• No results found

A multiscale approach for the coupled simulation of blood flow and thrombus formation in intracranial aneurysms

N/A
N/A
Protected

Academic year: 2021

Share "A multiscale approach for the coupled simulation of blood flow and thrombus formation in intracranial aneurysms"

Copied!
10
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Procedia Computer Science 18 ( 2013 ) 1006 – 1015

1877-0509 © 2013 The Authors. Published by Elsevier B.V.

Selection and peer review under responsibility of the organizers of the 2013 International Conference on Computational Science doi: 10.1016/j.procs.2013.05.266

International Conference on Computational Science, ICCS 2013

A multiscale approach for the coupled simulation of blood flow

and thrombus formation in intracranial aneurysms

Simon Zimny

a,∗

, Bastien Chopard

b

, Orestis Malaspinas

b

, Eric Lorenz

c

, Kartik Jain

a

,

Sabine Roller

a

, J¨org Bernsdorf

a

aGerman Research School for Simulation Sciences and RWTH Aachen University, Schinkelstr. 2a, Aachen, Germany bComputer Science Department, University of Geneva, 7 route de Drize,1227 Carouge, Switzerland

cComputational Science, Faculty of Science, University of Amsterdam, Netherlands

Abstract

This paper considers a multiscale description of thrombus formation and its simplified numerical implementation in the case of cerebral aneurysms. In particular, we extend previously introduced generic 2D models towards 3D patient specific aneurysm geometries. The multiscale amplification method contributes to considerably reducing simulation time. This allows us to achieve a mesh resolution high enough to resolve details of the stent geometry which is triggering flow conditions to induce clotting. Simulation results presented in this paper are qualitatively in a good agreement with clinical observations.

Keywords: Lattice Boltzmann Method; Scale Separation Map; Thrombus; Intracranial Aneurysms; Multiscale Simulation

1. Introduction and Motivation

In recent years the treatment of intracranial aneurysms (IA) has been studied at length and methods of treat-ment have improved significantly. Besides surgical clipping and endovascular coiling, the use of stents as flow diverters is becoming more and more popular. The idea behind this non-invasive treatment technique is to change the blood flow properties in order to trigger thrombosis in aneurysm cavities and, thus, its complete occlusion, eventually preventing the risk of rupture.

For the clinician it is a great challenge to decide the appropriate treatment strategy, once an IA has been detected. Ideally, the choice of a suitable stent is supported by a prediction of the occlusion probability of an IA, which requires an accurate estimation of the post-treatment flow properties. Currently, the clinician’s intuition and experience is the only guideline in this process. A secondary concern is the optimization of stent geometries, which requires a detailed insight into the complex interaction of flow and biological processes for various combinations of stent and aneurysm topologies.

In the recent years several projects simulated the blood flow in human brains, especially in IAs. The HemeLB project1 focussed on the prediction of the blood flow and on near vascular malformations while the @neurIST

Corresponding author. Tel.:+49-(0)241-80-99769; fax: +49-(0)241-80-6-99769.

E-mail address: s.zimny@grs-sim.de.

1http://www.2020science.net/software/hemelb

(2)

project2focus was on the rupture risk assessment of cerebral aneurysm. In addition a simple thrombosis model

using only the wall shear stress (WSS) as clotting criteria has been developed [1].

The aim of the present paper is to develop a concept for numerical simulation of the highly complex process of thrombus formation in patient specific intracranial aneurysms. This requires the concurrent simulation of both, blood flow in complex geometries, and biological processes such as blood coagulation under the influence of local flow properties. These results shall help the clinicians to choose the optimal stent design to trigger the complete occlusion of the patient specific aneurysm. A particular challenge arises from the fact that these processes occur over a large spectrum of time- and length scales, which prevent the application of classical simulation techniques due to the expected enormous demand of memory and simulation time.

Here we base our approach on an extension of the model proposed by Ouared et al. [2, 3] and Bernsdorf and Harrison et al. [4, 5, 6] by formalizing the multiscale part of the problem.

The remainder of this publication is structured as follows: First the bio-medical agents that are involved in the formation of a thrombus are explained in detail. Based on this a possible scale separation map (SSM) including various time (μs to weeks) and length scales (Å to dm) is defined. After definition of the SSM and its associated coupling patterns, a simplification strategy for a possible implementation is described.

