CFD Simulations AC7-03: Difference between revisions

From KBwiki
Jump to navigation Jump to search
m (Mike moved page Lib:CFD Simulations AC7-03 to CFD Simulations AC7-03 over redirect)
 
(176 intermediate revisions by one other user not shown)
Line 4: Line 4:
}}
}}
__TOC__
__TOC__
=Turbulent Blood Flow in a Ventricular Assist Device=
=Flow in a Ventricular Assist Device - Pump Performance & Blood Damage Prediction=
'''Application Challenge AC7-03'''   © copyright ERCOFTAC 2021
'''Application Challenge AC7-03'''   © copyright ERCOFTAC 2022
 
=CFD Simulations=
=CFD Simulations=
==Overview of CFD Simulations==
==Overview of CFD Simulations==
Various large-eddy simulations (LES) and unsteady Reynolds-averaged Navier-Stokes (URANS) computations were carried out using the commercial flow solver ANSYS CFX. All simulations were performed at the nominal operation (design) point and a partial operation point of the VAD. In total, five LES computations on different grid sizes were conducted for the verification of the LES results. The simulation at the finest grid was used as the reference case for the comparison with URANS. For the URANS cases, an extended grid convergence study was performed using seven URANS grids to analyze the influence of the spatial discretization on the main assessment parameters. Additionally, an URANS computation with a <math> k </math> - <math> \omega </math> - SST turbulence models were performed on the finest grid for the comparison with the LES results.
Large-eddy simulations (LES) and unsteady Reynolds-averaged Navier-Stokes (URANS) computations were carried out using the commercial flow solver ANSYS CFX. All simulations were performed at the nominal operation (design) point and a partial operation point of the VAD. In total, six LES computations on different grid sizes were conducted for the verification of LES. The simulation on the finest grid was used as the reference case for the comparison with URANS. For the URANS cases, an extended grid convergence study was performed using seven URANS grids to analyze the influence of the spatial discretization on the main assessment parameters. The URANS computation on the finest grid was performed for the comparison with the LES results.


==Computational Domain==
==Computational Domain==


The whole VAD was considered in the numerical analysis. A sketch of the computational domain can be seen in Fig. X. Inflow and outflow cannulas were included in the computational domain. The inlet and outlet of the cannulas were placed sufficiently far away (four and seven impeller diameters respectively) from the pump in order to minimze the influences of the boundary conditions on the results.
The whole VAD was considered in the numerical analysis. A sketch of the pump domain can be seen in Fig. 3.1. Inflow and outflow cannulas were included before and after this pump domain. The inlet and outlet of the cannulas were placed sufficiently far away (four and seven impeller diameters respectively; not shown here) from the pump in order to minimze the influences of the boundary conditions.


Block-structured grids with hexahedral-elements were created using ANSYS ICEM CFD. Since URANS and LES have different requirements for grid resolution and quality, two different, final meshes were created:
Block-structured grids with hexahedral-elements were created using ANSYS ICEM CFD. Since URANS and LES have different requirements for grid resolution and quality, two different, final meshes were created:
a.) The final mesh for the LES computation has a size of <math> 102 M </math> elements. The grid was built according to literature recommendations by Fröhlich and Menter for wall-resolving LES methods. Attention was paid that the near-wall grid fitted the upper limits of <math> \Delta_x^+ \leq 40 </math> for the grid with in flow direction and <math> \Delta_z^+ \leq 15 </math> for the grid width in spanwise direction. Furthermore, the first wall-normal node had a maximal dimensionless wall distance of <math> y^+_1 \leq 1 </math> and the grid growth factor was <math> r_g=1.05 </math>. This meshing strategy has been implemented throughout the whole domain. Grid angles were larger than 23<math>^{\circ} </math> and the volume change smaller than 5 with apect ratios smaller than 40 at the pumps wall. The aspect ratios were further reduced with increasing wall distance, so that the values range from 1 to 6 in the core flow region.
a.) The final mesh for the LES computation has a size of <math> 102 M </math> elements. The grid was built according to literature recommendations by Fröhlich [12] and Menter [13] for wall-resolving LES methods. Attention was paid that the near-wall grid observed the upper limits of <math> \Delta_x^+ \leq 40 </math> for the grid size in flow direction and <math> \Delta_z^+ \leq 15 </math> for the grid size in spanwise direction. Furthermore, the first wall-normal node had a maximal dimensionless wall distance of <math> y^+_1 \leq 1 </math> and the grid expansion factor was <math> r_g=1.05 </math> directly at the wall. This meshing strategy has been implemented throughout the whole domain. Grid angles were larger than 23<math>^{\circ} </math> and the volume change smaller than 5 with apect ratios smaller than 40 at the pump wall. The aspect ratios were further reduced with increasing wall distance, so that the values range from <math> 1 </math> to <math> 5 </math> in the core flow region. The gap was discretized with 28 and 74 elements for the height and width.  


[[Image: Figure 2.png|600px|centre|thumb|Fig.3.1 Computational grid of the LES. (a) Surface grid at the blades and the hub. (b) Volume grid in an axial cut-plane at the leading edge of the impeller blade.]]
[[Image: 2Mio_B1.png|800px|centre|thumb|Fig.3.1 Computational grid of the LES: surface grid at the blades and the hub of the rotor and guide vanes. The grid was reduced by a factor of four in each direction for better vizualisation.]]


