Jtam.dvi JOURNAL OF THEORETICAL AND APPLIED MECHANICS 42, 1, pp. 21-39, Warsaw 2004 COMPUTATIONS OF AN UNSTEADY VISCOUS FLOW IN A THREE DIMENSIONAL SYSTEM OF DUCTS. PART I: FORMULATION OF THE MATHEMATICAL PROBLEM AND NUMERICAL METHOD Jacek Szumbarski Piotr Olszewski Andrzej Styczek Jacek Rokicki Institute of Aeronautics and Applied Mechanics, Warsaw University of Technology e-mail: jasz@meil.pw.edu.pl; polsz@meil.pw.edu.pl; jack@meil.pw.edu.pl Numerical modeling of an unsteady flow of a viscous incompressible flu- id inside a branched pipe system is considered. The mathematical for- mulation is given with special emphasis on inlet/outlet conditions. The equivalentweak formof the initial-boundary value problem is presented. The numericalmethod based on solutions to particular Stokes problems is proposed and described in some details. Finally, some general remarks about the implementation issues within the framework of the spectral element discretization are made. Key words: Navier-Stokes equations, defective boundary conditions, spectral element method 1. Introduction Numerical simulations of time-dependent viscousflows inside complexduct systemshave recently become increasingly interesting forComputationalFluid Dynamics (CFD) community. This interest seems also to be stimulated mo- stly by medical applications. During the last decade, significant progress in CFD techniques applied to biological flows has been achieved. Undoubtedly, highly accurate numerical simulation of various types of motion of bio-fluids is a serious challenge. Much effort has been done, for instance, to develop computational models of the human cardiovascular system. These attempts 22 J.Szumbarski et al. are motivated mostly by medical needs – it is expected that future, reliable computational models would be useful in the optimization of various cardio- surgery procedures. The necessary condition for success in this respect is to develop computational models and techniques which give realistic results by taking into account all important features like pulsating character of motion, complicated geometry and compliance of branched vessels, and, at least, non- Newtonian rheology of blood. In this work, we consider the problem of numerical simulation of an unste- ady flow of a viscous incompressible fluid in a system of branched pipes. The focus is on the proper mathematical formulation of the problem with spe- cial emphasis on inlet/outlet conditions. These conditions are ”defective” in a sense that they are based on averaged values of the pressure and/or the volume flux. The presented formulation is a generalization of the approach recently proposed by Formaggia et al. (2000). Some remarks on the numerical implementation of the spectral elementmethod aremade, leaving detailed de- scription of the solver and the presentation of obtained results to the second part of the paper. 2. Mathematical formulation Weconsider three-dimensionalunsteady(pulsating)motionofaNewtonian fluid in a branched pipe system with certain number of inlet/outlet (I/O) sections (see Fig.1). The mathematical problem is to solve the Navier-Stokes and continuity equations ∂tv+∇v ·v=− 1 ρ ∇p+ν∇2v ∇·v=0 (2.1) subject to appropriate initial and boundary conditions. In the above, the sym- bols v, p, ρ and ν denote velocity, pressure,mass density and kinematic visco- sity, respectively. Since the mass density is a constant value, it is convenient to choose ρ=1. The boundary conditions are defined as follows. At the impermeable (ma- terial) partof theboundaryΓ , theno-slip condition for thevelocity is imposed, i.e. v|Γ = 0. At the I/O sections Si (i = 1, ...,N), the following variants of the boundary conditions are considered: Computations of an unsteady viscous flow... 23 Fig. 1. The computational domain • variant Volume Flux (VF) Φi(v)≡ ∫ Si v ·n ds=Fi(t) i=1, ...,NVF (2.2) Fi(t) – given, • variant Average Pressure (AP) 1 |SNVF+i| ∫ SNVF+i p ds=Pi(t) i=1, ...,NAP (2.3) Pi(t) – given. At each I/O section either VF orAPvariant of the boundary conditions is imposed, and NVF +NAP =N. Time variations of the volume fluxes or the averaged pressure are defined by the given functions Fi(t) (i=1, ...,NVF) or Pi(t) (i=1, ...,NAP), respectively. The integral boundary conditions formulated above are of much interest because they are natural inmost of practical situations. Indeed, the knowledge of temporal variations of either the section-averaged pressure or the volume flux can be usually assumed. In medical context, such data can be provided by measurements or by the lumped-parameter models of the cardiovascular system. On the other hand, the precise distributions of physical quantities (like pressure or velocity) at I/O sections are usually not available – they have to be evaluated in the solution process. 24 J.Szumbarski et al. The question arises whether such ”defective” boundary conditions can be incorporated in amathematically consistent way into an initial-boundary pro- blem to equations (2.1). Surprisingly enough, this problem has been given a rigorous mathematical treatment only recently. In their seminal paper, Hey- wood et al. (1996) showed that the boundary conditions formulated above can be accounted for in an appropriate variational formulation. Recently, Formag- gia et al. (2000) have presented an improved variant of this approach. They used the Lagrangemultipliers technique toVF-type boundary conditions that allowed for convenient simplification of the function spaces involved in the va- riational formulation. This approach is a theoretical basis of the currentwork. The novelty consists in admitting coexistence of the I/O sections withVFand AP boundary conditions, not considered in previous works. The newVF/AP variational formulation of the problem goes as follows: Find • the velocity field v∈V = {v∈ [H1(Ω)]3 : v|Γ =0} • the pressure field p∈Q=L2(Ω) • the (time-dependent) Lagrange multiplies λi ∈R, i=1, ...,NVF such that • for each v∈V : (∂tv+∇v ·v,v)+ν(∇v,∇v)+ NVF∑ i=1 λiΦi(v)+ + NAP∑ i=1 Pi(t)ΦNVF+i(v)− (p,∇·v)= 0 • for each q∈Q : (q,∇·v)= 0 • Φi(v)=Fi(t), i=1, ...,NVF • v|t=t0 =v0 (the initial condition). Heywood et al. (1996) showed that the smooth solutions to this variational problem satisfy the following ”classical” boundary conditions (p−ν∂nvn) ∣∣∣ Si =λi i=1, ...,NVF (p−ν∂nvn) ∣∣∣ SNVF+i =Pi(t) i=1, ...,NAP ∂nvτ ∣∣∣ Si =0 i=1, ...,N (2.4) Computations of an unsteady viscous flow... 25 In theabove,weuse thenotation ∂nvn =(∇v·n)·nand ∂nvτ =(∇v·n)·τ, where (∇v)ij = ∂vi/∂xj. It can be noticed that the corresponding boundary conditions do not have any direct physical interpretation. In particular, they are not formulated in terms of normal and tangent surface stress distributions. Indeed, the latter would involve only the symmetric part of the tensor ∇ · v. The conditions (2.4) are sometimes referred to as ”pseudo-traction” conditions. tangent component of the surface stress vanishes identically and the di- stributions of the normal component are uniform at each inlet/outlet section. Moreover, the normal stress distributions at I/O sections of AP