Abstract

The newly developed algorithm called the grid-by-grid inversion method is a very convenient method for converting an existing computer code for Newtonian flow simulations to that for viscoelastic flow simulations. In this method, the hyperbolic constitutive equation is split such that the term for the convective transport of stress tensor is treated as a source which is updated iteratively. This allows the stress tensors at each grid point to be expressed in terms of velocity gradient tensor at the same location, and the set of stress tensor components is found after inverting a small matrix at each grid point. To corroborate the robustness and accuracy of the grid-by-grid inversion method, we apply it to the 4 : 1 axisymmetric contraction problem. This algorithm is found to be robust and yields accurate results as compared with other finite volume methods. Any commercial CFD packages for Newtonian flow simulations can be easily converted to those for viscoelastic fluids exploiting the grid-by-grid inversion method.

1. Introduction

Contrary to the techniques of computational fluid dynamics for Newtonian fluids, the numerical algorithms for viscoelastic flows are not so matured. The hyperbolic nature of the constitutive equation incurs peculiar flow phenomena such as rod-climbing and extrudate swell as well as causes difficulties in numerical simulation [13]. Since the momentum balance equation is elliptic in steady state and parabolic in unsteady state, the complete set for viscoelastic flows is a mixed type, hyperbolic-elliptic, or hyperbolic-parabolic. This situation is difficult to treat numerically since it is difficult to devise a numerical algorithm that works for mixed systems. Another difficulty associated with the hyperbolic constitutive equation is the choice of appropriate boundary conditions for the stress field at the boundary of computational domain. One can impose nonslip boundary condition on the walls for the velocity field but there exist no apparent or natural boundary conditions for the stress components at the wall. Various numerical techniques for solving viscoelastic flows, such as finite volume methods, finite element methods, and spectral methods, are well documented in the references cited [49].

In the present investigation, a newly developed algorithm called the grid-by-grid inversion method [10] is employed to solve the viscoelastic flows through an axisymmetric contraction. Figure 1 shows the flow geometry for the axisymmetric contraction. Viscoelastic fluid passes from one circular tube into the other tube of smaller radius and generates a complex flow having a strong shear near the walls and uniaxial extension along the centerline. The existence of strong shearing zones and uniaxial extension zone makes this flow geometry a good test bed for numerical algorithms of viscoelastic fluids. Many investigators of rheology adopt this flow geometry as an important benchmark problem, especially the 4 : 1 contraction geometry [11]. We shall solve the 4 : 1 axisymmetric contraction problem employing the newly developed “grid-by-grid inversion method” [10], which is implemented based on a finite volume method [10].

The hyperbolic constitutive equations of viscoelastic fluids have a nonlocal character because of the term representing convective transport of stress tensor. If this term is assumed to be known, the constitutive equation becomes local and the stress tensor is easily evaluated for a given velocity gradient tensor at the same location. The six stress tensor components for the cases of a three-dimensional flow are found after inverting a six by six matrix at each grid point and are substituted into the Navier-Stokes equation as a source term. In this way, the numerical solution of viscoelastic flows becomes as straightforward as that of Newtonian fluids. We call this algorithm the grid-by-grid inversion method since the viscoelastic stress tensor is obtained by the grid-by-grid inversion of a matrix equation at each grid point. This algorithm can easily be implemented using finite volume methods, finite element methods and spectral methods. When applied to the 4 : 1 axisymmetric contraction problem, it is found that the grid-by-grid inversion method yields accurate results efficiently in comparison with numerical results of traditional algorithms.