In a first step towards simulation, a numerical model was derived from the macroscopic part of the SSM and implemented into a flow solver software package to simulate the thrombus formation in a patient specific aneurysm geometry. This model is depending on two flow properties, namely the wall shear stress (WSS) and the residence time of the blood in certain areas of the aneurysm. Due to the high complexity of the aneurysm geometry the blood flow is simulated using the lattice Boltzmann method, which is proved to be optimal for large numerical simulations. [7].

2. Multiscale Modeling

2.1. The biological process of thrombus formation

The formation of thrombus is a highly complex process depending on a large number of biochemical agents reacting with each other, rheological properties of blood plasma, vessel wall (endothelium, extracellular matrix) and blood cells like platelets or red blood cells (RBC). The biochemical agents can support (procoagulants) or hinder (anticoagulants) the thrombus formation. Currently, there is no full understanding of the detailed processes

that lead to thrombosis in aneurysm cavities. The THROMBUS European project 3 addresses this question by

combining new biological experiments and new clinical observations into a multiscale numerical simulation. In this paper we review the processes that are known to play a major role in clot formation, in general. The key ingredients are

• platelets including its activation, aggregation and adhesion (see Fig. 1, red dotted circle)

• the coagulation cascade leading to fibrin production including the incorporation of particles (e.g. RBCs) (see Fig. 1, blue dotted circle)

• the endothelium as a modulator of procoagulants [8].

Although all parts can initiate the formation of a blood clot on their own, the interaction between them is more likely to occur since platelets as well as coagulation factors and the endothelium are always present in human blood vessels.

Platelets in their inactive form have a discoid shape, a smooth rippled surface [9] and an average diameter of 1.5 - 2μm [10]. Their surface hosts various numbers of receptors that can be stimulated by pro- and anticoagulants. In addition the platelets organelles include different kinds of granules. The α-granules contain several procoagulants (e.g. factor V, von Willebrand factor (vWF) [11, 12]).

The activation of platelets can be triggered by • exposure to high shear stresses (∼ 5Pa [13, 14]) 2http://cilab2.upf.edu/aneurist1/index.php

(3)

platelets cells (e.g. RBC,monocytes) extrinsic materials(e.g. stent, coil) endothelium platelet activation platelet aggregation coagulation cascade clot formation changes in the clot structure clot lysis

Fig. 1. The biological processes leading to thrombus formation.

• extrinsic materials (e.g. stent, coil) [10]

• coagulation factors [10, 15] (e.g. collagen XII release from injured endothelium [16, 17]) • other activated platelets

Activated platelets change their shape from discoid to spherical [10] and start to expose proteins and growth factors from their granules. Receptors on their surface support the ability to aggregate to the subendothelial matrix of the vessel wall [10], exogenous surfaces and to each other. Note also that platelets can adhere without prior activation [18]. Hemoglobin that is known to hinder the natural platelet inhibition process [12] is emitted from damaged erythrocytes (red blood cells) at extremely high shear stresses (∼ 150Pa, [19]), which triggers the activation of platelets.

Under healthy conditions the pro- and anticoagulants are balanced. Due to external influences (e.g. trauma, inflammation) this balance is disturbed leading to thrombus formation. The influence of imbalanced coagulants can be described by the coagulation cascade. The coagulation cascade leads to the production of fibrin from its inactive form fibrinogen, which aggregates to fibrin strands forming a cross-linked mesh. In this mesh blood cells, e.g. RBCs, are caught and incorporated. Uchida et al. [20] characterizes the different types of thrombi by its color and by the level of platelets (white) and RBCs (red, brown) incorporated. White thrombi tend to be fibrin-rich while brown and red thrombi tend to be fibrin-poor [20].

The coagulation cascade consists of two major pathways (extrinsic and intrinsic), which lead to a common pathway in the end [21]. The intrinsic pathway (also known as contact activation pathway) is initiated by inflam-mations or the contact of blood to an extrinsic surface (e.g. stents). . The more important pathway to consider for the formation of a stent-induced thrombus is the extrinsic pathway (also known as Tissue Factor Pathway). This pathway is initiated by a dysfunctioning, or a damage to the vessel wall (e.g. inadequate flow properties, cut or stitch while stent deployment).

Among a various number of proteins (Protein C and S) and agents (vWF) the twelve coagulation factors I-XIII (factor VI is not defined) are major contributors to the fibrin production. Besides fibrin thrombin and its inactivated form prothrombin is the central link of the two pathways and activates fibrinogen to fibrin [13].

Koskinas et al. [8] pointed out that in areas of low WSS proinflammatory genes are upregulated. These areas appear in the bulge of stented aneurysms and often coincide with recirculation zones, in which the blood is trapped. Due to the higher residence time, the procoagulant factors in the blood increase and initiate the formation of a blood clot.