type are given directly as the functions Pi(t) (i = 1, ...,NAP), while at I/O sections of VF type these distributions are unknown and have to be determined so that the prescribed volume fluxes are achieved. It should be emphasized that the boundary conditions implied by the va- riational formulation at the AP inlets/outlets are not, in general, equivalent to the condition for the averaged pressure (2.3). Instead, one has the following equality ∫ Si [p−ν(∇v ·n) ·n] dS=Pi(t)|Si| (2.5) However, it can be shown that conditions (2.5) and (2.3) are equivalent if only the surface of the inlet/outlet section is flat, i.e. when it is obtained by a plane cut of the pipe. Indeed, in such a case, the following equality holds ∫ Si (∇v ·n) ·n dS =0 (2.6) In order to prove (2.6), we choose the coordinate system so that the first axis is perpendicular to the (flat) surface section Si. Then the normal vector at each point of Si is defined as n= [1,0,0], and the surface integral can be calculated as follows ∫ Si (∇v ·n) ·n dS = ∫ Si ∂v1 ∂x1 dS =− ∫ Si (∂v2 ∂x2 + ∂v3 ∂x3 ) dS = =− ∮ ∂Si (v2η2+v3η3) dℓ=0 In the above, we have used the continuity equation and the Green The- orem of the plane in order to transform the surface integral to the contour integral along ∂Si. The unitary vector η = [η2,η3] lies in the plane of the 26 J.Szumbarski et al. pipe section Si, and it is normal to the contour line ∂Si. Since the velocity vanishes identically at the pipe wall, the contour integral is equal to zero. Using (2.5) and (2.6), one finally concludes that Pi(t)= 1 |Si| ∫ Si p dS (2.7) Note again that, for the above proof to work, one has to assume sufficient regularity of the solution to ensure the existence of the velocity derivatives and the pressure at the inlet/outlet surfaces (in the sense of traces). In the end, the following comment should be made. If NAP > 0, i.e. at least one I/OsectionhasbeenequippedwithAPboundaryconditions then the pressure field is uniquely defined. On the other hand, one may like to impose I/O conditions of the VF type exclusively. If NVF = N then the functions Fi(t), i=1, ...,N, must obey the following constraint ∑N i=1Fi(t)≡ 0, implied by the volume (or mass) conservation. Moreover, the pressure field is defined up to an additive constant. It is actually more reasonable (and convenient) to set NVF = N − 1 so that NAP = 1, and choose P1(t) ≡ 0. The volume will be conserved (within achievable numerical accuracy) due to the continuity equation, and pressure ambiguity will not occur. 3. Time integration schemes and semi-discrete forms of the variational problem In order to solve anunsteadyflowproblemvarious timediscretization sche- mes can be applied. The popular choice is the explicit/implicit approach, i.e. the use of an implicit integration scheme for the linear part and an explicit scheme for the nonlinear part of the Navier-Stokes equation. The advantage of this approach is that symmetric and positive definite linear problems have to be solved at each time step. In contrast, methods using the implicit (or semi-implicit) time discretization to the nonlinear terms lead to large nonli- near problems or, at least, to linear but usually not symmetric and indefinite problemswhich aremuchmore difficult to solve efficiently. On the other hand, the explicit treatment of the nonlinear terms suffers from limitations due to stability requirements. This is why it is essential to use integration schemes with favorable stability properties. Sucha family of stiff-stablemulti-step sche- meshas beenproposedbyKarniadakis