2. Governing Equations and the Grid-by-Grid Inversion Method

We consider incompressible isothermal flows of viscoelastic fluids. The governing equations may be written in dimensionless variables as follows:𝐯=0,(2.1)Re𝜕𝐯𝜆𝜕𝑡+𝐯𝐯=𝝈,(2.2)𝝈=𝑃𝐈+2𝛽𝐃+𝝉,(2.3)𝜕𝝉𝜕𝑡+𝐯𝝉=2(1𝛽)𝐃+𝜆(𝐯)𝑇𝝉+𝝉𝐯𝝉.(2.4) In the above equations, 𝑃 is pressure, 𝐃 is the rate of deformation tensor, 𝝉 is the viscoelastic part of the stress tensor. The parameter 𝛽 is the ratio of the retardation and relaxation times of the fluid. Equation (2.4) is the constitutive equation of the Oldroyd-B model [9], 𝜆 is the dimensionless relaxation time or the Deborah number, and Re is the Reynolds number. The superscript 𝑇 in (2.4) is the transpose. The Reynolds number and the dimensionless Deborah number are defined byRe=𝜌𝑈𝐿𝜂𝜆,𝜆=1𝑈𝐿,(2.5) where 𝜌 is the density, 𝜆1 the dimensional relaxation time, 𝑈 the characteristic speed, and 𝐿 is the characteristic length. The characteristic velocity 𝑈 and the characteristic length 𝐿 are taken as the average velocity in the downstream tube and the radius of the downstream tube, respectively. To compare with the results of other investigators [9, 11], we take the value of 𝛽 to be 1/9.

Next, we consider the grid-by-grid inversion method as applied to the set of equations (2.1)–(2.4). Contrary to the Newtonian fluids where the stress field 𝝉 depends on the velocity gradient tensor 𝐯 locally, the stress depends on the velocity gradient tensor nonlocally due to the fact that constitutive equation is hyperbolic partial differential equation. The convective transport of the stress tensor, represented by 𝐯𝝉 in (2.4), takes care of the memory effect of 𝝉 in its dependence on 𝐯 and causes the functional dependence of 𝝉 on 𝐯 to be nonlocal. In the grid-by-grid inversion method, 𝐯𝝉 is assumed to be a known source term which shall be updated iteratively in the process. Then the constitutive equation (2.4) can be converted such that 𝝉 is represented as a local function of 𝐯 as in the case of Newtonian fluids. This arrangement renders the numerical solution of viscoelastic fluids as straightforward as that of Newtonian fluids. The steps for obtaining the velocity and the stress at time step 𝑛+1, 𝐯𝑛+1 and 𝝉𝑛+1, using known velocity and stress at time step 𝑛,𝐯𝑛 and 𝝉𝑛, proceed in an iterative manner as follows. First, define𝐂𝐯𝑛+1(𝑖𝑡)𝝉𝑛+1(𝑖𝑡),(2.6) where the superscript 𝑛+1(𝑖𝑡) indicates variable at time step 𝑛+1 in the 𝑖𝑡th iteration. Discretizing (2.4) in time implicitly and representing 𝐃 in terms of the velocity gradient 𝐯, we find the following local matrix equations defined at each grid point.𝜆𝝉Δ𝑡+1.0𝑛+1(𝑖𝑡+1)𝜆(𝐯)𝑇𝑛+1(𝑖𝑡)𝝉𝑛+1(𝑖𝑡+1)𝜆𝝉𝑛+1(𝑖𝑡+1)(𝐯)𝑛+1(𝑖𝑡)=𝜆𝝉Δ𝑡𝑛𝜆𝐂+(1𝛽)(𝐯)𝑛+1(𝑖𝑡)+(𝐯)𝑇𝑛+1(𝑖𝑡).(2.7) Once 𝐂 is evaluated using variables at 𝑛+1(𝑖𝑡), (2.7) can be solved for 𝝉𝑛+1(𝑖𝑡+1) by inverting a six-by-six matrix at each grid point for the six independent stress components for the case of three-dimensional flows. Equation (2.7) is also solved at the boundary grid points to find the boundary stress field. This suggests a natural method of imposing boundary conditions for the hyperbolic constitutive equations at all boundaries of the computational domain. This method had been employed to impose outflow boundary conditions by Papanastasiou et al. [12] and is called the open boundary condition. Using 𝝉𝑛+1(𝑖𝑡+1) obtained from (2.7), the velocity at the 𝑛+1 th time step in the 𝑖𝑡+1th iteration, 𝐯𝑛+1(𝑖𝑡+1), is found by solving (2.1)–(2.3) as follows:𝐯𝑛+1(𝑖𝑡+1)𝐯=0,(2.8)Re𝑛+1(𝑖𝑡+1)𝐯𝑛Δ𝑡+𝐯𝑛+1(𝑖𝑡+1)𝐯𝑛+1(𝑖𝑡+1)=𝑃𝑛+1(𝑖𝑡+1)+2𝛽2𝐯𝑛+1(𝑖𝑡+1)+𝝉𝑛+1(𝑖𝑡+1).(2.9) Although (2.8)-(2.9) can be solved using any numerical methods for incompressible Navier-Stokes equations, we employ a finite volume method based on the SIMPLE algorithm [13, 14]. To stabilize the numerical scheme for large values of 𝜆, (2.6) is evaluated using one of the various upwind schemes. In the present work, we adopt a higher-order upwind scheme, MinMod [14], to evaluate 𝐂.