The process of clot lysis is almost simultaneously initiated [12] and helps to curtail the clot growth. This process is dependent on the anticoagulants like the tissue-factor-pathway-inhibitor (TFPI) , antithrombin, proteins, plasmin and prostacyclin. Also physical quantities like high shear stresses may lead to partial or complete abortion or dilution of the thrombus.

(4)

Experiments by Kamocka et al. [16] lead to the conclusion that the thrombus itself goes through several stages. At first the thrombus grows rapidly. After this growth phase the volume decreases due to the contraction of the incorporated platelets. Kamocka et al. also suggests the hypothesis that the surface composition changes from platelet- to fibrin-rich forming a fibrin cap.

2.2. The complete scale separation map

The processes that lead to the formation of a thrombus are acting on a wide range of temporal (ms to weeks) and spatial scales (Å to several cm). Due to this wide spectrum the concept of a scale separation map (SSM) proposed by Hoekstra et al. [22] is used to describe the formation of a thrombus in intracranial aneurysms. A SSM is defined as a two dimensional map with the temporal scales on the x- and the spatial scales on the y-axis. The different processes are drawn as boxes within their specific temporal- and spatial range. The mutualffff interaction of the different processes is described by directed edges between the boxes.ffff

For the case of thrombus formation in IA at first a complete scale separation map (see Fig. 2) has been gener-ated directly from the biological processes described in section 2.1. The complete system was split up in several subsystems describing pure physical processes like the rheology (light green) and particles (yellow), biochemical processes like the formation (orange) and structure and lysis of a thrombus (light blue).

Fig. 2. The complete scale separation map.

The rheology contains of the bulk flow (BF) describing the blood flow itself and the boundary layer (BL) describing the influence of the moving vessel wall on the flow. The spatial bounds are based on the diameter of the basilar artery (∼ 4.3mm [23]) and the maximal size of a single strut of a stent (∼ 0.1mm [24]) while the temporal bounds are determined by the cardiac cycle (∼ 1.0s [25]) and turbulence effects (ffff ∼ 1.0ms). The boundary layer (BL) acts in spatial scales of approximately an order of magnitude less than the bulk flow (∼ 0.1mm, max. strut size to∼ 10m, min. strut size, [24]). A range of ∼ 10ms to ∼ 1min was chosen to define the temporal scales.

The last pure physical process shown in this version of the SSM is the particle transport (PT) referring to the transport of cells such as red blood cells (RBCs) and platelets (passive and activated) in the bulk flow. This subsystem is spatially bound by the size of platelets, the smallest cell of interest in the blood stream with a diameter of∼ 1.5 to 3μm and clusters of RBCs (diameter ∼ 7.5μm) and platelets. For the temporal bounds a range of ∼ 1s to∼ 1min was chosen.

The biochemical processes describing the thrombus formation are namely the coagulation cascade (CC), platelet aggregation/n adhesion (PlA), particle processes (PP) incl. platelet activation and dynamics of the thrombus (CSL) like change in structure, size and its lysis. The corresponding spatial and temporal scales are listed below:

(5)

• Coagulation Cascade: CC

spatial:0.1μm (α-granules size, 200-500 nm [11]) temporal:∼ 1ms up to ∼ several min

• Platelet Aggregation: PlA

spatial: severalμm (aggregated platelets) up to several cm (clot size)

temporal:<min (platelet-platelet aggregation) up to weeks/months (development of a full clot) • Particle Processes: PP

spatial:∼ 1μm to ∼ 10μ0 m (particle size)

temporal:∼ 0.1s (activation by coagulation factors, [15]) to ∼ several min (shear induced activation, [26]) • Clot Structure/e Lysis: CSL

spatial:∼ 1μm (one platelet) up to ∼ 0.1mm (several particles)

temporal:∼ 10ms (lysis of small aggregates) to ∼ several days (macroscopic changes in the clot structure) The arrows in Fig. 2 describe the kind of information exchanged between the different subsystems. The redffff arrows are based on physical properties like shear stress and particle densities, while the black arrows represent the biomedical influenced properties like rates of activated coagulation factors.

2.3. From the complete to the reduced scale separation map

Due to the sheer complexity of the complete SSM (see Fig. 2) a full implementation in a numerical multiscale model is beyond reach at the moment. In particular, the complexity caused by overlap of scales introduces compli-cations. In attempt to group processes according to common functional entities (molecules, cells, flow structures / tissue structure) we arrived at an SSM that contains only a few but well separated subprocesses (see Fig. 3). The arrows represent again the type of interaction. In the following subsections 2.3.1 to 2.3.6 all underlying subsystems and their spatial and temporal scales are presented.