et al. (1991). Consider, for presentation Computations of an unsteady viscous flow... 27 purposes, the model evolutionary equation with linear and nonlinear opera- tors, denoted by L and N, respectively z′(t)=L(t,z(t))+N(t,z(t)) (3.1) The mixed multi-step method is obtained by using the backward differentia- tion scheme (implicit) for the linear operator, and the linear extrapolation (explicit) scheme for the nonlinear operator. The general formula for the K- step method is 1 ∆t K∑ k=0 βkz (n+1−k) =L(n+1)+ K∑ k=1 αkN (n+1−k) (3.2) As usual, the upper index in the brackets indicates a time instant each term is evaluated at. The values of the coefficients {αk, k = 1, ...,K} and {βk, k = 0, ...,K} can be found byKarniadakis et al. (1991). As an example, we give the explicit formula for the 3rd order method 1 ∆t (11 6 z(n+1)−3z(n)+ 3 2 z(n−1)− 1 3 z(n−2) ) = (3.3) =L(n+1)+3N(n)−3N(n−1)+N(n−2) More details on other schemes of this sort and their stability characteristics can be found in the cited paper. When the multistep method, see (3.2), is used, variational problem (Pro- blem 2.1) takes the following semi-discretized form β0 ∆t ( v(m+1),v ) +ν ( ∇v(m+1),∇v ) − ( p(m+1),∇·v ) + NVF∑ i=1 Φi(v)λ (m+1) i = =− 1 ∆t K∑ k=1 βk ( v (m+1−k),v ) − 1 ∆t K∑ k=1 αk ( ∇v(m+1−k) ·v(m+1−k),v ) − − NAP∑ i=1 P (m+1) i ΦNVF+i(v) (3.4) ( q,∇·v(m+1) ) =0 Φi ( v (m+1) ) =Fi(t) i=1, ...,NVF A different mixed explicit/implicit integration scheme has been proposed by Maday et al. (1990). It consists in splitting the integration of nonlinear 28 J.Szumbarski et al. (advective) and linear parts of the Navier-Stokes equation, i.e. these parts are treated separately. Formally, the method is based on the construction of the operator integration factor for the advective part. However, it can be shown that this operator does not need to be constructed explicitly. In the sequel,wewill provide short and informal derivation of the operator- integration-factor splitting (OIFS) method. The reader should refer to the original paper (Maday et al., 1990) for amore rigorous anddetailed exposition. Consider model equation (3.1) with the assumption that the nonlinear part is autonomous, i.e. it does not depend explicitly on time. Let u(t) be an arbitrary function satisfying the differential equation du(t)/dt=N(u(t)). Then, we assume the existence of the operator integration factor Q(τ,t) such that Q(τ,τ)= Id, τ ­ t and d[Q(τ,t)u(t)]/dt=0. Assume also that the time instant t= t̃ has been fixed, and consider the following initial value problem d dt u(t)=N(u(t)) t∈ (t̃,τ) (3.5) u(t̃)= z(t̃) Note that the function z(t) satisfies original equation (3.1), i.e. the equation with both linear and nonlinear terms. In view of the postulated properties of the integration factor Q, the follo- wing equality holds Q(τ, t̃)z(t̃)=u(τ) (3.6) It can be also shown that d dt [Q(τ,t)z(t)] =Q(τ,t)L(z(t))+ [ d dt Q(τ,t) ] (z(t)−u(t))+ (3.7) +Q(τ,t)[N(z(t))−N(u(t))] Assuming the continuity of the nonlinear operator N, and taking into account the initial condition u(t̃)= z(t̃), the following equality holds at t= t̃ d dt [Q(τ,t)z(t)] ∣∣∣ t=t̃ =Q(τ, t̃)L(z(t̃)) (3.8) The two remaining components in the right-hand side of equation (3.7) vanish identically at t = t̃. Moreover, it is reasonable to neglect these terms also for t slightly larger than t̃ as they depend only on difference between the values of the functions u and z. In other words, we assume that simplified equation (3.8) is approximately valid over the short time interval (t̃ ≡ tn, Computations of