The case of 𝛽=0 requires a special deliberation before employing standard algorithms for incompressible Navier-Stokes equations in the grid-by-grid inversion technique. When 𝛽=0, we decompose the total stress 𝝈 into pure viscous and elastic parts 𝑃𝐈+2𝐃 and 𝚺, respectively, such that𝝈=𝑃𝐈+2𝐃+𝚺,(2.10) where𝚺𝝉2𝐃.(2.11) Then the momentum and constitutive equations for the case of 𝛽=0 may be written as𝐯=0,(2.12)Re𝜕𝐯𝜕𝑡+𝐯𝐯=𝑃+2𝜆𝐯+𝚺,(2.13)𝜕𝚺𝜕𝑡+𝐯𝚺(𝐯)𝑇𝚺𝚺(𝐯)+𝚺=2𝜆𝜕𝐃𝜕𝑡+𝐯𝐃(𝐯)𝑇.𝐃𝐃𝐯(2.14) This decomposition is called the EVSS formulation and was proposed by Perera and Walters [15]. The implementation of grid-by-grid inversion method to the set of (2.12)–(2.14) is straightforward. First, we evaluate𝐄𝐯𝑛+1(𝑖𝑡)𝚺𝑛+1(𝑖𝑡).(2.15) Next, convert (2.14) to the following local matrix equations for 𝚺 decoupled at each grid point:𝜆𝚺Δ𝑡+1.0𝑛+1(𝑖𝑡+1)𝜆(𝐯)𝑇𝑛+1(𝑖𝑡)𝚺𝑛+1(𝑖𝑡+1)𝜆𝚺𝑛+1(𝑖𝑡+1)(𝐯)𝑛+1(𝑖𝑡)=𝜆𝚺Δ𝑡𝑛1𝜆𝐄2𝜆𝐃Δ𝑡𝑛+1(𝑖𝑡)1𝐃Δ𝑡𝑛+𝐯𝐃(𝐯)𝑇𝐃𝐃𝐯𝑛+1(𝑖𝑡).(2.16) Then, (2.12) and (2.13) are discretized as follows: 𝐯𝑛+1(𝑖𝑡+1)𝐯(2.17)Re𝑛+1(𝑖𝑡+1)𝐯𝑛Δ𝑡+𝐯𝑛+1(𝑖𝑡+1)𝐯𝑛+1(𝑖𝑡+1)=𝑃𝑛+1(𝑖𝑡+1)+2𝐯𝑛+1(𝑖𝑡+1)+𝚺𝑛+1(𝑖𝑡+1).(2.18) The structure of (2.16) is the same as that of (2.7) and can be solved for 𝚺𝑛+1(𝑖𝑡+1) by inverting a six-by-six matrix at each grid point for a three-dimensional flow once (𝐯)𝑛+1(𝑖𝑡) is given. After obtaining 𝚺𝑛+1(𝑖𝑡+1), 𝐯𝑛+1(𝑖𝑡+1) is found by solving (2.17)-(2.18) using the SIMPLE algorithm [13]. The second-order derivative terms appearing in 𝐯𝐃 of (2.14) are evaluated using a one-sided finite difference formula at the boundaries.

The overall solution procedure for the grid-by-grid inversion method may be summarized as follows.(1)𝐯𝑛 and 𝝉𝑛 have obtained in the previous time step 𝑛.

