• Non ci sono risultati.

Asymptotic model of fluid–tissue interaction for mitral valve dynamics

N/A
N/A
Protected

Academic year: 2021

Condividi "Asymptotic model of fluid–tissue interaction for mitral valve dynamics"

Copied!
12
0
0

Testo completo

(1)

1 23

Cardiovascular Engineering and

Technology

ISSN 1869-408X

Volume 6

Number 2

Cardiovasc Eng Tech (2015) 6:95-104

DOI 10.1007/s13239-014-0201-y

Asymptotic Model of Fluid–Tissue

Interaction for Mitral Valve Dynamics

Federico Domenichini & Gianni

Pedrizzetti

(2)

1 23

Your article is protected by copyright and

all rights are held exclusively by Biomedical

Engineering Society. This e-offprint is for

personal use only and shall not be

self-archived in electronic repositories. If you wish

to self-archive your article, please use the

accepted manuscript version for posting on

your own website. You may further deposit

the accepted manuscript version in any

repository, provided it is only made publicly

available 12 months after official publication

or later and provided acknowledgement is

given to the original source of publication

and a link is inserted to the published article

on Springer's website. The link must be

accompanied by the following text: "The final

publication is available at link.springer.com”.

(3)

Asymptotic Model of Fluid–Tissue Interaction for Mitral Valve Dynamics

F

EDERICO

D

OMENICHINI1

and G

IANNI

P

EDRIZZETTI2

1

Dipartimento di Ingegneria Civile e Ambientale, Universita` di Firenze, Via S. Marta 3, 50139 Florence, Italy; and

2

Dipartimento di Ingegneria e Architettura, Universita` di Trieste, P. le Europa 1, 34127 Trieste, Italy

(Received 28 June 2014; accepted 3 November 2014; published online 7 November 2014)

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

Abstract—The vortex formation process inside the left ventricle is intrinsically connected to the dynamics of the mitral leaflets while they interact with the flow crossing the valve during diastole. The description of the dynamics of a natural mitral valve still represents a challenging issue, especially because its material properties are not measurable in vivo. Medical imaging can provide some indications about the geometry of the valve, but not about its mechanical properties. In this work, we introduce a parametric model of the mitral valve geometry, whose motion is described in the asymptotic limit under the assumption that it moves with the flow, without any additional resistance other than that given by its shape, and without the need to specify its material properties. The mitral valve model is coupled with a simple description of the left ventricle geometry, and their dynamics is solved numerically together with the equations ruling the blood flow. The intra-ventricular flow is analyzed in its relationship with the valvular motion. It is found that the initial valve opening anticipates the peak velocity of the Early filling wave with little influence of the specific geometry; while subsequent closure and re-opening are more dependent on the intraventricular vortex dynamics and thus on the leaflets’ geometry itself. The limitations and potential appli-cations of the proposed model are discussed.

Keywords—Left ventricle, Mitral valve, Vortex formation, Cardiovascular fluid dynamics.

INTRODUCTION

Blood motion in the left ventricle (LV) is the result of the interaction between fluid dynamical forces and the surrounding tissue. The overall mechanical phe-nomena are part of a complex physiological system that involves numerous processes, such as electrical activation and regulation of pressure distribution in

the circulation. However, when focusing on local phenomena, LV fluid dynamics can be described in relation with endocardial motion and valvular inter-action. Clinical, numerical, and experimental studies have described in details the main characteristics of the healthy intraventricular flow, which features the development of vortices from the mitral orifice that smoothly drive the redirection of the incoming dia-stolic jet towards the aortic track.9,10,16 It is believed that such vortices play a central role in the overall synchronicity of the beating heart,25and the quantifi-cation of the vorticity patterns has driven attempts to improve understanding of cardiac diseases.12,14

The natural and apparently harmonious behavior of the flowing blood modifies in presence of a dysfunc-tional mitral valve (MV) or by the insertion of pros-thetic ones.21,28In fact, the vortex formation process is intrinsically connected to the dynamics of the mitral leaflets while they interact with the blood crossing the valve during diastole. However, modeling the dynam-ics of a natural MV still represents a challenging issue. Notwithstanding the general agreement about its shape and function, a satisfactory description of the mitral dynamics during its interaction with the incoming flow is not available, yet. The detailed recording of the valve geometry during its motion is still not feasible in the clinical routine, substantially preventing a rigorous mathematical description and the possible numerical implementation at an individual level. The rheological properties of the tissue (say, a realistic model of the stress/strain relationship), neither homogeneous nor isotropic, as well as its active or passive behavior, are known only approximately. Medical imaging can provide some indications about the geometry of the valve during few instants of its motion, within the limitation given by the available space/time resolution, but cannot provide information on mechanical prop-erties.

