• No results found

Large eddy simulation of LDL surface concentration in a subject specific human aorta

N/A
N/A
Protected

Academic year: 2021

Share "Large eddy simulation of LDL surface concentration in a subject specific human aorta"

Copied!
33
0
0

Loading.... (view fulltext now)

Full text

(1)

Large eddy simulation of LDL surface

concentration in a subject specific human aorta

Jonas Lantz and Matts Karlsson

Linköping University Post Print

N.B.: When citing this work, cite the original article.

Original Publication:

Jonas Lantz and Matts Karlsson, Large eddy simulation of LDL surface concentration in a subject specific human aorta, 2012, Journal of Biomechanics, (45), 3, 537-542.

http://dx.doi.org/10.1016/j.jbiomech.2011.11.039 Copyright: Elsevier

http://www.elsevier.com/

Postprint available at: Linköping University Electronic Press http://urn.kb.se/resolve?urn=urn:nbn:se:liu:diva-72895

(2)

Large eddy simulation of LDL surface concentration in

a subject specific human aorta

Jonas Lantz and Matts Karlsson

Department of Management and Engineering

Link¨oping University

SE-581 83 Link¨oping

Sweden jonas.lantz@liu.se

Abstract

The development of atherosclerosis is correlated to the accumulation of lipids in the arterial wall, which, in turn, may be caused by the build-up of low-density lipoproteins (LDL) on the arterial surface. The goal of this study was to model blood flow within a subject specific human aorta, and to study how the LDL surface concentration changed during a cardiac cycle. With measured velocity profiles as boundary conditions, a scale-resolving technique (large eddy simulation, LES) was used to compute the pulsatile blood flow that was in the transitional regime. The relationship between wall shear stress (WSS) and LDL surface concentration was investigated, and it was found that the accumulation of LDL correlated well with WSS. In general, regions of low WSS corresponded to regions of increased LDL concentration and vice versa. The instantaneous LDL values changed significantly during a cardiac cycle; during systole the surface concentration was low due to increased convective fluid transport, while in diastole there was an increased accumulation of LDL on the surface. Therefore, the near-wall velocity was investigated at four representative locations, and it was concluded that in regions with disturbed flow the LDL concentration had significant temporal changes, indicating that LDL accumulation is sensitive to not only the WSS but also near-wall flow.

Keywords: low-density lipoprotein, wall shear stress, disturbed flow, atherosclerosis

(3)

1. Introduction

Atherosclerotic lesions have been found to develop at certain sites in the human arterial system, preferentially at the inner wall of curved segments and outer walls of bifurcations (Karino et al., 1987; Berceli et al., 1990). An increased level of low-density lipoprotein (LDL) has been shown to promote the accumulation of cholesterol within the intima layer of large arteries (Ross and Harker, 1976; Hoff and Wagner, 1986). There is a small flux of water from the blood to the arterial wall, driven by the arterial pressure difference, which can transport LDL to the arterial wall. However, the endothelium presents a barrier to LDL, creating a flow-dependent concentration boundary layer. This concentration polarization is interesting, as regions with elevated LDL are co-located with low shear regions (Ethier, 2002), suggesting a relation-ship between LDL accumulation and flow dynamics. Studies in humans and animals indicate that the flux of LDL from the plasma into the arterial wall depends both on the LDL concentration and the LDL permeability at the plasma-arterial wall interface (Nielsen, 1996).

The surface concentration of LDL has previously been investigated on a vascular scale in straight arteries (Wada and Karino, 1999), arteries with multiple bends (Wada and Karino, 2002b), the carotid arteries (Fazli et al., 2011; Soulis et al., 2010), coronary arteries (Olgac et al., 2009; Kolandavel et al., 2006) and in the aorta (Liu et al., 2009). Generic models of the vessels are common, and studies are normally performed with the assumption of laminar (and sometimes steady) flow. Simulating non-pulsating laminar flows can give useful results, especially in smaller arteries where the flow rate is low and the effect from pulsatility can be neglected. However, Sun et al. (2006) conducted a numerical study where they studied the transmural LDL transport in a idealized vessel with a stenosis, under both steady and pulsatile conditions. They found large differences between the steady and pulsating model, suggesting that the steady flow assumption was inadequate, at least in the post-stenotic region where the flow was unsteady and disturbed.

The flow in the aorta is pulsatile with Reynolds numbers in the transi-tional regime, which creates a very complex flow field (Morbiducci et al., 2009; Stalder et al., 2011). While systolic acceleration normally has a stabilizing effect on the flow, flow disturbances might appear during the deceleration phase. Stein and Sabbah (1976) measured the blood velocity in both healthy and patients with aortic valvular disease and found highly disturbed (and in one case also turbulent) flow during systole and the early part of diastole

(4)

in all normals, and fully turbulent flow in patients with aortic insufficiency. Peacock et al. (1998) conducted an experimental study where they measured the critical Reynolds number for the onset of turbulence under physiological conditions in a straight pipe. They found that the critical Reynolds num-ber for the flow normally found in the aorta was on the order of 5500, and since the peak Reynolds number usually is higher, their results predicted the existence of disturbed aortic flows. Stalder et al. (2011) used Magnetic Resonance Imaging (MRI) to study the aortic flow in a large cohort of 30 young healthy subjects and their conclusion was that flow instabilities were present in healthy subjects at rest, but not necessarily fully turbulent flow. They expected that turbulence effects might be more pronounced for higher cardiac outputs or in the presence of diseases.

Due to the transitional and disturbed nature of the flow it is very diffi-cult to model, and might be best treated with a turbulence model. As noted in a review article by Yoganathan et al. (2005), common RANS (Reynolds-Averaged Navier Stokes) turbulence models are not the ideal choice for dis-turbed cardiovascular flows, but instead, a scale-resolving turbulence model such as Large-Eddy Simulation (LES) would be more suitable, due to the finer resolution and its ability to handle transition. There have been a num-ber of studies using LES on idealized blood vessels with a constriction (Mittal et al., 2003; Paul et al., 2009; Tan et al., 2011), and very good agreement compared to experimental results was found, demonstrating the potential for modeling physiological, low-Re transitional flows using LES. However, to the authors’ knowledge, LES has never been used in the study of LDL transport in a subject specific human aorta under pulsatile flow conditions.

