Você está na página 1de 18

Cardiovascular Engineering and Technology, Vol. 4, No. 4, December 2013 ( 2013) pp.

374–391
DOI: 10.1007/s13239-013-0166-2

Results of FDA’s First Interlaboratory Computational Study of a Nozzle


with a Sudden Contraction and Conical Diffuser
SANDY F. C. STEWART,1 PRASANNA HARIHARAN,1 ERIC G. PATERSON,2 GREG W. BURGREEN,3
VARUN REDDY,2 STEVEN W. DAY,4 MATTHEW GIARRA,4 KEEFE B. MANNING,2 STEVEN DEUTSCH,2
MICHAEL R. BERMAN,1 MATTHEW R. MYERS,1 and RICHARD A. MALINAUSKAS1
1
Office of Science and Engineering Laboratories, Food and Drug Administration,
Silver Spring, MD, USA; 2Pennsylvania State University, University Park, PA, USA; 3Mississippi State University, Starkville, MS,
USA; and 4Rochester Institute of Technology, Rochester, NY, USA
(Received 6 May 2013; accepted 11 September 2013; published online 27 September 2013)

Associate Editor Ajit P. Yoganathan oversaw the review of this article.

Abstract—The U.S. Food and Drug Administration recently purposes need to be conducted following the best practices
hosted an interlaboratory study to assess the suitability and available, and that targeted experimental validation is
methodology of computational fluid dynamics (CFD) for essential. The results of this interlaboratory study are freely
demonstrating medical device safety in regulatory submis- available in an internet repository (https://fdacfd.nci.nih.gov).
sions. The benchmark study was performed in a generic We encourage the use of this model in further studies, and
medical device consisting of a 0.012 m diameter cylindrical support the development of additional benchmarks, better
nozzle with a sudden contraction and 10 conical diffuser, on modeling techniques, and consensus standards and guidelines
either side of a 0.04 m long, 0.004 m diameter throat. Results for using CFD in the evaluation of medical devices.
from 28 simulations from around the world were compared
to planar particle image velocimetry (PIV) measurements Keywords—CFD validation, Standards, Medical device reg-
performed at three laboratories. Five flow rates were chosen ulation, Best practices, Blood flow.
that produced laminar, transitional, and turbulent regimes.
In general, the CFD results showed modest agreement in
global and local flow behaviors. However, all CFD data sets
contained wide degrees of velocity variation in comparison to
INTRODUCTION
experiment and with each other, much of which could be
attributed to the turbulence models used. Some of the
velocity discrepancies in the sixteen three dimensional (3D)
Computational fluid dynamics (CFD) is a popular
simulations resulted from substantial flow asymmetries tool used to study and develop complex medical de-
within and downstream of the conical diffuser. Large vices such as endovascular stents,6 cannulae,8 pros-
differences in velocity symmetry were found even when using thetic heart valves,25,35,37 and ventricular assist
the same code and many of the same simulation parameters. devices.14,26,36 However, the usefulness of CFD in
In contrast, no velocity asymmetries of the same magnitude
were observed in any of the planar PIV experiments;
assessing safety in a way that can facilitate the regu-
relatively minor asymmetries in one experiment were linked latory process has received less study. In the cardio-
to slight asymmetries in the velocity profile at the inlet to the vascular area, the main fluid component of interest is
nozzle. CFD predictions of peak wall shear stress at the blood, which is extremely difficult to model. The ~50%
sudden contraction (normalized to that 0.015 m down- cellular content can still only be fully modeled at lim-
stream) varied over two orders of magnitude, with variations
of greater than 10 times even when the same software and
ited scale sizes, while the full chemistry of coagulation
turbulence model were used. This degree of shear stress is as yet too complex to be modeled without significant
variation will necessarily propagate through calculations of simplifications. One technique used to evaluate blood
blood damage in medical devices. We conclude that CFD damage, such as hemolysis, is by assessing the effects
simulations used in qualifying medical devices for regulatory of shear stress and exposure time in a continuum fluid.
Most blood damage models depend on empirical shear
stress/exposure time data gathered decades ago.
Address correspondence to Sandy F. C. Stewart, Office of Science
However, if the physics of the continuum blood flow is
and Engineering Laboratories, Food and Drug Administration, not modeled correctly, then the computed biological
Silver Spring, MD, USA. Electronic mail: sandy.stewart@fda.hhs.gov effects cannot be trusted.20 Therefore, this interlaboratory
374
1869-408X/13/1200-0374/0  2013 Biomedical Engineering Society (Outside the U.S.)
Results of FDA’s First Interlaboratory Computational Study 375

study had three main goals: (1) to take a snapshot of of goal #2, will be detailed in a future report. Data
the current practice in medical device fluid dynamics from this study is available on our website,
computations, (2) to compare hemolysis levels pre- https://fdacfd.nci.nih.gov, and is already being used in
dicted by the computations, and (3) provide bench- benchmark studies (goal #3).3,9,10,15,17,34,38
mark models for validation work in the medical device
arena. In this paper, we concentrate on goal #1.
This is a continuation of our earlier work which METHODS
evaluated flow through the nozzle in the opposite
direction, through a conical concentrator and sudden Participants in the computational interlaboratory
expansion.31 In that paper, we found variations in study were asked to model flow in a generic example of
velocities relative to the experiments that depended a tubular medical device containing a sudden con-
primarily on the turbulence models used. Recirculation traction and conical diffuser (Fig. 1), similar to models
in the post-expansion region was difficult to model at used previously.18,32 This model shares flow charac-
the higher Reynolds numbers. No one turbulence teristics with some common medical devices (e.g.,
model proved superior overall when applied to this catheters, needles) and medical device problems (e.g.,
problem. We also found wide variations in shear stress kinks in dialysis tubing). Nozzle throat Reynolds
distributions (particularly the wall shear stress). As numbers (Ret) were specified at 500, 2000, 3500, 5000,
mentioned above, this has particular relevance for and 6500 under steady flow conditions. A Newtonian
blood damage modeling: if the shear stresses cannot be viscosity of 0.0035 N-s/m2 and a density of 1056 kg/m3
accurately predicted, then reliable and accurate were specified. All other parameters (inlet/outlet
numerical prediction of shear-induced blood damage lengths, mesh density, cell/element type, axisymmetric
cannot be performed. We noted that higher mesh vs. 3D mesh, inlet boundary conditions, etc.) were left
densities did not necessarily provide better match to to the discretion of the individual participants. More
the velocities; thus careful experimental validation is details of the interlaboratory study can be found in our
needed, with monitoring of shear stresses in areas rel- previous report,31 and at our data repository website,
evant to blood damage. We found no particular cor- https://fdacfd.nci.nih.gov.
relation between self-rated expertise and accuracy, and The CFD simulations were compared to experi-
we concluded that there was little uniformity of prac- mental data obtained using 2D planar particle image
tice among the 28 users. CFD modeling of an actual velocimetry (PIV) in machined and polished acrylic
device must take into consideration many complica- nozzles as described previously.16 Steady flow veloci-
tions: short and/or curved entrance and exit conduits, ties were measured at the cross-sectional model loca-
pulsatile flow, transitional flows, secondary flows, tions shown in Fig. 1b, at three separate laboratories
sharp corners, and a range of viscosity and hematocrit (two experimental data sets obtained at one lab, one
under which the device is used. We concluded that each at the other two labs). The nozzle models were
CFD can be a useful part of regulatory evaluations, fitted with straight inlet tubes ~1.20 m long (~100 tube
but that successful computational simulation of this diameters) to ensure fully developed flow at the nozzle
nozzle model will not aid in a regulatory submission inlet. Outlet tubes ~1.0 m long were also used to pre-
where the model is not applicable. vent outlet flow conditions from influencing the reat-
The current geometric configuration allowed for the tachment point in the model.
examination of CFD performance in an environment The fluid used in the PIV experiments was com-
of a continuously varying adverse pressure gradient, posed of 50 wt% saturated sodium iodide (NaI) solu-
present in the conical diffuser. The diffuser geometry tion, 20 wt% glycerin, and 30 wt% water, formulated
allowed investigation into how different models pre- to match the index of refraction of the acrylic nozzles
dicted flow separation at different locations in the (~1.49) used in the experimental flow loops. The range
presence of the adverse pressure gradient, which may of dynamic viscosity and density was equal to 1.66–
be relevant to any device where flow decelerates. 1.77 9 103 kg/m3 and 6.74–7.77 9 103 N s/m2,
Likewise, the sudden contraction, not studied in the respectively. To ensure proper control of the flow
first paper, is relevant to implants containing small conditions, the dynamic viscosity and density of the
orifices, such as gaps between heart valve leaflets, or to fluid were measured before each PIV measurement.
diameter mismatches between cannulae and dialysis The flow rate was adjusted according to the fluid
tubing. Regions just upstream of the sudden contrac- property variations to match the specified throat
tion are challenging locations for the computation of Reynolds numbers. The PIV measured velocities were
wall shear stress, which are important for hemolysis then scaled to those in the computational simulations
computations. The actual hemolysis calculations, part using dynamic similarity.13 All axial velocities (PIV
376 STEWART et al.

