You are on page 1of 12

Comput. Methods Appl. Mech. Engrg.

196 (2007) 3049–3060


www.elsevier.com/locate/cma

Simulation of high-Reynolds number vascular flows


Paul F. Fischer a,*, Francis Loth b, Seung E. Lee c, Sang-Wook Lee d,
David S. Smith b, Hisham S. Bassiouny e
a
Mathematics and Computer Science Division, Argonne National Laboratory, 9700 S. Cass Avenue, Argonne, IL 60439, United States
b
Department of Mechanical Engineering, University of Illinois, Chicago, IL 60607, United States
c
Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA, United States
d
Mechanical & Industrial Engineering, University of Toronto, 5 King’s College Road, Toronto, ON, Canada
e
Section of Vascular Surgery, University of Chicago, Chicago, IL 60607, United States

Received 18 March 2006; received in revised form 10 October 2006; accepted 30 October 2006

Abstract

We describe a spectral element solution strategy for simulation of weakly turbulent vascular flows. Novel approaches for treatment of
outflow boundary conditions and flow division are presented. We demonstrate that several million gridpoints are required to obtain con-
verged rms statistics when uniform refinement is used and discuss possible approaches to cost reduction.
Ó 2007 Elsevier B.V. All rights reserved.

Keywords: Turbulence; Vascular flows; Parallel computing; Spectral elements

1. Introduction amenable to sub-grid-scale modeling required for large-


eddy or Reynolds-averaged Navier–Stokes simulations.
While much of the hemodynamics in a healthy human The only reliable simulation approach at present is to
body has low-Reynolds number, resulting in laminar flow, directly resolve all scales of motion. While the Reynolds
relatively high-Reynolds number flow is observed at some number is not high-(Re = 1000–2000, typ.), the physical
specific locations, which can cause transition to turbulence. dissipation is nonetheless small, and high-order methods
(The term ‘‘turbulence’’ refers to the motion of a fluid hav- are essential for efficiency. Moreover, turbulent blood flow
ing local velocities and pressures that fluctuate randomly.) exhibits a much broader range of scales than does its lam-
For example, the peak Reynolds number in the human inar (healthy) counterpart and thus requires an order of
aorta has been measured to be approximately 4000 [26]. magnitude increase in spatial and temporal resolution.
Surgical constructions such as the arteriovenous (AV) For example, recent work by Sherwin and Blackburn has
graft, which consists of a prosthetic graft material surgi- shown that roughly one to two million gridpoints are
cally attached between an artery and a vein, also results required for spectral/spectral-element-based simulations
in relatively high-Reynolds number flow (1000–3000) of turbulence in an idealized stenosis [43].
[8,41]. Complex geometries such as a severe stenosis also In this paper, we discuss temporal and spatial resolution
can cause turbulent flow in the vasculature [23]. requirements for direct numerical simulation in two cases
The simulation of turbulent vascular flows presents sig- where turbulence is commonly found in the vasculature,
nificant numerical challenges. Because such flows are only namely, in a stenosed carotid artery and in the venous
weakly turbulent, they lack an inertial sub-range that is anastomosis of an arteriovenous graft. We also describe
recent developments in the spectral element method
designed specifically for the simulation of high-Reynolds
*
Corresponding author. Tel.: +1 630 252 6018; fax: +1 630 252 5986. number vascular flows. The paper is organized as follows.
E-mail address: fischer@mcs.anl.gov (P.F. Fischer). Section 2 presents an outline of the spectral element

0045-7825/$ - see front matter Ó 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.cma.2006.10.015
3050 P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060