Address correspondence to Federico Domenichini, Dipartimento di Ingegneria Civile e Ambientale, Universita` di Firenze, Via S. Marta 3, 50139 Florence, Italy. Electronic mail: federico.domenichini@ unifi.it

Cardiovascular Engineering and Technology, Vol. 6, No. 2, June 2015 (2014)pp. 95–104 DOI: 10.1007/s13239-014-0201-y

1869-408X/15/0600-0095/02014 Biomedical Engineering Society

95

(4)

In previous works, the influence of the MV on the LV flow was accounted imposing kinematic boundary conditions at the inlet orifice.3,9,18,19,26 Results were able to reproduce the main features of the intraven-tricular flow, although the fluid–tissue interaction (FTI) was not accounted. Following the same approach, the outcome of a MV replacement proce-dure on the LV flow was mimicked by varying the inclinations of the trans-mitral velocity profile. Results showed evidence of a reversal in the main circulatory diastolic flow structure28,32 previously observed in clinical investigation.11 In the same way, within an axisymmetric approximation, a previous study ana-lyzed the influence of valvular opening by changing the inlet velocity profile, suggesting that the valve motion influences only the early accelerating phase.1Recently, a comparison between the trans-valvular flow as given by 3D Color Doppler measurements and numerical simulations was performed assuming fixed circular or rectangular shapes of the mitral orifice, still avoiding to model explicitly the presence of the MV.30Actually, all these models did not study the valve dynamics but tried to simulate its presence in terms of inlet velocity distribution to reproduce the flow in the LV.

Differently, a relevant effort was spent during the past years to investigate the influence of prosthetic valves on blood flow, because of the relevance of the topic, and given that such valves can be described accurately both in terms of geometry and mechanical properties. A greater number of works was devoted to the study of the aortic valve replacement2,5,6,17than on the MV. Early numerical work analyzed the pressure distribution close to prosthetic leaflets inserted between simplified cylindrical atrium and ventricle, in order to detect the possible presence of cavitation, which can be related to platelets damage.3 The same issue was investigated in a laboratory model of the LV.31 More recently, under the 2D approximation, a numerical patient-specific approach based on Mag-netic Resonance Imaging (MRI) data of the LV geometry and assuming rigid aortic and mitral valves has been developed, showing that the mitral leaflets opening dynamics influences the development of the intraventricular flow.33

Preliminary analysis on FTI in the natural MV was presented in an early version of the Immersed Boundary Method.29This and more recent works, that use data from MRI, studied the interaction between blood and MV modelled as an elastic body immersed in the fluid computational domain.13,20These probably represent the most accurate FTI solution still available, although the necessary approximations and the limited information available on MV properties do not ensure a greater accuracy of this approach with respect to carefully designed inlet boundary conditions.

Following what briefly recalled above, it should be clear that the complete fluid–tissue coupled dynamics is still an open issue, given the variability of geomet-rical and structural parameters, and the scarcity of information about the MV.

We propose here an alternative modeling that can partially reduce, when appropriate, the difficulty of retrieving punctual information on the MV. In this approach, the MV is provided with a given geometry, whose motion is described by few degrees of freedom; in the present example just one degree of freedom, that represents the degree of valvular opening. The FTI problem thus simplifies as it couples just one, or few, degree of freedom on the structural side. It is faced here, in an asymptotic limit, under the assumption that the valve moves with the flow, without any resistance other than that given by its shape, thus without the need to specify its material properties.

The intraventricular fluid dynamics and the vortex formation are analyzed in their relationship with the MV motion. Differences are evaluated with respect to the geometrical parameters, and quantified in terms of integral quantities.

METHODS

This work is focused on the MV, and the geometry of the model LV is taken as that of a simple prolate spheroid following previous works,9,27with the aortic outflow accounted with a simple open/closed behav-ior.8Attention is here posed on the definition of MV geometry and function.

Mathematical Definition of the Mitral Valve Geometry Let us consider that the mitral annulus has a cir-cular shape, with radius R, and that it lays on the horizontal plane at z = 0 of a Cartesian [x, y, z] (or cylindrical [r, h, z]) system of coordinates, centered in [0, 0, 0]. On the main longitudinal section, say y = 0, that crosses both the mitral and aortic orifices, the anterior and posterior leaflets are at h = 0 and h = p, with lengths (1 + e)R and (1  e)R, respectively. The dimensionless parameter e represents the asymmetry of the leaflets (Fig.1a). The valuee = 0 corresponds to a not-physiological valve (symmetric with respect to the x-axis in the mathematical model), with leaflets of equal lengths. Increasing e, the anterior leaflet pro-gressively becomes longer than the posterior one, until the unrealistic model value e = 1, that in principle corresponds to a single leaflet valve. Physiological values of e can be estimated in the range 0.25–0.45. On the orthogonal longitudinal plane x = 0 (h = p/2 and h = 3p/2), the valve is symmetric with respect to the

F. DOMENICHINI ANDG. PEDRIZZETTI 96

(5)

y-axis, being kR long; the dimensionless parameter k defines the ellipticity of the valvular edge (Fig.1b). During the motion, the above defined four points on the tip of the mitral valve are assumed to follow cir-cular trajectories described by a single variable, the opening angleu(t).

From the positions of these points, for any value of the angleu, the entire profile of the valvular edge xe(h)

is built in a simple way using sinusoidal functions of the circumferential parametric coordinate h:

xeð Þ ¼ R cos h 1  cos uh ð Þ  eR cos u

yeð Þ ¼ R sin h 1  k cos uh ð Þ zeð Þ ¼ h 1þ k 2 þ e cos h þ 1 k 2 cos 2h   Rsin u ð1Þ Therefore, for any value of the opening angle u(t), the horizontal [x, y] projection of the mitral edge is an ellipse centered in [eRcos u, 0], and with semi-axes [R(1 cos u), R(1  k cos u)].