FIGURE 1. Nozzle specifications: (a) dimensions of nozzle (inlet and outlet lengths unspecified); (b) cross-sectional cuts defined
for data submission for the Conical Diffuser.

and CFD) were then normalized by the mean axial to blood and glycerin, respectively.16 All the CFD and
velocity calculated from the velocity profile at the inlet experimental pressure values were normalized by the
(Fig. 1b, location 1, z = 0.048): downstream value at axial coordinate z = 0.09032 m.
uz
uz ¼ ð1Þ
uz
RESULTS
In Eq. (1), uz = normalized velocity, uz = velocity
from PIV or CFD, and uz = average velocity at the Axial Velocity Along Nozzle Centerline
inlet, calculated as follows:
At Ret = 500, the experiments showed that the
Qinlet normalized axial velocity along the centerline increased
uz ¼ ; ð2Þ
Ainlet abruptly at the sudden contraction (Fig. 2a), continued
to increase within the nozzle throat as the flow devel-
where Qinlet = flow rate at the inlet (calculated from
oped, then slowly decreased as a function of axial
the velocity profile determined from PIV or CFD), and
position within the conical diffuser. A jet formed
Ainlet = area of the 0.012 m diameter inlet tube.
within the conical diffuser associated with low velocity
Pressures were validated against measurements
recirculation zones at the wall (the diffuser angle being
made in a separate wall pressure tap model mounted in
too large to smoothly diffuse the flow at this Ret), and
the same flow loop, using an aqueous glycerin solution.
the jet widened and slowed as the diffuser area
Measured pressures were converted to those of the
increased with axial position downstream. While the
simulations by the following equation to preserve
PIV results demonstrated good interlaboratory com-
similarity according to the Euler number:
parability, the results from the simulations showed a
qb u2b high degree of variation. Many of the participants used
Pb ¼ Pg ; ð3Þ turbulence models (including variations of k-epsilon,
qg u2g
k-omega, and shear stress transport models) at
where P = pressure, q = density, u = axial velocity at Ret = 500, even though this is below the critical
inlet averaged over inlet area, and subscripts b and g refer Reynolds number for pipe flow. The wide degree of
Results of FDA’s First Interlaboratory Computational Study 377

variation of results (and use of turbulence models at errors in the datasets that could not be corrected. For
Ret = 500) was also observed in the nozzle model comparative purposes, the pressures for each case
when oriented with the sudden expansion at the exit (simulation and experimental) were referenced to that
(the Sudden Expansion orientation).31 Axial velocities at z = 0.09032 m, the most downstream point mea-
predicted by many of the k-omega/shear stress trans- sured in the wall tap experimental model (near location
port (SST) models matched those of the experiment 11 in Fig. 1b). As expected, both the simulations and
well, but only up to a point, then abruptly fell off experiments showed higher pressure gradients in the
between 0.08 and 0.12 m downstream of the sudden small diameter throat than in the inlet and outlet tubes.
contraction. However, even some of the laminar sim- An adverse pressure gradient at the wall was predicted
ulations did not accurately predict the centerline axial in many cases just downstream of the sudden con-
velocity. Two of the poorly performing laminar models traction, due to the presence of a vena contracta (seen
(labeled #1 and #3 in Fig. 2a) used outlet lengths that experimentally at Ret greater than 500; Figs. 3b–3d).
were likely too short for the flow to develop properly An adverse pressure gradient was also found within the
for the outlet boundary conditions used. Triangular diffuser as the fluid separated from the wall, with low
and tetahedral cells are thought to be less accurate for velocities in recirculation zones. The inset pictures in
CFD than quadrilateral and hexahedral cells of the Fig. 3 show the simulated centerline pressures. The
same size, and indeed some of the poorly performing most noticeable difference between the simulated wall
laminar simulations used triangular cells (#1, #3, and and centerline pressures is apparent just downstream
#5) or tetrahedral cells (#2 and #4). However, some of of the sudden contraction (z ‡ 0.0 m), where the wall
the other triangle and tetrahedral simulations per- pressure was somewhat lower than the centerline
formed better. Other than the short outlet length, pressure in many simulations. As flow enters a sudden
however, the problems in the laminar simulations were contraction, the velocities are higher near the wall (as
not readily identifiable. the flow accelerates around the corner), but lower at
At Ret = 2000 in the experiments (Fig. 2b), flow the centerline. The flow typically separates, forming a
was observed to be laminar within the inlet and nozzle small recirculation zone (Fig. 3e), with a correspond-
throat, and transitional within the cone and outlet. The ing increase in wall pressure. Just downstream of the
jet broke down within the outlet tube, the location of sudden contraction, the centerline velocity increases as
which varied considerably among the simulations. The the flow develops. Contour images supplied by par-
one Spalart–Allmaras (SA) model matched the cen- ticipants (not shown, but available at https://fdacfd.
terline velocity well, while the SST and k-omega (KO) nci.nih.gov) suggest that these recirculation zones may
models performed better than the k-epsilon (KE) not have been resolved in some of the simulations,
models. There was considerable variation among the particularly if turbulent wall functions with a large cell
SST, KO, and KE models. size in the near-wall layer were used.
The KE and SST models appeared to bracket the
experimental centerline velocity as the Ret increased
(Figs. 2c and 2d). At Ret = 6500, the various SST Volumetric Flow Rate
models matched the throat centerline velocity better
than did the KE models, while the reverse was true Volumetric flow rates were calculated by integrating
within the diffuser. In contrast, in the Sudden Expan- over the axial velocity profiles at each of the 12 cross-
sion orientation, the SST models always performed sectional cuts, to give a conservation of mass error
better in the nozzle throat and downstream of the metric as a function of axial position z (Fig. 4)31:
sudden expansion, although the KE models QCFD  Qtheory
approached the SST models as Ret increased.31 At all EQ ¼  100% ð4Þ
Qtheory
Ret (including Ret = 5000, not shown) there was little
overlap between the KE and SST models. In general, where QCFD = the volumetric flow rate computed
the SST models tended to transition to turbulence from the CFD axial velocity profiles, and Qtheory =
farther downstream than did the KE models, but they theoretical volumetric flow rate calculated from the
did not necessarily do better in matching the centerline throat Reynolds number and diameter. For axisym-
velocities at all Re. metric meshes, QCFD was calculated from zero radius
to the outer radius of the nozzle, along each of the
twelve axial velocity profiles. For 3D meshes, QCFD
Pressure Variations
was calculated across the diameter. The radius was
Figures 3a–3d show the wall pressures, as well as dependent on location (inlet, outlet, throat, or cone). A
the pressures along the centerline (insets). Two to three simple Riemann sum in the radial direction was used
simulations per Ret were omitted due to formatting for the numerical integration. When compared to the
378 STEWART et al.