The final URANS grid is coarser with a total size of <math> 20 M </math> elements. Mesh quality criteria were kept within the ranges of the ANSYS CFX guidelines with maximum aspect ratios around 100, volume changes smaller than 6, and a minimum angle of <math> 20^{\circ} </math>. The mesh near the rotor wall has an area-averaged and maximal wall distance of <math> y^+_1=0.7 </math> and <math> 2.1 </math>. All other pump parts have <math> y^+_1 </math>-values of one or smaller. Care was taken that the nearest wall layer contains more than 10 cells with a growth rate of <math> r_g = 1.2 </math>. An advanced mesh convergence study of Eça and Hoekstra was performed to check whether the mesh size is appropriate to reflect the fluid mechanical and hemodynamical parameter. Therefore, the final URANS mesh was coarsened to 6 coarser grids with a mesh coarsening factor between 1.06 and 1.15.
The final URANS grid is coarser having a total number of <math> 20 M </math> elements. Mesh quality criteria were kept within the ranges of the ANSYS CFX guidelines with maximum aspect ratios around 100, volume changes smaller than 6, and a minimum angle of <math> 20^{\circ} </math>. The mesh near the rotor wall has an area-averaged and maximal wall distance of <math> y^+_1=0.7 </math> and <math> 2.1 </math>. All other pump parts have <math> y^+_1 </math>-values of one or smaller. The mesh in the gap constists of 19 and 41 elements for the height and width. Additionally, the grid expansion rate was <math> r_g = 1.2 </math>. An advanced mesh convergence study of Eça and Hoekstra [14] was performed to check whether the mesh size is appropriate to reflect the fluid mechanical and hemodynamical parameter. For this, the final URANS mesh was coarsened to 6 coarser grids with a mesh coarsening factor between 1.06 and 1.15.


==Solution Strategy==
==Solution Strategy==
In order to guarantee a properly working numerical scheme,  several physical constraints were defined for the simulations. Blood was modeled as a Newtonian fluid with a dynamic viscosity of <math> \mu=3.5~mPas </math> and a density of <math> \rho=1050~kg/m^3 </math>. This is a valid simplification because blood viscosity is independent of stress when shear rates are larger than <math> 100~1/s </math> and this condition is fulfilled in the pump.
In order to guarantee a properly working numerical scheme,  several physical constraints were defined for the simulations. Blood was modeled as a Newtonian fluid with a dynamic viscosity of <math> \mu=3.5~mPas </math> and a density of <math> \rho=1050~kg/m^3 </math>. This is a valid simplification since blood viscosity is independent of stress when shear rates are larger than <math> 100~1/s </math> and this condition is fulfilled in the pump [19].
   
   
The impeller rotated at a constant nominal frequency of 7,900 rpm. Transient rotor-stator coupling was used with a time step equal to 0.36° and rotation of the impeller for LES and URANS, respectively. The resulting RMS Courant numbers were 0.6 for the Large-eddy and 3.4 for the Reynolds-averaged simulation. A total of 33 revolutions were calculated. Convergence was reached, when RMS residuals were at <math> 10^{-5} </math> and all monitored values were in a statistically steady condition. Time averaging was done for at least 13 revolutions for statistical analysis of the results.
The impeller rotated at a constant nominal frequency of <math> 7,900~rpm </math>. Transient rotor-stator coupling was used with a time step equal to <math> 0.36^{\circ} </math> and <math> 3^{\circ} </math> rotation of the impeller for LES and URANS, respectively. The transient rotor-stator coupling takes into account all of the transient flow characteristics by using a sliding interface, which allows for a smooth rotation between the rotary (rotor) and stationary (guide vane) part [20]. During the calculation, the cell zones slide (that is, rotate) relative to one another along the mesh interface in discrete steps [21].
 
The resulting RMS Courant numbers were 0.6 for the Large-eddy and 3.4 for the Reynolds-averaged simulation. A total of 33 revolutions were calculated. Convergence was reached, when RMS residuals were at <math> 10^{-5} </math> and all monitored values were in a statistically steady condition. Time averaging was done for at least 13 revolutions for statistical analysis of the results.


==Boundary Conditions==
==Boundary Conditions==
A constant, nominal flow rate of ܳ<math> Q= 4.5 ݈l/min </math> was given at the outlet in both simulations. All walls were assumed to be  hydraulically smooth and the no-slip condition applied. Total pressure was set to zero at the inlet. Since the Reynolds number is small at the inlet cannula (ܴ݁<math> Re \approx 2,300 </math>) and no perturbations are expected upstream, no turbulent inflow condition were specified.
A constant flow rate of <math> Q=4.5~l/min </math> or <math> Q=2.5~l/min </math> were given at the outlet in the simulations. All walls were assumed to be  hydraulically smooth and the no-slip condition applied. Total pressure was set to zero and a uniform velocity profile was given at the inlet. Since the Reynolds number is small at the inlet cannula (ܴ݁<math> Re \leq 2,300 </math>) and no perturbations are expected upstream, no turbulent inflow condition were specified for LES.


==Application of Physical Models==
==Application of Physical Models==
For LES, the governing equations were spatially and temporarly discretized by a bounded CDS scheme and a 2nd order Euler backward scheme. The dynamic Smagorinsky sub-grid scale model was used  for closure of the LES equations with a bounded Smagorinsky parameter between <math> C_s = 0.0 - 0.3 </math>.  
For LES, the governing equations were spatially and temporarly discretized by a bounded CDS scheme and a 2nd order Euler backward scheme. The dynamic Smagorinsky sub-grid scale model was used  for closure of the LES equations with a bounded Smagorinsky parameter between <math> C_s = 0.0 - 0.3 </math>.  


For URANS, we used a high-resolution scheme and a 2nd order Euler backward scheme for spatial and temporal discretization, respectively. The a <math> k </math>-<math> \omega </math>-SST model was applied as turbulence model. We expect that streamline curvature and transitional flow effects will influence  the physics within the pump. Therefore, the curvature correction model and a Γ-Θ-transition model were applied for the <math> k </math>- <math> \omega </math> and the SST model.
For URANS a 2nd order upwind scheme and a 2nd order Euler backward scheme for spatial and temporal discretization respectively, were used. The <math> k </math>-<math> \omega </math>-SST model was applied as turbulence model. It is to be expected that streamline curvature and transitional flow effects will influence  the physics within the pump. Therefore, the curvature correction model [15] and a <math>\Gamma</math>-<math>\Theta</math>-transition model [16] were applied for the <math> k </math>- <math> \omega </math>-SST model.


