Skip to main content
Log in

Modelling and Subject-Specific Validation of the Heart-Arterial Tree System

  • Published:
Annals of Biomedical Engineering Aims and scope Submit manuscript

Abstract

A modeling approach integrated with a novel subject-specific characterization is here proposed for the assessment of hemodynamic values of the arterial tree. A 1D model is adopted to characterize large-to-medium arteries, while the left ventricle, aortic valve and distal micro-circulation sectors are described by lumped submodels. A new velocity profile and a new formulation of the non-linear viscoelastic constitutive relation suitable for the {Q, A} modeling are also proposed. The model is firstly verified semi-quantitatively against literature data. A simple but effective procedure for obtaining subject-specific model characterization from non-invasive measurements is then designed. A detailed subject-specific validation against in vivo measurements from a population of six healthy young men is also performed. Several key quantities of heart dynamics—mean ejected flow, ejection fraction, and left-ventricular end-diastolic, end-systolic and stroke volumes—and the pressure waveforms (at the central, radial, brachial, femoral, and posterior tibial sites) are compared with measured data. Mean errors around 5 and 8%, obtained for the heart and arterial quantities, respectively, testify the effectiveness of the model and its subject-specific characterization.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Figure 1
Figure 2
Figure 3
Figure 4

Similar content being viewed by others

References

  1. Alastruey, J., A. W. Khir, K. S. Matthys, P. Segers, S. J. Sherwin, P. R. Verdonck, K. Parker, and J. Peirò. Pulse wave propagation in a model human arterial network: Assessment of 1-D visco-elastic simulations against in vitro measurements. J. Biomech. 44(12):2250–2258, 2011.

    Article  PubMed Central  PubMed  Google Scholar 

  2. Avolio, A. P., Multi-branched model of the human arterial system. Med. Biol. Eng. Comput. 18(6):709–718, 1980.

    Article  CAS  PubMed  Google Scholar 

  3. Bessems, D., C. G. Giannopapa, M. C. M. Rutten, and F. N. van de Vosse. Experimental validation of a time-domain-based wave propagation model of blood flow in viscoelastic vessels. J. Biomech. 41(2):284–291, 2008.

    Article  PubMed  Google Scholar 

  4. Bessems, D., M. Rutten, and F. N. van de Vosse. A wave propagation model of blood flow in large vessels using an approximate velocity profile function. J. Fluid. Mech. 580:145–168, 2007.

    Article  Google Scholar 

  5. Blanco, P. J., R. A. Feijóo, and S. Urquiza. A unified variational approach for coupling 3D–1D models and its blood flow applications. Comput. Methods Appl. Mech. 196(41–44):4391–4410, 2007.

    Article  Google Scholar 

  6. Blanco, P.J., R.A. Feijóo. A dimensionally-heterogeneous closed-loop model for the cardiovascular system and its applications. Med. Eng. Phys. 35(5):652–667, 2013.

    Article  CAS  PubMed  Google Scholar 

  7. Blanco, P. J., P. R. Trenhago, L. G. Fernandes, and R. A. Feijóo. On the integration of the baroreflex control mechanism in a heterogeneous model of the cardiovascular system. Int. J. Numer. Methods Biomed. Eng., pp. 412–433, 2012.

  8. Bollache, E., N. Kachenoura, A. Redheuil, F. Frouin, E. Mousseaux, P. Recho, and D. Lucor. Descending aorta subject-specific one-dimensional model validated against in vivo data. J. Biomech. 47(2):424–431, 2014.

    Article  CAS  PubMed  Google Scholar 

  9. Canic, S., C. Hartley, D. Rosenstrauch, J. Tambaca, G. Guidoboni, and A. Mikelić. Blood flow in compliant arteries: an effective viscoelastic reduced model, numerics, and experimental validation. Ann. Biomed. Eng. 34(4):575–592, 2006.

    Article  PubMed  Google Scholar 

  10. Caroli, A., S. Manini, L. Antiga, K. Passera, B. Ene-Iordache, R. Bogdan, S. Rota, G. Remuzzi, A. Bode, J. Leermakers, F.S. van de Vosse, R. Vanholder, M. Malovrh, J. Tordoir, and A. Remuzzi. Validation of a patient-specific hemodynamic computational model for surgical planning of vascular access in hemodialysis patients. Kidney Int. 84(6):1237–1245, 2013.