FIGURE 2. Axial velocity along nozzle centerline, normalized to inlet flow rate (see text): (a) Ret 5 500; (b) Ret 5 2000; (c)
Ret 5 3500; (d) Ret 5 6500. Lines are CFD with color and style representing the turbulence model used. Experimental data
(symbols) are plotted as the mean 6 95% confidence interval (n 5 4). In (a), Ret 5 500, laminar CFD results labeled 1, 3, and 5 were
reported to be axisymmetric, with triangular elements. Laminar CFD results labeled 2 and 4 were reported to be 3D tetrahedral.

theoretical flow rate, a considerable degree of variation non-zero velocities beyond the wall at the throat radius
was observed in a number of the simulations, as well as (r < 0.002, r > 0.002 m) in some of the data. Errors
in some of the experiments. In addition, several 3D in the experiments at z = 0 m were due to subpotimal
simulations demonstrated substantial flow asymme- visualization through the perpendicular wall at the sud-
tries within the conical diffuser, causing errors in the den contraction, and due to residual machining marks.
flow rate calculated from Eq. (4), which assumes
symmetric flow.31 The asymmetries were likely due to
Flow Asymmetry
instabilities resulting from the adverse pressure gradi-
ent visible within the conical diffuser in Fig. 3. To Flow asymmetry in the 3D simulations was char-
differentiate conservation of mass errors from asym- acterized by calculating the ratio of the volumetric flow
metries, only the errors upstream of the diffuser were rate (calculated as described above) on one side of the
evaluated (i.e., locations 1–3 and 5–7 in Fig. 1b). Mass centerline to the other, to produce a number £1.0, with
conservation errors at the sudden contraction (location a value of 1.0 indicating perfect symmetry16 (Fig. 5).
4, z = 0 m) were also neglected because of errors in The degree of asymmetry was highly variable from
extracting data from the participants’ solutions, giving simulation to simulation, and reached negative values
non-zero interpolated velocities along the perpendicu- in some cases (representing net negative flow on one
lar wall of the sudden contraction. Examples of such side of the centerline). Asymmetries developed within
errors are visible in panel c of Figs. 6–9, which show the conical diffuser, but resolved further downstream.
Results of FDA’s First Interlaboratory Computational Study 379

FIGURE 3. Wall pressure (main figures) referenced to value at z 5 0.09032 m (red lines over nozzle outlines): (a) Ret 5 500; (b)
Ret 5 2000; (c) Ret 5 3500; (d) Ret 5 6500. Lines are CFD results plotted by families of turbulence model used. Experimental data
(symbols) from wall tap measurements are plotted as the mean 6 95% confidence interval (n 5 3). Inset plots are corresponding
CFD pressures along centerline also compared to experimental wall tap measurements. Panel (e) is a contour plot of one simu-
lation at Ret 5 3500, showing a localized low pressure zone adjacent to wall just downstream of sudden contraction.

Asymmetries were also calculated from the experi- mass errors £10% upstream of the conical diffuser. At
ments, but were lower in magnitude than in the sim- Ret = 500, the experimental axial velocity profiles
ulations. We return to this subject below, and in the demonstrated laminar flow throughout the nozzle, as
‘‘Study limitations’’ section. would be expected from the low Reynolds number at
the model inlet (e.g., Reinlet = 167 for Ret = 500;
Fig. 6a). Despite many of the participants submitting
Axial Velocity Profiles
turbulent simulations, most of the simulations were
For clarity, the figures in this section are limited to able to accurately predict velocity in the nozzle
the 64-68% of the simulations with conservation of inlet (Figs. 6a and 6b) and throat (Figs. 6d and 6e).
380 STEWART et al.

FIGURE 4. Experimental and simulated flow rates in Conical Diffuser: (a) Ret 5 500; (b) Ret 5 2000; (c) Ret 5 3500; (d)
Ret 5 6500. Lines are calculated from CFD results plotted by families of turbulence model used. Symbols are calculated from
experimental PIV measurements. Dotted black lines are the theoretical flow rates calculated from the Ret.

However, all of the simulations underestimated the than two of the laminar simulations. In the throat, the
velocities by about 25% in most of the central area of predictions were fairly good (Figs. 7d and 6e); however,
the sudden contraction (Fig. 6c), with some overesti- the central velocities were again underestimated by all
mation near the wall (also note the low calculated flow the simulations at the sudden contraction (Fig. 7c). In
rates at z = 0 in Fig. 4a). As mentioned above, a few the diffuser, the turbulent models began to fail, with the
of the simulations showed non-zero velocities at radii KO and KE models performing worse than the SST and
greater than 0.002 m, possibly due to errors in data SA models (Figs. 7f and 6g). Just downstream of the
extraction. Within and downstream of the diffuser diffuser (Fig. 7h), only the laminar simulations matched
(Fig. 6f through i), simulations using the KE models the measurements; however, further downstream the
diverged markedly from measurements, while the experimental jets were observed to be in transition, so
laminar, KO, and SST models generally performed that the turbulence models performed better than the
better, even predicting the recirculation zones in the laminar simulations (Fig. 7i). Thus the quality of pre-
diffuser and just downstream (Figs. 6f–6h). One of the diction of mean velocities in the transition to turbulence
laminar simulations did poorly, however, possibly due is very much dependent on the Reynolds number, the
to too-short an outlet length. Further downstream, position of the jet breakdown, and the turbulence model
nearly all of the turbulent simulations performed used. Furthermore, the KO model (a steady flow 3D
poorly (Figs. 6h and 6i). study with hexagonal elements) showed a striking
At Ret = 2000, many of the simulations did not asymmetry that was not observed in the experiments or
accurately predict the parabolic velocity profile at the other simulations (Figs. 7f–7i).
inlet observed in the experiments (Fig. 7a). Interest- At Ret = 3500, the inlet flow was observed to be
ingly, three of the SST simulations performed better laminar, while most of the turbulent models assumed a
Results of FDA’s First Interlaboratory Computational Study 381