The procedure for the time step 𝑛+1 begins as follows.(2) Assume 𝐯𝑛+1(𝑖𝑡) and 𝝉𝑛+1(𝑖𝑡). For the first iteration (𝑖𝑡=1), 𝐯𝑛+1(𝑖𝑡)=𝐯𝑛 and 𝝉𝑛+1(𝑖𝑡)=𝝉𝑛.(3) Evaluate 𝐂 of (2.6) (𝛽0) or 𝐄 of (2.15) (𝛽=0) using an upwind scheme.(4) Using 𝐯𝑛+1(𝑖𝑡), solve (2.7) (𝛽0) or (2.16) (𝛽=0) for 𝝉𝑛+1(𝑖𝑡+1) by inverting a six-by-six matrix at each grid point including the boundary grids.(5) Using 𝝉𝑛+1(𝑖𝑡+1),

(𝛽0) solve the momentum equation, (2.9), and continuity equation to find 𝐯𝑛+1(𝑖𝑡+1).

(𝛽=0) solve the momentum equation, (2.18), and continuity equation to find 𝐯𝑛+1(𝑖𝑡+1).(6) Convergence check for 𝝉𝑛+1 and 𝐯𝑛+1. If not converged, go to step (2). Otherwise, update the time step and go to step (1).

Usually convergence is attained in two or three iterations. The novelty of the grid-by-grid inversion method is its easiness of numerical implementation. As noted in the above procedure, one can easily convert any existing computer code for Newtonian flow simulations to that for viscoelastic flow simulation by adding a subroutine that solves the viscoelastic constitutive equation using the grid-by-grid inversion method, summarized as steps (3)~(4), to evaluate the viscoelastic source terms in the Navier-Stokes equation, 𝝉. Adding a source term to the Navier-Stokes equation is an easy procedure whether we employ a finite volume method or a finite element method. In the subroutine for the viscoelastic constitutive equation, one inverts a small matrix at each grid point, which can be performed cheaply. Therefore, one can easily convert any commercial CFD package for Newtonian fluid flows to that for viscoelastic fluids employing the grid-by-grid inversion method. Its robustness and accuracy are corroborated in the next section, where the grid-by-grid inversion method is employed to solve the 4 : 1 viscoelastic axisymmetric contraction problem. Although Phillips and Williams [16] solve the viscoelastic constitutive equation by converting small matrix equations in their semi-Lagrange method, it is very difficult to convert an existing Newtonian code to a viscoelastic code using the semi-Lagrange method since it treats the momentum and the constitutive equation simultaneously.