The complete shape of the valve xv(h, s) is then

defined introducing a second parametric coordinate s, that spans from the annulus, s = 0, where xv(h,

0) = [Rcosh, Rsinh, 0], to the tip of the mitral edge, s = 1 where xv(h, 1) = xe(h). The entire mitral shape is

constructed by the regularity condition that the valve

connects to the mitral annulus with horizontal tangent, that is ¶zv/¶s = 0 at s = 0. The simplest option, in

absence of further constraints, is that the xv and yv

coordinates have a linear dependence on s, and zv a

quadratic one.

In synthesis, the mathematical shape of this MV model can be written in Cartesian coordinates

xvðh; sÞ ¼ R cos h 1  s cos uð Þ  eRs cos u

yvðh; sÞ ¼ R sin h 1  sk cos uð Þ zvðh; sÞ ¼ s2 1þ k 2 þ e cos h þ 1 k 2 cos 2h   Rsin u ð2Þ This artificial MV shape, one example shown in Fig. 1c, was intended to provide a very simple repre-sentation in analytical terms and to test the perfor-mance of the FTI procedure. We are aware of some discrepancies with respect to real valves, like the flat annulus, that could be easily included on top of this representation, or different other ones. The mathe-matical expression (Eq. (2)) is also useful because it allows the analytic calculation of the fundamental integral quantities describing the valvular motion that will be used in the dynamics later on. If the valve shape was known only numerically, the same quantities could be computed similarly by numerical integration.

From Eq. (2), the mitral valve velocity vv can be

obtained by time derivative

vvxðh; sÞ ¼ Rsðcos h þ eÞ sin u

du dt vvyðh; sÞ ¼ R sin hsk sin u du dt vvzðh; sÞ ¼ Rs2 1þ k 2 þ e cos h þ 1 k 2 cos 2h   cos udu dt ð3Þ It is useful to define the following scalar time dependent quantity

IðtÞ ¼ Z

Smv

ðvv nÞdSmv; ð4Þ

the integral of the valvular velocityvvorthogonal to its

surface Smv, that is the rate of volume spanned by the

moving valve. The vector n is the unitary normal to the valve surface. The integral in Eq. (4) can be rewritten in terms of the time-invariant system of coordinates (s, h) giving, after tedious but straightforward manipula-tions, IðtÞ ¼ I0 du dt ¼ ðIxþ Iyþ IzÞ du dt ; ð5Þ

where the functionIk(t) can be written, being the time

dependence given by u( t), FIGURE 1. Sketch of the valve definition on the y 5 0 plane

(a), the x 5 0 plane (b), three-dimensional geometry (c). Graphs refer to e 5 0.25; k 5 0.6, u¼ p=4. Top view of the model (d). MV mitral valve, AO aortic orifice, d is the eccen-tricity parameter.

Mitral Valve Model 97

(6)

IxðtÞ ¼ pR3 2 3e 2þ1 2þ k 6 k 4ð3e 2þ 2Þ cos u   sin2u; ð6aÞ IyðtÞ ¼ pR3 1þ 3k 6 kþ k 4ðe 2 2kÞ cos u   sin2u; ð6bÞ IzðtÞ ¼ pR3 c1cos2u c2cos3u   : ð6cÞ

The coefficients in Eq. (6c) arec1¼ e2=3þ 1  kð Þ2=

12þ 1 þ kð Þ2=6 and c2¼ 1 þ k þ e 2k=4. It must be

noticed that when the valve is closed, u¼ 0, Ix= Iy= 0, while Iz¼ pR3ðc1 c2Þ. The quantities

defined in Eqs. (3)–(6) will be used below. Left Ventricle Geometry

The mathematical definition of the model LV geometry strictly follows that reported in previous works.9,27The endocardial boundary is assumed to be half of a prolate spheroid, whose geometry is defined by two functions of time, that of the equatorial plane diameter D(t) and that of its major semi-axis H(t). The moving endocardial surface represents the inner kinematic forcing condition during the numerical integration, as specified below. Data about H(t) and D(t) were collected from a young healthy subject. Just for reference, the (minimum) tele-systolic D0diameter

is about 3.5 cm and the tele-diastolic one about 5 cm; the major semi-axis length ranges from 5.4 to 6.4 cm. The main indicator of the ventricular function, the Ejection Fraction (EF) is 56%. Given the functions D(t) and H(t), the ventricular volume VLV can be

written as VLV ¼ pHD2=6, whose time derivative gives

the flow rate QLV entering/exiting the ventricular

chamber.

An additional geometrical parameter, d, provides the eccentricity between the centers of the mitral plane and mitral annulus (Fig.1d). During the systolic phase (QLV< 0) the aortic track is simulated with a circular

opening at z = 0, centered at y = 0, spanning along x from the mitral annulus (x = R) to the ventricle wall x = D(t)/2 + dD0(Fig. 1d).

Just for reference, the problem can be made dimensionless assuming as time-scale the heart-beat periodT, as length-scale the end-systolic diameter D0

of the mitral plane, and as velocity scale the maximum of the averaged vertical velocity across the mitral annulus during the diastole Umax= QLVmax/(pR

2

).9 From the dimensional values, accounting that T is around 1 s, we have a Stokes number b = D02/(mT) of

about 370, and a Strohual number St = D0/(UmaxT)

around 0.08.