an unsteady viscous flow... 29 τ = t̃+∆t≡ tn+1). Now, we integrate this equation numerically, performing one step with the implicit method based of the backward differentiation of theKth order 1 ∆t K∑ k=0 βkQ(tn+1, tn+1−k)z(tn+1−k)=Q(tn+1, tn+1)L(z(tn+1)) (3.9) With the use of equality (3.6), one can get rid of any explicit reference to the operator Q in (3.9) and finally obtain the following linear equation β0 ∆t z(tn+1)−L(z(tn+1))=− 1 ∆t K∑ k=1 βkuk(tn+1) (3.10) In the above, the functions uk (k=1, ...,K) are the solutions to the following initial value problems d dt uk(t)=N(uk(t)) t∈ (tn+1−k, tn+1) (3.11) uk(tn+1−k)= z(tn+1−k) Initial value problems (3.11) are solved using a suitable explicit integration scheme. In hydrodynamic applications, where the nonlinearity is due to the advective part of the fluid acceleration, the 4th order explicit Runge-Kutta method is preferred. Usually, only a few steps are done to march over the time interval ∆t. Note that the number of the evaluation of the operator N becomes quite large when the number of sub-steps as well as the order of the backward differentiation get larger. If, for instance, K =3 and five RK4 sub- steps are used per each time interval ∆t then the number of the evaluations of the operator N is 120. In the context of the flow problem considered in this study, the OIFS method described above can be written as follows β0 ∆t ( v (m+1),v ) +ν ( ∇v(m+1),∇v ) − ( p(m+1),∇·v ) + NVF∑ i=1 Φi(v)λ (m+1) i = =− 1 ∆t K∑ k=1 βk ( v̂ (m+1) k ,v ) − NAP∑ i=1 P (m+1) i ΦNVF+i(v) (3.12) ( q,∇·v(m+1) ) =0 Φi ( v(m+1) ) =Fi(t) i=1, ...,NVF In the above, the vector fields v̂ (m+1) k are defined as v̂ (m+1) k ≡ v̂k(tm+1) where v̂k, k = 1, ...,K, are obtained as the approximate solutions to the 30 J.Szumbarski et al. following initial value problems ∂ ∂t v̂k =−∇v̂k · v̂k t∈ (tm+1−k, tm+1) (3.13) v̂k(tm+1−k)=v(tm+1−k)≡v (m+1−k) Couzy (1995) tested systematically the OIFS method and compared then themulti-stepmethodsproposedbyKarniadakis et al. (1991). TheOIFS sche- mes proved to be superior in terms of stability properties (except for cases of low Reynolds numbers where the multi-stepmethods become unconditionally stable). On the other hand, themulti-stepmethods are free from the splitting error and usually provide better accuracy. Superior stability properties of the OIFS schemes make them more suitable for computations of laminar flows with larger Reynolds numbers, blood flows in large vessels being an example. 4. Space discretization In order to obtain a tractable algebraic problem, one has to set up a space discretization. The first step of this procedure is to define appropriate finite dimensional function spaces. These spaces are spanned by a finite number of basic functions. The velocity is approximated as a linear combination of the 3NV basic vector fields from the function space V (see the Section 2) defined as v1 = [w1,0,0] v2 = [w2,0,0] · · · vNV = [wNV ,0,0] vNV+1 = [0,w1,0] vNV+2 = [0,w2,0] · · · v2NV = [0,wNV ,0] v2NV+1 = [0,0,w1] v2NV+2 = [0,0,w2] · · · v3NV = [0,0,wNV ] (4.1) Analogously, the pressure field will be approximated within the finite dimen- sional subspace of Q spanned by the basic (scalar) functions q1,q2, ...,qNP . With the use of the discrete function spaces defined above, the unknown ve- locity and pressure field can be expressed as v(m+1) = NV∑ j=1 ( u (m+1) 1 ) j vj + NV∑ j=1 ( u (m+1) 2 ) j vNV+j + NV∑ j=1 ( u (m+1) 3 ) j v2NV+j (4.2) p(m+1) = NQ∑ j=1 ( π(m+1) ) j qj Computations of an unsteady viscous flow... 31 Let us define the