3. Viscoelastic Flow through an Axisymmetric Contraction

In this section, we solve the flow of an Oldroyd-B fluid through a 4 : 1 axisymmetric contraction geometry using the grid-by-grid inversion method. The schematic representation of the flow geometry is depicted in Figure 1. The viscoelastic flow through the circular contraction geometry has served as a standard benchmark problem for numerical algorithms for viscoelastic fluids. The presence of a singularity in the entry-flow geometry as well as regions of high shear and extension in the flow have been a significant challenge for a long time, and many investigators have struggled to develop accurate and robust algorithms for viscoelastic flows through contraction geometries [17, 18]. The contraction ratio 4 : 1 is the standard one in this benchmark problem. A good review of old works may be Trebotich et al. [18]. Recently, Phillips and Williams [11, 16] used a finite volume method to solve the flow of an Oldroyd-B fluid through planar and axisymmetric contractions. Among these two geometries, the axisymmetric contraction yields more dramatic viscoelastic effects. In the present investigation, we will solve the flow of Oldroyd-B fluid through the axisymmetric geometry using the grid-by-grid inversion method based on the SIMPLE algorithm to corroborate the accuracy and robustness of the grid-by-grid inversion method. When solving (2.8)-(2.9) or (2.17)-(2.18) using the SIMPLE algorithm, we adopt a collocated mesh and the pressure oscillation induced by the collocated mesh is eliminated using the Rhie-Chow method [19], and a higher-order upwind scheme, MinMod [14], is used to treat the inertia force term in the momentum equation and the convective transport of stress tensor in the constitutive equation. For the 4 : 1 axisymmetric contraction geometry depicted in Figure 1, the number of relevant components of stress tensor is four because the 𝝉 term in (2.9) contains a circular stress component 𝜏𝜃𝜃 as follows:(𝝉)𝑟=1𝑟𝜕𝜕𝑟(𝑟𝜏𝑟𝑟𝜕)+𝜏𝜕𝑧𝑟𝑧𝜏𝜃𝜃𝑟,(𝝉)𝑧=1𝑟𝜕𝜕𝑟(𝑟𝜏𝑟𝑧𝜕)+𝜏𝜕𝑧𝑧𝑧.(3.1) Although the relevant velocity components for axisymmetric flows are (𝑣𝑟,𝑣𝑧), it is necessary to find the four stress components, 𝜏𝑟𝑟, 𝜏𝑟𝑧, 𝜏𝑧𝑧, and 𝜏𝜃𝜃 which appear in (3.1). The constitutive equation (2.4) may be written for an axisymmetric geometry as follows:𝜆𝜕𝜏𝑟𝑟𝜕𝑡+𝑣𝑟𝜕𝜏𝑟𝑟𝜕𝑟+𝑣𝑧𝜕𝜏𝑟𝑟𝜕𝑧=2(1𝛽)𝜕𝑣𝑟2𝜕𝑟+𝜆𝜕𝑣𝑟𝜏𝜕𝑟𝑟𝑟+2𝜕𝑣𝑟𝜏𝜕𝑧𝑟𝑧𝜏𝑟𝑟𝜆,(3.2)𝜕𝜏𝑟𝑧𝜕𝑡+𝑣𝑟𝜕𝜏𝑟𝑧𝜕𝑟+𝑣𝑧𝜕𝜏𝑟𝑧𝜕𝑧=(1𝛽)𝜕𝑣𝑧+𝜕𝑟𝜕𝑣𝑟𝜕𝑧+𝜆𝜕𝑣𝑟𝜏𝜕𝑟𝑟𝑧+𝜕𝑣𝑟𝜏𝜕𝑧𝑧𝑧+𝜏𝑟𝑟𝜕𝑣𝑧𝜕𝑟+𝜏𝑟𝑧𝜕𝑣𝑧𝜕𝑧𝜏𝑟𝑧,𝜆(3.3)𝜕𝜏𝑧𝑧𝜕𝑡+𝑣𝑟𝜕𝜏𝑧𝑧𝜕𝑟+𝑣𝑧𝜕𝜏𝑧𝑧𝜕𝑧=2(1𝛽)𝜕𝑣𝑧2𝜕𝑧+𝜆𝜕𝑣𝑧𝜏𝜕𝑟𝑟𝑧+2𝜕𝑣𝑧𝜏𝜕𝑧𝑧𝑧𝜏𝑧𝑧𝜆,(3.4)𝜕𝜏𝜃𝜃𝜕𝑡+𝑣𝑟𝜕𝜏𝜃𝜃𝜕𝑟+𝑣𝑧𝜕𝜏𝜃𝜃𝑣𝜕𝑧=2(1𝛽)𝑟𝑟2𝑣+𝜆𝑟𝑟𝜏𝜃𝜃𝜏𝜃𝜃.(3.5) It is to be noted that equations for (𝜏𝑟𝑟,𝜏𝑟𝑧,𝜏𝑧𝑧) are coupled with each other, while that for 𝜏𝜃𝜃 is decoupled. The relevant components of 𝐂 of (2.6) are𝐶𝐼𝐽=𝑣𝑟𝜕𝜏𝐼𝐽𝜕𝑟+𝑣𝑧𝜕𝜏𝐼𝐽𝜕𝑧,(3.6) where 𝐼𝐽=𝑟𝑟,𝑟𝑧,𝑧𝑧,𝜃𝜃. Assuming known values of 𝐶𝐼𝐽, (3.2)–(3.4) may be represented as a set of matrix equations at each grid point including the boundary grids as follows:𝜆Δ𝑡+12𝜆𝜕𝑣𝑟𝜕𝑟2𝜆𝜕𝑣𝑟0𝜕𝑧𝜆𝜕𝑣𝑧𝜆𝜕𝑟Δ𝑡+1𝜆𝜕𝑣𝑟+𝜕𝑟𝜕𝑣𝑧𝜕𝑧𝜆𝜕𝑣𝑟𝜕𝑧02𝜆𝜕𝑣𝑧𝜆𝜕𝑟Δ𝑡+12𝜆𝜕𝑣𝑧𝜕𝑧𝑛+1(𝑖𝑡)𝑖𝑘𝜏𝑟𝑟𝜏𝑟𝑧𝜏𝑧𝑧𝑛+1(𝑖𝑡+1)𝑖𝑘=𝜆𝜏𝑟𝑟(𝑛)Δ𝑡𝜆𝐶𝑟𝑟+2(1𝛽)𝜕𝑣𝑟𝜕𝑟𝜆𝜏𝑟𝑧(𝑛)Δ𝑡𝜆𝐶𝑟𝑧+(1𝛽)𝜕𝑣𝑧+𝜕𝑟𝜕𝑣𝑟𝜆𝜕𝑧𝜏Δ𝑡𝑧𝑧(𝑛)𝜆𝐶𝑧𝑧+2(1𝛽)𝜕𝑣𝑧𝜕𝑧𝑛+1(𝑖𝑡)𝑖𝑘.(3.7) On the other hand, from (3.5), 𝜏𝜃𝜃𝑖𝑘 can be found at each grid point within the computational domain and on the boundary as follows:𝜏𝜃𝜃𝑛+1(𝑖𝑡+1)𝑖𝑘=𝜆𝜏Δ𝑡𝜃𝜃𝑛𝜆𝐶𝜃𝜃𝑣+2(1𝛽)𝑟𝑟𝑛+1(𝑖𝑡)𝑖𝑘𝜆𝑣Δ𝑡+12𝜆𝑟𝑟𝑛+1(𝑖𝑡)𝑖𝑘.(3.8) Once (𝜏𝑟𝑟𝑖𝑘,𝜏𝑟𝑧𝑖𝑘,𝜏𝑧𝑧𝑖𝑘,𝜏𝜃𝜃𝑖𝑘)𝑛+1(𝑖𝑡+1) are obtained from (3.7)-(3.8), they are used to solve (2.8)-(2.9) to find (𝑣𝑟,𝑣𝑧)𝑛+1(𝑖𝑡+1)𝑖𝑘 in the 4 : 1 axisymmetric contraction geometry.