Fig. 3. The reduced scale separation map.

2.3.1. Macroscopic flow and advection-diffi usionffff

This process is referring to the macroscopic blood flow where the blood is treated as a continuum. Addition-ally the advection-diffusion processes of several substances, which are crucial for the thrombus formation, likeffff coagulation factors, cofactors, proteins and nucleotides or lysis, like anticoagulants, of a thrombus are considered.

(6)

The spatial scales are chosen based on the diameter of the basilar artery (∼ 4.3mm, [23]) up to the diameter of giant intracranial aneurysms (< 2.5cm, [27]). The temporal scales describe the range of turbulence influences (∼ 1ms) up to a cardiac cycle (∼ 1s, [25]).

2.3.2. Clotting dynamics and clot formation

The formation of a thrombus can generally be characterized by two phases [16]. At first the volume of the thrombus increases rapidly forming a rather soft structure. This is followed by a volume reduction as a result of platelet contraction and leads to a stable structure. A key part of the stability is the development of a cross-linked fibrin mesh, which is a product of the coagulation cascade [12]. The lysis of a clot can be due to anticoagulants (e.g. tissue factor pathway inhibitor (TFPI) or physical impact from the flow, like high shear stresses [12]. The spatial scales are based on the full occlusion of a giant intracranial aneurysm (< 2.5cm, [27]) and geometrical change with an effect on the macroscopic flow (∼ 0.1mm). The temporal scales are described by the time required for such a change (∼ several s) up to the fully developed clot (∼ several days).

2.3.3. Microscopic flow, particle transport and adhesion/aggregation

Platelets are often seen as the main ingredient for the initiation and the formation of a thrombus. Both, the aggregation of platelets and the adhesion to the vessel wall, act together in this formation process. The aggregation of platelets can be induced by a great number of agents (like ADP, epinephrine, collagen and thrombin) [10]. Platelets adhere to a variety of extracellular matrix connective tissue components (e.g. fibrin, laminin and collagen [10]) using several receptors. Collagen can be exposed by endothelial damage or when the endothelium is subject to both high [17] or low shear rates [28]. High shear rates are also related to receptor unlocking, to the break up of platelet aggregants [12] and the damage of red blood cells which lead to the exposure of hemoglobin, hindering the natural platelet activation inhibition [19].

2.3.4. Molecular receptors

The major reactions in the cascade take place on the surface of cells . The platelet receptor glycoprotein IIb/IIIa (GP IIb/IIIa) plays an important role in the aggregation of platelets [29], using fibrinogen as the connecting agent. There is also a large number of other receptors for procoagulants (e.g. GPIb/IX/V for vWF, GP VI, GP Ia/IIa collagen [30]), promoting the aggregation and adhesion of platelets.

2.3.5. Coagulation Cascade

The coagulation cascade is a subcellular process, which leads to the production of cross-linked fibrin meshes, a key part of a stable thrombus formation. As described previous in section 2.1 the coagulation cascade is based on a large number of procoagulants in an inactive and active form as well as on anticoagulants acting against the production of fibrin strands. In addition to these nucleotides, proteins and cofactors take part in the coagulation cascade.

2.3.6. Platelet activation

The second key part of thrombus formation are activated platelets. Platelets in their inactive form have a discoid shape with a smooth rippled surface [10] and an average diameter of 1.5 to 2 m [9]. The activation can be triggered by physical impact (e.g. high shear stresses [13, 14]), extrinsic materials [12], other activated platelets, certain proteins or coagulation factors [10]. After the activation the platelets change their shape from discoid to spherical [9] and start to expose proteins and growth factors (e.g. tissue factor, vWF).

2.4. The interaction scheme

In a numerical multi-scale simulation the different subsystems working on independent spatial- and temporal scales have to communicate with each other to exchange necessary data like flow properties, geometry updates or particle distributions. The mutual coupling between the different parts of the reduced SSM is described using a graphical representation (see Fig. 3) to achieve a better overview. The subsystems are shown as boxes linked to each other by directed edges describing what has to be communicated (see Fig. 4). This representation helps finding a possible communication structure for a numerical multi-scale simulation. Furthermore, it is a good way of finding suitable simplifications of the simulation framework such that it can be executed in reasonable time on available HPC systems.