FIGURE 5. Flow rate asymmetry calculated from axial velocities. Lines are calculated from CFD axial velocities (3D simulations
only), plotted by families of turbulence model used. Symbols are calculated from experimental axial velocities. (a) Ret 5 500; (b)
Ret 5 2000; (c) Ret 5 3500; (d) Ret 5 6500.

turbulent velocity profile (Figs. 8a and 8b). Predictions models performing better than the others in the inlet
at the sudden contraction followed trends similar to and throat (Figs. 9a–9e), and the KE models per-
Ret = 500 and 2000. Errors in one of the experimental forming better in the diffuser and outlet (Fig. 9g–9i).
profiles caused large confidence intervals on one side. Again, several of the simulations showed marked
Within the nozzle throat, the SST models predicted the asymmetries in the outlet. These trends were also seen
flow profiles fairly well (Figs. 8d and 8e). The SST, SA, at Ret = 5000 (not shown).
and laminar models performed well in the diffuser, Similar simulation parameters can lead to markedly
while the KE simulations began to fail there (Figs. 8f different levels of asymmetry. Two such cases are
and 8g). Obvious flow asymmetries were observed in compared in Fig. 10. The axial velocity profiles ori-
some cases, in the diffuser and outlet (Figs. 8f–8i). ented at 0, 45, 90, and 135 are plotted for
At Ret = 6500 (Fig. 9), the flow rate in one set of Ret = 6500, just downstream of the conical diffuser
experiments was 10% too high, which tripped the flow (z = 0.072 m, location 10 in Fig. 1b). The parameters
at the inlet from the laminar to the turbulent regime; for both simulations are shown in Table 1. Despite
the other experiments were laminar at the inlet many similarities—same code, same turbulence model
(Fig. 9a). (This experimental error resulted in the lar- (SST), same cell shape (hexahedral)—only one of the
ger confidence intervals in Fig. 9.) Among the simu- two simulations showed patently asymmetric flow.
lations, similar results to Ret = 3500 were found; only Figure 10 also shows the meshes used in the two sim-
the laminar simulation predicted a parabolic inlet ulations, as well as the user-provided color velocity
velocity profile. Inconsistent CFD predictions were contour plots, with the locations of the velocity con-
found with the turbulent models overall, with the SST tours superimposed. One significant difference between
382 STEWART et al.

FIGURE 6. Normalized axial velocity profiles for throat Reynolds number 5 500: (a) z 5 20.024 m (nozzle walls are at radial
locations 20.006 and 0.006 m), (b) z 5 20.004 m, (c) z 5 0.0 m (nozzle walls are at radial locations 20.002 and 0.002 m), (d)
z 5 0.004 m, (e) z 5 0.032 m, (f) z 5 0.048 m, (g) z 5 0.056 m, (h) z 5 0.072 m, and (i) z 5 0.100 m. Lines are CFD with color and
style representing the turbulence model used. The experimental data (symbols) are plotted as the mean 6 95% confidence interval
(n 5 4). Simulations with conservation of mass errors >10% upstream of the conical diffuser were omitted.

the two simulations is that the symmetric study was pffiffiffiffiffiffiffiffiffiffi


performed in steady flow, while the asymmetric one jsj ¼ 2:0l Sij Sij ; ð5Þ
was a transient study. However, the asymmetric KO where Sij = components of strain-rate tensor,
results observed at Ret = 2000, 3500, and 6500 in l = dynamic viscosity, and repeated indices indicate
Figs. 7–9 (as well as at Ret = 5000, not shown) were summation. Computed wall shear stress values along
from simulations performed under steady flow condi- the nozzle are shown in Fig. 11. CFD data was re-
tions, using a hexahedral mesh. Thus steady vs. tran- quested from participants as a function of axial dis-
sient conditions cannot be the sole reason for the tance only, so that, in 3D simulations, the angular
difference. position (h) on the wall was not known. Figure 11a
shows all CFD data for Ret = 500; for clarity, the
Wall Shear Stress (WSS)
remaining panels only show 0.5% of the data points,
For clarity, the figures in this section are limited to because some of the variations in WSS with h were
simulations with conservation of mass errors £10%. large enough to obscure the rest of the data. In general,
Computational shear stress magnitudes, jsj, were re- the most consistent simulations were those using SST
quested from participants performing simulations, and models; however, even within the SST models, the
defined as31: variations in WSS within the nozzle throat appeared to
Results of FDA’s First Interlaboratory Computational Study 383

FIGURE 7. Normalized axial velocity profiles for throat Reynolds number 5 2000: (a) z 5 20.024 m, (b) z 5 20.004 m, (c)
z 5 0.0 m, (d) z 5 0.004 m, (e) z 5 0.032 m, (f) z 5 0.048 m, (g) z 5 0.056 m, (h) z 5 0.072 m, and (i) z 5 0.100 m. Lines are CFD
with color and style representing the turbulence model used. The experimental data (symbols) are plotted as the mean 6 95%
confidence interval (n 5 4). Simulations with conservation of mass errors >10% upstream of the conical diffuser were omitted.

increase with Ret. Salient features include a sharp peak and gradient artifacts.4,5,12,19,22,24,33 With PIV being
in WSS at the sudden contraction, as well as a local used as a common validation tool for CFD, our future
minimum just downstream of the contraction as evi- studies will focus on addressing some of these issues in
dence of a recirculation zone. an effort to improve the accuracy of WSS measure-
Wall shear stress values estimated from PIV were ments.
not reported in Fig. 11 because the inter-laboratory Variations in the WSS results from only the SST
variability in the experimental WSS was very high models are shown in greater detail in Fig. 12, which
(~100%). The primary reason for the large variability compares simulations using two commercial CFD
in PIV WSS was inadequate spatial resolution in the codes (codes A and B). In this figure, the WSS has been
PIV images which prevented us from resolving the normalized to the value at z = 0.015 m (i.e., 0.015 m
viscous sub-layer (near wall region), especially for downstream of the sudden contraction and within the
turbulent flows. In addition, the presence of large near- throat region). The location of the reattachment point
wall velocity gradients and the uncertainties involved (indicated by the minimum WSS) as well as the size of
in identifying the exact wall-location also contributed the peak WSS at the sudden contraction varied con-
to the measurement uncertainty. Previous studies have siderably from simulation to simulation, despite the
shown that some of these issues could potentially be use of the same CFD code. See Table 2 for details of
addressed using high resolution images and by using the simulations presented in Fig. 12a. The information
special near-wall algorithms that minimize the imaging in Table 2 reveals no obvious relationship between the
384 STEWART et al.