4. Results

In this section, we compare the results from the grid-by-grid inversion method with those of Phillips and Williams [11] to corroborate the accuracy and robustness of the grid-by-grid inversion method. To make the comparison meaningful, we adopt the same flow geometry and inlet conditions as those employed by Phillips and Williams [11]. Namely, the contraction ratio is 4 : 1, the length of the larger channel is 27𝑅, and that of smaller channel is 49𝑅 when 𝑅 is the radius of the smaller channel (cf. Figure 1). At the inlet, the parabolic Poiseuille flow is adopted as Phillips and Williams [11] have done.(-) inlet velocity𝑣𝑧1(𝑟)=6416𝑟2,𝑣𝑟=0,0𝑟4,(4.1) where 𝑟𝑟/𝑅 and 𝑟 is the dimensional radial distance. Contrary to Phillips and Williams [11], we do not impose inlet boundary conditions for the stress field, since the grid-by-grid inversion method does not require them and simply solve (3.7)-(3.8) at the inlet grid points to find the stress field at the inlet location. This is also an important merit of the grid-by-grid inversion method over other algorithms for the viscoelastic flows. Although Papanastasiou et al. [12] employed this method of imposing open outflow boundary conditions previously and named it open boundary condition method or free boundary condition method, it appears very naturally in the grid-by-grid inversion method.

We have employed three sets of grid system to ensure grid convergence of numerical results. For the three sets of grids, that is, 14,580 (Grid A), 32,535 (Grid B), and 57,960 (Grid C), the extrastress component 𝜏𝑧𝑧 is plotted when 𝜆=1.0 for Re=0.0 along 𝑟=1 in Figure 2(a), and along 𝑟=0.064 in Figure 2(b). For there three mesh systems, the numbers of cells along 𝑟 in the small tube are 30, 45, and 60, respectively. For Grid A, the minimum Δ𝑟 is 0.005 at the wall and it grows gradually until it becomes 5.1×102 at the center. For Grid B, (Δ𝑟)min is 0.003 and (Δ𝑟)max is 3.4×102 at the center, and for Grid C, (Δ𝑟)min=0.002 at the wall and (Δ𝑟)max=2.3×102 at the center. The results of the present work are also compared with those of Phillips and Williams [11]. Though the grid number increases from 32,535 to 57,960, there is not appreciable change in 𝜏𝑧𝑧. Therefore, we adopt Grid B in the subsequent computations. Figure 3 shows the extrastress component 𝜏𝑧𝑧 at Re=0.0 for 𝜆=0.5,1.0, and 1.5 along 𝑟=1.0 (Figure 2(a)) and along 𝑟=0.064 (Figure 2(b)). The grid-converged data for 𝜏𝑧𝑧 obtained from the grid-by-grid inversion method are somewhat smaller than the benchmark data of Phillips and Williams [11]. At the corner of the contractor, 𝜏𝑧𝑧 has a sharp overshoot, which increases with respect to 𝜆, and it settles down to a downstream value rapidly. Near the center of the contracted channel (𝑟=0.064), a smaller overshoot of 𝜏𝑧𝑧 appears, which increases as 𝜆 increases, and it settles down to a downstream value much slowly as compared to that at the wall (𝑟=1.0).