Mathematical Definition of the Mitral Valve Motion The basic idea is that the mitral valve motion is almost completely driven by the blood dynamics. Thus, we assume that the valve ‘‘moves with the flow’’. In other words, assuming as negligible the difference between the densities of blood and tissue, the leaflets are regarded as infinitesimally thin buoyant elements moving inside the surrounding fluid that are con-strained to the geometrical shapes, with one degree of freedom, defined in Eq. (2).

At every instant the motion of the valve is driven by imposing the conservation of mass at an integral level, thus ensuring that the rate of volume of fluid crossing the valve position

IfðtÞ ¼

Z

Smv

ðv  nÞdSmv; ð7Þ

where v is the local blood velocity, is equal to that allowed by valve motion, Eq. (4). With the aid of the integrals defined in Eqs. (5) and (7), the temporal variation of the angle u(t) must obey the following differential equation

I0ðtÞ

du

dtðtÞ ¼ IfðtÞ: ð8Þ

Mathematical Definition of the Fluid Problem and Numerical Method

Numerical simulations of the ventricular blood flow, assumed a Newtonian incompressible fluid, were performed solving the Navier–Stokes and continuity equations @v @tþ vrv ¼  rp q þ mr 2v; ð9Þ r  v ¼ 0

where t, v, p, q and m are the time, velocity, pressure, density (1060 kg/m3) and kinematic viscosity (3.3 9 106m2/s), respectively.

The above system of equations was solved in a computational bi-periodic box with a standard frac-tional step method using a mixed spectral-finite dif-ferences scheme, with the LV wall and the MV immersed therein. The Navier–Stokes equation is dis-cretized on a 3D staggered grid using second order centered finite differences and advanced in time with a third order Runge–Kutta method. The mass conser-vation is then satisfied solving the Poisson’s equation for the instantaneous irrotational correction of the pressure field. This second step, given the assumption of periodicity of the numerical box along the xand y directions, is performed in the Fourier space, allowing

F. DOMENICHINI ANDG. PEDRIZZETTI 98

(7)

fast solutions. At the boundaries along the z direction, free slip conditions are enforced for the tangential (x and y) components of the velocity. For the z (normal) component of the velocity, a null value is imposed at the top of the computational box, and a free flow is imposed at the bottom side. These conditions are congruently implemented for the irrotational part of the flow, resulting in a Neumann and a Dirichlet conditions for the irrotational correction and for the pressure field itself, at the top and bottom of the domain, respectively.7 Different combinations of the boundary conditions along the z direction on the ver-tical component of the velocity did not lead to appre-ciable differences in the solutions.

The forcing term of the system is represented by the imposed motion of the LV wall. Its motion, that of the mitral valve, and the presence of the closed part of the mitral annulus are accounted using an Immersed Boundary Method. The numerical time-marching algorithm is as follows. At a generic instant t, the position and velocity of the LV wall are given by the imposed motion of the endocardium model, that is the forcing of the system, and the geometry and velocity of the valve, as well as the fluid velocities, are known. Then the flow and the MV leaflets are advanced in time by the Navier–Stokes Equation (Eq. (9)) and the actual value of du=dt, respectively, with the velocity boundary conditions on all the walls, including the MV leaflets. The incompressibility constraint is then enforced adding a proper irrotational correction (Fractional Step Method) and the value of du=dt is also updated by Eq. (8). To this aim, the integralIfis

evaluated numerically. This is performed transforming the integral in Eq. (7) in the Lagrangian system of coordinates (h, s), where the domain of integration is time-independent. Integral (Eq. (7)) can be rewritten, omitting the subscript mv for brevity,

Ifð Þ ¼t Z S vinidS¼ Z S vidSi¼ Z 1 0 ds Z 2p 0 viJidh;

where dSi are the Cartesian projections of the MV

surface, and Jiare the determinants of the matrices of

transformations between the Cartesian and parametric definition of the valve surface, that can be easily derived from Eq. (2). The fluid velocity in Eq. (7) is evaluated interpolating the values calculated in the computational points close to the MV surface, and the integration is then performed numerically.

It is easy to show that att = 0, Eq. (8) tends to the

analytical condition

du=dtðt ¼ 0Þ ¼ QLVðt ¼ 0Þ= cð 1 c2Þ=pR3. During the

flow evolution, the angle u tð Þ is left to evolve satisfying Eq. (8), with the constraint, for physical reasons, to remain between the maximum opening u¼ p=2, and

the complete closure u¼ 0. Particular attention is posed in solving the flow through the mitral orifice, apart for the first instant of computation at t = 0, when the blood velocity profile must be equal to that of the valve (Eq. (3)). When the valve is open, the orifice corresponds to points internal to the computational domain, and the velocity is left free to modify satisfy-ing Eq. (9). When the valve is closed, a zero value is imposed to the velocity, accordingly to the IBM pro-cedure. The same method is used for the aortic orifice. Several preliminary runs have been performed to ensure the independence of the solution on the numer-ical parameters, such as the dimensions of the compu-tational box and the numerical details. The results presented in what follows have been obtained assuming