    Article  PubMed  Google Scholar 

  11. Chen, P., A. Quarteroni, and G. Rozza. Simulation-based uncertainty quantification of human arterial network hemodynamics. Int. J. Numer. Methods Biomed. Eng. 29(6):698–721, 2013.

    Article  Google Scholar 

  12. Cockburn, B. An introduction to the discontinuous galerkin method for convection-dominated problems. In: Advanced Numerical Approximation of Nonlinear Hyperbolic Equations, edited by A. Quarteroni, Lecture Notes in Mathematics, vol. 1697, pp 151–268. Berlin: Springer, 1998.

  13. Crouse, J. R., U. Goldbourt, G. Evans, J. Pinsky, A. R. Sharrett, P. Sorlie, W. Riley, G. Heiss, and for the ARIC Investigators. Risk factors and segment-specific carotid arterial enlargement in the atherosclerosis risk in communities (aric) cohort. Stroke 27(1):69–75, 1996.

  14. Davis, A., C. Holloway, A. J. Lewandowski, N. Ntusi, R. M. Nethononda, A. Pitcher, J. M. Francis, P. Leeson, S. Neubauer, and O. J. Rider. Diameters of the normal thoracic aorta measured by cardiovascular magnetic resonance imaging; correlation with gender, body surface area and body mass index.J. Cardiovasc. Magn. Reson. 15(1):E77, 2013.

    Article  PubMed Central  Google Scholar 

  15. Fitchett, D. H. Lv-arterial coupling: interactive model to predict effect of wave reflections on lv energetics. Am. J. Physiol. Heart Circ. 261(4):1026–1033, 1991.

    Google Scholar 

  16. Formaggia, L., D. Lamponi, and A. Quarteroni. One dimensional models for blood flow in arteries. J. Eng. Math. 47(3–4):251–276, 2003.

    Article  Google Scholar 

  17. Guyton, A., and J. Hall. Textbook of Medical Physiology, 11th edition. Philadelphia: Elsevier Saunders, 2004.

    Google Scholar 

  18. Hamburg, N. M., M. M. Mott, S. J. Bigornia, M.-A. Duess, M. A. Kluge, D. T. Hess, C. M. Apovian, J. A. Vita, and N. Gokce. Maladaptive enlargement of the brachial artery in severe obesity is reversed with weight loss. Vasc. Med. 15(3):215–222, 2010

  19. Holenstein, A., P. Niederer, and M. Anliker. A viscoelastic model for use in predicting arterial pulse waves. J. Biomech. Eng. 102:318–326, 1980.

    Article  CAS  PubMed  Google Scholar 

  20. Huberts, W., C. de Jonge, W. van der Linden, M. Inda, J. Tordoir, F. van de Vosse, and E. Bosboom. A sensitivity analysis of a personalized pulse wave propagation model for arteriovenous fistula surgery. part a: Identification of most influential model parameters. Med. Eng. Phys. 35(6):810–826, 2013.

    Article  CAS  PubMed  Google Scholar 

  21. Kalita, P., and R. Schaefer. Mechanical Models of Artery Walls. Arch. Comput. Methods Eng. 15(1):1–36, 2007.

    Article  Google Scholar 

  22. Kivity, Y., and R. Collins. Non linear wave propagation in viscoelastic tubes: application to aortic rupture. J. Biomech. 7:67–76, 1974.

    Article  CAS  PubMed  Google Scholar 

  23. Korakianitis, T., and Y. Shi. Numerical simulation of cardiovascular dynamics with healthy and diseased heart valves.J. Biomech. 39(11):1964–1982, 2006.

    Article  PubMed  Google Scholar 

  24. Krejza, J., M. Arkuszewski, S. E. Kasner, J. Weigele, A. Ustymowicz, R. W. Hurst, B. L. Cucchiara, and S. R. Messe. Carotid artery diameter in men and women and the relation to body and neck size. Stroke, 37(4):1103–1105, 2006.

    Article  PubMed  Google Scholar 

  25. Ku D. N. Blood flow in arteries. Annu. Rev. Fluid. Mech. 29:399–434, 1997.

    Article  Google Scholar 

  26. Langewouters, G. Visco-elasticity of the human aorta in vitro in relation to pressure and age. PhD thesis, Free University, Amsterdam, 1982.

  27. Liang, F., S. Takagi, R. Himeno, and H. Liu. Biomechanical characterization of ventricular-arterial coupling during aging: a multi-scale model study. J. Biomech. 42(6):692–704, 2009.