The goal of this study is therefore to compute the surface concentration of LDL on a human aorta in a disturbed flow using a LES model. As the flow is pulsatile, the changes of LDL during a cardiac cycle and the sensitivity to local flow dynamics are investigated.

2. Method

The method is only briefly described here; details on MRI acquisition, mesh generation, and how the blood flow and LDL transport were modeled using LES are found in the supplementary materials section. Also, a com-parison between a laminar model and the LES was made, and is discussed in the supplementary materials, together with a motivation for the use of LES.

(5)

2.1. MRI Acquisition and Geometry Reconstruction

Geometry and flow data were acquired using a 1.5 T Philips Achieva MRI scanner and segmented using a 3D level-set algorithm (Heiberg et al., 2010). High quality hexahedral meshes were constructed in ANSYS ICEM CFD 13.0 (ANSYS Inc. Canonsburg, PA, USA), with special attention paid to near-wall resolution.

2.2. Modeling Blood Flow

From the velocity profiles in the MRI measurements, the Reynolds num-ber based on inlet diameter ranged from 150 at late diastole to 6500 at peak systole, with a mean of 1200. A subject specific velocity profile boundary condition was set in the ascending aorta, while mass flow rates where spec-ified in the branches leaving the arch, see figure 1. The flow was computed using LES with the WALE sub-grid model which accounts for both near-wall velocity and transition (Nicoud and Ducros, 1999). Due to the transient nature of the LES model, 55 cardiac cycles were computed and phase aver-ages of WSS and LDL were computed using the last 50 cycles. This ensured results that were independent of sudden transient effects. The simulations were run at National Supercomputer Centre (NSC), Link¨oping, Sweden.

2.3. Modeling LDL Transport

The transport of LDL was modeled as a passive non-reacting scalar, with inlet concentration set to 1.2 mg/ml following Stangeby and Ethier (2002), and a zero concentration gradient was assumed at all outlets. A boundary condition that allows for mass transfer through the wall was set on the arterial surface: CwVw−D ∂C ∂n w = KwCw (1)

where Cw is the concentration of LDL at the wall, Vw the water

infiltra-tion velocity, D the kinematic diffusivity, ∂C/∂n the concentrainfiltra-tion gradient normal to the wall, and Kw a permeability coefficient for the wall.

3. Results

3.1. WSS and LDL Distributions

The instantaneous WSS are elevated in the branches compared to the rest of the aorta during the acceleration phase, see the upper row of figure 2. On the inner curvature of the arch and in the vicinity of the branches, areas

(6)

of both low and high WSS are present, indicating regions with large WSS gradients. The descending and thoracic aorta shows only minor variations in WSS, due to the structured flow and relatively simple geometry. During the deceleration phase and beginning of diastole when smaller vortices are present in the flow, the WSS level drops significantly with only small areas close to the branches and inner curvature of the arch with elevated shear stress.

The normalized distributions of LDL during the same times in the cardiac cycle are displayed in the lower row of figure 2. The LDL seems to be inversely proportional to the WSS value; areas with low WSS value tend to have an elevated LDL concentration and vice-versa. This is especially true in the branches during systole, where an elevated WSS level corresponds to a low region of LDL. During diastole this effect is less apparent because of the low flow rate. Instead, accumulation of LDL increases on the surface due to the decrease of wall parallel convective transport of LDL in the fluid during diastole. To investigate this further, the time-averaged WSS and LDL values for each computational node on the aorta were plotted against each other, see figure 3. Generally, regions of low LDL correspond to regions of elevated WSS and vice-versa. The results are in very good agreement with findings from Wada and Karino (2002b) and Soulis et al. (2010), who described the surface concentration of LDL as inversely proportional to the WSS magnitude. However, as there is a significant scatter among the values, the LDL concentration might not only be dependent on the WSS, but also on other parameters such as local geometry features and the near-wall flow.

3.2. Near-wall Velocity and LDL

To further investigate how local flow features affects the LDL concen-tration, the near-wall velocity at four representative locations on the aorta were extracted. A line starting at the wall and going 5 mm in normal di-rection into the vessel was used. In figures 4 to 7 the corresponding WSS and LDL distributions are plotted in the upper row and the local near-wall velocity magnitude is plotted in the lower row, where abbreviations MA, PS, MD, and BD stands for max acceleration, peak systole, max deceleration and the beginning of diastole, respectively. In the thoracic aorta the near-wall velocity profile is well structured with a well pronounced acceleration and deceleration phase during systole, figure 4. There is a region on the order of 0.5-0.75 mm from the wall where the velocity is very low, but significant velocity gradients exists less than 1 mm from the wall. A small increase

(7)

in velocity is seen at the beginning of diastole, which is the remaining of a disturbed flow structure that formed in the descending aorta at max accel-eration. Again, there seems to be a correlation between the WSS and LDL, with decreased LDL levels during systole.

At the outer side of the descending arch the flow profile is only slightly disturbed during the deceleration phase, indicating an overall well-structured flow, see figure 5. However, here the near-wall velocity is significantly higher throughout the entire systolic phase, with only a thin region of about 0.25 mm with decreased velocity. This was expected as the flow rate is higher in the outer curvature due to the bending of the aorta. The LDL level decreases during systole when the WSS is high, but recovers during diastole.

At the inner curvature of the descending arch there are velocity fluctua-tions present, indicated by the disturbed flow pattern in figure 6. Also, the near wall velocity is significantly lower compared to the flow in the thoracic aorta. There is a large region with almost stationary flow, starting at the wall and going 2 mm into the vessel, which is caused by the formation of a separation zone during systole. This separation zone moves back and forth, which adds to the flow disturbances. The WSS and LDL levels become very affected by the fluctuations, as seen by the oscillatory curves in figure 6.

Moving 10 mm downstream, the flow situation is even more complex, with more flow disturbances in the vicinity of the wall, see figure 7. A short segment of the accelerating phase is seen, but it breaks up around peak systole when a large recirculation zone forms. Regions with stationary flow close to the wall are mixed with fluctuating velocities, creating a very complex flow pattern. This is again reflected in the oscillating LDL concentration, predominantly during systole.