FIGURE 8. Normalized axial velocity profiles for throat Reynolds number 5 3500: (a) z 5 20.024 m, (b) z 5 20.004 m, (c)
z 5 0.0 m, (d) z 5 0.004 m, (e) z 5 0.032 m, (f) z 5 0.048 m, (g) z 5 0.056 m, (h) z 5 0.072 m, and (i) z 5 0.100 m. Lines are CFD
with color and style representing the turbulence model used. The experimental data (symbols) are plotted as the mean 6 95%
confidence interval (n 5 4). Simulations with conservation of mass errors >10% upstream of the conical diffuser were omitted.

simulation parameters and the variations in reattach- laminar simulations, which assumed laminar flow
ment point location or peak WSS. everywhere, and the turbulence models, which
assumed turbulent flow everywhere, had difficulty
predicting the velocities over the entire computational
DISCUSSION domain. At Ret = 3500, three of the SST models
(Fig. 8, red dashed lines) performed well throughout
Many of the same observations reported here for the domain, with realistic parabolic input profiles,
CFD simulations of the Conical Diffuser orientation of blunted flow within the throat, and accurate recircu-
the nozzle interlaboratory study were also made for the lation zone predictions within the diffuser. Two of
Sudden Expansion orientation, including the wide these three SST simulations were axisymmetric (with
variety of methods used, common errors made, and quadrilateral cells), both using the same commercial
variations in results obtained.31 The most obvious er- program. The one 3D simulation used hexahedral cells
ror appeared to be the use of turbulence models at and a second commercial program. The 3D simulation
Ret = 500, where experimentally the flow was and one of the axisymmetric simulations did not use
observed to be laminar throughout. At higher Ret, the wall functions. The minimum cell size varied from
experimental flows were observed to be laminar at the 1.2 9 1013 to 7.9 9 109 m, and the number of cells
inlet, with a transition from laminar to turbulent flow varied from ~52,000 for the two axisymmetric simu-
occurring at the sudden contraction. Thus both the lations to 1.4 9 106 for the 3D simulation. Thus, these
Results of FDA’s First Interlaboratory Computational Study 385

FIGURE 9. Normalized axial velocity profiles for throat Reynolds number 5 6500: (a) z 5 20.024 m, (b) z 5 20.004 m, (c)
z 5 0.0 m, (d) z 5 0.004 m, (e) z 5 0.032 m, (f) z 5 0.048 m, (g) z 5 0.056 m, (h) z 5 0.072 m, and (i) z 5 0.100 m. Lines are CFD
with color and style representing the turbulence model used. The experimental data (symbols) are plotted as the mean 6 95%
confidence interval (n 5 4). Simulations with conservation of mass errors >10% upstream of the conical diffuser were omitted.

three SST simulations with good results were obtained Sudden Expansion orientation.31 This may in part be
using a variety of methods. In contrast, all of the KE due to data interpolation/extraction errors, an obvious
models at Ret = 3500 (Fig. 8, blue dash-dot lines) example of which can be seen at the sudden contrac-
underpredicted the central flow everywhere except for tion itself (Fig. 6c), where velocities at the wall should
the downstream region of the throat and the outlet be zero due to the no-slip condition. Any non-zero
(Figs. 8e and 8i). The SST models were less successful velocities interpolated erroneously to the wall of the
at Ret = 6500. Because there are so many variables sudden contraction (r < 0.002 m, r > 0.002 m) were
that could influence the results, any two simulations neglected in our volumetric flow rate calculation, and
may predict similar results for vastly different reasons. thus contributed to some of the low flow rates at
Thus there is no one-size-fits-all solution for this par- z = 0.0 m in Fig. 4. Although this did not explain all
ticular problem and range of flow rates, emphasizing the errors (there were far fewer flow rate errors at the
the need for making well-informed choices in simula- sudden expansion with flow going in the opposite
tions. Such variation among turbulence models (albeit direction31), it does suggest that extrapolation/inter-
at higher Reynolds numbers) have been observed in polation errors within the flow domain may have
diffusers in previous studies.1,21 contributed to violations of the conservation of mass.
Conservation of mass did not appear to be obeyed Several of the 3D simulations of the Conical Dif-
in some of the simulations; we also observed this in the fuser nozzle configuration showed significant flow
386 STEWART et al.

FIGURE 10. Axial velocity profiles at z 5 0.072 m downstream of sudden contraction (as in Fig. 9g) for two 3D simulations at
Ret 5 6500. Upper left shows normalized axial velocity profiles at four angular cuts: dashed lines, Simulation 1 (asymmetric
results); solid lines, Simulation 2 (symmetric results, all of which superimpose). Upper right: comparison of user-supplied
(hexahedral) mesh plots for same two simulations. Bottom row: comparison of user-supplied images of cross-sectional cuts of
axial velocities; lines refer to placement of profiles in upper left plot. Refer to Table 1 for comparison of simulation parameters.

TABLE 1. Downstream symmetry: model parameters of two simulations at Ret 5 6500 shown in Fig. 10.

Parameter Simulation 1 Simulation 2

Code Code A Code A


Turbulence model SST SST
Z coordinate of inlet 0.12 0.06
Z coordinate of outlet 0.18 0.30
Total Number of cells 6.6 9 105 13.4 9 105
Dominant cell type Hex Hex
Wall functions Standard wall functions Automatic
Time integration Transient (time step = 9.5 9 104) Steady
Upstream velocity profile Turbulent profile Parabolic profile
Outlet boundary condition Averaged pressure = 0 Not given
Self-ascribed expertise Expert Beginner
Flow symmetry within diffuser Asymmetric Symmetric

asymmetries developing in the area of adverse pressure symmetric sudden expansions and diffusers, both
gradient in the diffuser, despite the symmetric geome- experimentally7,11 and numerically.23,27 The asymme-
try and flow inlet boundary conditions. Similar asym- tries observed in the 3D simulations of this study may
metries have been reported in laminar flow in have resulted from the evolution of the unstable flow
Results of FDA’s First Interlaboratory Computational Study 387

FIGURE 11. Wall shear stress from computations as a function of axial coordinate, z: (a) Ret 5 500; (b) Ret 5 2000; (c)
Ret 5 3500; (d) Ret 5 6500. Lines are CFD results plotted against turbulence model used. Simulations with conservation of mass
errors >10% upstream of the conical diffuser are omitted. In panel a, all points are shown regardless of azimuth, h. In panels b, c,
and d, only a subset of points are shown for clarity; some CFD simulations at higher Ret had a wide variation of WSS with azimuth,
due to asymmetric flow.