==Numerical Accuracy==
==Numerical Accuracy==


Various numerical accuracy studies were performed with different verification methods for the URANS and LES computations. These studies has been already published in different papers of the author (see XXX). Two verification methods, one for LES and one for URANS, will be presented below.  
Various numerical accuracy studies were performed with different verification methods for the URANS and LES computations (e.g., in Refs. [2] or [17]). Some of them will be presented below.  


'''Extended grid convergency study for URANS'''
'''Extended Grid Convergency Study for URANS'''


The discretization error is a consequence of the approximation made in finite volume methods due to the transformation of the governing equations into a system of algebraic equations, which are discretely solved on a computational
The discretization error is a consequence of the approximation made in finite volume methods due to the transformation of the governing equations into a system of algebraic equations, which are discretely solved on a computational
grid. An examination of the result’s discretization error is highly relevant and can be done for RANS methods by grid convergence studies. In this AC, the discretization error is estimated for URANS by means of an advanced grid convergence study, which determines an uncertainty interval as an indicator for the discretization error. A summary of the procedure is given below. For a detailed derivation, please see the study of Eça and Hoekstra. Generally, the discretization error can be estimated by:
grid. An examination of the resulting discretization error is highly relevant and can be done for RANS methods by grid convergence studies. In this AC, the discretization error is estimated for URANS by means of an advanced grid convergence study, which determines an uncertainty interval <math> U_\phi </math> as an indicator for the discretization error. A brief explanation of the procedure is given below. For a detailed derivation, please see Refs. [17] or [18].
 
<center><math> \epsilon_\phi \widetilde{=} \phi_i - \phi_0 = \alpha h_i^p \qquad\qquad\qquad\qquad(11) </math></center>
 
<math> \phi_i </math> is an arbritary flow quantity, <math> \phi_0 </math> is the estimate of the exact result, <math> \alpha </math> is a constant, <math> h_i=(\inc_x \inc_y \inc_z)^{1/3} </math> is the cell size, and <math> p </math> is the observed order of convergence. Values of <math> \alpha </math>,<math> p </math> and <math> \phi_0 </math> are needed for the error estimation. These values can be determined by using at least four simulations on different grid sizes and a least-squares error estimation using the minimum of the following equation:


<center><math> S(\phi_0, \alpha, p) = \sqrt{\sum_{i=1}^N (\phi_i-(\phi_0+\alpha h_i^p))^2} \qquad\qquad\qquad\qquad(12) </math></center>
Generally, the discretization error can be estimated by:


The resulting parameter <math> (\phi_0+\alpha h_i^p) </math> characterizes a polynomial function, which is called the fit. This fit approximates the development of the computed CFD results. The deviation between the CFD result <math> \phi_i </math> and the fit is specified by the fit’s standard deviation <math> \sigma </math>.
<center><math> \epsilon_\phi \widetilde{=} \phi_i - \phi_0 = \alpha h_i^p, \qquad\qquad\qquad\qquad(11) </math></center>


The discretization error of the CFD result is indicated by an uncertainty <math> U_\phi </math>, which is a 95% confidence interval, wherein the exact solution of the URANS equations is expected. The determination of <math> U_\phi </math> depends on several fit and convergence properties, which are explained in detail in the study of Eça and Hoekstra. An example of a “good” error estimation is given by Eq. (13), which uses the fit’s data and an additional safety factor <math> F_s </math>.
where <math> \phi_i </math> is an arbritary flow quantity, <math> \phi_0 </math> is the estimate of the exact result, <math> \alpha </math> is a constant, <math> h_i=(\Delta_x \Delta_y \Delta_z)^{1/3} </math> is the cell size, and <math> p </math> is the observed order of convergence.


<center><math> U_\phi(\phi_i) = F_s \cdot \mid \alpha h_i^p \mid + \sigma + \mid \phi_i - \phi_{fit} \mid \qquad\qquad\qquad\qquad(13) </math></center>
The result of the presented method is an uncertainty interval <math> U_\phi </math>, which is an indicator for this discretization error <math> \epsilon_\phi </math>. The uncertainty is a 95% confidence interval, wherein the exact solution of the URANS equations (solution on an infinitely fine grid) is expected. The determination of <math> U_\phi </math> depends on several simulation and convergence properties, which are explained in detail in the study of Eça and Hoekstra [18].


The resulting uncertainty interval <math> U_\phi </math> is an indicator for the deviation between the computed result <math> \phi_i </math> on the (typical) cell size <math> h_i </math> and the estimated, exact URANS solution <math> \phi_0 </math>. Furthermore, it is possible to examine whether a computational grid is sufficiently fine or still too coarse to indicate a grid-independent and reliable solution for the analyzed parameter of interest.  
It is possible to examine whether a computational grid is sufficiently fine or still too coarse to indicate a grid-independent and reliable solution for the analyzed parameter of interest. In total, seven meshes were used for the advanced grid convergence study. These meshes are summarised in Tab. 3.1. The term <math> D_2 / (1/N \cdot V_{pump})^{1/3} </math> indicates the dimensionless ratio of the pump's diameter <math> D_2 </math> to the representative grid size <math> h_r=(1/N \cdot V_{pump})^{1/3} </math> of a specific simulation.