Figure 4 shows comparison of streamlines for 𝜆=0.0, 0.5, 1.0, and 1.5 obtained from the grid-by-grid inversion method with those from the Phillips and Williams [11] when Re=0.0. The grid-by-grid inversion method yields accurate results as compared with the benchmark data for the range of 𝜆 values considered. The size of the corner vortex increases as 𝜆 increases. Figure 5 depicts comparison of streamlines at various values of 𝜆 when Re=1.0. The predictions of the grid-by-grid inversion method are also in good agreement with those of the benchmark data [11]. Comparing with the streamlines of Re=0.0, it is found that the vortex sizes are diminished due to the inertia effect [11]. The size of the corner vortex can be measured using a parameter 𝐿1, as suggested by Phillips and Williams [11] and Oliveira et al.[8], which is the distance of the upstream separation point from the salient corner. Figure 6 shows the variation of 𝐿1 with respect to 𝜆 for Re=0.0 and Re=1.0. For Re=1.0, the grid-by-grid inversion method is found to yield results similar to those of the benchmark data of Phillips and Williams [11]. For Re=0.0, the results of the grid-by-grid inversion method are compared with those of Oliveira et al. [8] as well as those of Phillips and Williams [11]. It is shown that both the grid-by-grid inversion and Phillips and Williams [11] yield 𝐿1 somewhat larger than those of Oliveira et al. [8]. Since Oliveira et al. [8] do not consider the case Re=1.0, we cannot compare the grid-by-grid inversion method with Oliveira et al. when Re=1.0. To corroborate the accuracy of the grid-by-grid inversion method further, we compare the results for the vortex intensity Ψ𝑅, which is defined as the flow rate in recirculation divided by inlet flow rate, in Figure 7 when Re=0.0. As in previous comparison, the grid-by-grid inversion method yields results similar to those of Phillips and Williams [11], while Oliveira et al. [8] predicts results somewhat smaller than the grid-by-grid inversion method. The grid-by-grid inversion method yields results for 𝜆 larger than 1.5 and Re larger than 1.0, which demonstrates the robustness of the grid-by-grid inversion method as compared to other finite volume methods or finite element methods [9, 11, 16, 18].

5. Conclusion

The grid-by-grid inversion method [10] has been applied to the 4 : 1 circular contraction problem in the present investigation. Viscoelastic flows through the contraction generate complex flows exhibiting strong shear and uniaxial extension, which is a good test bed for the robustness and accuracy of a new numerical algorithms. In the grid-by-grid inversion method [10], the hyperbolic constitutive equation is split such that the term for the convective transport of stress tensor is treated as a source, which is updated iteratively. This allows the stress tensor at each grid point to be expressed in terms of velocity gradient tensor at the same location, and the set of stress tensor components is found after inverting a small matrix at each grid point. The grid-by-grid inversion method can be implemented easily in any commercial CFD packages for Newtonian fluid flows to convert them to be used for viscoelastic fluid flows. The grid-by-grid inversion method is found to yield accurate results as compared with the benchmark data of Phillips and Williams [11]. It predicts accurately the variation of corner vortex size and the 𝜏𝑧𝑧 overshoot with respect to 𝜆 at Re=0.0 and Re=1.0.

Acknowledgment

This paper was supported by the Human Resources Development of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) Grant funded by the Korea Government Ministry of Knowledge and Economy (no. 20114010203090).