Lx = Ly= 6 cm, and Lz= 8 cm, the dimensional

lengths of the computational box with a number of computational grid points along the three directions Nx= Ny= Nz= 128. The time step has been fixed to

satisfy the stability conditions; a typical value is T/2048, being T the heartbeat period. Additional numerical parameters are those required for the spatial discreti-zation of the moving inner boundaries. The circumfer-ential number of Lagrangian points describing both the LV wall and the valve has been fixed as Nh= 384. Along

the longitudinal direction, a number Ns= 192 of points

has been used; in this case, the subscript s stands for the apex-mitral description of the LV wall and for the annulus-edge discretization of the mitral valve geome-try. Further numerical details are out of the scope of the present study and were reported in previous works.8,22–24

RESULTS

In this study we maintained constant the LV geo-metrical parameters, including the radius of the mitral annulus R = D0/3 = 1.2 cm, and the eccentricity

d = 0.15D0= 0.54 cm. In the MV geometry, the

ellipticity parameter was kept fixed to a realistic value k = 1/3, and we studied the effect of the leaflet asym-metry by varying the parameter e between e = 0, cor-responding to symmetric leaflets, to a maximum value set to e = 0.6.

Figures2 and 3 show exemplary results, corre-sponding to e = 0.2, in terms of distribution of the normal vorticity and the in-plane velocities on the longitudinal (symmetry) plane y = 0 and of the three-dimensional vorticity structure described by the second invariant k2 of the velocity gradient,15 respectively.

During the early diastolic phase, the mitral valve opens very rapidly offering, as by it moves with the fluid, a little resistance to the accelerated flow. It is worth mentioning that, in the very early phase, the valve

Mitral Valve Model 99

(8)

moves slightly more rapidly than the nearby flow and exhibits a brief birth of small opposite vorticity patches at the tip of the leaflets’ edges, a phenomenon previ-ously noticed in the case of 2D rigid leaflets.27 At the end of the opening phase the valve reaches its maxi-mum angle before the peak E-velocity. The rapid valve deceleration, coupled with the high flow velocity, leads to the separation of the boundary layer and the for-mation of the well-known LV vorticity structure downstream to the valve that deeply penetrates the ventricular chamber,Figs. 2a and3a att = 0.25T, end of the E-wave. The typical vorticity pattern, with a dominant counter-clockwise vortex from the anterior leaflet and a weaker clockwise one, Fig.2a, corre-sponds to a deformed vortex ring9which also presents longitudinal streaks in correspondence of the leaflets’ tips (Fig. 3a). This growing vortex motion reflects in an upstream fluid dynamical thrust on the valve that gets to its partial closure during the final phase of the E-wave. During diastasis, with a slow inflow, the valve presents a nearly constant degree of opening while the main vorticity structure dissipates for viscous effects, Figs.2b and3b att = 0.375T, end of the diastasis. The second diastolic jet during the A-wave produces an-other ring-like vortex that is not able to reach the apical region, Figs.2c and 3c att = 0.5T, end of the A-wave, that is accompanied by the second complete opening of the valve. The early systolic phase shows how the vorticity helps in redirecting the flow towards the aortic orifice,27 while the almost irrotational sys-tolic jet begins to wash-up the LV,Figs. 2d and3d at

t = 0.5625T, early systole. It can be observed that the mitral valve accelerates its backward motion with flow inversion and reaches the complete closure shortly after the beginning of systole.

Flow evolution and valve dynamics are not quali-tatively altered when varying the asymmetry parameter

e. Quantitative differences can be depicted in Fig. 4, where the temporal evolution of the opening angleu is reported for ten different values of e. Two principal phenomena are common to all the cases. First, the complete initial opening of the valve anticipates the peak value of the E-wave with very small differences between the various cases. This can be imputable to the extremely rapid motion of the valve that is substan-tially driven by mass conservation and orifice size, which does not depend on e, and weakly affected by the effective profile of the entering velocity.1The fol-lowing evolution is much more affected by the valve geometry, as it depends on the asymmetry of the LV vorticity that interacts with leaflets having different lengths. Differences in the valvular position and mo-tion immediately reflect on the creamo-tion of vorticity that, in turn, immediately affects the fluid dynamical action on the valve itself. This is demonstrated by the

timing of the second complete opening of the valve, related to the diastolic A-wave. Apart from the ex-treme values e = 0 and e = 0.6, the other cases do not show any trace of regular dependence on the asym-metry parameter, even if the differences in the values of this second timing are around the three per cent of the heartbeat period T. In any case, the valve motion does not appear to be synchronous with the diastolic inflow. This becomes further clear looking at the closing phase of the valve. In all the cases analyzed the total closure of the valve is delayed with respect to the beginning of the systole. It is worthwhile to notice that the cases characterized by intermediate values of e show an al-most common behavior.

The results reported in Fig.4 are synthesized in Fig. 5, where the instants of the first and second complete opening and that of complete closure of the valve are reported in function ofe, and with respect to the temporal evolution of the flow rate QLV (right

panel). The first complete opening of the mitral valve always anticipates the E-wave peak, while the complete closure of the valve is always delayed with respect to the beginning of the systolic contraction, giving a substantially e—independent timing of this phenome-non, with the possible exception of the limiting non-physiological values e = 0 and e = 0.6. The timing of the second opening does not show any sign of regular dependence on the leaflets’ asymmetry, being the intermediate coupled valve-flow dynamics less con-strained by the global mass conservation balance dri-ven by the imposed motion of the dri-ventricular wall, and therefore more influenced by the local details of the solution.