In total, seven meshes were used for the advanced grid convergence study. These meshes are summarised in Tab. 3.1. Since the mesh influence on the computation of, e.g., stresses is one major aspect of this AC, the results of the grid convergency study will be presented in the evaluation [[Evaluation AC7-03|Evalution]].
{|border="1" cell padding="25" cell spacing="3" align="center"
{|border="1" cell padding="25" cell spacing="3" align="center"
|+align="bottom"|Table 3.1 Names and grid size properties of the seven URANS grids.
|+align="bottom"|Table 3.1 Names and grid size properties of the seven URANS grids.
!''Grid'' !! colspan="1"| ''Million grid elements <math> N </math>'' !! colspan="1"| ''Typical cell size (m)'' !! colspan="1"| ''Grid refinement parameter'' !! colspan="1"| ''Time and space averaged <math> y^+_1 </math>-value over blades and maximum''  
!''Grid'' !! colspan="1"| ''Million grid elements <math> N </math>'' !! colspan="1"| ''<math> D_2 / (1/N \cdot V_{pump})^{1/3} </math>'' !! colspan="1"| ''Grid refinement parameter'' !! colspan="1"| ''Time and space-averaged <math> y^+_1 </math>-value over blades and maximum''  
|-
|-
|UR-1 || 3.3 || <math>3 \cdot 2.11^{-4}</math> || 1.06 ||  1.43(3.58)
|UR-1 || 3.3 || <math> 76 </math> || 1.06 ||  1.43 (3.58)
|-
|-
|UR-2 || 4.0 || <math>3 \cdot 1.98^{-4}</math> || 1.15 ||  1.25(3.29)
|UR-2 || 4.0 || <math> 81 </math> || 1.15 ||  1.25 (3.29)
|-
|-
|UR-3 || 5.9 || <math>3 \cdot 1.73^{-4}</math> || 1.06 ||  1.12(3.00)
|UR-3 || 5.9 || <math> 93 </math> || 1.06 ||  1.12 (3.00)
|-
|-
|UR-4 || 7.0 || <math>3 \cdot 1.63^{-4}</math> || 1.11 ||  1.00(2.74)
|UR-4 || 7.0 || <math> 99 </math> || 1.11 ||  1.00 (2.74)
|-
|-
|UR-5 || 9.6 || <math>3 \cdot 1.47^{-4}</math> || 1.12 ||  0.89(2.49)
|UR-5 || 9.6 || <math> 110 </math> || 1.12 ||  0.89 (2.49)
|-
|-
|UR-6 || 13.6 || <math>3 \cdot 1.31^{-4}</math> || 1.12 ||  0.79(2.31)
|UR-6 || 13.6 || <math> 123 </math> || 1.12 ||  0.79 (2.31)
|-
|-
|UR-7 || 19.4 || <math>3 \cdot 1.16^{-4}</math> || - ||  0.68(2.14)
|UR-7 || 19.4 || <math> 138 </math> || - ||  0.68 (2.14)
|-
|-
|}
|}


Figure 3.2. shows the pressure heads via the VAD impeller and the whole pump and the uncertainties experienced. The uncertainty intervals of total pressure heads were up to 4.8% for the finest grid. From an engineering point of view, these uncertainties are in an acceptable range for the VAD design. In addition, the uncertainty for the pressure head via the whole VAD (4.8%) is higher than for the impeller alone (1.7%). The reason for this is that turbulent phenomena, e.g. detachment in the outlet guide vane, affect the pressure increase via the whole pump. Those effects may have a significant grid sensitivity and thus affect the uncertainty of the pressure head for the entire VAD. Furthermore, the relatively small uncertainties for the pressure heads suggest that the finest grid resolution is enough to guarantee a nearly grid-independent solution and no further grid refinement seems to be required based on these results.


'''Power-Loss-Analysis (PLA) for LES'''
[[Image:uncertainty_pressure.png|400px|center|thumb|Fig. 3.2. Pressure head via the impeller (top). Pressure head via the VAD (bottom). The error bars mark the numerical uncertainties (deviations in percent).]]
 
The power loss analysis is an LES quality assessment method, which was developed by the authors of this study. It can be used to investigate whether the performed simulation is adequately capable of computing the internal power losses due to TKE production and dissipation within the flow. These processes, espacially the dissipation, are connected with the computed stresses (see Eq. (6) in the [[Description AC7-03|Description]]). The method is graphically shown in Figure 3.2. In the PLA, three definitions of the total power losses <math> P_{tot} </math> for the mean (time-averaged) flow field are calculated and afterwards, the results from the three total power losses are compared. The rationale and derivations of the PLA are comprehensively explained in Ref. X. In the
following, just the most important informations regarding the PLA will be explained.
 
[[Image: PLA_Methode.jpg|600px|centre|thumb|Fig.3.2 Sketch of the PLA method.]]
 
 
The first total power loss <math> P_{tot,1} </math> is determined by using the pump characteristics of the VAD. The power loss can be calculated by taking the difference between the power at the coupling <math> P </math> (blade's torque <math> M_{bl} </math> multiplied with angular velocity <math> 2 \pi n </math>) and the increase in hydraulic power <math> P_{hyd} </math> via the VAD, see Equation (14):
 
<center><math> P_{tot,1} = P - P_{hyd} = M_{bl} \cdot 2 \pi n - \Delta p_{tot} \cdot Q. \qquad\qquad\qquad\qquad(14) </math></center>
 
The total power loss can also be determined by integrating and summing up all internal power losses within the VAD. Three loss parts must be considered as internal losses for LES. The first internal loss <math> P_{dir} </math> (Equation (15)) is due to direct dissipation <math> \varepsilon_{dir} </math> and describes the transfer of mean flow energy directly into heat.
 