3.3. Dynamic Behavior of LDL Concentration

Clearly, the LDL concentration is affected by the pulsatile flow and there seems to be a correlation with the WSS. Therefore, the instantaneous WSS and LDL values on each node on the entire aorta were analyzed. As an example, consider a point on the outer curvature of the arch after the three branches, where the LDL surface concentration changes significantly. During the systolic acceleration phase the WSS increases due to the increase in mass flow rate, and the surface concentration drops due to increased convection in the flow, see the left graph in figure 8. After peak systole the flow decelerates and there is an increase in LDL which follows the same path back to the initial

(8)

value. During diastole, the remaining disturbed flow structures causes only minor fluctuations in WSS which yields small changes in LDL concentration. This shape is representative for the entire aorta, and if all instantaneous WSS and LDL values are considered, the graph to the right in figure 8 is obtained. It consists of data from 50 cardiac cycles (approximately 2.5 Billion surface data points) and represents the aggregated effect of adding curves like the one in the left figure onto each other. A pattern emerges where it is clearly seen than there exists a maximum value of LDL for each WSS value, and that low values of LDL generally correspond to high values of WSS and vice-versa. The maximum value can for this case be described by the equation:

LDL ≤ α

WSSβ + γ (2)

where the coefficients can be determined by a nonlinear least-square fit. In this simulation α, β and γ were found to be 0.20, 0.33 and 0.98, respectively. 4. Discussion

Previous studies (Nielsen, 1996; Stangeby and Ethier, 2002) have found that high arterial wall permeability to LDL and an elevated LDL concentra-tion at the wall may directly cause accelerated development of atherosclerosis and the filtration flow will increase. An elevated wall permeability to LDL is likely explained by a dysfunctional endothelial layer subjected to both chemical and mechanical features (Tarbell, 2003). As noted by Vincent et al. (2010), cellular scale mechanisms such as the inclusion of an endothelial gly-cocalyx layer (EGL) may be as important as the effects of blood flow when modeling the concentration polarization effect. They concluded that the dif-fusivity and depth of LDL in the EGL were the two most critical parameters and called for experimental studies to investigate the importance of these parameters further. It is believed that increased thickness of the EGL de-creases the LDL accumulation in the intima (Liu et al., 2011a). Also, the water flux to the artery is spatially heterogeneous on the subcellular level which might affect the LDL surface concentration, as investigated by Vin-cent et al. (2009) and Wada and Karino (2002a). Besides the spatial scales, the development and genesis of atherosclerosis is a long-term process with large temporal scales. On the other hand, transport of LDL from the blood to the arterial wall represents a process which is present on scales less than a second (Sun et al., 2006). Clearly, the role of LDL transport and its con-nection to atherosclerosis is a complex phenomena, present not only at two

(9)

very different time scales, but also at two very different length scales. Studies have shown that atherosclerotic plaques localizes in sites where the WSS is low, but the actual interaction between WSS, LDL surface concentration, and LDL infiltration rate remains unidentified (Stangeby and Ethier, 2002). In the present study, disturbed aortic flow was modeled using a scale-resolving turbulence model, which can account for disturbed and transitional effects. Studies where the flow has been treated as laminar, have shown that there are only small temporal changes of the surface concentration during a cardiac cycle (Fazli et al., 2011; Liu et al., 2011b; Kolandavel et al., 2006). However, when disturbed flow is present, e.g. in a post-stenotic region, the surface concentration can be significantly affected by the pulsatile flow (Sun et al., 2006; Liu et al., 2011b). We found that disturbed flow was present in the aorta when the peak Reynolds numbers is on the order of 5-6000, and that the disturbances had a significant effect on the LDL surface concentration. The disturbances formed predominantly in the descending aorta during the latter part of systole, but were still present in the beginning of diastole.

In general, an elevated WSS level resulted in a lowered LDL surface con-centration as apparent in figure 2, which is in line with results from other studies (Liu et al., 2009; Fazli et al., 2011). These low values of LDL were present in regions where the flow was found to be disturbed, e.g. around branches and the inner curvature of descending aorta. This is believed to be caused by the increased convective fluid transport at those regions, creating a thinner concentration boundary layer which is more sensitive to the effects of the pulsatile flow. In regions where the flow is well structured the concen-tration boundary layer is thicker and the surface concenconcen-tration less sensitive to the dynamics of the flow. This theory is supported by the fact that the LDL level increases during diastole, when the flow rate and convective effects are lowered. Besides local flow disturbances, the pulsating flow also affected the LDL concentration; large temporal variations were seen during systole with decreasing LDL levels during systolic acceleration and increasing levels during systolic deceleration.

The reason why atherosclerosis development occurs at preferential sites is still not fully understood, but these regions coincide with disturbed flow and fluctuating WSS, and as shown in this article, also fluctuating LDL concen-tration. These mechanical variations may adversely affect the endothelium, in such way that it promotes leaky cell junctions (which allows for LDL to diffuse into the wall), or even damages the endothelial layer. The type of flow greatly affects the endothelial permeability; Phelps and DePaola (2000)

(10)

reported that areas of enhanced endothelial permeability coincide with the presence of disturbed flow, where the shear stress magnitude is low but the spatial and temporal gradients in shear stress are large.

Hence, from a flow perspective, a possible explanation to why atheroscle-rosis develops at certain sites may depend on two flow features: WSS and near-wall flow. The WSS affects the endothelial permeability for LDL trans-port into the arterial wall, while the near-wall velocity affects the LDL con-centration boundary layer. Increased LDL levels might also increase the endothelial permeability, as discussed by Guretzki et al. (1994), resulting in a non-linear feedback loop that increases the LDL transport into the wall. Therefore, our results indicate that the amount of LDL is not only depen-dent by the value of WSS, but also by the near-wall flow patterns, such as disturbances and high near-wall velocities.

5. Acknowledgments

This work was supported by grants from the Swedish research council, VR 2007-4085 and VR 2010-4282. The Swedish National Infrastructure for Computing (SNIC) is acknowledged for computational resources provided by the National Supercomputer Centre (NSC) under grant No. SNIC022/09-11. Dr. Tino Ebbers at the Department of Medicine and Care at Link¨opings University is acknowledged for the MRI measurements. This work has been conducted in collaboration with the Center for Medical Image Science and Vi-sualization (CMIV, http://www.cmiv.liu.se/) at Link¨oping University, Swe-den. CMIV is acknowledged for provision of financial support and access to leading edge research infrastructure.

6. Conflict of Interest Statement

Neither of the authors have any commercial or non-commercial relation-ship that might lead to a conflict of interest.