through the adverse pressure gradient in the diffuser. successfully captured in many cases despite suboptimal
This decelerating flow is highly sensitive to any per- predictions of velocity and shear stress. Note that we
turbations generated by mechanisms such as round-off did not request data about whether unsteady simula-
errors or mesh nonuniformities. The sudden contrac- tions were performed as first-order, second-order, or
tion geometry just upstream is an additional compli- hybrid; this would be good candidate data to collect in
cation not included in earlier studies. future interlaboratory studies. Whether this level of
Complex unsteady flows due to flow instabilities in detail is necessary for evaluating computational studies
areas of adverse pressure gradients have also been in regulatory submissions is unknown.
found in computational and experimental models of a Some of the PIV experiments also showed minor flow
total cavopulmonary connection in repair of single asymmetries that were much less significant than most of
ventricle heart disease,30 in which four vessels are sur- those of the computational simulations and were
gically connected into a cross-shaped junction, with attributed to slight inlet flow asymmetries. Experimental
two inflow branches (superior and inferior vena cava) flow asymmetries may also have been due to imperfect
and two outflow branches (right and left pulmonary axial alignment of the inlet tubing, nozzle throat, cone,
arteries). Good agreement in global results (control and outlet sections of the acrylic models (although de-
volume power losses) were found between experiment tailed standard operating procedures were in place to
and steady simulations, but accurate description of the minimize these16). Numerical artifacts leading to
detailed velocity field required a second-order accurate asymmetries may have included asymmetric, non-flow-
unsteady solution on a very fine computational mesh aligned meshes (e.g. tetrahedral meshes) that produce
that could resolve (but not exactly duplicate) the dis- asymmetric regions of artificial dissipation, or charac-
turbed flow in the center of the junction.30 In the cur- teristics associated with specific turbulence models. For
rent study, global pressure drops across the model were unsteady calculations, such instabilities may arise as a
388 STEWART et al.

transient start-up anomaly and persist for an unex-


pectedly large number of time steps before ever con-
vecting out of the domain. Numerical round-off errors
in the geometry, mesh, and solvers (including the tur-
bulence models used) may also have contributed to
instabilities. Because of the unstable nature of deceler-
ating flow, even minor perturbations occurring in a
perfectly symmetric situation (whether experimental or
numerical) can be amplified leading to visible asymme-
tries. Asymmetries are likely to occur in medical devices
due to transient flow pulses, changes in flow geometries,
and curvature effects. Therefore, we contend that in
both computations of flow in medical devices, and in
experimental validations thereof, vigilance is needed to
identify and understand instabilities and asymmetries,
and to carefully assess their real-world importance.
Flow of blood past corners and edges is a concern
because of the possibility of high shear stresses in these
areas. In this nozzle model, the sudden contraction is the
location of the highest shear stresses and is the prime
location where blood damage and/or platelet activation
might occur. This is especially true near the wall, where
the shear stress is high but the absolute velocities are low,
so that the exposure time is also high (although less fluid
volume would be expected to be exposed because of the
lower velocities). Furthermore, the small recirculation
zone just downstream of the sudden contraction may
allow activated platelets to aggregate. As large varia-
tions in the wall shear stress magnitudes and distribu-
tions near the sudden contraction occurred in the
simulations, the prediction of blood damage or platelet
FIGURE 12. Wall shear stress (WSS) downstream of sudden aggregation near such features must be carefully vali-
contraction for SST computations at Ret 5 3500, for CFD dated before its clinical impact can be determined.
codes A (top) and B (bottom), normalized to the value at
z 5 0.015 m for clarity. Symbols indicate 3D simulations where
A brief summary of CFD ‘‘Best Practices’’ identified
submitted WSS values were not uniform as a function of h at a in this study can be found in the Appendix, to com-
single z location. *TF 5 Transitional flow option enabled. plement those listed in our previous report.31 Note that

TABLE 2. Details of wall shear stress (WSS) at sudden contraction for seven selected SST studies (all performed with Code A, as
shown in Fig. 12, top panel), at Ret 5 3500.

Plot # in Fig. 12, top panel, Code A

Data from file header 1 2 3 4 5 6 7

Mesh type 3D 3D 3D 3D Axisym. Axisym. Axisym.


Cell type Hexahedral Hexahedral Hexahedral Tetrahedral Quadrilateral Quadrilateral Triangle
Experience (1 = novice, 10 = expert) 8 2 4 4 6 10 7
Inlet length (m) 0.120 0.060 0.090 0.250 0.298 0.090 0.180
Outlet length (m) 0.183 0.303 0.120 0.313 0.270 0.153 0.220
#Cells 661200 1339392 2998272 1897091 20486 130232 342450
Cell minimum volume (m3) 1.20E12 1.22E13 6.09E11 3.23E13 4.40E15 6.00E11 3.95E15
Cell maximum volume (m3) 3.10E10 2.27E10 1.89E13 9.01E09 2.20E10 7.49E08 3.85E11
Turbulence model SST SST SST SST SST* SST SST
Turbulence wall functions Standard Automatic Standard Automatic None Standard None
Time integration Transient Steady Steady Steady Steady Steady Steady
Max normalized WSS 1.64 6.94 1.08 2.98 16.10 1.07 21.51
z coordinate of minimum WSS (m) 0.00346 0.00675 0.00275 0.00047 0.00400 0.00400 0.00620

*Study #5 specified ‘‘RANS k-omega SST’’ in file header; all others specified ‘‘SST’’ or ‘‘Shear Stress Transport’’.
Results of FDA’s First Interlaboratory Computational Study 389