<center><math> P_{dir}=\rho \int_V \epsilon_{dir} dV=\mu \int_V \langle \frac{\partial u_i}{\partial x_j} \rangle \bigg( \langle  \frac{\partial u_i}{\partial x_j} \rangle + \langle \frac{\partial u_j}{\partial x_i} \rangle \bigg)dV \qquad\qquad\qquad(15) </math></center>


The second loss part is the energy transfer from mean flow into turbulent motion and is called turbulence production <math> TKE_{prod} </math>. The corresponding loss term <math> P_{prod} </math> is defined in Eq. (16). It is taken from the directly resolved flow field of a flow computation.
The uncertainties for the stress-dependent MIH index (Fig. 3.3, bottom right) indicate a higher but acceptable uncertainty for the finest grid with <math> \pm </math> 8%. In contrast, the coarsest grid has a two times higher uncertainty for MIH (30%) than the uncertainty for the pressure head (15%). On the other hand, the uncertainties concerning the volumes, in which certain stress thresholds are exceeded, indicate larger uncertainty intervals. These uncertainties are up to 4 times larger than those for the pump characteristics, as can be seen in the figures. Of course, the uncertainties of the different blood damage indicators will decrease with higher grid resolutions so that the absolute values will converge to a final state. But even for the finest grid, a decay of the slope of the fit in Fig. 3.3. is not obvious in the range of data obtained from the flow computations.  


<center><math> P_{prod}= \rho \int_V TKE_{Prod} dV = \rho \int_V - \langle u_i' u_j' \rangle \bigg\langle \frac{\partial u_i}{\partial x_j} \bigg\rangle dV  \qquad\qquad\qquad(16) </math></center>
The URANS computations were performed with grids whose mesh sizes are comparable to current URANS blood pump simulations (e.g., in [3], [5], [33]). In our case, there are still discernible uncertainties for the shear-dependent quantities for these mesh sizes, although a small uncertainty is indicated for the pressure heads.
Since the LES does not resolve all of the turbulent motions and losses, the modeled loss contribution from the smallest scales must be included into the PLA. In general, the LES turbulence model must ensure that a proper amount of turbulent dissipation from the smallest scales is provided. Due to the equilibrium between turbulence production and dissipation at the smallest scales, the modeled turbulent losses from the applied Smagorinsky model can be written as in Eq. (17):


<center><math> P_{mod}=\rho \int_V \epsilon_{mod} dV=\rho \int_V \langle \nu_t \frac{\partial u_i}{\partial x_j} \bigg( \frac{\partial u_i}{\partial x_j} + \frac{\partial u_j}{\partial x_i} \bigg) \rangle dV. \qquad\qquad\qquad(17) </math></center>
[[Image:uncertainty_blooddamage.png|600px|center|thumb|Fig. 3.3. Stress-dependent variables. Upper left: Volume in the pump where stresses of 9 Pa are exceeded. Upper right:  Volume in the pump where stresses of 50 Pa are exceeded. Bottom left: Volume in the pump where stresses of 150 Pa are exceeded. Bottom right: Stress-dependent MIH index. ]]


Hence, the second total power loss <math> P_{tot,2} </math> is stated in Eq. (18) by summing up the loss parts from Eq. (15), (16) and (17).
'''Evaluation of the Flow Resolution of LES Using Time-Averaged Quantities'''


<center><math> P_{tot,2}=P_{dir} + P_{prod} + P_{mod} \qquad\qquad\qquad\qquad(18) </math></center>
The first step was to check whether the near-wall grid is fine enough to allow no-slip conditions at the walls. This was done by determining the grid step sizes in wall units, <math> \Delta_x^+ </math> , <math> \Delta_z^+ </math> , <math>y^+ </math>, using time-averaged wall shear stresses. These quantities are given above and satisfy the recommendations given in Ref. [12] and [13].
The losses from the resolved turbulent flow part in Eq. (16) can also be expressed by another flow variable, which acts on different turbulent scales. It is the resolved turbulent dissipation rate <math> \epsilon_{turb} </math>. This variable expresses the energy loss from the small turbulent motions into heat and is defined in terms of a power loss <math> P_{diss} </math> in Eq. (19):


<center><math> P_{diss}=\rho \int_V \epsilon_{turb} dV=\mu \int_V \langle \frac{\partial u'_i}{\partial x_j} \bigg( \frac{\partial u'_i}{\partial x_j} + \frac{\partial u'_j}{\partial x_i} \bigg) \rangle dV. \qquad\qquad\qquad\qquad(19) </math></center>
Then, the degree of flow resolution was estimated in the core flow region. One way to assess the fineness of the grid is to estimate the Kolmogorov length <math> \eta_{k} </math> by the relation:
The summation of Eq. (15), (17) and (19) leads to a third formulation of a total power loss <math> P_{tot,3} </math> for a turbulent flow field, see Eq. (20):


<center><math> P_{tot,3}=P_{dir} + P_{diss} + P_{mod}. \qquad\qquad\qquad\qquad(20) </math></center>  
<center><math> \eta_{k} = \Bigl( \frac{\nu^3}{\epsilon_{turb}} \Bigl) ^\frac{1}{4}, \qquad\qquad\qquad\qquad(12) </math></center>


The PLA method makes it possible to compare the direct losses in the mean flow (due to <math> \epsilon_{dir} </math>) with the turbulent losses due to the turbulent energy cascade (<math> TKE_{prod} </math> and <math> \epsilon_{turb} </math>). From a numerical point of view, the PLA can be used as a quality assessment method, since the total power losses from Eq. (14), (18) and (20) should be identical per definition. If this is fulfilled, it can be verified that the power losses in form of direct and turbulent dissipation rates as well as turbulence production were adequately computed by a simulation.
which follows from the theory of isotropic turbulence [22]. The local ratio <math> \Delta/ \eta_{k} </math> gives information about the degree of resolution of dissipative scales through the grid [12]. <math> \Delta </math> was calculated in every grid cell by <math> \Delta=(\Delta_x \Delta_y \Delta_z)^{\frac{1}{3}} </math>. The ratio was <math> \leq 5 </math> throughout the whole flow domain, which is smaller than the ratios of <math> \Delta/ \eta_{k}  < (12-25) </math> defined as desirable for LES from Refs. [12] and [23].  