(7)

systole flow macro

advection-diffusion

micro particle transport adhesion / aggregation micro flow platelet contraction platelet activation coagulation cascade molecular receptors clotting dynamics (including clot lysis clot formation) velocity field shea r rate as activation criteria > sp ecies concentration < changes in the concentrations > shea r rates as rupture criteria < geometry up date <> up dated pa rticle/aggregation distribution (new platelets from ruptures) >species concentration <adhesion / aggregation particle distributions < changes in species concentration > concentration of coagulants as act. criteria species concen tration > platelet activated < platelet distribution concentration of pro-and anti-coagulants

Fig. 4. The interaction scheme of processes contributing to the thrombus formation.

3. From multiscale model to numerical simulation

In a full simulation all subsystems described in the reduced SSM (see Fig. 3) should be modeled. In a first step towards this complete multiscale simulation only the tissue levels are considered here, namely the “systole-flow/ macro advection-diffusion” and “clotting dynamics/ clot formation”.

3.1. The systole-flow/macro advection-diffusion subsystem

The systole-flow as well as the transport of blood particle in the flow is simulated using the lattice Boltzmann method (LBM). We refer the unfamiliar reader to the abundant literature on LBM. See for instance [31].

Due to its easy mesh generation, the independence of run-time from the complexity of the flow geometry and very good parallel performance the LBM is well suitable to simulate fluid flows in highly complex geometries using several hundred-thousand cores on modern supercomputers [7, 32]. Several powerful implementations of the LBM exists, in particular the Palabos software4that is intensively used in the THROMBUS project to simulate

thrombus formation in cerebral aneurysms.

3.2. The clotting dynamics/clot formation subsystem

To abstract the sub tissue level processes in the reduced SSM presented in Fig. 3, namely the yellow and orange boxes, macroscopic rheological properties are used. The wall shear stress (WSS) plays a critical role in the formation of a blood clot as stated by Chopard et al. [2, 3] and Harrison et al. [5, 6]. Using the WSS as an upper threshold for the formation of a thrombus is underlayed by common theories describing embolism [8].

The second key parameter for the formation of a blood clot is the time blood resides in certain regions. The so-called residence time model [4, 5] was derived in analogy to experiments of the clotting process of enzymatically activated milk. The main idea is that usually enough procoagulants and activated platelets are present in the blood, and coagulation occurs when they have sufficient time to aggregate. The residence time (equivalent to ’age of the fluid’) is computed by means of a passive scalar whose value is increased at every iteration to estimate the time from the local concentration. At the inlet the concentration of this tracer is initialized with 0, slowly increasing while the fluid travels downstream towards the outlet.

(8)

To prevent clots from developing in non-physical areas like near the outlet and inside the flow itself we are using the proximity [2, 3, 5] condition. This condition allows elements to solidify only in near wall regions, preventing non-physical insulated clot formation in the bulk of the flow.

In case age threshold and near-wall conditions are fulfilled for a computational fluid element, it will turn into a solid (optional with a given probability p) depending on the level of the WSS.

The above rules are with slight modifications a combination of those proposed in [2, 3] for the case of thrombus formation in simple 2D and in [5, 6] for 3D geometries.

4. Numerical simulation

4.1. Amplification for numerical simulation

To resolve the aneurysm with sufficient accuracy the mesh size and such the time step has to be very small. With a time step in the order of 10−5s, more than 8 billion iterations have to be performed for simulating a single day in real time. As can be seen from the reduced SSM (Fig. 3) the upper temporal bound for the clot formation is in the order of weeks. This leads to the conclusion that simulating the whole process would be far too expensive even on the largest supercomputers.

To overcome these problems the technique of amplification [33] can be used. The main idea is to amplify the effect of the long-term process of thrombus formation and only simulate one or two cardiac cycles for the short-term process of blood flow, as proposed in [3].

In the macroscopic model based on the reduced SSM proposed beforehand, the short-term process identifies with the “systole-flow/macro advection-diffusion” subsystem, while the long-term process with the “clotting dy-namics/clot formation” subsystem. Due to this temporal separation of the two systems, the “clotting dydy-namics/clot formation” subsystem can be amplified e.g. by increasing the probability p of an element to solidify or the growth rate.