  28. Lighthill, J. Mathematical Biofluiddynamics. Philadelphia, PA: Society for Industrial and Applied Mathematics, 1975.

    Book  Google Scholar 

  29. Malossi, A. C. I., P. J. Blanco, and S. Deparis. A two-level time step technique for the partitioned solution of one-dimensional arterial networks. Comput. Methods Appl. Mech. 237–240:212–226, 2011.

    Google Scholar 

  30. Matthys, K.S., J. Alastruey, J. Peirò, A. W. Khir, P. Segers, P. R. Verdonck, K. Parker, and S. J. Sherwin. Pulse wave propagation in a model human arterial network: assessment of 1-d numerical simulations against in vitro measurements.J. Biomech. 40(15):3476–86, 2007.

    Article  PubMed  Google Scholar 

  31. Müller, L.O., and E. F. Toro. A global multiscale mathematical model for the human circulation with emphasis on the venous system. Int. J. Numer. Methods Biomed. Eng. 2014.

  32. Mynard, J. P., M. R. Davidson, D. J. Penny, and J. J. Smolich. A simple, versatile valve model for use in lumped parameter and one-dimensional cardiovascular models. Int. J. Numer. Methods Biomed. Eng. 28: 626–641, 2012.

    Article  CAS  Google Scholar 

  33. Nichols, W., and M. F. O’Rourke. McDonald’s Blood Flow in Arteries. Theoretical, Experimental and Clinical Principles. Oxford, UK: Oxford University Press, 2005.

  34. Nicosia, S. and G. Pezzinga. Mathematical models of blood flow in the arterial network.J. Hydraul. Res. 45(2):188–201, 2007.

    Article  Google Scholar 

  35. Noordergraaf, A., P. Verdouw, and H. Boom. The use of an analog computer in a circulation mode. Prog. Cardiovasc. Dis. 5:419–439, 1963.

    Article  CAS  PubMed  Google Scholar 

  36. Olufsen, M.S. Structured tree outflow condition for blood flow in larger systemic arteries. Am. J. Physiol. Heart Circ., 1999, pp. 257–268.

  37. Olufsen, M. S., C. S. Peskin, W. Y. Kim, E. M. Pedersen, A. Nadim, and J. Larsen. Numerical simulation and experimental validation of blood flow in arteries with structured-tree outflow conditions. Ann. Biomed. Eng. 28(11):1281–99, 2000.

    Article  CAS  PubMed  Google Scholar 

  38. Ottesen, J., M. S. Olufsen, and J. K. Larsen. Applied Mathematical Models in Human Physiology, Vol. 33. Philadelphia, PA: Society for Industrial and Applied Mathematics, 2005.

    Google Scholar 

  39. Quarteroni, A. Cardiovascular Mathematics: Modeling and Simulation of the Circulatory System. MSA. Milano, Italy: Springer-Verlag, 2009.

  40. Raghu, R., I. E. Vignon-Clementel, C. A. Figueroa, and C. A. Taylor. Comparative study of viscoelastic arterial wall models in nonlinear one-dimensional finite element simulations of blood flow. J. Biomech. Eng. 133(August):1–11, 2011.

  41. Reymond, P., Y. Bohraus, F. Perren, F. Lazeyras, and N. Stergiopulos. Validation of a patient-specific one-dimensional model of the systemic arterial tree. Am. J. Physiol. Heart Circ. 301(3):1173–1182, 2011.

    Article  Google Scholar 

  42. Reymond, P., F. Merenda, F. Perren, D. Rüfenacht, and N. Stergiopulos. Validation of a one-dimensional model of the systemic arterial tree. Am. J. Physiol. Heart Circ. 297(1):208–222, 2009.

    Article  Google Scholar 

  43. Ruan, L., W. Chen, S. R. Srinivasan, M. Sun, H. Wang, A. Toprak, and G. S. Berenson. Correlates of common carotid artery lumen diameter in black and white younger adults: the bogalusa heart study. Stroke 40(3):702–707, 2009.

    Article  PubMed  Google Scholar 

  44. Sagawa, K. Cardiac Contraction and the Pressure–Volume Relationship. Oxford, UK: Oxford University Press, 1988.

    Google Scholar 

  45. Segers, P., N. Stergiopulos, N. Westerhof, P. Kolh, and P. R. Verdonck. Systemic and pulmonary hemodynamics assessed with a lumped-parameter heart-arterial interaction model. J. Eng. Math. 47:185–199, 2003.