following indexed structures: (MV )ij =(wi,wj)≡ ∫ Ω wiwj dx i,j =1, ...,NV (KV )ij =(∇wi,∇wj)≡ ∫ Ω ∇wi ·∇wj dx i,j =1, ...,NV A= β0 ∆t MV +νKV (4.3) (ΛF1 )ij =Φi(vj) i=1, ...,NVF , j=1, ...,NV (ΛP1 )ij =ΦNF+i(vj) i=1, ...,NAP , j=1, ...,NV (ΛF2 )ij =Φi(vj) i=1, ...,NVF , j=NV +1, ...,2NV (ΛP2 )ij =ΦNF+i(vj) i=1, ...,NAP , j=NV +1, ...,2NV (ΛF3 )ij =Φi(vj) i=1, ...,NVF , j=2NV +1, ...,3NV (ΛP3 )ij =ΦNF+i(vj) i=1, ...,NAP , j=2NV +1, ...,3NV (Dα)ij =−(qi,∂Xαwj) i=1, ...,NQ, j=1, ...,NV , α=1,2,3 (P)i =Pi i=1, ...,NAP (F)i =Fi i=1, ...,NVF Now, the algebraic problem to be solved at each time step of the numerical simulation can be written in the form of   A 0 0 (D1) ⊤ (ΛF1 ) ⊤ 0 A 0 (D2) ⊤ (ΛF2 ) ⊤ 0 0 A (D3) ⊤ (ΛF3 ) ⊤ D1 D2 D3 0 0 ΛF1 Λ F 2 Λ F 3 0 0     u1 u2 u3 π λ   (m+1) =   r1− (Λ P 1 ) ⊤P r2− (Λ P 2 ) ⊤P r3− (Λ P 3 ) ⊤P 0 F   (m+1) (4.4) The detailed structure of the right-hand side vector depends on the time discretization scheme used. In the Kth order multi-step method, the vectors r (m+1) (∗) are defined as ( r (m+1) (∗) ) i =− 1 ∆t K∑ k=1 βk (( v(m+1−k) ) (∗) ,wi ) + (4.5) − K∑ k=1 αk (( ∇v(m+1−k) ·v(m+1−k) ) (∗) ,wi ) i=1, ...,NV 32 J.Szumbarski et al. while in the case of the OIFS scheme, ( r (m+1) (∗) ) i =− 1 ∆t K∑ k=1 βk (( v̂ (m+1) k ) (∗) ,wi ) i=1, ...,NV (4.6) where v̂ (m+1) k have been defined by (3.13). In formulas (4.5) and (4.6), the subscript (∗) indicates theCartesian com- ponent, i.e. it stands for either 1, 2 or 3. 5. Construction of the solution at each time step Assuming a fixed time step in the integration scheme, we propose the following method of dealing with algebraic problem (4.4). Consider the following linear systems of equations (k=1, ...,NVF +NAP)   A 0 0 (D1) ⊤ 0 A 0 (D2) ⊤ 0 0 A (D3) ⊤ D1 D2 D3 0      u {k} 1 u {k} 2 u {k} 3 π{k}    =    −(ΛF1 ) ⊤λ{k} −(ΛF2 ) ⊤λ{k} −(ΛF3 ) ⊤λ{k} 0    (5.1) where λ {k} j = { 0 if j 6= k 1 if j= k The solutions to the above systems will be referred to as the particular Stokes solutions. For each such solution, we introduce a vector containing values of the volume flux through all inlets/outlets f {k} = [ ΛF1 ΛP1 ] u {k} 1 + [ ΛF2 ΛP2 ] u {k} 2 + [ ΛF3 ΛP3 ] u {k} 3 k=1, ...,NVF +NAP (5.2) If thegeometryof theflowdomainandthe time steparefixed, all particular solutions can be computed once and forever. Next, we define an additional problem as follows   A 0 0 (D1) ⊤ 0 A 0 (D2) ⊤ 0 0 A (D3) ⊤ D1 D2 D3 0     u {0} 1 u {0} 2 u {0} 3 π{0}   =   r1 r2 r3 0   (5.3) Computations of an unsteady viscous flow... 33 Again, the vector of the volume fluxes is computed as f{0} = [ ΛF1 ΛP1 ] u {0} 1 + [ ΛF2 ΛP2 ] u {0} 2 + [ ΛF3 ΛP3 ] u {0} 3 (5.4) Note that, in contrast to systems (5.1), the set of equations (5.3) has to be solved at each time step of the flow simulation. The solution to the full Stokes problem (4.4) can be now expressed as the following superposition   u1 u2 u3 π   =   u {0} 1 u {0} 2 u {0} 3 π{0}   + NVF∑ k=1 λk   u {k} 1 u {k} 2 u {k} 3 π{k}   + NAP∑ k=1 Pk   u {NVF+k} 1 u {NVF+k} 2 u {NVF+k} 3 π{NVF+k}   (5.5) The last term in the right-hand side of expression (5.5) contains the given values of the averaged pressures Pk (k = 1, ...,NAP). The multipliers λk, k=1, ...,NVF , are evaluated at the given time step from the following linear system TFλ=F −f {0} VF −TPP (5.6) where TF =   f {1} 1 · · · f {NVF} 1 ... ... ... f {1} NVF · · · f {NVF} NVF   TP =   f {NVF+1} 1 · · · f {NVF+NAP} 1 ... ... ... f {NVF+1} NVF · · · f {NVF+NAP} NVF   (5.7) and f {0} VF ≡ f {0}(1 : NVF). Note that the matrix TF is nonsingular as long as NVF