4.2. Simulation setup

Based on the Model presented in subsection 3.2 a blood clotting simulation in a patient specific stented aneurysm (see Fig. 5) has been performed. The grid size was chosen to beδx = 6.5 · 10−5m resulting in ap-prox. 45 million fluid elements, the time step to beδt = 2.89 · 10−5s the inlet velocity to be uin= 9.435 · 10−2m/s. Using the inlet diameter as the characteristic length and the mean velocity umean = uin/2 as the characteristic velocity resulting in a Reynolds number of 50. The density of blood was set to ρblood = 1025kg/m3 and the

kinematic viscosity to beνblood= 3.8 · 10−6m2/s.

The simulations have been performed on the superMUC at the LRZ-Munich using the APES framework [32]. Due to its efficient representation of the fluid elements and their neighbor relations it is highly suitable to solve large scale systems on hundred-thousand cores. The geometry has been provided by the THROMBUS project in STL format. At first the geometry has been read and meshed by the SEEDER. On the resulting grid the simulation has been performed using the MUSUBI solver and postprocessed by the HARVESTER.

Fig. 5. (a) Complete aneurysm; (b) Stent deployed in aneurysm (provided by THROMBUS project).

As mentioned in subsection 3.2 the simulation had to be amplified to reduce the number of iterations. Therefore the probability for an element was chosen to be constant p= 1 and the growth rate was amplified by letting only whole elements solidify. After the convergence of the flow and the passive scalar tracer the clotting model was switched on with a wall shear stress threshold of 1.037 · 10−1Pa and a residence time threshold of 14.5s.

(9)

4.3. Simulation results

Fig. 6. Temporal evolution of the thrombus formation in the aneurysm bulge for different time steps (105, 3 · 105, 4 · 105, 1.3 · 106iterations).

Fig. 6 shows the thrombus formation in the aneurysm bulge at different time steps in terms of solidified fluid elements (marked in red). The thrombus mainly starts developing from the right face of the aneurysm bulge. Additionally, very few clots appear on the left face. Clots continue to grow in layers, eventually resulting in complete occlusion after 1.3 · 106 iterations. A comparison between these simulation results and those from the

non-stented aneurysm with the same parameters lead to the conclusion, that the insertion of the stent increases the thrombus formation in the aneurysm bulge drastically.

5. Conclusion and Outlook

The model presented in this paper is a first step towards multi-scale simulation of thrombus formation in intracranial aneurysms based on the amplification approach. They extend the work of Ouared et al. [2, 3] and Bernsdorf and Harrison et al. [4, 5, 6] towards realistic 3D patient specific stented aneurysm geometries. The results obtained from the numerical simulations qualitatively match the expected process of thrombus growth.

Using the amplification method we were able to reduce the number of iterations from expected billions to millions for simulating the whole process. Without considering a multi-scale approach, such a simulation would be impossible even on largest super-computers.

The simulations performed for this paper were done with steady flow conditions at a Reynolds number of Re=50. An extension towards physiological flow conditions employing pulsatile flow at higher Reynolds numbers will be the next step. Further extensions of the model will be adapting the growth rate of the thrombus and the probability of solidification in the macroscopic model. In addition, the local concentration of platelets in an active and inactive form can be modeled as passive scalars. The platelet activation process might be induced by flow properties, mainly WSS, and procoagulants.

In the microscopic part of the simulation (see Fig. 3, yellow and orange boxes) cells in the blood, like RBCs and platelets, can be simulated using a suspension code where the transport and shape of the particles are fully resolved. This highly complex approach [34] is currently considered in the THROMBUS project.

The most important coagulation factors like fibrin and thrombin concentrations as well as their interaction, the endothelial and blood cells can be simulated explicitly. This will eventually lead to a fully coupled multiscale sim-ulation of thrombus formation in intracranial aneurysms. A full clotting model, including all the above ingredients is under development in collaboration with biologists, clinicians and computer scientists within the THROMBUS project. Due to the large number of subsystems (Fig. 3) involving spatial and temporal scales spanning over sev-eral orders of magnitude, advanced multiscale techniques have to be employed. The use of a distributed coupling framework (e.g. MAPPER5) might be extremely helpful to achieve the ambitious task of a realistic quantitative

simulation of thrombus formation. Acknowledgements

The research work presented in this paper was funded by the European Commission in the Seventh Framework Programme in the area of Virtual Physiological Human (ICT-2009.5.3, project reference 269966).

(10)

References

[1] Y. Ventikos, E. Holland, T. Bowker, P. Watton, N. Kakalis, M. Megahed, F. Zhu, P. Summers, J. Byrne, Computational modelling for cerebral aneurysms: risk evaluation and interventional planning, British Journal of Radiology 82 (Special Issue 1) (2009) 62–71. [2] R. Ouared, B. Chopard, Lattice Boltzmann simulations of blood flow: non-Newtonian rheology and clotting processes, Journal of

Statistical Physics 121 (1) (2005) 209–221.

