quantitatively characterized by the POD coefficients. For example, at z/d = 6, a larger a 1 corresponds to a smaller length of flow reversal zone and a stronger downwash flow. At z/d = 3.5 and 1, two typical flow modes can be identified from a 1 and a 2. While large a 1 and/or a 2 correspond to anti-symmetrical vortex shedding, as in a 2D cylinder wake, small a 1 and a 2 lead to symmetrical vortex shedding. Any values between the large and small a 1 and/or a 2 correspond to the flow structure between these two typical flow modes. As such, the probability of occurrence of a flow structure may be determined from the distribution of the POD coefficients.  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