7. References

Berceli, S.A., Warty, V.S., Sheppeck, R.A., Mandarino, W.A., Tanksale, S.K., Borovetz, H.S., 1990. Hemodynamics and low density lipoprotein metabolism. Rates of low density lipoprotein incorporation and degrada-tion along medial and lateral walls of the rabbit aorto-iliac bifurcadegrada-tion. Arteriosclerosis 10, 686–694.

(11)

Ethier, C.R., 2002. Computational modeling of mass transfer and links to atherosclerosis. Ann Biomed Eng 30, 461–471.

Fazli, S., Shirani, E., Sadeghi, M.R., 2011. Numerical simulation of LDL mass transfer in a common carotid artery under pulsatile flows. J Biomech 44, 68–76.

Guretzki, H.J., Gerbitz, K.D., Olgemoller, B., Schleicher, E., 1994. Athero-genic levels of low density lipoprotein alter the permeability and composi-tion of the endothelial barrier. Atherosclerosis 107, 15–24.

Heiberg, E., Sjogren, J., Ugander, M., Carlsson, M., Engblom, H., Arheden, H., 2010. Design and validation of Segment–freely available software for cardiovascular image analysis. BMC Med Imaging 10, 1.

Hoff, H.F., Wagner, W.D., 1986. Plasma low density lipoprotein accumula-tion in aortas of hypercholesterolemic swine correlates with modificaaccumula-tions in aortic glycosaminoglycan composition. Atherosclerosis 61, 231–236. Karino, T., Goldsmith, H.L., Motomiya, M., Mabuchi, S., Sohara, Y., 1987.

Flow patterns in vessels of simple and complex geometries. Ann. N. Y. Acad. Sci. 516, 422–441.

Kolandavel, M.K., Fruend, E.T., Ringgaard, S., Walker, P.G., 2006. The effects of time varying curvature on species transport in coronary arteries. Ann Biomed Eng 34, 1820–1832.

Liu, X., Fan, Y., Deng, X., 2011a. Effect of the endothelial glycocalyx layer on arterial LDL transport under normal and high pressure. J. Theor. Biol. 283, 71–81.

Liu, X., Fan, Y., Deng, X., Zhan, F., 2011b. Effect of non-Newtonian and pulsatile blood flow on mass transport in the human aorta. J Biomech 44, 1123–1131.

Liu, X., Pu, F., Fan, Y., Deng, X., Li, D., Li, S., 2009. A numerical study on the flow of blood and the transport of LDL in the human aorta: the physiological significance of the helical flow in the aortic arch. Am. J. Physiol. Heart Circ. Physiol. 297, 163–170.

(12)

Mittal, R., Simmons, S.P., Najjar, F., 2003. Numerical study of pulsatile flow in a constricted channel. Journal of Fluid Mechanics 485, 337–378. Morbiducci, U., Ponzini, R., Rizzo, G., Cadioli, M., Esposito, A., De

Co-belli, F., Del Maschio, A., Montevecchi, F.M., Redaelli, A., 2009. In vivo quantification of helical blood flow in human aorta by time-resolved three-dimensional cine phase contrast magnetic resonance imaging. Ann Biomed Eng 37, 516–531.

Nicoud, F., Ducros, F., 1999. Subgrid-scale stress modelling based on the square of the velocity gradient tensor. Flow, Turbulence and Combustion 62, 183–200.

Nielsen, L.B., 1996. Transfer of low density lipoprotein into the arterial wall and risk of atherosclerosis. Atherosclerosis 123, 1–15.

Olgac, U., Poulikakos, D., Saur, S.C., Alkadhi, H., Kurtcuoglu, V., 2009. Patient-specific three-dimensional simulation of LDL accumulation in a human left coronary artery in its healthy and atherosclerotic states. Am. J. Physiol. Heart Circ. Physiol. 296, 1969–1982.

Paul, M.C., Mamun Molla, M., Roditi, G., 2009. Large-Eddy simulation of pulsatile blood flow. Med Eng Phys 31, 153–159.

Peacock, J., Jones, T., Tock, C., Lutz, R., 1998. The onset of turbulence in physiological pulsatile flow in a straight tube. Experiments in Fluids 24, 1–9.

Phelps, J.E., DePaola, N., 2000. Spatial variations in endothelial barrier function in disturbed flows in vitro. Am. J. Physiol. Heart Circ. Physiol. 278, H469–476.

Ross, R., Harker, L., 1976. Hyperlipidemia and atherosclerosis. Science 193, 1094–1100.

Soulis, J.V., Fytanidis, D.K., Papaioannou, V.C., Giannoglou, G.D., 2010. Wall shear stress on LDL accumulation in human RCAs. Med Eng Phys 32, 867–877.

Stalder, A.F., Frydrychowicz, A., Russe, M.F., Korvink, J.G., Hennig, J., Li, K., Markl, M., 2011. Assessment of flow instabilities in the healthy aorta using flow-sensitive MRI. J Magn Reson Imaging 33, 839–846.

(13)

Stangeby, D.K., Ethier, C.R., 2002. Computational analysis of coupled blood-wall arterial LDL transport. J Biomech Eng 124, 1–8.

Stein, P.D., Sabbah, H.N., 1976. Turbulent blood flow in the ascending aorta of humans with normal and diseased aortic valves. Circ. Res. 39, 58–65. Sun, N., Wood, N.B., Hughes, A.D., Thom, S.A., Xu, X.Y., 2006. Fluid-wall

modelling of mass transfer in an axisymmetric stenosis: effects of shear-dependent transport properties. Ann Biomed Eng 34, 1119–1128.

Tan, F.P., Wood, N.B., Tabor, G., Xu, X.Y., 2011. Comparison of LES of steady transitional flow in an idealized stenosed axisymmetric artery model with a RANS transitional model. J Biomech Eng 133, 051001. Tarbell, J.M., 2003. Mass transport in arteries and the localization of

atherosclerosis. Annu Rev Biomed Eng 5, 79–118.

Vincent, P.E., Sherwin, S.J., Weinberg, P.D., 2009. The effect of a spatially heterogeneous transmural water flux on concentration polarization of low density lipoprotein in arteries. Biophys. J. 96, 3102–3115.