    Article  Google Scholar 

  46. Shi, Y., P. Lawford, and R. Hose. Review of zero-D and 1-D models of blood flow in the cardiovascular system. Biomed. Eng. Online 10(1):33, 2011.

    Article  PubMed Central  PubMed  Google Scholar 

  47. Shishido, T., K. Hayashi, K. Shigemi, T. Sato, M. Sugimachi, and K. Sunagawa. Single-beat estimation of end-systolic elastance using bilinearly approximated time-varying elastance curve. Circulation 102(16):1983–1989, 2000.

    Article  CAS  PubMed  Google Scholar 

  48. Smulyan, H., S. J. Marchais, B. Pannier, A. P. Guerin, M. E. Safar, and G. M. London. Influence of body height on pulsatile arterial hemodynamic data.J. Am. Coll. Cardiol. 31(5):1103–1109, 1998.

  49. Stergiopulos, N., D. F. Young, and T. R. Rogge. Computer simulation of arterial flow with applications to arterial and aortic stenoses. J. Biomech. 25(12):1477–1488, 1992.

    Article  CAS  PubMed  Google Scholar 

  50. Valdez-Jasso, D., D. Bia, Y. Zocalo, R. Armentano, M. A. Haider, and M. S. Olufsen. Linear and nonlinear viscoelastic modeling of aorta and carotid pressure-area dynamics under in vivo and ex vivo conditions. Ann. Biomed. Eng. 39(5):1438–1456, 2011.

    Article  PubMed Central  PubMed  Google Scholar 

  51. van de Vosse, F.N., and N. Stergiopulos. Pulse wave propagation in the arterial tree. Annu. Rev. Fluid. Mech. 43(1):467–499, 2011.

    Article  Google Scholar 

  52. Westerhof, N., F. Bosman, C. De Vriens, and A. Noordergraaf. Analog studies of the human systemic arterial tree. J. Biomech. 2(2):121–134, 1969.

    Article  CAS  PubMed  Google Scholar 

  53. Westerhof, N., N. Stergiopulos, and M. I. Noble. Snapshots of Hemodynamics: An Aid for Clinical Research and Graduate Education. New York: Springer, 2010.

    Book  Google Scholar 

  54. Wildman, R. P., V. Mehta, T. Thompson, S. Brockwell, and K. Sutton-Tyrrell. Obesity is associated with larger arterial diameters in caucasian and African-American young adults. Diabetes Care 27(12):2997–2999, 2004.

    Article  PubMed  Google Scholar 

  55. Xiao, N., J. Alastruey, and C. A. Figueroa. Systematic comparison between 1-D and 3-D hemodynamics in compliant arterial models. Int. J. Numer. Methods Biomed. Eng. 30:204–231, 2014.

    Article  PubMed  Google Scholar 

  56. Zebekakis, P. E., T. Nawrot, L. Thijs, E. J. Balkestein, J. van der Heijden-Spek, L. M. Van Bortel, H. A. Struijker-Boudier, M. E. Safar, and J. A. Staessen. Obesity is associated with increased arterial stiffness from adolescence until old age. J. Hypertens. 23(10):1839–1846, 2005.

Download references

Acknowledgments

Franco Veglio, Alberto Milan, and Dario Leone of the of the Hypertension Center, Department of Medical Sciences, University of Turin are acknowledged for the data measure and their irreplaceable help.

Conflict of interest

There are no conflicts of interest.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andrea Guala.

Additional information

Associate Editor Diego Gallo oversaw the review of this article.

Appendices

Appendix

Coefficients of the constitutive Eq. (13)

$$\begin{aligned}&B_1=-\frac{1}{a_3^3} \bigl (a_5^3+PWV^6 \rho ^3+3 PWV^4 \rho ^2 a_5+3 a_5^2 PWV^2 \rho \bigr ) \quad \Biggl [\frac{\text {N}}{\text {m}^2}\Biggr ]\\&B_2=\frac{3\rho PWV^2}{A_0 a_3^3}\bigl (\rho ^2 PWV^4 +2 \rho a_5 PWV^2 +a_5^2\bigr ) \quad \Biggl [\frac{\text {N}}{\text {m}^4}\Biggr ]\\&B_3=-\frac{3 \rho ^2 PWV^4}{A_0^2 a_3^3} \bigl (a_5+\rho PWV^2 \bigr ) \quad \Biggl [\frac{\text {N}}{\text {m}^6}\Biggr ],\quad B_4=\Biggl (\frac{\rho PWV^2}{a_3 A_0}\Biggr )^3 \quad \Biggl [\frac{\text {N}}{\text {m}^8}\Biggr ]. \end{aligned}$$

Numerical Method

The model described in the second section is solved using the Runge-Kutta Discontinuous-Galerkin method12: first, space is discretized by a Discontinuous-Galerkin approach, then time evolution is described by a second-order Runge-Kutta scheme.