method, including a discussion of the transport properties a2 ¼ ÿ7=3, and a3 ¼ 2=3, which has a stability region that
relevant to high-Reynolds number flow simulation. Section encompasses a part of the imaginary axis, similar to third-
3 discusses the imposition of flow division and treatment of order Adams–Bashforth [21]. As an alternative to (2), we
outflow boundary conditions for turbulent flows. Section 4 frequently use the operator-integration-factor scheme of
briefly describes our mesh generation procedure. Section 5 Maday et al. [36] that circumvents the CFL (Courant–
presents the results of grid convergence studies for two tur- Friedrichs–Lewy) stability constraints by setting NLn ¼ 0
bulent flow cases, including experimental validation results. and replacing the left-hand side of (2) with an approxima-
Section 6 closes with a brief summary. tion to the material derivative of u. In either case, one ob-
tains an unsteady Stokes problem of the form
2. Navier–Stokes discretization Hun ÿ rpn ¼ f n
ð3Þ
We consider numerical solution of incompressible r  un ¼ 0;
Navier–Stokes equations in X, to be solvedÿ implicitly.For k ¼ 2; H is the Helmholtz oper-
ou 1 ator H :¼ 2Dt 3 1
ÿ Re r2 . In Section 3, we will formally refer
þ u  ru ¼ ÿrp þ r2 u; r  u ¼ 0; ð1Þ to (3) in operator form S us ðun Þ ¼ f n . In concluding our tem-
ot Re
poral discretization overview, we note that we often stabi-
subject to appropriate initial and boundary conditions. lize high-Re cases by filtering the velocity at each step
Here, u is the velocity field, p is the pressure normalized ðun ¼ F ðun ÞÞ, using the high-order filter described in
by the density, and Re = UD/m is the Reynolds number [10,13].
based on the characteristic velocity U, length scale D, Spatial discretization of (3) is based on the PN ÿ PN ÿ2
and kinematic viscosity m. For blood flow, the Newtonian spectral element method of Maday and Patera [35]. The
assumption is valid for shear rates approximately 100 sÿ1 SEM is a high-order weighted residual approach similar
and above, which generally holds in larger vessels where to the finite element method (FEM). In the SEM, functions
transition is expected to take place. The use of a rigid do- are approximated as tensor-product Lagrange polynomials
main follows current practice in the field, and its validity of degree N on each of E sub-domains (elements), Xe,
varies from vessel to vessel. For example, in arteriovenous e ¼ 1; . . . ; E, leading to n  EN d unknown basis coefficients
grafts, wall motion is on the order of 1% of the vessel dia- for each velocity component in Rd ; d ¼ 2 or 3, with N = 4–
meter [33], so a rigid assumption is a reasonable starting 16 being typical. In the PN ÿ PN ÿ2 method, the pressure is
point for analysis of the flow transition process. Such an approximated as a tensor-product polynomial of degree
assumption, however, precludes incorporation of any en- N ÿ 2 and is discontinuous, leading to np ¼ EðN ÿ 1Þ
d
ergy storage and exchange mechanism between the flowing basis coefficients for p.
blood and the elastic wall. While we anticipate studying The relatively high-polynomial degree of the SEM is
such phenomena in the near future, our focus here is on enabled by the use of tensor-product bases having the form
the numerical algorithms related to (1). Our discretization (in 2D)
of (1) is based on the spectral element method (SEM) which
is described in detail elsewhere (e.g., [6,9,10]). Here, we give N X
X N
uðxe ðr; sÞÞjXe ¼ ueij hNi ðrÞhNj ðsÞ; ð4Þ
a brief outline of the SEM and numerical timestepping
i¼0 j¼0
scheme to provide a context for the features that are spe-
cific to the simulation of high-Reynolds number vascular where the ueij s are the nodal basis coefficients on element Xe
flows. and hNi 2 PN is the Lagrange polynomial based on the
For the temporal discretization, we employ a semi- Gauss–Lobatto quadrature points, fnNj gNj¼0 (the zeros of
implicit formulation in which the nonlinear terms are trea- ð1 ÿ n2 ÞL0N ðnÞ, where LN is the Legendre polynomial of de-
ted explicitly and the remaining linear Stokes problem is gree N). Here xe ðr; sÞ is the coordinate mapping from
treated implicitly. The time derivative in (1) is approxi- ^ ¼ ½ÿ1; 1Šd to Xe, implying that the elements are curvi-
X
mated by using a kth-order backwards difference formula linear quadrilaterals in 2D or hexahedra in 3D.
(BDFk, k = 2 or 3), which for k = 2 reads In the SEM, all of the operator evaluations for explicit
timestepping and iterative solution of implicit sub-steps
3un ÿ 4unÿ1 þ unÿ2
¼ Sðun Þ þ NLn : ð2Þ are performed in matrix-free form. As first suggested by
2Dt Orszag [39], this approach leads to a reduction in memory
Here, un ÿ q represents the velocity at time tn ÿ q, q = 0, . . ., 2, and operation counts from OðEN 6 Þ (in 3D) to OðEN 3 Þ and
and Sðun Þ is the linear symmetric Stokes operator that OðEN 4 Þ, respectively. Unstructured data accesses are
implicitly incorporates the divergence-free constraint. The required at the global level (i.e., for e ¼ 1; . . . ; E), but local
term NLn approximates the nonlinear terms at time level data accesses within Xe are structured in i–j–k form. In
tn P and is given by the extrapolant NLn :¼ particular, differentiation—a central kernel in operator
ÿ j aj unÿj  runÿj . For k = 2, the standard extrapolation evaluation—can be implemented as a P cache-efficient
would use a1 = 2 and a2 = ÿ1. Typically, however, we matrix–matrix product. For example, uer;ij ¼ p D b ip ue , with
pj
use a three-term second-order formulation with a1 ¼ 8=3, b ip :¼ h0 ðni Þ would return the derivative of (4) with respect
D p
P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060 3051

to the computational coordinate r at the points ðni ; nj Þ. Dif- typical (and feasible, because of the use of matrix-free oper-
ferentiation with respect to x is obtained by the chain rule ator evaluation [6]). These high-orders lead to excellent
[6]. transport (minimal numerical diffusion and dispersion)
Inserting the SEM basis (4) into the weak form of (3) for a significantly larger fraction of the resolved modes
and applying numerical quadrature, we obtain the discrete than is possible with the FEM. This point is illustrated in
unsteady Stokes system Fig. 1, which shows the error, k, for eigenvalues associated
with the model convection problem ut þ ux ¼ 0 on ½0; 2pŠ
H un ÿ DT pn ¼ Bf n ; Dun ¼ 0: ð5Þ
versus the fraction of resolvable modes, k=k max . Here
Here, H ¼ Re 1 3
A þ 2Dt B is the discrete equivalent of H; ÿA is k max ¼ n=2, according to the Nyquist criterion, n ¼ EN is
the discrete Laplacian, B is the (diagonal) mass matrix the number of degrees of freedom for this one-dimensional
associated with the velocity mesh, D is the discrete diver- problem, and k :¼ j~k ÿ kj=k. The approximate eigenvalue
gence operator, and f n accounts for the explicit treatment is computed as ~k :¼ ð/0k ; D/k ÞN =ð/0k ; /0k ÞN ; where
of the nonlinear terms. Note that the Galerkin approach /k ðxÞ :¼ cosðkxÞ, D is the spectral element derivative oper-
implies that the governing system in (5) is symmetric and ator associated with E uniformly sized elements of order N,
that the matrices H, A, and B are all symmetric positive and (. , .)N refers to quadrature on the N + 1 Gauss–Lob-
definite. atto Legendre nodal points within each domain that also
The Stokes system (5) is solved approximately, by using correspond to the Lagrange interpolation points. Fig. 1a
the kth-order operator splitting analyzed in [3,36,40]. The shows the errors for n = 512, N = 1, 2, 4, 8, 16, and
splitting is applied to the discretized system so that ad E :¼ n=N . Taking 1% as an acceptable error threshold
hoc boundary conditions are avoided. For k = 2, one first (indicated by the dashed line in Fig. 1a), we see that 10%
solves of the modes are well resolved with linear elements
(N = 1), whereas approximately half of the modes are well
H ^u ¼ Bf n þ DT pnÿ1 ; ð6Þ resolved for N P 8. Thus, the SEM provides roughly a
fivefold reduction in the required number of gridpoints
which is followed by a pressure correction step
per space dimension to properly propagate waves at typical
Edp ¼ ÿD^u; un ¼ u^ þ DtBÿ1 DT dp; pn ¼ pnÿ1 þ dp; ð7Þ engineering tolerances. Note that, because the abscissa is
scaled by kmax, the curves in Fig. 1a exhibit little material
where E :¼ 23 DtDBÿ1 DT is the Stokes Schur complement change with increased resolution; as n increases, one
governing the pressure in the absence of the viscous term. resolves more waves, but the relative fraction remains
Sub-steps (6) and (7) are solved with preconditioned conju- unchanged. By the same token, one cannot circumvent
gate gradient (PCG) iteration. Jacobi preconditioning is the Nyquist sampling criterion by simply increasing N. In
sufficient for (6) because H is strongly diagonally domi- fact, as N ! 1, one can resolve at most ð2=pÞk max waves,
nant. E is less well-conditioned and is solved either by because of the spacing of stable (Gauss-type) point distri-
the multilevel overlapping Schwarz method developed in butions [6]. The use of moderate values of N is motivated
[9,12] or by more recent Schwarz-multigrid methods by the fact that one resolves nearly this number of waves
[11,34]. The solution of (7) constitutes the most compute- with N  8–16.
intensive sub-step of our Navier–Stokes time advancement. The benefits of a minimally dispersive/dissipative spatial
discretization are illustrated by the two-dimensional con-
2.1. Convergence properties vection problem in Fig. 1b and c, which show the solution
after an initially pointed cone is subjected to plane rotation
The primary distinction of the SEM is that it is designed on a pair of n ¼ 32  32 grids [6,19]. The second-order
for much higher approximation orders than are typically case, with E ¼ 16  16 elements, exhibits significant
used with the FEM. With the SEM, orders N = 4–16 are numerical dispersion after a single rotation. This dispersion