Vincent, P.E., Sherwin, S.J., Weinberg, P.D., 2010. The effect of the endothe-lial glycocalyx layer on concentration polarisation of low density lipopro-tein in arteries. J. Theor. Biol. 265, 1–17.

Wada, S., Karino, T., 1999. Theoretical study on flow-dependent concen-tration polarization of low density lipoproteins at the luminal surface of a straight artery. Biorheology 36, 207–223.

Wada, S., Karino, T., 2002a. Prediction of LDL concentration at the luminal surface of a vascular endothelium. Biorheology 39, 331–336.

Wada, S., Karino, T., 2002b. Theoretical prediction of low-density lipopro-teins concentration at the luminal surface of an artery with a multiple bend. Ann Biomed Eng 30, 778–791.

Yoganathan, A.P., Chandran, K.B., Sotiropoulos, F., 2005. Flow in pros-thetic heart valves: state-of-the-art and future directions. Ann Biomed Eng 33, 1689–1694.

(14)

0 0.2 0.4 0.6 0.8 1 −0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 Time [s] Mass Flow [kg/s] 0 0.2 0.4 0.6 0.8 1 80 85 90 95 100 105 110 115 120 125 130 Pressure [mmHg]

Ascending Mass Flow (MRI) Descending Mass Flow (MRI) Brachiocephalic Mass Flow Left Common Carotid Mass Flow Left Subclavian Mass Flow Windkessel Pressure

Figure 1: Measured mass flow curves in the ascending and descending aorta together with the resulting mass flow rates in the branches leaving the arch. The pressure on the outlet (computed with a Windkessel model) in the thoracic aorta is also plotted for reference.

(15)

Figure 2: Upper row: WSS distribution on the arterial surface. Lower row: normalized LDL distribution on the arterial surface. From left to right: max acceleration, max de-celeration, beginning of diastole, and mid diastole, respectively. Regions of elevated WSS seems to correspond to decreased values of LDL.

(16)

0 2.5 5 7.5 10 12.5 15 1 1.05 1.1 1.15 1.2 1.25 Time−averaged WSS [Pa] Time−averaged LDL [C w /C 0 ]

Figure 3: Scatter plot of time-averaged LDL vs time-averaged WSS. Data is taken from the entire aortic arch and consists of approximately 60 000 data points which were phase-averaged over 50 cardiac cycles. Here it is apparent that increased values of WSS cor-responds to decreased values of LDL surface concentration, while regions with low WSS takes on the whole spectrum of LDL values. This indicates that LDL surface concentration does not only depend on WSS values, but also on other flow features such as near-wall velocity and disturbances.

(17)

Figure 4: A combination of WSS and LDL distributions during a cardiac cycle, and also the near wall velocity magnitude along a line normal the wall and 5 mm into the vessel, located on the inner side of the thoracic aorta (red dot). Abbreviations MA, PS, MD, and BD stands for max acceleration, peak systole, max deceleration, and the beginning of diastole, respectively. There is a correlation between WSS and LDL during systole; the WSS increases during systolic acceleration while the LDL concentration decreases, and during the deceleration phase the WSS decreases due to lowered flow rate while the LDL level increases. The flow profile is well structured with a pronounced increase in velocity during the acceleration phase which gradually decreases during the deceleration phase. A small increase in near-wall velocity is seen at the beginning of diastole, which comes from small flow disturbance that originated in the aortic arch at max acceleration.

(18)

Figure 5: A combination of WSS and LDL distributions during a cardiac cycle, and also the near wall velocity magnitude along a line normal the wall and 5 mm into the vessel, located on the outer side of the descending aorta (red dot). Same abbreviations as in figure 4. The flow profile is relatively well structured with some fluctuating velocities during the deceleration phase. Noticeably is that the region of low velocity close to the wall during systole is very thin; the distance from the wall to maximum velocity is on the order of 0.25 mm. A clear correlation between WSS and LDL is seen, especially during systole.

(19)

Figure 6: A combination of WSS and LDL distributions during a cardiac cycle, and also the near wall velocity magnitude along a line normal the wall and 5 mm into the vessel, located on the inner side of the descending aorta (red dot). Same abbreviations as in figure 4. The flow profile is less well structured with fluctuating velocities, due to a dynamic separation zone that moves back an forth during systole, which in turn affects the WSS and LDL levels.

(20)

Figure 7: A combination of WSS and LDL distributions during a cardiac cycle, and also the near wall velocity magnitude along a line normal the wall and 5 mm into the vessel, located on the inner side of the descending aorta (red dot). Same abbreviations as in figure 4. The location is 10 mm downstream of the location in figure 6, and a large difference in flow profile behavior is seen. After the initial acceleration the flow breaks up and large regions of either stagnant or disturbed flow is seen.

0 0.5 1 1.5 2 2.5 3 3.5 1 1.02 1.04 1.06 1.08 1.1 1.12 1.14 1.16 WSS [Pa] LDL [C w /C 0 ] Systolic acceleration Systolic deceleration Diastole Peak systole Start systole

Figure 8: Left: example of how LDL and WSS changes in one location on the aortic arch during one cardiac cycle. The effect from the pulsatile flow is apparent. Right: instantaneous WSS and LDL values on the entire aortic arch. In total, 50 cardiac cycles were used, yielding approximately 2.5 Billion surface data points.

(21)

Appendix A. Supplementary Materials Appendix A.1. MRI Acquisition

MRI measurements were performed on a young healthy male, who gave proper consent to participate in the study. Geometry and flow data were ac-quired using a 1.5 T Philips Achieva MRI scanner. The complete aorta was obtained within a breath hold and the 3D volume data was recon-structed to a resolution of 0.78x0.78x1.00 mm3

. To retrieve a physiologi-cal inlet velocity profile, time-resolved aortic flow profiles were obtained by performing throughplane 2D velocity MRI acquisition placed supracoronary and perpendicular to the flow direction and was reconstructed to 40 time-frames per cardiac cycle with a spatial resolution of 1.37x1.37 mm2

. The throughplane where the velocity profiles were measured is indicated as sec-tion A-A in figure C.1. Aortic geometry from the MRI images was extracted with a 3D level set algorithm using the freely available software Segment (http://segment.heiberg.se) (Heiberg et al., 2010).