Conservative Form

In the conservative form, model (1) and (17) reads

$$\begin{aligned} \frac{\partial \mathbf {U}}{\partial t}+\frac{\partial \mathbf {F}(\mathbf {U})}{\partial x}+\mathbf {S}(\mathbf {U})=0 \end{aligned}$$
(29)

where \(\mathbf {U}=[A(x,t),Q(x,t)]^T\) are the conservative variables, whereas the corresponding fluxes and source terms are

$$\begin{aligned} \mathbf {F}(\mathbf {U})= \left[ \begin{array}{l} Q \\ \beta \frac{Q^2}{A} +\displaystyle \sum _{j=2}^{4}\frac{j\!-\!1}{\rho {j}}A^jB_j -\frac{B_5}{\rho } \sqrt{A} \frac{\partial Q}{\partial x} \end{array}\right],\quad \mathbf {S}(\mathbf {U})=\left[\begin{array}{l} 0\\ \displaystyle \sum _{j=1}^{4}\frac{A^j}{\rho {j}}\frac{\text{ d }B_j}{\text{ d }x} +\frac{B_5}{\rho \sqrt{A}} \frac{\partial Q}{\partial x} \frac{\partial A}{\partial x}\!-\!N_4\!-\!A\,b_x \end{array}\right], \end{aligned}$$

where \(b_x=g \sin (\gamma )\) is the longitudinal projection of the gravitational acceleration.

To solve system (29) in a one-dimensional generic domain \(\Omega \) discretized into \(N_{el}\) elemental non-overlapping regions \(\Omega _e=[x_e^l,x_e^r]\)—such that \(x_{e+1}^l=x_e^r\) for \(e=1,\ldots,N_{el}\) and \( \cup _{e=1}^{N_{el}} \Omega _e=\Omega \)—the weak form of Eq. (29) is written

$$\begin{aligned} \Biggl (\frac{\partial \mathbf {U}}{\partial t},\varvec{\psi } \Biggr )_\Omega +\Biggl (\frac{\partial \mathbf {F}}{\partial x},\varvec{\psi } \Biggr )_\Omega +\Bigl (\mathbf {S},\varvec{\psi } \Bigr )_\Omega =0 \end{aligned}$$
(30)

where \(\varvec{\psi }\) is a set of arbitrary test functions in \(\Omega \) and \( (\mathbf {v},\mathbf {p})_\Omega =\int _\Omega \mathbf {v}\mathbf {p} \text {dx} \) is the standard \(\mathbf {L}^2(\Omega )\) inner product between two generic function \(\mathbf {v}\) and \(\mathbf {p}\) whose domain is the closed interval \(\Omega \). Decomposing the integrals in Eq. (30) into the elemental regions and integrating by part the second term, one obtains

$$\begin{aligned} \sum _{e=1}^{N_{el}} \Biggl [\Biggl (\frac{\partial \mathbf {U}}{\partial t},\varvec{\psi } \Biggr )_{\Omega _e} -\Biggl (\mathbf {F},\frac{d \varvec{\psi }}{d x} \Biggr )_{\Omega _e} +\bigl [ \mathbf {F} \varvec{\psi }\bigr ]_{\partial \Omega _e} +\Bigl (\mathbf {S},\varvec{\psi } \Bigr )_{\Omega _e}\Biggr ] =0. \end{aligned}$$
(31)

In order to discretize the problem in space, U is assumed to belong to the finite dimensional space of the \(\mathbf {L}^2(\Omega )\) functions—that are polynomials of degree 1 on each element \(\Omega _e\)—obtaining the approximate solution \(\mathbf {U}_h\) (subscript \(h\) marks an element of such a space). Even if the solution \(\mathbf {U}_h\) can be discontinuous at boundaries between elemental regions, information propagates by upwinding the fluxes \(\mathbf {F}\) on the third term of Eq. (31).