computational modeling is not required in device pre- Future Work


market notifications and applications submitted to the
The goal of this study is to identify issues with and
US FDA. However, modeling may play a useful role as
provide information on the use of CFD in regulating
part of a device submission when the goals of the study
medical devices. We are currently using this informa-
are carefully designed within a regulatory context.
tion in the development of regulatory guidance for the
Despite our use of turbulence models to help organize
reporting of computational studies to the FDA. As
the data presented herein, we do not endorse or rec-
members of a number of standards groups (e.g.,
ommend the use of any particular turbulence model
ASME V&V 40 for Medical Devices), we will provide
over any other.
any knowledge gained from this study to the stan-
dards-developing community.
Study Limitations As these studies are published, all of the experi-
mental and computational data from this project will
All the limitations identified in our previous report
be posted at https://fdacfd.nci.nih.gov, so as to be
apply.31 In addition, we note that the 2D PIV derived
available to anyone who wishes to use it for further
velocities were measured from only one plane in the
development. We plan to run additional experimental
nozzle; thus any flow asymmetries were not fully
and computational interlaboratory studies in the fu-
characterized. Also, for Fig. 5, only the asymmetry
ture. Please visit the website or contact us for more
from one plane was calculated for each 3D simulation
information.
(an example of looking at multiple planes is shown in
Fig. 10). Thus in both experimental and computational
cases, any out-of-plane asymmetry (that is symmetric
within the plane) may have led only to a reduction in APPENDIX: BEST PRACTICES (CONTINUED
the calculated flow rate (as in Fig. 4), rather than to a FROM REFERENCE31)
flow asymmetry <1.0 (as in Fig. 5). This demonstrates
Additional CFD ‘‘Best Practices’’ identified in this
the indeterminancy of the azimuthal position of an
study are listed below. These are not necessarily all-
asymmetry in both computations and experiment.
inclusive, nor do they guarantee an accurate simula-
Despite these limitations, it is clear that the experi-
tion.
mental asymmetries that were observed were much
smaller than those seen in some of the simulations. 18. Asymmetric or disturbed flow can occur in
We did not attempt to use an error metric in this areas of decelerating flow (adverse pressure
study, because errors were highly dependent on loca- gradients), both in computations and experi-
tion in our previous analysis.31 Error metrics have their mental data. Such disturbances may occur in
uses, and we recommend they be employed in, e.g., laminar, transitional, or turbulent regimes.
determining the best combination of simulation These disturbances may be truly physically
parameters (i.e., sensitivity analysis) per the methods present, or they may be spurious and arti-
described by Oberkampf and Barone.28 factual, particularly in numerical computa-
In the PIV experiments, larger than expected 95% tions. Validation in these areas is challenging
confidence intervals were observed in some locations at because the nature and origin of these dis-
most Reynolds numbers (Figs. 6–9), due to a combi- turbances may differ subtly among bench
nation of factors: (1) errors in the inlet flow rate experiments, computer simulations, animal
(especially in one experiment) which changed the models, and the clinical setting. True physical
character of the flow, especially in the inlet and outlet asymmetries will not be accurately computed
regions; (2) optical interference by the forward facing if an axisymmetric simulation is run, or may
step at the sudden contraction; (3) minor asymmetries not be accurately captured if the entire flow
in the diffuser geometry; and (4) difficulties in accurate volume or flow regime is not adequately
velocity measurement near the nozzle walls. In the sampled both in a simulation and/or experi-
latter case, experimental WSS measurements in the mental setting. As stated by Oberkampf
throat region were difficult to measure reliably due to et al.29: ‘‘When grids are sufficiently refined,
issues such as optical interference, inadequate pixel completely new processes or characteristics on
resolution and large velocity gradients near wall, and the smaller spatial scales can develop. It is not
so were omitted in this paper. A future report will uncommon that a steady flow field can
focus on developing better near-wall measurement become unsteady as the grid is refined and
algorithms to improve the accuracy of WSS measure- very small-scale phenomena can develop that
ments. did not exist on the coarser grids.’’
390 STEWART et al.