{|border="1" cell padding="25" cell spacing="3" align="center"
Furthermore, frequency spectra were also used as additional tools to evaluate the physical resolution of the LES on the finest grid. These are shown, for example, in Refs. [25] and [27] for various locations in the computed flow of the rotor and outlet guide vane.
|+align="bottom"|Table 3.2 Results of the PLA analysis for the finest LES grid.


!''PLA variable & operator'' !! colspan="1"| ''Value''
[[Image:meshAnalysis.png|600px|center|thumb|Fig. 3.4. Analysis of flow variables on different grid sizes. Left: Pressure head <math> H </math>. Right:  Volume-integrated, resolved turbulent kinetic energy (TKE) <math> k </math> in the impeller domain. ]]
|-
| <math> P_{bl} ~- </math> || <math> 2.231~W </math>
|-
| <math> P_{hyd}~= </math> || <math> 0.739~W </math>
|-
|<math> {P_{tot,1}} </math> || <math> {1.492~W} </math>
|-
| <math> \bm{(P_{tot,2}-P_{tot,1})/P_{tot,1}\cdot 100} </math> || <math> \bm{-1.5 \%} </math>
|-
|<math> P_{dir}~+ </math> || <math> 1.159~W~(78.8 \%) </math>
|-
| <math> P_{prod}~+ </math> || <math> 0.286~W~(19.5 \%) </math>
|-
| <math> P_{mod}~= </math> || <math> 0.025~W~(1.7 \%) </math>
|-
|<math> {P_{tot,2}} </math> || <math> {1.470~W}~(100 \%) </math>
|-
| <math> \bm{(P_{tot,2}-P_{tot,1})/P_{tot,1}\cdot 100} </math> || <math> \bm{-1.5 \%} </math>
|-
| <math> {(P_{tot,2}-P_{tot,1})/P_{tot,1}\cdot 100} </math> || <math> {-1.5 \%} </math>
|-
| <math> \bm{(P_{tot,2}-P_{tot,1})/P_{tot,1}\cdot 100} </math> || <math> \bm{-1.5 \%} </math>
|-
|-
| <math> P_{dir}~+ </math> || <math> 1.159~W~(82.2 \%) </math>
|-
|<math> P_{diss}~+</math> || <math> 0.225~W~(16.0 \%) </math>
|-
|<math> P_{mod}~= </math> || <math> 0.025~W~(1.8 \%) </math>
|-
| <math> {P_{tot,3}} </math> || <math> {1.409~W}~(100 \%) </math>
|-
| <math> \bm{(P_{tot,2}-P_{tot,1})/P_{tot,1}\cdot 100} </math> || <math> \bm{-1.5 \%} </math>
|-
| <math> {(P_{tot,3}-P_{tot,1})/P_{tot,1}\cdot 100} </math> ||  <math> {-5.6 \%} </math>
|-
|}
Table 3.2 summarizes the results of the PLA for the turbulent production rates <math> TKE_{prod} </math> as well as for the turbulent dissipation rates <math> \epsilon_{turb} </math> for the finest LES grid. From a numerical point of view, the relative deviation between the total power losses is smaller using the turbulence production <math> P_{tot,2} </math> (<math>-1.5\%</math>) as with the turbulent dissipation <math>P_{tot,3}</math> (<math>-5.6\%</math>). Furthermore, both loss definitions shows that the majority of the turbulent losses is directly resolved within the flow <math>(16.0\%-19.5\%</math>) instead of being modeled <math>1.7 \%-1.8 \%</math>).


The high dependency on the spatial discretization can be analyzed using Fig. 3.3. The subfigures show some PLA results, which were ascertained on different grid sizes. It can be seen from subfigure~a.) that the relative deviations between the total power losses decrease with increasing grid size to <math>-1.5\%</math> and <math>-5.6\%</math>, respectively. Furthermore, it is noticeable from subfigure~b.) that the total power loss <math> P_{tot,1} </math> (based on the pump characteristics, Eq. (14)) is relatively constant from a grid size of 10M nodes with a value of <math> P_{tot,1}\approx 1.50~W </math>. In addition, the internal loss part <math> P_{dir} </math> from direct dissipation <math> \epsilon_{dir} </math> shows a nearly constant progression with a value of <math> P_{dir}\approx 1.16~W </math> between 30M and 102M node points. Therefore, the reduction between the power losses <math> P_{tot,2} </math>, respectively <math> P_{tot,3} </math>, and <math> P_{tot,1} </math> in subfigure a.) must be a result of a more accurate computation of the turbulent losses based on the turbulent dissipation and turbulence production rates.
In addition, the variations of characteristic flow variables on different grid sizes were evaluated. This is shown in figure 3.4., where the pressure head <math> H </math> and the volume-integral of the resolved, turbulent kinetic energy (TKE) <math> k </math> was plotted against the number of grid points. The resolved TKE is defined as:


Despite the fact that the finest grid cannot resolve the turbulent dissipation rates in total, the relative deviation to the reference value (<math> P_{tot,1} </math>) is relatively small with <math> \approx 5 \% </math>. Contrariwise, that means that $95\%$ of the dissipative losses can be investigated within the flow field of the simulation. In addition, the simulation on the 102M grid is also able to reproduce nearly all losses from the turbulence production <math> P_{prod} </math>. Together with the fact that the majority of the losses are directly resolved instead of being modeled, it can be concluded that the 102M grid simulation is appropriate for the flow investigation in the VAD.
<center><math> k = \frac{1}{2}\langle {u_i' u_i'} \rangle. \qquad\qquad\qquad\qquad(13) </math></center>


[[Image: PLAres.png|350px|centre|thumb|Fig.3.3 PLA results. (A) Relative deviations between <math> P_{tot,2} </math> resp. <math> P_{tot,3} </math> and <math> P_{tot,1} </math>. (B) Total power loss <math> P_{tot,1} </math> on different grid sizes. Also (B), loss part due to direct dissipation on different grid sizes. (C) Loss part due to turbulence production rates and dissipation rates on different gird sizes. ]]
As can be seen from the figure, the variation of the results on the fine grid sizes (<math> N>30M </math> grid elements) is small, indicating a convergent solution with nearly constant variables on the finest grids.  