Fig. 1. (a) Relative error in the 1D spectrum of u ¼ kux versus fraction of resolvable modes for n = 512, E :¼ n=N , and N = 1 (d), 2 (s), 4 (+), 8 (*), and
16 (h); (b–c) spectral element solution of convected cone problem after a single plane-rotation on a 32 · 32 grid, (b) ðE; N Þ ¼ ð16  16; 2Þ, (c) ð4  4; 8Þ [6].
3052 P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060

is dramatically reduced as the order is increased to N ¼ 8 ~un satisfies S us ð~un Þ ¼ f n and u~0 satisfies S us ð~u0 Þ ¼ 0 but with
(E ¼ 4  4). The improvement is striking in light of the fact different boundary conditions, then un :¼ ~un þ ~u0 will sat-
that classical theory implies that the asymptotic conver- isfy S us ðun þ ~u0 Þ ¼ f n with boundary conditions un joX ¼
gence rate for a high-order method is no better than for ~un joX þ ~u0 joX . With this principle, the flow split for a simple
a low-order method if the problem is not smooth (e.g., bifurcation (one common inflow, two daughter outflow
well-resolved) [19]. In the present case, the benefits derive branches) is imposed as follows. In a preprocessing step:
from the fact that the high-order case is better able to prop-
agate those components of the solution that are resolved. It (i) Solve S us ð~u0 Þ ¼R 0 with a prescribed inlet profile hav-
is the advection properties rather than exponential conver- ing flux Q e :¼ ~u  ndA, and no flow (i.e., homoge-
inlet 0
gence that make the spectral element method attractive for neous Dirichlet conditions) at the exit of one of the
the simulation of turbulent flows in which one must prop- daughter branches. Use Neumann (natural) bound-
agate small scale structures over relatively long distances ary conditions at the other branch. Save the resultant
(cf. Fig. 4b and Fig. 8). velocity–pressure pair ð~u0 ; ~p0 Þ.
(ii) Repeat the above procedure with the role of the
3. Boundary conditions daughter branches reversed, and call the solution
ð~u1 ; ~p1 Þ.
Boundary conditions for the simulation of transition in
vascular flow models present several challenges not found Then, at each timestep:
in classical turbulence simulations. As velocity profiles
are rarely available, our usual approach at the vessel inflow (iii) Compute ð~un ; ~pn Þ satisfying (3) with homogeneous
is to specify a time-dependent Womersely flow that Neumann conditions on each daughter branch, R and
matches the first 20 Fourier harmonics of measured flow compute the associated fluxes Q e n :¼ u
~ n
 ndA,
i oXi
waveform. In some cases, it may be necessary to augment i = 0, 1, where oX0 and oX1 are the respective active
such clean profiles with noise in order to trigger transition exits in (i) and (ii).
at the Reynolds numbers observed in vivo. At the outflow, (iv) Solve the following for (a0 ; a1 ) to obtain the desired
our standard approach is to use the natural boundary con- flow split Qn0 :Qn1 :
ditions (effectively, p = 0 and ou
on
¼ 0) associated with the
variational formulation of (3). This outflow boundary e n þ a0 Q
Qn0 ¼ Q e ðdesired flux on branch 0Þ ð8Þ
0
treatment is augmented in two ways for transitional vascu- Qn ¼ Q e
e n þ a1 Q ðdesired flux on branch 1Þ ð9Þ
1 1
lar flows, as we now describe.
0 ¼ a0 þ a1 ðchange in flux at inletÞ ð10Þ
3.1. Fast implicit enforcement of flow division P
(v) Correct thePsolution by setting un :¼ ~un þ i ai ~
ui and
pn :¼ ~pn þ i ai ~pi .
Imposition of proper flow division (or flow split) is cen-
tral to accurate simulation of vascular flows through bifur-
cations (sites prone to atherogenesis). The distribution of Remarks
volumetric flow rate through multiple daughter branches
is usually available through measured volume flow rates. The above procedure provides a fully implicit iteration-
A typical distribution in a carotid artery bifurcation, for free approach to applying the flow split that readily extends
example, is a 60:40 split between the internal and external to a larger number of branches by expanding the system
carotid arteries. The distribution can be time-dependent, (8)–(10). Condition (10) ensures that the net flux at the inlet
and the method we outline below is applicable to such is unchanged and, for a simple bifurcation, one needs only
cases. A common approach to imposing a prescribed flow to store the difference between the auxiliary solutions. We
split is to apply Dirichlet velocity conditions at one outlet note that Sus is dependent on the timestep size Dt and that
and standard outflow (Neumann) conditions at the other. the auxiliary solutions ð~ui ; ~pi Þ must be recomputed if Dt (or
The Dirichlet branch is typically artificially extended to m) changes. (One must also recompute the auxiliary
diminish the influence of spurious boundary effects on the solutions if the geometry changes, as would be the case
upstream region of interest. Here, we present an approach when computing fluid-structure interaction. In such situa-
to imposing arbitrary flow divisions among multiple tions, the iterative approach of Gin et al. might be more
branches that allows one to use Neumann conditions at appropriate [16].) The amount of viscous diffusion that can
each of the branches, thus reducing the need for extraordi- take place in a single application of the unsteady Stokes
nary extensions of the daughter branches. operator is governed by Dt, and one finds that the auxiliary
Our flow-split scheme exploits the semi-implicit solutions have relatively thin boundary layers with a broad
approach outlined in the preceding section. The key obser- flat core. The intermediate solutions obtained in (iii) have
vation is that the unsteady Stokes operator, which is trea- inertia and so nearly retain the proper flow split, once
ted implicitly and which controls the boundary conditions, established, such that the magnitude of ai will be relatively
is linear and that superposition therefore applies. Thus, if small after just a few timesteps. It is usually a good idea to
P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060 3053