19. Shear stresses near corners, edges, and walls are CONFLICT OF INTEREST
difficult to model accurately and to validate
None.
experimentally, especially when turbulence or
transition is involved. Mesh refinement studies
near corners (or modeling corners with their
measured radii) is usually beneficial. Adequate REFERENCES
validation may require advanced experimental
techniques.4,5,12,19,22,24,33 1
Apsley, D. D., and M. A. Leschziner. Advanced turbulence
20. Modeling a spatial transition from laminar to modelling of separated flow in a diffuser. Flow Turbul.
turbulent flow poses many numerical chal- Combust. 63:81–112, 1999.
2
lenges. Some success is possible by advanced Bhushan, S., and D. K. Walters. A dynamic hybrid Rey-
Reynolds-averaged Navier–Stokes (RANS) nolds-averaged Navier Stokes–Large eddy simulation
modeling framework. Phys. Fluids 24:1–7, 2012.
turbulence models, such as the Shear Stress 3
Bhushan, S., D. K. Walters, and G. W. Burgreen. Laminar,
Transport models or dynamic hybrid RANS/ turbulent, and transitional simulations in benchmark cases
Large Eddy Simulation models.2 More sophis- with cardiovascular device features. Cardiovasc. Eng.
ticated non-RANS models (e.g., large eddy Technol. 2013. doi:10.1007/s13239-013-0155-5.
4
simulations) may have utility here,9 but have yet Charonko, J., S. Karri, J. Schmieg, S. Prabhu, and P.
Vlachos. In vitro, time-resolved PIV comparison of the
to be fully explored in the medical device field effect of stent design on wall shear stress. Ann. Biomed.
due to their higher computational cost. Eng. 37:1310–1321, 2009.
5
21. A basic understanding of the underlying Charonko, J., S. Karri, J. Schmieg, S. Prabhu, and P.
numerical algorithms of post-processing is Vlachos. In vitro comparison of the effect of stent config-
essential. Data extraction/interpolation needs uration on wall shear stress using time-resolved particle
image velocimetry. Ann. Biomed. Eng. 38:889–902, 2010.
to be carefully performed to prevent non- 6
Chen, Z., Y. Fan, X. Deng, and Z. Xu. A new way to
physical results, as demonstrated by non-zero reduce flow disturbance in endovascular stents: a numerical
velocities at (and beyond) the wall of the study. Artif. Organs 35:392–397, 2011.
7
sudden contraction observed in some simula- Cherdron, W., F. Durst, and J. H. Whitelaw. Asymmetric
tions. flows and instabilities in symmetric ducts with sudden
expansions. J. Fluid Mech. 84:13–31, 1978.
8
De Bartolo, C., A. Nigro, G. Fragomeni, F. M. Colacino,
D. Wang, C. C. Jones, and J. Zwischenberger. Numerical
and experimental flow analysis of the Wang-Zwische
double-lumen cannula. ASAIO J. 57:318–327, 2011.
ACKNOWLEDGMENTS 9
Delorme, Y. T., K. Anupindi, and S. H. Frankel. Large
eddy simulation of FDA’s idealized medical device.
We recognize and appreciate the support of the Cardiovasc. Eng. Technol. 2013. doi:10.1007/s13239-013-
worldwide CFD community. In particular, the authors 0161-7.
10
wish to thank our study participants: Yared Alemu, K. Down, L. A. Computational investigations of red blood
Amano, J. Ashton, Mehdi Behbahani, Catrin Blud- cell mechanical trauma and of diseased renal artery
szuweit-Philipp, Danny Bluestein, Parnian Boloori- hemodynamics. Ph.D. thesis. University of Oklahoma,
2011..
Zadeh, Alistair Brown, Scott Corbett, Julien de Cha- 11
Durst, F., A. Melling, and J. H. Whitelaw. Low Reynolds
rentenay, D. de Zelicourt, Wade Dummer, K. Fraser, number flow over a plane symmetric sudden expansion. J.
Leonid Goubergrits, Kurt Graichen, Linden Heflin, Fluid Mech. 64:111–128, 1974.
12
Darren Hitt, Marcus Hormes, Marc Horner, Xueying Etebari, A., and P. P. Vlachos. Improvements on the
Huang, S. Krishnan, S. Kühne, Patricia Lawford, accuracy of derivative estimation from DPIV velocity
measurements. Exp. Fluids 39:1040–1050, 2005.
Ming Li, Z. Li, William Louisos, Takehisa Mori, Alex 13
Fung, Y. C. Biomechanics: Circulation (2nd ed.). New
Medvedev, Joerg Mueller, Sei Murakami, Mariagrazia York: Springer, 1997.
14
Pilotelli, M. Polster, N. Reuel, T. Schauer, E. Sirois, Giridharan, G. A., C. Lederer, A. Berthe, L. Goubergrits,
Ken Solen, Ulrich Steinseifer, Wei Sun, S. Takahashi, J. Hutzenlaub, M. S. Slaughter, R. D. Dowling, P. A.
Masaaki Tamagawa, K. Tan, Dalin Tang, Ertan Ta- Spence, and S. C. Koenig. Flow dynamics of a novel
counterpulsation device characterized by CFD and PIV
skin, Z. Teng, R. Toyao, Dan White, Z. Jon Wu, M. modeling. Med. Eng. Phys. 33:1193–1202, 2011.
Xenos, and Ajit Yoganathan. This study was sup- 15
Goubergrits, L., J. Osman, and K. Affeld First experience
ported by the Food & Drug Administration’s Critical with an FDA critical path initiative: CFD and hemolysis.
Path Initiative Program. The mention of commercial In: Proceedings of the XXXVI Annual ESAO Congress.
products, their source, or their use in connection with Compiègne, France; 2009. p. 398.
16
Hariharan, P., M. Giarra, V. Reddy, S. Day, K. B. Man-
material reported herein is not to be construed as either ning, S. Deutsch, S. F. C. Stewart, M. R. Myers, M. R.
an actual or implied endorsement of such products by Berman, G. W. Burgreen, E. G. Paterson, and R. A.
the U.S. Department of Health and Human Services. Malinauskas. Multilaboratory particle image velocimetry
Results of FDA’s First Interlaboratory Computational Study 391
28
analysis of the FDA benchmark nozzle model to support Oberkampf, W. L., and M. F. Barone. Measures of
validation of computational fluid dynamics simulations. J. agreement between computation and experiment: valida-
Biomech. Eng. 133:041002, 2011. tion metrics. J. Comput. Phys. 217:5–36, 2006.
17 29
Heflin, L. A., D. V. Papavassiliou, and E. A. O’Rear. Oberkampf, W. L., T. G. Trucano, and C. Hirsch. Verifi-
Simulation of hemolysis experiments: relating hemolysis cation validation and predictive capability in computa-
and turbulent kinetic energy dissipation. Biomedical Engi- tional engineering and physics. Appl. Mech. Rev. 57:345–
neering Society Annual Meeting, Pittsburgh, PA, Oct. 7– 384, 2004.
30
10, 2009. Pekkan, K., D. de Zélicourt, L. Ge, F. Sotiropoulos, D.
18
Hinds, M. T., Y. J. Park, S. A. Jones, D. P. Giddens, and Frakes, M. A. Fogel, and A. P. Yoganathan. Physics-dri-
B. R. Alevriadou. Local hemodynamics affect monocytic ven CFD modeling of complex anatomical cardiovascular
cell adhesion to a three-dimensional flow model coated flows—a TCPC case study. Ann. Biomed. Eng. 33:284–300,
with E-selectin. J. Biomech. 34:95–103, 2001. 2005.
19 31
Hochareon, P., K. B. Manning, A. A. Fontaine, J. M. Stewart, S. F., E. G. Paterson, G. W. Burgreen, P. Harih-
Tarbell, and S. Deutsch. Wall shear-rate estimation within aran, M. Giarra, V. Reddy, S. W. Day, K. B. Manning, S.
the 50 cc Penn State artificial heart using particle image Deutsch, M. R. Berman, M. R. Myers, and R. A. Malin-
velocimetry. J. Biomech. Eng. 126:430–437, 2004. auskas. Assessment of CFD performance in simulations of
20
Hund, S. J., J. F. Antaki, and M. Massoudi. On the rep- an idealized medical device—results of FDA’s first com-
resentation of turbulent stresses for computing blood putational interlaboratory study. Cardiovasc. Eng. Technol.
damage. Int. J. Eng. Sci. 48:1325–1331, 2010. 3:139–160, 2012.
21 32
Iaccarino, G. Predictions of a turbulent separated flow Umezu, M., H. Fujimasu, T. Yamada, T. Fujimoto, M.
using commercial CFD codes. J. Fluid Eng. Trans. ASME Ranawake, A. Nogawa, and T. Kijima. Fluid dynamic
123:819–828, 2001. investigation of mechanical blood hemolysis. In: Heart
22
Kähler, C. J., U. Scholz, and J. Ortmanns. Wall-shear- Replacement Artificial Heart 5, edited by T. Akutso, and
stress and near-wall turbulence measurements up to single H. Koyanagi. Tokyo: Springer, 1996, pp. 327–335.
33
pixel resolution by means of long distance micro-PIV. Exp. Westerweel, J., P. F. Geelhoed, and R. Lindken. Single-
Fluids 41:327–341, 2006. pixel resolution ensemble correlation for micro-PIV appli-
23
Kameneva, M. V., G. W. Burgreen, K. Kono, B. Repko, J. cations. Exp. Fluids 37:375–384, 2004.
34
F. Antaki, and M. Umezu. Effects of turbulent stresses White, A. T., and C. K. Chong. Rotational invariance in
upon mechanical hemolysis: experimental and computa- the three-dimensional lattice Boltzmann method is depen-
tional analysis. ASAIO J. 50:418–423, 2004. dent on the choice of lattice. J. Comput. Phys. 230:6367–
24
Karri, S., J. Charonko, and P. P. Vlachos. Robust wall 6378, 2011.
35
gradient estimation using radial basis functions and proper Xenos, M., G. Girdhar, Y. Alemu, J. Jesty, M. Slepian, S.
orthogonal decomposition (POD) for particle image ve- Einav, and D. Bluestein. Device Thrombogenicity Emula-
locimetry (PIV) measured fields. Meas. Sci. Technol. tor (DTE)—design optimization methodology for cardio-
20:045401-1–045401-14, 2009. vascular devices: a study in two bileaflet MHV designs. J.
25
Kaufmann, T. A., T. Linde, E. Cuenca-Navalon, C. Sch- Biomech. 43:2400–2409, 2010.
36
mitz, M. Hormes, T. Schmitz-Rode, and U. Steinseifer. Yang, X. C., Y. Zhang, X. M. Gui, and S. S. Hu. Com-
Transient, three-dimensional flow field simulation through putational fluid dynamics-based hydraulic and hemolytic
a mechanical, trileaflet heart valve prosthesis. ASAIO J. analyses of a novel left ventricular assist blood pump. Artif.
57:278–282, 2011. Organs. 35(10):948–955, 2011.
26 37
Medvitz, R. B., D. A. Boger, V. Izraelev, G. Rosenberg, Yoshida, M., P. D. Wearden, O. Dur, K. Pekkan, and V.
and E. G. Paterson. Computational fluid dynamics O. Morell. Right ventricular outflow tract reconstruction
design and analysis of a passively suspended Tesla pump with bicuspid valved polytetrafluoroethylene conduit. Ann.
left ventricular assist device. Artif. Organs 35:522–533, Thorac. Surg. 91:1235–1238, 2011.
38
2011. Yu, H., D. Thévenin, and G. Janiga. Numerical prediction
27
Nabavi, M. Three-dimensional asymmetric flow through a of hemolysis based on computational fluid dynamics. In:
planar diffuser: effects of divergence angle, Reynolds Proceedings of the ECCOMAS Thematic International
number and aspect ratio. Int. Commun. Heat Mass Transf. Conference on Simulation and Modeling of Biological
37:17–20, 2010. Flows (SIMBIO 2011). Brussels; 2011. pp. 1–4.

Você também pode gostar