An additional integral parameter is the Vortex Formation Time (VFT).12The VFT is here computed as4 VFTðtÞ ¼ Z t 0 VfzmðsÞ  VezmðsÞ 2RmeðsÞ ds: ð10Þ

In Eq. (10), Rme is the instantaneous geometrical

mean radius of curvature of the elliptic z-projection of the valvular edge, having area Ame¼ pR2

1 cos u

ð Þ 1  k cos uð Þ and the numerator is the rel-ative fluid velocity: Vezmis the mean value (along h) of

the z-component of the velocity at the tip of the valve vvz(h, s = 1), Vezm¼ 0:5R 1 þ kð Þ cos udu=dt, and

Vfzmis the mean vertical velocity of the fluid. Given the

non-planar shape of the valve, and considering that Vfzm must represent an average value of the fluid

velocity crossing the valvular edges, it is evaluated numerically considering the cylindrical column of the fluid domain included between the minimum and maximum value of ze, and spanning horizontally on

the area Ame. In Eq. (10) we assume thatt = 0 corre-F. DOMENICHINI ANDG. PEDRIZZETTI

100

(9)

sponds to the beginning of the diastole. Results are reported inFig. 6. Computations have been performed both for the E-wave only and for the whole diastole. In both cases, higher values of the asymmetry parameter are associated to lower VFT. This can be imputable to the different arrangement of the flow field that, at higher values ofe and obeying the same integral bal-ance, presents little lower values of the vertical com-ponent of the velocity used in Eq. (10) due to a major horizontal deviation of the flow. In any case, the re-sults are in complete agreement with those reported in the literature for healthy subjects,12 showing little higher values in correspondence of intermediate values of e in the range 0.1–0.2. The VFT computed on the basis of the diastolic flow rate and the fixed geometry of the mitral annulus are also reported for reference, showing how the valvular motion must be taken into account in properly estimating the VFT.4

DISCUSSION

A one degree of freedom, parametric model of the mitral valve geometry has been introduced. Based on this, the flow-valve interaction is expressed in integral terms based on the conservation of mass without the need of defining the material characteristic of the MV tissue.

To put it differently, a complete FTI problem suffers from the limited knowledge and not measurability of the mechanical properties of the tissue; this ignorance is overcome here by assuming the knowledge of the valve geometry and implicitly assuming that unknown mechanical stresses are those that maintain such a valvular shape.

Through this model, it was possible to draw some initial conclusions about the flow-driven dynamics of the MV. Details of the flow field do depend on the

FIGURE 2. Flow fields on the vertical plane y 5 0 in the case e 5 0.2 and d/D050.15. Distribution of the normal y-component of

the vorticity field (color coded) and in-plane velocity vectors. From (a) to (d): t/T 5 [0.25 0.375 0.5 0.5625]. Vorticity values are in the range 2180 s21(blue) and +180 s21(red). The reference velocity vector at the bottom of Fig.2a corresponds to 50 cm/s.

FIGURE 3. Three-dimensional flow structure the case e 5 0.2 and d/D050.15, visualized by the isosurfaces of the second

invariant of the velocity gradient,15k

25 21000 s22. From (a) to (d): t/T 5 [0.25 0.375 0.5 0.5625]. Dark surfaces correspond to the

mitral valve during the diastole, and to the aortic orifice during the systolic phase.

Mitral Valve Model 101

(10)

valve geometry; however, its main features are not al-tered by only differences in leaflets’ length. It is found that the valve opening is not in phase with the imposed diastolic flow, this was expected because the valve is opened by transvalvular pressure gradient which is a result of gradient of kinetic energy (in phase with velocity) and acceleration (in quadrature with it). The initial opening is a pure mass driven phenomenon and it appears to be weakly dependent on the valve asym-metry; partial closure on diastasis and second opening depend on the characteristics of the intraventricular flow and therefore are more influenced by the MV geometry. The valve closure is also a rapid phenome-non that completes after the beginning of systole, therefore a little physiological regurgitation appears physically unavoidable during MV closure. Within the limitation of the model, it must be outlined that such a

delay of the MV closure is about the 3% of the heartbeat period, with the exception of the extreme not physiological cases e = 0 and e = 0.6.

This study was limited to a simple model of the LV in normal conditions and showed that leaflets’ asymmetry, maintaining the same orifice size, does not alter signifi-cantly the global properties of the intraventricular flow. It however demonstrated the feasibility of this integral approach for FTI opening a different perspective to the study of the real cardiac valves dynamics.

Several simplifications have been adopted in devel-oping the model; among them, the assumption of a single degree of freedom of the valvular motion, and the shape of the valve itself with a fixed and circular mitral annulus. It must pointed out that any of these can be easily removed. In principle, any other natural or dys-functional shape may be used. Eventually requiring a numerically based evaluation of the integral properties. The model presented here can be extended to two (one per leaflet) or more degrees of freedom, and accordingly introduce sub-domains of integration for the mass balance. The same balance Eq. (8) can be extended with the inclusion of additional contributions accounting for inertial or viscoelastic properties that may be used to further characterize the valvular mechanical behavior in global terms. This approach represents a sort of re-verse-engineering problem where the inaccessible mechanical properties of the MV are reconstructed from the observation of the functional and geometric information.