gradually ramp up application of the correction if the the standard (uncorrected Neumann) condition near the
initial condition is not near the desired flow split. outflow boundary of an internal carotid artery at
Otherwise, one runs the risk of having reversed flow on Re  1400 (based on the peak flow rate and stenosis diam-
portions of the outflow boundary and subsequent instabil- eter). Inward-pointing velocity vectors can be seen at the
ity, as discussed in the next section. Moreover, to accom- exit boundary, and the simulation becomes catastrophi-
modate the exit ‘‘nozzle’’ ðr  u > 0Þ condition introduced cally unstable within 100 timesteps beyond this point.
below, which changes the net flux out of the exit, we The center panel shows the flow field computed with the
compute Q e n at an upstream cross-section where r  u ¼ 0. outflow correction. The flow is leaving the domain at all
i
points along the outflow boundary and the simulation is
stable for all time. The difference between the two cases
3.2. Turbulent outflow boundary conditions
(right) shows that the outflow treatment does not pollute
the solution upstream of the boundary.
Turbulent flows can generate vortices of sufficient
strength to create a (locally) negative flux at the outflow
boundary. Because the Neumann boundary condition does 4. Mesh generation
not specify flow characteristics at the exit, a negative flux at
the outflow can rapidly lead to instability, with cata- Spectral element mesh generation shares much in com-
strophic results. One way to eliminate incoming character- mon with its FE counterpart. Several important distinc-
istics is to force the exit flow through a nozzle, effectively tions, however, place constraints on the SE meshes. First,
adding a mean axial component to the velocity field. The the use of matrix-free operator evaluation, which reduces
advantage of using a nozzle is that one can ensure that the storage from OðEN 6 Þ to OðEN 3 Þ and work per evalua-
the characteristics at the exit point outward under a wide tion from OðEN 6 Þ to OðEN 4 Þ, is dependent upon the tensor-
range of flow conditions. By contrast, schemes based on product forms (4). This reduction is most effectively
viscous buffer zones require knowledge of the anticipated achieved with hexahedral elements, which may be curved
space and time scales to ensure that vortical structures through the use of iso- or sub-parametric mappings from
are adequately damped as they pass through the buffer ^ to Xe [6]. Second, the fact that typical orders are in the
X
zone. range N = 8–16 implies roughly a thousandfold reduction
Numerically, a nozzle can be imposed without change to in the number of elements required compared to an FE
the mesh geometry by imparting a positive divergence to mesh at comparable resolution. The reduction, while
the flow field near the exit (in the spirit of a supersonic noz- advantageous in reducing the size of the input files, places
zle). In our simulations, we identify the layer of elements significant constraints on the mesh generation scheme.
adjacent to the outflow and there impose a divergence Consequently, we have developed an SE mesh generation
DðxÞ that ramps from zero at the upstream end of the layer scheme that is specific to vascular geometries [28,44].
to a fixed positive value at the exit. Specifically, we set Our mesh generation scheme employs a sweeping algo-
DðxÞ ¼ C½1 ÿ ðx? =L? Þ2 Š, where x? is the distance normal rithm, in which disc-shaped slabs are meshed by using a
to the boundary and L? is maximum thickness of the last standard O-grid configuration, as illustrated in Fig. 3.
layer of elements. By integrating the expression for D from The slab surfaces are identified with iso-surfaces of conduc-
x? =L? ¼1 to 0, one obtains the net gain in mean velocity tion (potential) solutions satisfying homogeneous Neu-
over the extent of the layer. We typically choose the con- mann conditions along the vessel wall. The iso-surfaces
stant C such that the gain is equal to the mean velocity are computed by numerically solving a sequence of Laplace
prior to the correction. One could, however, increase the equations, one for each branch, in the computational
gain if stronger fluctuations are encountered. domain on a preliminary mesh comprising tetrahedral ele-
Results for the nozzle-based outflow condition are illus- ments. Because of the robustness of the conduction prob-
trated in Fig. 2. The left panel shows the velocity field for lem, the preliminary mesh can be of almost arbitrary

Fig. 2. Velocity vectors near the outflow of an internal carotid artery: (left) uncorrected, (center) corrected, and (right) corrected minus uncorrected.
3054 P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060

Fig. 3. Swept templates, projected onto isopotential surfaces, for hexahedral mesh generation in a carotid bifurcation geometry. The closeup on the right
shows the principal iso-surfaces from the three conduction problems that provide a uniquely determined trisection of the domain into independently swept
branches [44].