[3] R. Ouared, B. Chopard, B. Stahl, D. A. R¨ufenacht, H. Yilmaz, G. Courbebaisse, Thrombosis modeling in intracranial aneurysms: a lattice Boltzmann numerical algorithm, Computer Physics Communications 179 (1) (2008) 128–131.

[4] J. Bernsdorf, S. E. Harrison, S. M. Smith, P. V. Lawford, D. R. Hose, Concurrent numerical simulation of flow and blood clotting using the lattice boltzmann technique, in: J. Ma, L. Yang (Eds.), Proceedings of the 2005 11th International Conference on Parallel and Distributed Systems (ICPADS’05), Vol. II, IEEE, 2005, pp. 336–340.

[5] S. E. Harrison, S. S. M, J. Bernsdorf, D. R. Hose, P. V. Lawford, Application and validation of the lattice Boltzmann method for modelling flow-related clotting, Journal of biomechanics 40 (13) (2007) 3023–3028.

[6] S. Harrison, J. Bernsdorf, D. Hose, P. Lawford, A lattice Boltzmann framework for simulation of thrombogenesis, Progress in Compu-tational Fluid Dynamics 8 (1-4) (2008) 121–128.

[7] J. Zudrop, H. Klimach, M. Hasert, K. Masilamani, S. Roller, A fully distributed CFD framework for massively parallel systems, Pro-ceedings of the Cray User Group CUG 2012, April 29 - May 3, 2012, Stuttgart, Germany.

[8] K. C. Giannoglou, Y. S. Chatzizisis, A. P. Antoniadis, K. C. Koskinas, Role of Endothelial Shear Stress in Stent Restenosis and Throm-bosis, Journal of the American college of Cardiology 59 (15) (2012) 1337–1349.

[9] W. Siess, Molecular mechanisms of platelet activation, Physiological Reviews 69 (1) (1989) 58–178.

[10] S. Kamath, A. Blann, G. Lip, Platelet activation: assessment and quantification, European Heart Journal 22 (17) (2001) 1561–1571. [11] P. Harrison, E. M. Cramer, Plateletα-granules, Blood reviews 7 (1993) 52–62.

[12] M. Anand, K. Rajagopal, K. R. Rajagopal, A model incorporating some of the mechanical and biochemical factors underlying clot formation and dissolution in flowing blood, Journal of Theoretical Medicine 5 (3-4) (2003) 183–218.

[13] M. Anand, K. Rajagopal, K. R. Rajagopal, A model for the formation, growth, and lysis of clots in quiescent plasma. A comparison between the effects of antithrombin III deficiency and protein C deficiency, Journal of Theoretical Biology 253 (4) (2008) 725–738. [14] J. M. Ramstack, L. Zuckermann, L. F. Mockros, Shear-induced activation of platelets, J Biomechanics 12 (1979) 113–125. [15] G. Born, P. D. Richardson, Activation time of blood platelets, Journal of Membrane Biology 57 (1980) 87–90.

[16] M. M. Kamocka, J. Mu, X. Liu, N. Chen, A. Zollman, B. Sturonas-Brown, K. Dunn, Z. Xu, D. Z. Chen, M. S. Alber, E. D. Rosen, Two-photon intravital imaging of thrombus development, Journal of Biomedical Optics 15 (2010) 1–7.

[17] O. Traub, B. C. Berk, Laminar shear stress: mechanisms by which endothelial cells transduce an atheroprotective force, Arteriosclerosis, Thrombosis, and Vascular Biology 18 (1998) 677–685.

[18] J. Yip, Y. Shen, R. K. Andrews, Primary platelet adhesion receptors, Australian Biochemist 35 (2004) 4.

[19] L. B. Leverett, J. Hellums, C. P. Alfrey, E. C. Lynch, Red blood cell damage by shear stress, Biophysical journal 12 (3) (1972) 257–273. [20] Y. Uchida, Y. Uchida, T. Sakurai, M. Kanai, S. Shirai, T. Morita, Characterization of Coronary Fibrin Thrombus in Patients With Acute

Coronary Syndrome Using Dye-Staining Angioscopy, Arteriosclerosis, Thrombosis, and Vascular Biology 31 (6) (2011) 1452–1460. [21] S. Lee, A. M. Pham, S. G. Pryor, T. Tollefson, J. M. Sykes, Efficacy of Crosseal fibrin sealant (human) in rhytidectomy, Archives of

Facial Plastic Surgery 11 (1) (2009) 29–33.