ACKNOWLEDGMENTS

This work has been supported by MIUR (Italian Ministry of University and Research) under the Grant PRIN 2012HMR7CF.

FIGURE 4. Top panel: Temporal evolution of the opening angle u for d/D050.15 varying the asymmetry parameter e.

Bottom panel: temporal evolution of the entering/exiting flow rate QLV, cm3/s. Markers correspond to the four instants

reported in Figs.2and3. Positive QLV: diastole; negative one:

systole.

FIGURE 5. Left panel: Instants of the first (blue) and second (green) complete opening, and that of complete closure of the valve (red) as a function of the asymmetry parameter e, d/D050.15. Right panel: temporal evolution of the entering/

exiting flow rate QLV, cm3/s. Onset of systole is illustrated by

the horizontal dashed line.

FIGURE 6. Vortex formation time (VFT) as a function of the asymmetry parameter e. Red continuous line (and markers), blue continuous line (and markers): VFT computed as defined in Eq. (10) for the E-wave only and for the complete diastole, respectively. Dashed lines correspond to the VFT for the same periods computed on the basis of the flow rate QLV, and the

fixed area of the mitral annulus.

F. DOMENICHINI ANDG. PEDRIZZETTI 102

(11)

STATEMENT OF HUMAN STUDIES Authors declare that human subjects were not involved in this study.

STATEMENT OF ANIMAL STUDIES Authors declare that animals were not involved in this study.

CONFLICT OF INTEREST

Authors declare that there are no conflicts of interest to disclose.

REFERENCES

1Baccani, B., F. Domenichini, and G. Pedrizzetti. Model

and influence of mitral valve opening during the left ven-tricular filling. J. Biomech. 36(3):355–361, 2003.

2

Borazjani, I., L. Ge, and F. Sotiropoulos. Curvilinear immersed boundary method for simulating fluid structure interaction with complex 3D rigid bodies. J. Comput. Phys. 227:7587–7620, 2008.

3

Cheng, R., Y. G. Lai, and K. B. Chandran. Three-dimensional fluid–structure interaction simulation of bileaflet mechanical heart valve flow dynamics. Ann. Bio-med. Eng.32(11):1471–1483, 2004.

4

Dabiri, J. O. Optimal vortex formation as a unifying principle in biological propulsion. Annu. Rev. Fluid Mech. 41:17–33, 2009.

5

de Tullio, M. D., A. Cristallo, E. Balaras, and R. Verzicco. Direct numerical simulation of the pulsatile low trough an aortic bileaflet mechanical heart valve. J. Fluid Mech. 622:259–290, 2009.

6de Tullio, M. D., G. Pedrizzetti, and R. Verzicco. On the

effect of aortic root geometry on the coronary entry-flow after a bileaflet mechanical heart valve implant: a numer-ical study. Acta Mech. 216:147–163, 2011.

7

Domenichini, F. Three-dimensional impulsive vortex for-mation from slender orifices. J. Fluid Mech. 666:506–520, 2011.

8

Domenichini, F., and G. Pedrizzetti. Intraventricular vor-tex flow changes in the infarcted left ventricle: numerical results in an idealised 3D shape. Comput. Methods Bio-mech. Biomed. Eng.14:95–101, 2011.

9

Domenichini, F., G. Pedrizzetti, and B. Baccani. Three-dimensional filling flow into a model left ventricle. J. Fluid Mech.539:179–198, 2005.

10

Domenichini, F., G. Querzoli, A. Cenedese, and G. Pe-drizzetti. Combined experimental and numerical analysis of the flow structure into the left ventricle. J. Biomech. 40(9):1988–1994, 2007.

11Faludi, R., M. Szulik, J. D’hooge, P. Herijgers, F.

Rade-makers, G. Pedrizzetti, and J.-U. Voigt. Left ventricular flow patterns in healthy subjects and patients with pros-thetic mitral valves: an in vivo study using echocardio-graphic particle image velocimetry. J. Thorac. Cardiovasc. Surg.139:1501–1510, 2010.

12

Gharib, M., E. Rambod, A. Kheradvar, D. J. Sahn, and J. O. Dabiri. Optimal vortex formation as an index of cardiac health. Proc. Natl. Acad. Sci. U.S.A. 109:6205–6308, 2006.

13

Griffith, B. E., X. Luo, D. M. mcQueen, and C. S. Peskin. Simulating the fluid dynamics of natural and prosthetic heart valves using the immersed boundary method. Int. J. Appl. Mech.1(1):137–177, 2009.

14

Hong, G. R., G. Pedrizzetti, G. Tonti, P. Li, P. Wei, J. K. Kim, A. Bawela, S. Liu, N. Chung, H. Houle, J. Narula, and M. A. Vannan. Characterization and quantification of vortex flow in the human left ventricle by contrast echo-cardiography using vector particle image velocimetry. J. Am. Coll. Cardiol. Imaging1:705–717, 2008.

15Jeong, J., and F. Hussain. On the identification of a vortex.