quality. The iso-surfaces define a set of natural coordinate converge only in the mean and higher-order statistics, which
systems that have the desirable properties of being orthog- require long-time simulations to eliminate natural fluctua-
onal to the vessel walls and of being guaranteed to produce tions as a source of variance. For pulsatile turbulent flows,
nonintersecting cross-sections that could otherwise lead to it is necessary to use phase averaging, in which one samples
vanishing Jacobians associated with the transformation the solution at a certain phase of the cardiac cycle over a
from X ^ to Xe. Through a judicious choice of conduction large number of cycles and then averages these results.
problems, one can identify three principal iso-surfaces that Experiments in the transitional flow regime have shown that
naturally trisect the bifurcation geometry, as depicted in relatively slow pulsatility coupled to rapidly evolving turbu-
Fig. 3(right). This trisection leads to a decomposition of lence merely acts as a switch that turns the turbulence on or
the bifurcation into three branches, each of which is indi- off and does not materially change the turbulent state [1,25].
vidually meshed with the sweeping algorithm. The advan- Thus, given the length of the cardiac cycle (>10 flow-
tages of the conduction-based approach is that it can be through times for a typical bifurcation model) and the num-
automated (starting with a segmented stack, mesh genera- ber of samples required to reach a statistically stationary
tion requires a matter of minutes on a workstation [44]) state, it is preferable to test for grid convergence by simply
and it produces high-quality all-hexahedral meshes with a using steady inlet conditions at the peak Reynolds numbers.
minimum number of topology-induced defects. (For exam- One can then exploit temporal homogeneity and ergodicity
ple, there is only a single 3-valence vertex in the bifurca- to obtain mean and rms velocity distributions that can be
tion-plane mesh cross-section and only four 3-valence used for spatial convergence tests.
vertices in any axial cross-section.) As with its global spectral counterpart, the standard
Given a generated hex mesh, we curve the lumen surface convergence procedure for the spectral element method is
using bicubic patches that match surface edges and surface to increase the polynomial degree N for a fixed number
normals at each vertex. This provides C1 continuity at the of elements E. It is also possible, however, to use an adap-
vertices and, for smooth surfaces, converges cubically with tive procedure in which one refines the mesh and varies the
respect to the surface patch diameter. Higher continuity polynomial degree locally to obtain optimal convergence
spline patches could also be used, but we have found this rates, as is done with hp finite element methods [22,37]
approach adequate for the cases that we have considered. (see Section 5.3). Our general approach has been to con-
The cubic patch deformation is extended into the element struct a mesh that is finest in the region of interest, starting
volumes using the transfinite interpolation algorithm of with relatively low-degree (typ., N = 4), and to then
Gordon and Hall [18]. increase N as we increase the Reynolds number. We typi-
cally construct the mesh such that, at the target Reynolds
5. Resolution requirements number, N = 8–12 will be sufficient. This range of N is typ-
ically optimal for the performance of our spectral element
An important component of any numerical study is the code and, as illustrated in Fig. 1, provides a significant frac-
establishment of adequate resolution. This is particularly tion of the maximal benefit to be derived from high-order
challenging in turbulent flows because one can expect to approximations.
P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060 3055

In this section, we examine the resolution requirements flow and leads to high-WSS. This high-WSS and the pres-
for direct numerical simulation of turbulence in two cases, ence of turbulence may damage the endothelial cells [14,20]
flow in a stenosed carotid artery and flow in an arterio- and play a role in platelet activation [42] or plaque rupture
venous graft. As a measure of convergence, we present [15]. In addition, the large spatial WSS gradient associated
time-averaged and rms velocity distributions as a function with turbulent flow is thought to enhance mass transport
of resolution (N) for a fixed number of elements. We note into the arterial wall proximal to the stenosis, which may
that convergence of these profiles is really a minimal weaken the plaque and make it vulnerable to rupture [5].
requirement for spatial convergence. For example, Ethier High-temporal shear gradients are also shown to stimulate
et al. [7] have pointed out that accurate determination of endothelial cell proliferation [45]. The ability to accurately
wall shear stress or wall shear stress gradient distributions predict the hemodynamics in such an environment is of
represent even more stringent convergence criteria, even in interest to clinicians and researchers alike.
the case of laminar flows. We have undertaken a numerical study of pulsatile flow
in a carotid bifurcation model that exhibits a severe steno-
5.1. Stenosed carotid artery flow sis (75% area reduction) in the internal carotid artery
(ICA) [29]. The computational mesh, shown in Fig. 4a
Atherothrombotic carotid stenoses, followed by ische- and b, comprises E = 2544 spectral elements generated by
mic stroke, is one of the leading causes of mortality and using the potential-based approach described in Section
morbidity in Western countries [38]. The presence of 4. A time-dependent Womersely profile, synthesized from
high-grade stenosis manifested by plaque deposits can phase averages of measured flow waveforms, was imposed
result in transition to turbulent flow, which may produce at the inlet to the common carotid artery (CCA), and a
an audible sound (bruit) discernible by a physician. 41:59 split was imposed between the external (ECA) and
Many studies have been conducted to characterize local internal (ICA) branches throughout the 0.75 s cardiac cycle
hemodynamics and their role on the early development of by using the technique described in Section 3. The Rey-
atherosclerosis in arteries. Atherosclerotic-prone sites are nolds number, based on the ICA diameter and bulk veloc-
often localized at bifurcations, junctions, and regions of ity, ranged from 380 at diastole to 1320 at peak systole. For
high-curvature, which are also regions of low-wall shear N = 9, a timestep of Dt = 5 · 10ÿ6 s (150 000 steps/cycle)
stress (WSS) and flow disruption [2,4,17,27,46]. guaranteed a Courant–Freidrichs–Lewy (CFL) number of
In contrast, regions downstream of severe constrictions <0.5 throughout the cycle, which ensured stability of the
(post-stenotic regions) experience a significantly different semi-implicit time advancement. Each cardiac cycle
biomechanical environment than do healthy vessels required 20 h of CPU time on 256 processors of the
because of the presence of transitional and turbulent flow. TCS1 parallel computer at the Pittsburgh Supercomputer
Because flow resistance is primarily controlled by smaller Center.
vessels downstream (arterioles and capillaries), the pres- Although our primary focus is on the resolution require-
ence of a stenosis does not materially change the flow rate. ments for turbulent vascular flows, we present some flow
The area reduction within a stenosis thus accelerates the details here in order to indicate where resolution is

Fig. 4. (a) and (b) E = 2544 element mesh for stenosed carotid artery simulations and (c) coherent vortical structures at systolic mid-deceleration phase
identified by using the k2 criterion of Jeong and Hussain [24].
3056 P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060