<br/>
<br/>
Line 177: Line 117:
}}
}}


© copyright ERCOFTAC 2021
© copyright ERCOFTAC 2022

Latest revision as of 10:52, 11 January 2023

Front Page

Description

Test Data

CFD Simulations

Evaluation

Best Practice Advice

Flow in a Ventricular Assist Device - Pump Performance & Blood Damage Prediction

Application Challenge AC7-03   © copyright ERCOFTAC 2022

CFD Simulations

Overview of CFD Simulations

Large-eddy simulations (LES) and unsteady Reynolds-averaged Navier-Stokes (URANS) computations were carried out using the commercial flow solver ANSYS CFX. All simulations were performed at the nominal operation (design) point and a partial operation point of the VAD. In total, six LES computations on different grid sizes were conducted for the verification of LES. The simulation on the finest grid was used as the reference case for the comparison with URANS. For the URANS cases, an extended grid convergence study was performed using seven URANS grids to analyze the influence of the spatial discretization on the main assessment parameters. The URANS computation on the finest grid was performed for the comparison with the LES results.

Computational Domain

The whole VAD was considered in the numerical analysis. A sketch of the pump domain can be seen in Fig. 3.1. Inflow and outflow cannulas were included before and after this pump domain. The inlet and outlet of the cannulas were placed sufficiently far away (four and seven impeller diameters respectively; not shown here) from the pump in order to minimze the influences of the boundary conditions.

Block-structured grids with hexahedral-elements were created using ANSYS ICEM CFD. Since URANS and LES have different requirements for grid resolution and quality, two different, final meshes were created: a.) The final mesh for the LES computation has a size of elements. The grid was built according to literature recommendations by Fröhlich [12] and Menter [13] for wall-resolving LES methods. Attention was paid that the near-wall grid observed the upper limits of for the grid size in flow direction and for the grid size in spanwise direction. Furthermore, the first wall-normal node had a maximal dimensionless wall distance of and the grid expansion factor was directly at the wall. This meshing strategy has been implemented throughout the whole domain. Grid angles were larger than 23 and the volume change smaller than 5 with apect ratios smaller than 40 at the pump wall. The aspect ratios were further reduced with increasing wall distance, so that the values range from to in the core flow region. The gap was discretized with 28 and 74 elements for the height and width.

Fig.3.1 Computational grid of the LES: surface grid at the blades and the hub of the rotor and guide vanes. The grid was reduced by a factor of four in each direction for better vizualisation.

The final URANS grid is coarser having a total number of elements. Mesh quality criteria were kept within the ranges of the ANSYS CFX guidelines with maximum aspect ratios around 100, volume changes smaller than 6, and a minimum angle of . The mesh near the rotor wall has an area-averaged and maximal wall distance of and . All other pump parts have -values of one or smaller. The mesh in the gap constists of 19 and 41 elements for the height and width. Additionally, the grid expansion rate was . An advanced mesh convergence study of Eça and Hoekstra [14] was performed to check whether the mesh size is appropriate to reflect the fluid mechanical and hemodynamical parameter. For this, the final URANS mesh was coarsened to 6 coarser grids with a mesh coarsening factor between 1.06 and 1.15.

Solution Strategy

In order to guarantee a properly working numerical scheme, several physical constraints were defined for the simulations. Blood was modeled as a Newtonian fluid with a dynamic viscosity of and a density of . This is a valid simplification since blood viscosity is independent of stress when shear rates are larger than and this condition is fulfilled in the pump [19].

The impeller rotated at a constant nominal frequency of . Transient rotor-stator coupling was used with a time step equal to and rotation of the impeller for LES and URANS, respectively. The transient rotor-stator coupling takes into account all of the transient flow characteristics by using a sliding interface, which allows for a smooth rotation between the rotary (rotor) and stationary (guide vane) part [20]. During the calculation, the cell zones slide (that is, rotate) relative to one another along the mesh interface in discrete steps [21].

The resulting RMS Courant numbers were 0.6 for the Large-eddy and 3.4 for the Reynolds-averaged simulation. A total of 33 revolutions were calculated. Convergence was reached, when RMS residuals were at and all monitored values were in a statistically steady condition. Time averaging was done for at least 13 revolutions for statistical analysis of the results.

Boundary Conditions

A constant flow rate of or were given at the outlet in the simulations. All walls were assumed to be hydraulically smooth and the no-slip condition applied. Total pressure was set to zero and a uniform velocity profile was given at the inlet. Since the Reynolds number is small at the inlet cannula (ܴ݁) and no perturbations are expected upstream, no turbulent inflow condition were specified for LES.

Application of Physical Models

For LES, the governing equations were spatially and temporarly discretized by a bounded CDS scheme and a 2nd order Euler backward scheme. The dynamic Smagorinsky sub-grid scale model was used for closure of the LES equations with a bounded Smagorinsky parameter between .

For URANS a 2nd order upwind scheme and a 2nd order Euler backward scheme for spatial and temporal discretization respectively, were used. The --SST model was applied as turbulence model. It is to be expected that streamline curvature and transitional flow effects will influence the physics within the pump. Therefore, the curvature correction model [15] and a --transition model [16] were applied for the - -SST model.

Numerical Accuracy

Various numerical accuracy studies were performed with different verification methods for the URANS and LES computations (e.g., in Refs. [2] or [17]). Some of them will be presented below.