J. Fluid Mech.285:69–94, 1995.

16Kilner, P. J., G. Z. Yang, A. J. Wilkes, R. H. Mohiaddin,

D. N. Firmin, and M. H. Yacoub. Asymmetric redirection of flow through the heart. Nature 404:759–761, 2000.

17

Le, T. B., and F. Sotiropoulos. Fluid-structure interaction of an aortic heart valve prosthesis driven by an ani-mated anatomic left ventricle. J. Comput. Phys. 244:41–62, 2013.

18

Lemmon, J. D., and A. P. Yoganathan. Three-Dimensional computational model of left heart diastolic function with fluid–structure interaction. J. Biomech. Eng. 122:109–117, 2000.

19

Lemmon, J. D., and A. P. Yoganathan. Computational modeling of left heart diastolic function: examination of ventricular dysfunction. J. Biomech. Eng. 122:297–303, 2000.

20

Ma, X., H. Gao, B. E. Griffith, C. Berry, and X. Luo. Image-based fluid–structure interaction model of the human mitral valve. Comput. Fluids 71:417–425, 2013.

21Machler, H., M. Perthel, G. Reiter, U. Reiter, M. Zink, P.

Bergmann, A. Waltensdorfer, and J. Laas. Influence of bileaflet prosthetic mitral valve orientation on left ventric-ular flow—an experimental in vivo magnetic resonance imaging study. Eur. J. Cardiothorac. Surg. 26:747–753, 2004.

22

Mangual, J. O., A. De Luca, E. Kraigher-Krainer, L. Toncelli, A. Shah, S. Solomon, G. Galanti, F. Domeni-chini, and G. Pedrizzetti. Comparative numerical study on left ventricular fluid dynamics after dilated cardiomyopa-thy. J. Biomech. 46:1611–1617, 2013.

23

Mangual, J. O., F. Domenichini, and G. Pedrizzetti. Describing the highly three dimensional right ventricle flow. Ann. Biomed. Eng. 40(8):1790–1801, 2012.

24

Mangual, J. O., F. Domenichini, and G. Pedrizzetti. Three dimensional numerical assessment of the right ventricular flow using 4D echocardiography boundary data. Eur. J. Mech. B/Fluids25:25–30, 2012.

25Narula, J., M. A. Vannan, and A. N. De Maria. Of that

waltz in my heart. J. Am. Coll. Cardiol. 49:917–920, 2007.

26

Pedrizzetti, G., and F. Domenichini. Nature optimizes the swirling flow in the human left ventricle. Phys. Rev. Lett. 95:108101, 2005.

27

Pedrizzetti, G., and F. Domenichini. Asymmetric opening of a simple bi-leaflet valve. Phys. Rev. Lett. 98:214503, 2007.

28

Pedrizzetti, G., F. Domenichini, and G. Tonti. On the left ventricular vortex reversal after mitral valve replacement. Ann. Biomed. Eng.38(3):769–773, 2010.

29

Peskin, C. S. Flow patterns around heart valves: a numerical method. J. Comput. Phys. 10:252–271, 1972.

30

Quaini, A., S. Canic, G. Guidoboni, R. Glowinski, S. R. Igo, C. J. Hartley, W. A. Zoghbi, and S. H. Little. A three-dimensional computational fluid dynamics model of regurgitant mitral valve flow: validation against in vitro Mitral Valve Model 103

(12)

standards and 3D Color Doppler methods. Cardiovasc. Eng. Technol.2(2):77–89, 2011.

31

Querzoli, G., S. Fortini, and A. Cenedese. Effect of the prosthetic mitral valve on the vortex dynamics and turbu-lence of the left ventricular flow. Phys. Fluids 22(4):1–10, 2010.

32

Seo, J. H., and R. Mittal. Effect of diastolic flow patterns on the function of the left ventricle. Phys. Fluids 25:110801, 2013.

33

Su, B., L. Zhong, X.-K. Wang, J.-M. Zhang, R. S. Tan, J. C. Allen, S. K. Tan, S. Kim, and H. L. Leo. Numerical simulation of patient-specific left ventricular model with both mitral and aortic valves by FSI approach. Comput. Methods Programs Biomed.113:474–482, 2014.

F. DOMENICHINI ANDG. PEDRIZZETTI 104

Riferimenti

Documenti correlati

In the first specification, we have asymmetric conformism: while in [4] we have a symmetric strategic complementarities towards both possible choices, in our setting we have

Da parte loro, i ragazzi continuano ad avere fiducia nella scienza, ma sono anche convinti che “la scienza moderna non può accentrarsi solo sul modello descrittivo della realtà,

Anzitutto, per effetto di una combinazio- ne di pregiudizi plurisecolari - quello sociale che distingue la carne come cibo per i ricchi e quello medico-umorale in base a cui

In this case the main characterizing properties of a time crystal are robust to perturbations, namely, (i) the presence of Floquet states with long-range correlations, (ii)

Among the NEG alloys, the alloy made from titanium (Ti), zirconium (Zr) and vanadium (V) has the lowest activation temperature at 180 degrees [11]. The surface impedance of

Another very important advantage related to the use of the Parametric Approach in Urban Design is the adaptability of generated models to changes in urban data, either referred to

Per i tumori che coinvolgono lo scheletro assiale come la colonna e il bacino, può essere utilizzata la tecnica di resezione conservativa, ma la prognosi è

Finally, a transfer print method, which consists of transferring RS and RS/yeast film layers onto a self-adherent paraffin substrate, was used for the realization of