required. (A more detailed description of this set of simula- satile case. Simulations with N = 7, 9, and 11 (n = 850 000,
tions is provided in [29].) Fig. 4c shows a set of typical vor- 1 820 000, and 3 338 000 points, respectively) were run from
tical structures, identified by the k2 criterion of Jeong and the same turbulent initial condition for 33 flow-through
Hussein [24], in the mid-deceleration phase past peak sys- times, with statistics collected over the last 30 flow-through
tole. The presence of transverse rolls, stretched by the mean times. The average and rms velocity profiles in the ICA
shear, and the increase in small-scale structure downstream mid-plane are shown in Figs. 6 and 7, respectively. The
of the stenosis are hallmarks of the transition process. results show that N = 9 and 11 give nearly identical mean
Nearer to the systolic peak (not shown) the flow is signifi- and rms distributions, whereas N = 7 shows significant
cantly more complex, with less clearly identifiable struc- deficiencies, particularly for the rms.
ture. The ICA is to be contrasted with the ECA, where
the flow exhibits one or two dominant axial vortices that 5.2. Transition in an arteriovenous graft
do not break down—the flow in the ECA remains laminar.
Further evidence of the spatio-temporal resolution require- Arteriovenous (AV) grafts consist of a 15 cm section
ments is given in Fig. 5(center), which shows time traces of of 6 mm i.d. synthetic tubing that is surgically implanted
axial velocity at points (a)–(f), indicated in Fig. 5(left). The to provide an arterial-to-vein round-the-clock short circuit.
passage of small-scale structures registers as high-frequency Because they connect a high-pressure vessel to a low-pres-
variation in velocities in the post-stenotic region. A Fourier sure one, high-flow rates are established that make AV
transform of the time trace for (f), shown in Fig. 5(right), grafts efficient dialysis ports for patients suffering from
reveals significant spectral peaks in the range 150– poor kidney function. The high-speed flow is normally
500 Hz, which is much higher than the order-unity frequen- accompanied by transition to a weakly turbulent state,
cies associated with the cardiac cycle. manifested as a 200–300 Hz vibration at the vein wall
To address the question of resolution, we have under- [30,33]. This high-frequency excitation is thought to lead
taken a series of runs with steady inlet conditions at a flow to intimal hyperplasia, which can lead to complete occlu-
rate corresponding to that just beyond the systolic peak, sion of the vein and graft failure within six months of
where maximum turbulence intensity is observed in the pul- implant. We are investigating the mechanisms leading to

Fig. 5. Pulsatile velocity results for a stenosed carotid artery: (left) time history points, (center) time traces of axially aligned velocity for four cardiac
cycles, and (right) axial velocity energy spectra for point (f).

Fig. 6. Comparison of time-averaged velocity profiles for transitional flow under steady inlet conditions: (a) N = 7, (b) N = 9, (c) N = 11, (d) N = 9 and 11
overlaid.
P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060 3057

Fig. 7. Comparison of rms profiles of axial velocity for transitional flow under steady inlet conditions: (a) N = 7, (b) N = 9, (c) N = 11, (d) N = 9 and 11
overlaid.

transition in subject-specific AV-graft models with the aim However, the rms of velocity showed noticeable change
of reducing turbulence through improved geometries. between polynomial orders, even in the results of N = 12.
Detailed comparisons with laser Doppler anemometry This dependency could be attributed to statistical variance
(LDA) measurements are presented in [32]. Results for a over the collection period or indicate the need for still
pulsatile flow study are given in [30], and the influence of higher grid resolution. However, the results with N = 12
the flow division between the proximal venous segment are adequate for demonstrating the primary features of
(PVS) and distal venous segment (DVS) is discussed in [31]. the flow since the rms of velocity is observed to be bounded
Fig. 8 shows a typical turbulent case when there is a in Fig. 10 with increasing polynomial order.
70:30 split between the PVS and DVS. Significant small- Based on these grid independence studies, the numerical
scale structures are generated downstream (toward the results with N = 12 were used for validation with experimen-
heart) of the anastomosis in the PVS, which channels the tal measurements obtained using laser Doppler anemometry
majority of the flow. Steady graft inlet conditions are (LDA), as described in [32]. Fig. 11 shows a comparison of
imposed, with a mean (cross-sectional average) velocity mean and rms cross-sectional velocity distributions in the
of U0. The Reynolds number based on the graft diameter turbulent PVS flow with a 70:30 flow split at Re = 1200.
is Re :¼ U 0 D=m ¼ 1200. The SEM solution was computed Measurements were taken at stations x/D = 1.34, 2.34,
with E = 2640 elements of order 12 (4.5 million gridpoints). and 3.34 (see Fig. 9). It is clear that the N = 12 case is able
A grid independence study was performed at Reynolds to qualitatively predict both the mean flow distribution
number 1200 with a flow split of 70:30, corresponding to and the rms fluctuations. A more quantitative analysis of
the conditions shown in Fig. 8. Comparisons of time-aver- these results is presented in [32]. Quantitative comparisons
aged velocity and root-mean-square (rms) of the velocity for an idealized AV-graft model are given in [33].
fluctuation with N ¼ 7, 10, and 12 for the 70:30 flow divi-
sion are shown in Figs. 9 and 10, respectively. Statistics 5.3. Discussion of resolution requirements
were gathered for 1 s after an initial transient of 0.15 s,
which was not included in the statistics. (For comparison, Several remarks are in order regarding the resolution
the mean flow-through time is 0.1 s in vivo units; requirements for these flows. First, the coherent structures
D = 6 mm and U0 = 649 mm/s). The time-averaged veloc- in Figs. 4 and 8 are indicative of the resolution require-
ity did not show significant change for N = 7, 10 or 12. ments for these weakly turbulent flows when contrasted

Fig. 8. Coherent structures in an AV graft at Re = 1200 with a 70:30 (PVS:DVS) flow split. The simulation employed E = 2640 elements of order N = 12
and dt = 5 · 106 s. The structures are visualized using the k2 criterion of Jeong and Hussein [24].
3058 P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060

Fig. 10. Comparison of rms velocity profiles in the AV graft for


transitional flow under steady inlet conditions with a 70:30 (PVS:DVS)
flow split: (a) urms, (b) vrms, and (c) wrms.

for the AV graft are not grid converged at N = 10. It is


possible that, despite collecting statistics over ten flow-
through times, a lack of converged temporal statistics is
the cause of the discrepancies in the AV graft profiles
Fig. 9. Comparison of time-averaged velocity profiles in the AV graft for
transitional flow under steady inlet conditions with a 70:30 (PVS:DVS)
(i.e., that they result from different realizations rather than
flow split: (a) u, (b) v, and (c) w. lack of spatial convergence). In both geometries, the mean
profiles are well-converged (within a few percent) even at
N = 7, which corresponds to 900 000 points for these con-
to their laminar counterparts, where the majority of the figurations. It is possible that even fewer points could be
structures are on the scale of the vessel diameter. Second, used for converged mean profiles and converged mean wall
we note that grid convergence is established in the case of shear stresses. This study, however, did not explore that
the carotid model for N = 9, but that the rms statistics limit.