The approximate solution \(\mathbf {U}_h=\mathbf {U}_h(x,t)\) on the element \(\Omega _e\) can be represented as

$$\begin{aligned} \mathbf {U}_h(x_e(\xi ),t)=\sum _{i=1}^2 \alpha _i(t) \phi _i(\xi ) \end{aligned}$$

where \(\alpha _i\) are the time dependent unknown weights, \(\phi _i(\xi )\) are the trial functions on the reference element \(\hat{\Omega }=[-1,1]\), and the following affine mapping between \(\hat{\Omega }\) and \(\Omega _e\) is introduced

$$\begin{aligned} x_e(\xi )=x_e^l \frac{1-\xi }{2}+x_e^r \frac{1+\xi }{2}. \end{aligned}$$

We choose first-degree Lagrange functions as expansion polynomial basis, \(\phi _1(\xi )\!=\!(1-\xi )/2\), \(\phi _2(\xi )\!=\!(1+\xi )/2\), and, following the usual Galerkin approach, the discrete test functions \(\varvec{\psi }_h\) are chosen in the same discrete space as the numerical solution \(\mathbf {U}_h\).

Introducing the approximate solution \(\mathbf {U}_h\) and the discrete test function into relation (31), one obtains the weak formulation as

$$\begin{aligned} \sum _{e=1}^{N_{el}} \Biggl [\Biggl (\frac{\partial \mathbf {U}_h}{\partial t},\varvec{\psi }_h \Biggr )_{\Omega _e}-\Biggl (\mathbf {F}(\mathbf {U}_h),\frac{d \varvec{\psi }_h}{d x}\Biggr )_{\Omega _e} +\bigl [ \mathbf {F}_{LF} \,\varvec{\psi }_h\bigr ]_{\partial \Omega _e}+\Bigl (\mathbf {S}(\mathbf {U}_h),\varvec{\psi }_h\Bigr )_{\Omega _e}\Biggr ] =0 \end{aligned}$$
(32)

where \(\mathbf {F}_{LF}\) are the upwinded numerical fluxes. They are obtained using the Lax-Friedrichs method as

$$\begin{aligned} \mathbf {F}_{LF}=\frac{1}{2} (\mathbf {F}(\mathbf {U}^-)+\mathbf {F}(\mathbf {U}^+)-\lambda _{max} (\mathbf {U}^+-\mathbf {U}^-)) \end{aligned}$$

where \(\mathbf {U}^+\) (\(\mathbf {U}^-\), resp.) are the variables on the right (left, resp.) side of the boundary and \(\lambda _{max}\) is the maximum eigenvalue of the matrix \(\mathbf {H}(\mathbf {U})\) of the quasi-linear form (see eq. (7.2.1), below).

The formulation (32) is then advanced in time by a second-order Runge-Kutta explicit method. The choice of such a low-order time-advancing scheme is coherent with the spatial discretization, in which the order of the expansion polynomial is one. This method is stable if the well-known Courant-Friedrichs-Lewy (CFL) condition is satisfied.

Boundary Conditions

When characteristics have slopes of opposite signs—as in the present situation, see below –, the differential system (1) and (17) needs only one physical condition at each domain boundary. However, the solution of the numerical problem requires to prescribe both variables at each boundary. For this reason, extra relations (called compatibility equations) have to be written at each boundary by projecting the differential equations along the outgoing characteristics.

Quasi-Linear Form

In order to formulate the compatibility conditions, system (29) is written in the quasi-linear form

$$\begin{aligned} \frac{\partial \mathbf {U}}{\partial t}+ \mathbf {H}(\mathbf {U})\frac{\partial \mathbf {U}}{\partial x}+\mathbf {S}_2(\mathbf {U})=0 \end{aligned}$$
(33)

where the terms

$$\begin{aligned} \mathbf {H}(\mathbf {U})=\left[\begin{array}{ll} 0&1 \\ \frac{1}{\rho }\displaystyle \sum _{j=1}^4jA^jB_{j+1} -\beta \frac{Q^2}{A^2}&2 \beta \frac{Q}{A} \end{array}\right],\quad \mathbf {S}_2(\mathbf {U})=\left[\begin{array}{l} 0\\ \frac{1}{\rho }\displaystyle \sum _{j=1}^4jA^j\frac{\text{ d }B_{j}}{\text{ d }x} -N_4 -A\,b_x \end{array}\right] \end{aligned}$$

are obtained by neglecting viscoelasticity. Notice that the characteristic variables move along the corresponding characteristic directions for a spatial distance equal to the product of the time step by the local wave celerity. Being both time step and celerity small, the spatial distance is very small thus allowing the viscoelastic effect to be neglected.