Appendix A.2. Mesh Generation

The segmented geometry obtained from the MRI measurement was cut in the ascending aorta at section A-A and in the thoracic aorta at section B-B, see figure C.1. The brachiocephalic, left common carotid, and left subcla-vian arteries were extended 30 mm to minimize numerical problems at the outlets. A high quality hexahedral computational mesh was constructed in ANSYS ICEM CFD 13.0 (ANSYS Inc. Canonsburg, PA, USA). The dimen-sionless wall distance y+

had a maximum value of 1.0 and a mean value of 0.25 during a cardiac cycle, which ensured a good resolution close to the wall. The y+

term is a dimensionless distance from the wall and is normally used to check where the first mesh node is located in the boundary layer. It is defined as y+

= yu∗/ν where y is the normal distance from the wall

to the first mesh node, u∗ the (wall) friction velocity, and ν the viscosity.

The friction velocity is defined as u = pτw/ρ where τw is the wall shear

stress and ρ the density of the fluid. It has been shown that the near-wall region can be divided into three layers; the innermost layer called the viscous sublayer, where viscosity plays an important role in momentum, energy and mass transfer. The outermost layer is called the defect layer where turbulence plays an important role. Between the two layers is the log layer where vis-cosity and turbulence are equally important. A y+

value of 1 means that the first mesh node is well inside the viscous sublayer and that consecutive mesh

(22)

nodes will resolve the rest of the viscous sublayer, the log-layer and the defect layer. Mesh independency tests were carried out with 2.8, 4.8, and 12.1 Mil-lion cells and both velocity, WSS and LDL quantities were considered. The LDL concentration boundary layer was resolved with at least 5 mesh nodes (locally even more) which was found to be sufficient. Differences between the two larger meshes were at most 2.5 % when regarding WSS and LDL levels. The smallest mesh gave similar results but the near-wall resolution was lower and it was therefore decided to use the 4.8 Million mesh.

Appendix A.3. Modeling Blood Flow

The fluid was set to resemble blood with a constant viscosity of 3.5e-3 Pa s and a density of 1080 kg/m3

. One assumption was that the fluid was New-tonian, which might not be the case for real blood. However, Liu et al. (2011) simulated aortic blood flow when the fluid was modeled both as New-tonian and with the non-NewNew-tonian Carreau model. They found a maximum difference of 20 % between the models in time-averaged WSS and LDL dis-tribution. The largest differences were located in the decending arch, while only minor differences were found in the ascending and thoracic aorta. From the velocity profiles in the MRI measurements, the Reynolds number based on inlet diameter, ranged from 150 at late diastole to 6500 at peak systole, with a mean of 1200. A subject specific velocity profile boundary condition based on MRI velocity measurements was set in the ascending aorta (see fig-ures C.1 and C.2). In the branches leaving the arch, mass flow rates based on the measured difference in ascending and descending flow times a scale fac-tor were used. The scale facfac-tors were based on local cross-sectional area and were 10/16, 1/16, 5/16, for the brachiocephalic, left common carotid, and left subclavian artery, respectively. In the descending aorta a three-element Windkessel model was used, which describes the relationship between the aortic outflow and aortic pressure. The measured as well as the resulting mass flow rates together with the Windkessel pressure are plotted in fig-ure 1 in the article. For further details on the Windkessel implementation, see Lantz et al. (2011).

As the flow is pulsating and the Reynolds number is in the range of transitional flow, both a laminar and a LES simulation were run on the same mesh to investigate the effect of including a scale-resolving turbulence model. LES is a technique that separates between large and small scales in the flow, and the scales larger than a filter width (normally the grid spacing) are resolved while the smaller scales are handled by a sub-grid model. The

(23)

governing equations are obtained by filtering the time-dependent continuity and Navier-Stokes equations, which then becomes:

∂ui ∂xi = 0, (A.1) ∂ui ∂t + ∂ ∂xj (uiuj) = − 1 ρ ∂p ∂xi + ν ∂ ∂xj  ∂ui ∂xj + ∂uj ∂xi  − ∂τij ∂xj , (A.2)

where overbar on a variable (e.g. ui) means that it is filtered. An extra term

−∂τij

∂xj appears in equation A.2, where τij denotes the subgrid-scale stresses

which includes the effect from the small scales. The subgrid-stresses are related to the large-scale strain rate tensor Sij through the eddy-viscosity

hypothesis:

τij −

1

3δijτkk= 2ντSij, (A.3)

where ντ is the eddy viscosity and Sij is the resolved strain-rate. The eddy

viscosity was modeled with the wall-adapted local eddy-viscosity (WALE) LES model by Nicoud and Ducros (1999). It was designed to eliminate the subgrid contribution in shear flows and to return the correct near-wall scaling of the subgrid eddy-viscosity (ντ = O(y3)) without any damping functions.

Nicoud and Ducros also showed that the model can handle transition regimes. The eddy-viscosity in the WALE model is defined as:

ντ = (Cwale∆)2

(Sd

ijSijd)3/2

(SijSij)5/2+ (SijdSijd)5/4

, (A.4)

where Cwaleis a model constant set to 0.5 based on results from homogeneous

isotropic turbulence, ∆ is the cube root of the computational cell volume, and Sd

ij is the traceless symmetric part of the square of the velocity gradient

tensor. The Sd

ij tensor can be rewritten in terms of (filtered) strain-rate Sij

and vorticity Ωij as:

Sd

ij = SikSkj+ ΩikΩkj −

1

3δij(SmnSmn−ΩmnΩmn). (A.5) To ensure that initial transient effects had disappeared and that the con-centration boundary layer had stabilized, 15 consecutive cardiac cycles were needed for the laminar model to stabilize the LDL concentration boundary layer; results were taken for the last cycle. Due to the transient nature of

(24)

the LES model, 55 cardiac cycles were computed and phase averages of WSS and LDL were computed using all 50 cycles. This ensures results that are independent of sudden transient effects. The simulations were performed with ANSYS CFX 13.0 (ANSYS Inc. Canonsburg, PA, USA). A maximum root-mean-square (RMS) residual of 10−6 was defined for the momentum

and continuity equations, and domain imbalances were converged well below 0.05 %. Independence tests were carried out on the convergence criteria to ensure an accurate solution. Temporal discretization was performed with a second order backward Euler scheme and the spatial discretization used sec-ond order central differencing. The time step was 1e-4 s yielding a maxium Courant number of 0.9. The simulations were run at National Supercomputer Centre (NSC), Link¨oping, Sweden.