Fig. 11. Numerical (SEM)/experimental (LDA) validation for AV graft flow with a 70:30 flow split and Re = 1200: cross-sectional mean and rms velocity
distributions (m/s) at x/D = 1.34 (A), 2.34 (B), and 3.34 (C) (see Fig. 9).
P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060 3059

Given the complexity of the flow regimes and the rela- [2] C.G. Caro, J.M. Fitz-Gerald, R.C. Schroter, Arterial wall shear and
tively small physical dissipation, it seems likely that the distribution of early atheroma in man, Nature 223 (211) (1969) 1159–
1160.
optimal solution approach for this class of problems will [3] W. Couzy, Spectral element discretization of the unsteady Navier–
be hp adaptive methods. The p-type refinement strategy Stokes equations and its iterative solution on parallel computers,
(increased polynomial order p, corresponding to the SEM Ph.D. thesis, Swiss Federal Institute of Technology-Lausanne, 1995,
N) exhibits the same dispersion and dissipation properties Thesis no. 1380.
as the SEM and is therefore a good choice in the high-Re [4] M.J. Davies, N. Woolf, P.M. Rowles, J. Pepper, Morphology of the
endothelium over atherosclerotic plaques in human coronary arteries,
limit. However, it is natural to build adaptivity into such Br. Heart J. 60 (6) (1988) 459–464.
methods, based on local error estimates. One could thus [5] N. DePaola, M.A. Gimbrone, P.F. Davies, C.F. Dewey, Vascular
avoid costly over-resolution in regions that are relatively endothelium responds to fluid shear stress gradients, Arterioscler.
quiescent. Such adaptivity could save roughly a factor of Thromb. 12 (11) (1992) 1254–1257.
three in the case of the carotid, since little resolution is [6] M.O. Deville, P.F. Fischer, E.H. Mund, High-order methods for
incompressible fluid flow, Cambridge University Press, Cambridge,
required in the CCA and ECA, and perhaps 30–50% in 2002.
the AV graft, which requires little resolution in the inlet. [7] C.R. Ethier, S. Prakash, D.A. Steinman, R.L. Leask, G.G. Couch, M.
For pulsatile simulations, the savings for the carotid can Ojha, Steady flow separation patterns in a 45 degree junction, J. Fluid
be even more significant because the turbulence is only Mech. 411 (2000) 1–38.
present over one fifth of the cardiac cycle (see Fig. 5). Such [8] M.F. Fillinger, E.R. Reinitz, R.A. Schwartz, D.E. Resetarits, A.M.
Paskanik, C.E. Bredenberg, Beneficial effects of banding on venous
dramatic savings will not likely be realized for AV grafts, intimal-medial hyperplasia in arteriovenous loop grafts, J. Vasc. Surg.
however, as they typically sustain high-flow rates with min- 11 (4) (1990) 556–566.
imal pulsatility [33]. [9] P.F. Fischer, An overlapping Schwarz method for spectral element
solution of the incompressible Navier–Stokes equations, J. Comput.
6. Summary Phys. 133 (1997) 84–101.
[10] P.F. Fischer, G.W. Kruse, F. Loth, Spectral element methods for
transitional flows in complex geometries, J. Sci. Comput. 17 (2002)
We have presented methodology and convergence 81–98.
results for the application of a high-order spectral element [11] P.F. Fischer, J.W. Lottes, Hybrid Schwarz-multigrid methods for the
method to the simulation of vascular flows. We have dem- spectral element method: Extensions to Navier–Stokes, in: R.
onstrated a fast implicit approach to imposition of arbi- Kornhuber, R. Hoppe, J. Périaux, O. Pironneau, O. Widlund, J.
Xu (Eds.), Domain Decomposition Methods in Science and Engi-
trary flow division among daughter branches that avoids neering Series, Springer, Berlin, 2004.
the need for extraordinary extensions or iteration to deter- [12] P.F. Fischer, N.I. Miller, H.M. Tufo, An overlapping Schwarz
mine exit pressures. We have also developed an outflow method for spectral element simulation of three-dimensional
treatment for turbulent flows that eliminates incoming incompressible flows, in: P. Bjørstad, M. Luskin (Eds.), Parallel
characteristics that can destabilize the simulation. We have Solution of Partial Differential Equations, Springer, Berlin, 2000, pp.
158–180.
demonstrated some of the desirable transport properties of [13] P.F. Fischer, J.S. Mullen, Filter-based stabilization of spectral
high-order discretizations that are relevant to turbulent element methods, Comptes rendus de l’Académie des sciences, Sér. I
flows. The results indicate that roughly 2–4 million points – Anal. Numér. 332 (2001) 265–270.
are required for direct numerical simulation of turbulent [14] D.L. Fry, Acute vascular endothelial changes associated with
vascular flows in bifurcation geometries at clinically rele- increased blood velocity gradients, Circ. Res. 22 (2) (1968) 165–197.
[15] S.D. Gertz, W.C. Roberts, Hemodynamic shear force in rupture of
vant Reynolds numbers, using the uniform refinement coronary arterial atherosclerotic plaques, Am. J. Cardiol. 66 (19)
(increased N) approach employed here. The results have (1990) 1368–1372.
been validated through comparisons with LDA [16] R. Gin, A.G. Straatman, D.A. Steinman, A dual-pressure boundary
measurements. condition for use in simulations of bifurcating conduits, J. Biomech.
Engrg. 124 (2002) 617–619.
[17] S. Glagov, C.K. Zarins, D.P. Giddens, D.N. Ku, Hemodynamics and
Acknowledgements atherosclerosis. insights and perspectives gained from studies of
human arteries, Arch. Pathol. Lab. Med. 112 (1988) 1018–1031.
This work was supported by the National Institutes of [18] W.J. Gordon, C.A. Hall, Transfinite element methods: blending-
Health, RO1 Research Project Grant (2RO1HL55296- function interpolation over arbitrary curved element domains,
04A2), by Whitaker Foundation Grant (RG-01-0198), Numer. Math. 21 (1973) 109–129.
[19] D. Gottlieb, S.A. Orszag, Numerical analysis of spectral methods:
and by the Mathematical, Information, and Computa- Theory and applications, SIAM-CBMS, Philadelphia, 1977.
tional Sciences Division sub-program of the Office of Ad- [20] J.D. Hellums, The resistance to oxygen transport in the capillaries
vanced Scientific Computing Research, US Department relative to that in the surrounding tissue, Microvasc. Res. 13 (1)
of Energy, under Contract DE-AC02-06CH11357. (1977) 131–136.
[21] L.W. Ho, A Legendre spectral element method for simulation
of incompressible unsteady viscous free-surface flows, Ph.D. the-
References sis, Massachusetts Institute of Technology, 1989, Cambridge, MA,
USA.
[1] N. Arslan, Experimental characterization of transitional unsteady [22] L.C. Hsu, C. Mavriplis, Adaptive meshes for the spectral element
flow inside a graft-to-vein junction, Ph.D. thesis, University of method, in: P. Bjørstad, M. Espedal, D. Keyes (Eds.), Domain
Illinois, Chicago, 1999, Department of Mechanical Engineering. Decomposition 9th Proc., John-Wiley, New York, 1997, pp. 374–381.
3060 P.F. Fischer et al. / Comput. Methods Appl. Mech. Engrg. 196 (2007) 3049–3060