Let \(\varvec{\Lambda }\) and \(\mathbf {L}\) be the eigenvalue and left eigenvector matrices of \(\mathbf {H}(\mathbf {U})\), respectively, so that \(\mathbf {L} \mathbf {H} \mathbf {L}^{-1} = \varvec{\Lambda }\). They read

$$\begin{aligned} \varvec{\Lambda }(\mathbf {U})=\left[\begin{array}{ll} \lambda _1&0 \\ 0&\lambda _2 \end{array}\right] \quad \mathbf {L}(\mathbf {U})=\left[\begin{array}{ll} -\lambda _2&0 \\ -\lambda _1&0 \end{array}\right]. \end{aligned}$$
(34)

In the present case, the eigenvalues read

$$\begin{aligned} \lambda _1=\beta \frac{Q(x,t)}{A(x,t)}-c(x,t), \qquad \qquad \lambda _2=\beta \frac{Q(x,t)}{A(x,t)}+c(x,t) \end{aligned}$$

where

$$\begin{aligned} c(x,t)=\sqrt{\frac{A}{\rho }(B_2+2 B_3 A+3 B_4 A^2)+(\beta -1) \beta \frac{Q^2}{A^2}} \end{aligned}$$

is the celerity of the propagation when the fluid is at rest. At least in physiological condition, the system results strictly hyperbolic because \(c\) is always greater than \(\beta Q/A\), entailing \(\lambda _1<0\) (backward propagation) and \(\lambda _2>0\) (forward propagation).

Compatibility Conditions

The characteristic variables \(\mathbf {W}\) are defined as

$$\begin{aligned} \frac{\partial \mathbf {W}}{\partial \mathbf {U}}=\mathbf {L}(\mathbf {U}). \end{aligned}$$
(35)

Due to the structure of the matrix \(\mathbf {L}\), it is not possible to obtain the characteristic variables analytically. Therefore, pseudo-characteristic variables are introduced by linearising Eq. (35).

By pre-multiplying the vectorial Eq. (33) by the left eigenvector matrix, one obtains

$$\begin{aligned} \mathbf {L}(\mathbf {U}) \frac{\partial \mathbf {U}}{\partial t}+ \mathbf {L}(\mathbf {U}) \mathbf {H}(\mathbf {U})\frac{\partial \mathbf {U}}{\partial x}+\mathbf {L}(\mathbf {U}) \mathbf {S}_2(\mathbf {U})=0 \end{aligned}$$

and then

$$\begin{aligned} \mathbf {L}(\mathbf {U}) \frac{\partial \mathbf {U}}{\partial t}+\varvec{\Lambda }(\mathbf {U}) \mathbf {L}(\mathbf {U}) \frac{\partial \mathbf {U}}{\partial x}+\mathbf {L}(\mathbf {U}) \mathbf {S}_2(\mathbf {U})=0. \end{aligned}$$
(36)

In order to introduce the total derivative, the first term of Eq. (36) is rewritten as

$$\begin{aligned} \mathbf {L}(\mathbf {U}) \frac{\partial \mathbf {U}}{\partial t}=\frac{\partial }{\partial t} \Bigl ( \mathbf {L}(\mathbf {U}) \mathbf {U} \Bigr )- \mathbf {U} \frac{\partial \mathbf {L}(\mathbf {U})}{\partial t} \end{aligned}$$

and the second term as

$$\begin{aligned} \varvec{\Lambda }(\mathbf {U}) \mathbf {L}(\mathbf {U}) \frac{\partial \mathbf {U}}{\partial x}=\varvec{\Lambda }(\mathbf {U}) \frac{\partial }{\partial x} \Bigl ( \mathbf {L}(\mathbf {U}) \mathbf {U} \Bigr )- \varvec{\Lambda }(\mathbf {U}) \mathbf {U} \frac{\partial \mathbf {L}(\mathbf {U})}{\partial x}, \end{aligned}$$

obtaining

$$\begin{aligned} \frac{\partial }{\partial t} \Bigl ( \mathbf {L} (\mathbf {U})\mathbf {U}\Bigr )+ \varvec{\Lambda }(\mathbf {U}) \frac{\partial }{\partial x} \Bigl ( \mathbf {L}(\mathbf {U}) \mathbf {U} \Bigr )- \varvec{\Lambda }(\mathbf {U}) \mathbf {U} \frac{\partial \mathbf {L}(\mathbf {U})}{\partial x}- \mathbf {U} \frac{\partial \mathbf {L}(\mathbf {U})}{\partial t}+ \mathbf {L}(\mathbf {U}) \mathbf {S}_2(\mathbf {U})=0. \end{aligned}$$