Modeling LDL Transport

The inlet concentration was set to 1.2 mg/ml following Stangeby and Ethier (2002), and a zero concentration gradient was assumed at all outlets. The transport of LDL was modeled as a passive non-reacting scalar, where the transport equation for the laminar model can be described as:

∂C

∂t + ∇ · (UC) = ∇ · (D∇C), (A.6)

where C is the concentration, U is the fluid velocity obtained from the flow simulation, and D the kinematic diffusivity. The first term on the left hand side is a transient term, accounting for the change of C within each control volume. The second term is the transport of C due to convection and the right hand term is the transport due to diffusion. Here D was assumed isotropic and set to a constant value of 5 · 10−12 m2

/s following Wada and Karino (1999, 2002); Olgac et al. (2008). For the LES model, filtering the transport equation yields an extra term:

∂C ∂t + ∇ · (UC) = ∇ ·  D + ντ Sct  ∇C  (A.7) where ντ is the eddy viscosity obtained from the LES subgrid model and Sct

the turbulent Schmidt number, which was set to the default value of 0.9 (AN-SYS Inc., 2010).

As a wall-free model was used in the simulations, a boundary condition which allows for mass transfer through the wall was set on the arterial sur-face. The net transport of LDL from the blood to the wall is modeled as the

(25)

difference between the amount of LDL carried to the wall as water filtration and the amount of LDL diffusing back to the bulk flow. Due to the filtration flow of water into the wall and LDL rejection from the endothelium, a con-centration polarization effect appears at the luminal side of the vessel wall. This effect can be modeled as:

CwVw−D ∂C ∂n w = KwCw (A.8)

where Cw is the concentration of LDL at the wall, Vw the water infiltration

velocity, D the kinematic diffusivity, ∂C/∂n the concentration gradient nor-mal to the wall, and Kw a permeability coefficient for the wall. The values

for Vw and Kw were set to 4·10−6cm/s (Wada and Karino, 1999, 2002; Olgac

et al., 2008; Kolandavel et al., 2006), and 2 · 10−10 m/s (Wada and Karino,

2002; Prosi et al., 2005), respectively.

Appendix A.4. Impact of Sub-Grid Scale Model

The role of the WALE sub-grid scale (SGS) model used in the LES sim-ulations was investigated by computing the eddy-viscosity ratio ντ/ν, which

is the ratio of eddy-viscosity added by the sub-grid model to the dynamic viscosity of the fluid. It can also be used to visualize the disturbed flow structures in the flow (Tan et al., 2011). The model is designed to introduce additional dissipation in regions where the flow is disturbed and turbulent fluctuations are present, while unsteady but laminar regions are ignored. In figure C.3 iso-surfaces of the eddy viscosity ratio is plotted at peak systole, max deceleration and end of systole. The maximum value is about 2, indicat-ing that the model locally adds up to 200% dissipation, to handle the energy in the scales smaller than the mesh (LES filter) size. The sub-grid model should not produce turbulent viscosity close to the wall, and at a distance 0.5 mm from the wall the maximum value is 0.5% which can be regarded as insignificant.

At peak systole the flow is still well organized, due to the stabilizing effect of the accelerating fluid, but small disturbances starts to grow in the aortic arch. The lack of any added eddy viscosity in the descending aorta indicates that the flow there is well structured. However, in the systolic deceleration phase, the flow becomes disorganized which is indicated by the growth of smaller structures, predominantly in the arch and descending aorta. Eddy viscosity ratio values on the order of 2 are present, indicating a disturbed

(26)

flow field. These effects are present throughout the rest of systole but dis-appears during diastole when the energy, and thus the disturbances, in the flow decreases and returns to a well structured state.

Appendix A.5. Motivation For the Use of LES

Since LES is rarely used in cardiovascular flows, partly due to it being computationally expensive and partly because smaller vessels often are as-sumed to be in the laminar regime, a comparison between a laminar model and the LES model was made.

First, a steady-state laminar model with LDL transport similar to the one used by Liu et al. (2011) was simulated. Our simulation used the same boundary conditions, but differed in numerical code used and, of course, in the patient-specific geometry. Using constant inlet velocity, blood density and Newtonian viscosity from Liu et al. (2011), and assuming an inlet diam-eter of Reynolds number 2.5 cm, the Reynolds number in the steady state simulation was 750. Despite geometrical differences, wall shear stress and LDL distributions agreed very well with those found by Liu et al. (2011). For example, our model predicted areas of low shear stress on the inner curva-ture of the descending aorta while the same region at the same time exhibited increased levels of LDL concentration, see figure C.4. Overall, the general appearance of the WSS and LDL distributions are very similar to the results seen in figure 3a and figure 4a in Liu et al. (2011), which made us confident that our transport model for LDL worked and could reproduce the results of other authors. We also simulated the same flow using a LES model, which gave exactly the same WSS and LDL distribution as the laminar model. No additional eddy viscosity was introduced by the sub-grid model, which meant that the flow was indeed purely laminar, and hence, there was no need for LES modeling.

However, the time-averaged Reynolds number measured by MRI in our pulsatile flow was 1200, with the peak Reynolds number on the order of 6000. It was therefore expected to find locally disturbed or even transitional flows, which made the choice of LES modeling appealing. It was therefore decided to compare a pulsatile laminar model with LES, to investigate any possible differences on the LDL concentration distribution caused by the turbulence model. Common RANS models such as k − ω or k − ǫ were not considered, as they do not resolve the turbulence properly and their (in general) poor performance in transitional cardiovascular flows (Yoganathan et al., 2005).

(27)

It was found that at these Reynolds numbers, where disturbed flow is present, the laminar model increases the effective transport of LDL, which, in turn, decreases the LDL concentration of the arterial surface significantly compared to the results from the LES, see figure C.5 (notice the different color scales). This is probably due to the laminar model’s inability to han-dle the disturbed flow accurately during systole. The LES sub-grid model became activated during large parts of systole and the beginning of dias-tole, indicating that small scale motions were present, which can be seen as an indicator of flow disturbances. When a periodic LDL concentration was reached, the laminar model seemed to be less affected by the flow field compared to the LES model, as the LDL concentration is almost stationary for the laminar model during a cardiac cycle, while it changes significantly in the LES.