Extended Grid Convergency Study for URANS

The discretization error is a consequence of the approximation made in finite volume methods due to the transformation of the governing equations into a system of algebraic equations, which are discretely solved on a computational grid. An examination of the resulting discretization error is highly relevant and can be done for RANS methods by grid convergence studies. In this AC, the discretization error is estimated for URANS by means of an advanced grid convergence study, which determines an uncertainty interval as an indicator for the discretization error. A brief explanation of the procedure is given below. For a detailed derivation, please see Refs. [17] or [18].

Generally, the discretization error can be estimated by:

where is an arbritary flow quantity, is the estimate of the exact result, is a constant, is the cell size, and is the observed order of convergence.

The result of the presented method is an uncertainty interval , which is an indicator for this discretization error . The uncertainty is a 95% confidence interval, wherein the exact solution of the URANS equations (solution on an infinitely fine grid) is expected. The determination of depends on several simulation and convergence properties, which are explained in detail in the study of Eça and Hoekstra [18].

It is possible to examine whether a computational grid is sufficiently fine or still too coarse to indicate a grid-independent and reliable solution for the analyzed parameter of interest. In total, seven meshes were used for the advanced grid convergence study. These meshes are summarised in Tab. 3.1. The term indicates the dimensionless ratio of the pump's diameter to the representative grid size of a specific simulation.

Table 3.1 Names and grid size properties of the seven URANS grids.
Grid Million grid elements Grid refinement parameter Time and space-averaged -value over blades and maximum
UR-1 3.3 1.06 1.43 (3.58)
UR-2 4.0 1.15 1.25 (3.29)
UR-3 5.9 1.06 1.12 (3.00)
UR-4 7.0 1.11 1.00 (2.74)
UR-5 9.6 1.12 0.89 (2.49)
UR-6 13.6 1.12 0.79 (2.31)
UR-7 19.4 - 0.68 (2.14)

Figure 3.2. shows the pressure heads via the VAD impeller and the whole pump and the uncertainties experienced. The uncertainty intervals of total pressure heads were up to 4.8% for the finest grid. From an engineering point of view, these uncertainties are in an acceptable range for the VAD design. In addition, the uncertainty for the pressure head via the whole VAD (4.8%) is higher than for the impeller alone (1.7%). The reason for this is that turbulent phenomena, e.g. detachment in the outlet guide vane, affect the pressure increase via the whole pump. Those effects may have a significant grid sensitivity and thus affect the uncertainty of the pressure head for the entire VAD. Furthermore, the relatively small uncertainties for the pressure heads suggest that the finest grid resolution is enough to guarantee a nearly grid-independent solution and no further grid refinement seems to be required based on these results.

Fig. 3.2. Pressure head via the impeller (top). Pressure head via the VAD (bottom). The error bars mark the numerical uncertainties (deviations in percent).

The uncertainties for the stress-dependent MIH index (Fig. 3.3, bottom right) indicate a higher but acceptable uncertainty for the finest grid with 8%. In contrast, the coarsest grid has a two times higher uncertainty for MIH (30%) than the uncertainty for the pressure head (15%). On the other hand, the uncertainties concerning the volumes, in which certain stress thresholds are exceeded, indicate larger uncertainty intervals. These uncertainties are up to 4 times larger than those for the pump characteristics, as can be seen in the figures. Of course, the uncertainties of the different blood damage indicators will decrease with higher grid resolutions so that the absolute values will converge to a final state. But even for the finest grid, a decay of the slope of the fit in Fig. 3.3. is not obvious in the range of data obtained from the flow computations.

The URANS computations were performed with grids whose mesh sizes are comparable to current URANS blood pump simulations (e.g., in [3], [5], [33]). In our case, there are still discernible uncertainties for the shear-dependent quantities for these mesh sizes, although a small uncertainty is indicated for the pressure heads.

Fig. 3.3. Stress-dependent variables. Upper left: Volume in the pump where stresses of 9 Pa are exceeded. Upper right: Volume in the pump where stresses of 50 Pa are exceeded. Bottom left: Volume in the pump where stresses of 150 Pa are exceeded. Bottom right: Stress-dependent MIH index.

Evaluation of the Flow Resolution of LES Using Time-Averaged Quantities

The first step was to check whether the near-wall grid is fine enough to allow no-slip conditions at the walls. This was done by determining the grid step sizes in wall units, , , , using time-averaged wall shear stresses. These quantities are given above and satisfy the recommendations given in Ref. [12] and [13].

Then, the degree of flow resolution was estimated in the core flow region. One way to assess the fineness of the grid is to estimate the Kolmogorov length by the relation:

which follows from the theory of isotropic turbulence [22]. The local ratio gives information about the degree of resolution of dissipative scales through the grid [12]. was calculated in every grid cell by . The ratio was throughout the whole flow domain, which is smaller than the ratios of defined as desirable for LES from Refs. [12] and [23].

Furthermore, frequency spectra were also used as additional tools to evaluate the physical resolution of the LES on the finest grid. These are shown, for example, in Refs. [25] and [27] for various locations in the computed flow of the rotor and outlet guide vane.

Fig. 3.4. Analysis of flow variables on different grid sizes. Left: Pressure head . Right: Volume-integrated, resolved turbulent kinetic energy (TKE) in the impeller domain.

In addition, the variations of characteristic flow variables on different grid sizes were evaluated. This is shown in figure 3.4., where the pressure head and the volume-integral of the resolved, turbulent kinetic energy (TKE) was plotted against the number of grid points. The resolved TKE is defined as:

As can be seen from the figure, the variation of the results on the fine grid sizes ( grid elements) is small, indicating a convergent solution with nearly constant variables on the finest grids.




Contributed by: B. Torner — University of Rostock, Germany

Front Page

Description

Test Data

CFD Simulations

Evaluation

Best Practice Advice

© copyright ERCOFTAC 2022