By introducing the total derivative \(\mathrm{D} / \mathrm{D} t\), the last equation becomes

$$\begin{aligned} \frac{D\mathbf {L}(\mathbf {U}) \mathbf {U}}{D t}- \varvec{\Lambda }(\mathbf {U}) \mathbf {U} \frac{\partial \mathbf {L}(\mathbf {U})}{\partial x}- \mathbf {U} \frac{\partial \mathbf {L}(\mathbf {U})}{\partial t}+ \mathbf {L}(\mathbf {U}) \mathbf {S}_2(\mathbf {U})=0. \end{aligned}$$
(37)

If time is now discretized so that \(\Delta t=t_{n+1}-t_{n}\) is the time step (the subscript \(n\) denotes the \(n\)-th time instant), the total derivative can be approximated as

$$\begin{aligned} \frac{D}{D t}\Bigl ({\mathbf {L}}_{n+1}{\mathbf {U}}_{n+1}\Bigr )\approx \frac{{\mathbf {L}}_{n+1}{\mathbf {U}}_{n+1}-{\mathbf {L}}_{n}^{\star }{\mathbf {U}}_{n}^{\star }}{\Delta t}=\frac{{\mathbf {L}}_{n} {\mathbf {U}}_{n+1}-{\mathbf {L}}_{n}^{\star }{\mathbf {U}}_{n}^{\star }}{\Delta t}+\frac{\bigl ({\mathbf {L}}_{n+1} -{\mathbf {L}}_{n}\bigr ){\mathbf {U}}_{n+1}}{\Delta t} \end{aligned}$$
(38)

where the superscript \(\star \) refers to the point located at the distance from the boundary equal to \(c \cdot \Delta t\). By assuming that

$$\begin{aligned} \frac{\bigl (\mathbf {L}_{n+1} -\mathbf {L}_{n}\bigr )\mathbf {U}_{n+1}}{\Delta t} \approx \mathbf {U} \frac{\partial \mathbf {L}(\mathbf {U})}{\partial t} \end{aligned}$$

the departure of the pseudo-characteristic from a generic state \(\mathbf {U}_n^{\star }\) is taken as

$$\begin{aligned} \mathbf {L}_n \mathbf {U}_{n+1}=\mathbf {L}_n^{\star } \mathbf {U}_{n}^{\star }+\Delta t \, \Bigl ( \varvec{\Lambda }_n^{\star } \frac{\partial \mathbf {L}_n^{\star }}{\partial x} \mathbf {U}_{n}^{\star }-\mathbf {L}_n^{\star } \mathbf {S}_2(\mathbf {U}_n^{\star }) \Bigl ) \end{aligned}$$
(39)

where \(\mathbf {L}_n=\mathbf {L}(\mathbf {U}_n)\) and \(\varvec{\Lambda }_n=\varvec{\Lambda }(\mathbf {U}_n)\).

Finally, because the pseudo-characteristics at the initial state are set null, Eq. (39) needs to be balanced by subtracting the initial condition to the dependent variables,29 i.e.

$$\begin{aligned} \mathbf {L}_n (\mathbf {U}_{n+1}-\mathbf {U}_{0})=\mathbf {L}_n^{\star } (\mathbf {U}_{n}^{\star }-\mathbf {U}_{0}^{\star })+\Delta t \, \Bigl (\varvec{\Lambda }_n^{\star } \frac{\partial \mathbf {L}_n^{\star }}{\partial x} (\mathbf {U}_{n}^{\star }-\mathbf {U}_{0}^{\star })-\mathbf {L}_n^{\star } (\mathbf {S}_2(\mathbf {U}_n^{\star })-\mathbf {S}_2(\mathbf {U}_0^{\star })) \Bigl ) \end{aligned}$$
(40)

where the subscript \(0\) refers to the initial condition. Since the reference condition (uniform pressure of 100 mmHg and no flow) is assumed as initial condition, the same subscript \(0\) is used to define both quantities.

The formulation (40) has to be coupled with the physical boundary conditions at every (internal and external) boundary. The corresponding non-linear systems are solved by an adaptive trust-region-dogleg method.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guala, A., Camporeale, C., Tosello, F. et al. Modelling and Subject-Specific Validation of the Heart-Arterial Tree System. Ann Biomed Eng 43, 222–237 (2015). https://doi.org/10.1007/s10439-014-1163-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10439-014-1163-9

Keywords

Navigation