[23] K.J. Hutchison, E. Karpinski, In vivo demonstration of flow [35] Y. Maday, A.T. Patera, Spectral element methods for the Navier–
recirculation and turbulence downstream stenoses in canine arteries, Stokes equations, in: A.K. Noor, J.T. Oden (Eds.), State-of-the-Art
J. Biomech. 18 (1985) 285–296. Surveys in Computational Mechanics, ASME, New York, 1989, pp.
[24] J. Jeong, F. Hussain, On the identification of a vortex, J. Fluid Mech. 71–143.
285 (1995) 69–94. [36] Y. Maday, A.T. Patera, E.M. Rønquist, An operator-integration-
[25] P.S. Klebanoff, W.G. Cleveland, K.D. Tidstrom, On the evolution of factor splitting method for time-dependent problems: Application to
a turbulent boundary layer induced by a three-dimensional roughness incompressible fluid flow, J. Sci. Comput. 5 (1990) 263–292.
element, J. Fluid Mech. 92 (1992) 101–187. [37] C. Mavriplis, Adaptive mesh strategies for the spectral element
[26] D.N. Ku, Blood flow in arteries, Annu. Rev. Fluid Mech. 29 (1997) method, Comput. Methods Appl. Mech. Engrg. 116 (1994) 77–86.
399–434. [38] C.J. Murray, A.D. Lopez, Alternative projections of mortality and
[27] D.N. Ku, D.P. Giddens, C.K. Zarins, S. Glagov, Pulsatile flow and disability by cause 1990–2020: Global burden of disease study, Lancet
atherosclerosis in the human carotid bifurcation, Arteriosclerosis 5 (3) 349 (9064) (1997) 1498–1504.
(1985) 293–302. [39] S.A. Orszag, Spectral methods for problems in complex geometry,
[28] S.E. Lee, Solution method for transitional flow in a vascular J. Comput. Phys. 37 (1980) 70–92.
bifurcation based on in vivo medical images, Master’s thesis, Univ. [40] J.B. Perot, An analysis of the fractional step method, J. Comput.
of Illinois, Chicago, 2002, Department of Mechanical Engineering. Phys. 108 (1993) 51–58.
[29] S.E. Lee, S.W. Lee, P.F. Fischer, H.S. Bassiouny, F. Loth, Direct [41] S.J. Ram, A. Magnasco, S.A. Jones, A. Barz, L. Zsom, S. Swamy,
numerical simulation of transitional flow in a stenosed carotid W.D. Paulson, In vivo validation of glucose pump test for measure-
bifurcation, J. Biomech., submitted for publication. ment of hemodialysis access flow, Am. J. Kidney Dis. 42 (4) (2003)
[30] S.W. Lee, P.F. Fischer, F. Loth, T.J. Royston, J.K. Grogan, H.S. 752–760.
Bassiouny, Flow-induced vein-wall vibration in an arteriovenous [42] J.M. Ramstack, L. Zuckerman, L.F. Mockros, Shear-induced acti-
graft, J. Fluids Struct. 20 (2005) 837–852. vation of platelets, J. Biomech. 12 (2) (1979) 113–125.
[31] S.W. Lee, D.S. Smith, F. Loth, P.F. Fischer, H.S. Bassiouny, [43] S.J. Sherwin, H.M. Blackburn, Three-dimensional instabilities and
Importance of flow division on transition to turbulence within an transition of steady and pulsatile axisymmetric stenotic flows, J. Fluid
arteriovenous graft, J. Biomech., in press. Mech. 533 (2005) 297–327.
[32] S.W. Lee, D.S. Smith, F. Loth, P.F. Fischer, H.S. Bassiouny, [44] C.S. Verma, P.F. Fischer, S.E. Lee, F. Loth, An all-hex meshing
Numerical and experimental simulation of transitional flow in a blood strategy for bifurcation geometries in vascular flow simulation, in:
vessel junction, Numer. Heat Transfer 51 (2007) 1–22. Proc. of the 14th Int. Meshing Roundtable Conf., San Diego, 2005.
[33] F. Loth, N. Arslan, P.F. Fischer, C.D. Bertram, S.E. Lee, T.J. [45] C.R. White, M. Haidekker, X. Bao, J.A. Frangos, Temporal
Royston, R.H. Song, W.E. Shaalan, H.S. Bassiouny, Transitional gradients in shear, but not spatial gradients, stimulate endothelial
flow at the venous anastomosis of an arteriovenous graft: Potential cell proliferation, Circulation 103 (20) (2001) 2508–2513.
relationship with activation of the ERK1/2 mechanotransduction [46] C.K. Zarins, D.P. Giddens, B.K. Bharadvaj, V.S. Sottiurai, R.F.
pathway, ASME J. Biomech. Engrg. 125 (2003) 49–61. Mabon, S. Glagov, Carotid bifurcation atherosclerosis. quantitative
[34] J.W. Lottes, P.F. Fischer, Hybrid multigrid/Schwarz algorithms for correlation of plaque localization with flow velocity profiles and wall
the spectral element method, J. Sci. Comput. 24 (2005) 45–78. shear stress, Circ. Res. 53 (4) (1983) 502–514.

You might also like