[22] A. Hoekstra, E. Lorenz, J. Falcone, B. Chopard, Towards a complex automata framework for multi-scale modeling: Formalism and the scale separation map, Computational Science–ICCS 2007 (2007) 922–930.

[23] W. Busch, Beitrag zur Morphologie und Pathologie der Arteria basialis (Untersuchungsergebnisse bei 1000 Gehirnen), European Archives of Psychiatry and Clinical Neurosciences.

[24] C. Lally, P. J. Prendergast, D. Kelly, Stents, in: M. Akay (Ed.), Wiley Encyclopaedia of Biomedical Engineering, John Wiley & Sons, pp. 3345–3355.

[25] P. Jehenson, D. Duboc, T. Lavergne, L. Guize, F. Guerin, M. Degeorges, A. Syrota, Change in human cardiac rhythm induced by a 2-T static magnetic field., Radiology 166 (1) (1988) 227–230.

[26] L. Badimon, J. Badim´on, A. Galvez, J. H. Chesebro, V. Fuster, Influence of arterial damage and wall shear rate on platelet deposition. Ex vivo study in a swine model, Arteriosclerosis, Thrombosis, and Vascular Biology 6 (3) (1986) 312–320.

[27] T. Higa, H. Ujiie, K. Kato, H. Kamiyama, T. Hori, Basilar artery trunk saccular aneurysms: morphological characteristics and manage-ment, Neurosurgical Review 32 (2) (2008) 181–191.

[28] P. Davis, Flow-medicated endothelial mechanotransduction, Physio. Rev. (1995) 519–560.

[29] D. Varga-Szabo, I. Pleines, B. Nieswandt, Cell Adhesion Mechanisms in Platelets, Arteriosclerosis, Thrombosis, and Vascular Biology 28 (3) (2008) 403–412.

[30] K. Broos, H. B. Feys, S. F. De Meyer, K. Vanhoorelbeke, H. Deckmyn, Platelets at work in primary hemostasis, YBLRE 25 (4) (2011) 155–167.

[31] S. Succi, The Lattice Boltzmann Equation, For Fluid Dynamics and Beyond, Oxford University Press, 2001.

[32] S. Roller, J. Bernsdorf, H. Klimach, M. Hasert, D. Harlacher, M. Cakircali, S. Zimny, K. Masilamani, L. Didinger, J. Zudrop, An adaptable simulation framework based on a linearized octree (2012) 93–105.

[33] A. G. Hoekstra, A. Caiazzo, E. Lorenz, J.-L. Falcone, B. Chopard, Complex automata: Multi-scale modeling with couples cellular automata, in: A. G. Hoekstra, J. Kroc, P. M. A. Sloot (Eds.), Simulating Complex Systems by Cellular Automata, Understanding Complex Systems, Springer, 2010, Ch. 3, pp. 29–57.

[34] L. Mountrakis, E. Lorenz, A. Hoekstra, Where do the platelets go? A simulation study of fully resolved blood flow through aneurysmal vessels, accepted for publication in Interface Focus Royal Society.

Referenties

GERELATEERDE DOCUMENTEN

Welke sport- en recreatiefaciliteiten zijn er gerealiseerd voor de Spelen in het Olympisch park in Stratford en hoe worden deze gebruikt door de bevolking.. Hoe worden

If a seizure is detected, information from the GPS device together with cell-ID information is used to determine the location of the patient so that

The increasing complexity in both the multidisciplinary cancer care and translation- al research requires a new type of closer collaboration between centers to put the most

Hij gaf aan dat door de aanwezigheid van een derde remlicht de detectie van remmen bij mist wordt vergroot en tevens dat een eventuele verwarring met twee mist- achterlichten

Wanneer daar dus gepoog word om die relatiewe gevoeligheid van verskillende spesies te bepaal deur verskillende response op blootstelling aan ’n chemiese stof te vergelyk, moet

Aangezien er geen archeologische waardevolle vindplaatsen zijn aangetroffen, kan deze vraag niet beantwoord worden. - Wat is de vastgestelde en verwachte

Tijdens de prospectie met ingreep in de bodem door middel van proefsleuven aan de Van de Wervelaan in Vorselaar zijn 176 sporen aangetroffen, waarvan de meerderheid

Er is tijdens de interviews ook navraag gedaan naar de invloed van de door te voeren arbeidstijdverkorting: één bedrijf zegt daar nauwelijks problemen mee te