Clearly, in this range of Reynolds numbers, there is a significant dif-ference between treating the flow as laminar or modeling it with a scale-resolving turbulence model. As the peak Reynolds number is in the tran-sitional regime (Peacock et al., 1998; Stalder et al., 2011), and the flow is disturbed during a substantial part of systole and the beginning of diastole (indicated by the local increase of eddy viscosity by the LES sub-grid model), one can conclude that the flow is best treated with scale-resolving turbulence model to accurately account for all flow features.

For aortic flows with a lower Reynolds number the laminar assumption might be valid, if the amount of disturbances generated is low. But, as in this case, when modeling LDL transport in a flow when the Reynolds number is in the transitional regime, the use of a scale-resolving technique such as LES is needed.

Appendix B. References

ANSYS Inc., 2010. ANSYS CFX-Solver Modeling Guide. 275 Technology Drive, Canonsburg, PA 15317, USA.

Heiberg, E., Sjogren, J., Ugander, M., Carlsson, M., Engblom, H., Arheden, H., 2010. Design and validation of Segment–freely available software for cardiovascular image analysis. BMC Med Imaging 10, 1.

Kolandavel, M.K., Fruend, E.T., Ringgaard, S., Walker, P.G., 2006. The effects of time varying curvature on species transport in coronary arteries. Ann Biomed Eng 34, 1820–1832.

(28)

Lantz, J., Renner, J., Karlsson, M., 2011. Wall shear stress in a subject spe-cific human aorta - Influence of Fluid-structure interaction. International Journal of Applied Mechanics 3.

Liu, X., Fan, Y., Deng, X., Zhan, F., 2011. Effect of non-Newtonian and pulsatile blood flow on mass transport in the human aorta. J Biomech 44, 1123–1131.

Nicoud, F., Ducros, F., 1999. Subgrid-scale stress modelling based on the square of the velocity gradient tensor. Flow, Turbulence and Combustion 62, 183–200.

Olgac, U., Kurtcuoglu, V., Poulikakos, D., 2008. Computational modeling of coupled blood-wall mass transport of LDL: effects of local wall shear stress. Am. J. Physiol. Heart Circ. Physiol. 294, H909–919.

Peacock, J., Jones, T., Tock, C., Lutz, R., 1998. The onset of turbulence in physiological pulsatile flow in a straight tube. Experiments in Fluids 24, 1–9.

Prosi, M., Zunino, P., Perktold, K., Quarteroni, A., 2005. Mathematical and numerical models for transfer of low-density lipoproteins through the arterial walls: a new methodology for the model set up with applications to the study of disturbed lumenal flow. J Biomech 38, 903–917.

Stalder, A.F., Frydrychowicz, A., Russe, M.F., Korvink, J.G., Hennig, J., Li, K., Markl, M., 2011. Assessment of flow instabilities in the healthy aorta using flow-sensitive MRI. J Magn Reson Imaging 33, 839–846.

Stangeby, D.K., Ethier, C.R., 2002. Computational analysis of coupled blood-wall arterial LDL transport. J Biomech Eng 124, 1–8.

Tan, F.P., Wood, N.B., Tabor, G., Xu, X.Y., 2011. Comparison of LES of steady transitional flow in an idealized stenosed axisymmetric artery model with a RANS transitional model. J Biomech Eng 133, 051001. Wada, S., Karino, T., 1999. Theoretical study on flow-dependent

concen-tration polarization of low density lipoproteins at the luminal surface of a straight artery. Biorheology 36, 207–223.

(29)

Wada, S., Karino, T., 2002. Theoretical prediction of low-density lipoproteins concentration at the luminal surface of an artery with a multiple bend. Ann Biomed Eng 30, 778–791.

Yoganathan, A.P., Chandran, K.B., Sotiropoulos, F., 2005. Flow in pros-thetic heart valves: state-of-the-art and future directions. Ann Biomed Eng 33, 1689–1694.

(30)

Appendix C. Additional Figures

Figure C.1: Left: Maximum Projection Image (MIP) of the aorta. Section A-A shows where flow profiles were measured in the ascending and descending aorta, and it also marks the inlet in the CFD-model. Section B-B indicate where the model ends in the thoracic aorta. Right: the hexahedral mesh used in the CFD-model.

(31)

0 0.5 1 Inlet velocity [m/s] 0 0.5 1 0 0.5 1 0 0.5 1

Figure C.2: Examples of measured velocity profiles in the ascending aorta which were used in the CFD-simulations. From top left to lower right: max acceleration, peak systole, max deceleration and mid diastole.

(32)

Figure C.3: Iso surfaces of eddy viscosity ratio ντ/ν at peak systole, max deceleration and end systole.

Figure C.4: WSS and LDL distribution in the steady state model at Reynolds number = 700. The results are very similar to the results in figure 3a and figure 4a in the article by Liu et al. (2011), which made us confident that our model for LDL worked and could reproduce the results of other authors.

(33)

Figure C.5: LDL distribution for the laminar and LES models at peak systole, max de-celeration and late diastole. The concentration level is significantly lower in the laminar model compared to the LES results, as well as the temporal changes; in the LES model the pulsatile flow clearly affects the LDL concentration while in the laminar model the concentration level seems to be almost stationary. Notice the different color scale for the two models.

References

Related documents

Stöden omfattar statliga lån och kreditgarantier; anstånd med skatter och avgifter; tillfälligt sänkta arbetsgivaravgifter under pandemins första fas; ökat statligt ansvar

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

Exakt hur dessa verksamheter har uppstått studeras inte i detalj, men nyetableringar kan exempelvis vara ett resultat av avknoppningar från större företag inklusive

To exclude that PCSK9 cell surface binding was inhibited in an unspecific manner by heparinase-generated heparan sulfate fragments, we first treated cells with heparinase then

Drivers of spatiotemporal variability in CO2 concentration and flux in the inflow area of a large boreal lake, Limnol... Abstract

In this paper we present a method that can be used to measme particle mass fractions in i multiphase flow consisting of water and iron ore particles.. The method is

Emellertid skulle det kunna tolkas som att det log-linjära sambandet inte gäller för väldigt låga LDL-nivåer eller som att det finns pleiotropa effekter hos statiner som gör

Parametric Study of a Mitral Valve Model for Blood Flow Simulation in the Left Ventricle of the