Reducing the number of parameters in 1D arterial blood flow modeling: less is more for patient-specific simulations

Sally Epstein, Marie Willemet, Phil J Chowienczyk, Jordi Alastruey, Sally Epstein, Marie Willemet, Phil J Chowienczyk, Jordi Alastruey

Abstract

Patient-specific one-dimensional (1D) blood flow modeling requires estimating model parameters from available clinical data, ideally acquired noninvasively. The larger the number of arterial segments in a distributed 1D model, the greater the number of input parameters that need to be estimated. We investigated the effect of a reduction in the number of arterial segments in a given distributed 1D model on the shape of the simulated pressure and flow waveforms. This is achieved by systematically lumping peripheral 1D model branches into windkessel models that preserve the net resistance and total compliance of the original model. We applied our methodology to a model of the 55 larger systemic arteries in the human and to an extended 67-artery model that contains the digital arteries that perfuse the fingers. Results show good agreement in the shape of the aortic and digital waveforms between the original 55-artery (67-artery) and reduced 21-artery (37-artery) models. Reducing the number of segments also enables us to investigate the effect of arterial network topology (and hence reflection sites) on the shape of waveforms. Results show that wave reflections in the thoracic aorta and renal arteries play an important role in shaping the aortic pressure and flow waves and in generating the second peak of the digital pressure and flow waves. Our novel methodology is important to simplify the computational domain while maintaining the precision of the numerical predictions and to assess the effect of wave reflections.

Keywords: 1D modeling; aortic pulse wave; digital pulse wave; hypertension; windkessel model.

Copyright © 2015 the American Physiological Society.

Figures

Fig. 1.
Fig. 1.
A: flow waveform prescribed at the aortic root as a reflective boundary condition (black), and simulated flow waveform at the digital artery (gray) for the 67-artery model. B: simulated pressure waveforms at the aortic root (black) and digital artery (gray) for the 67-artery model. C: schematic representation of the 55-artery 1-dimensional (1D) model (black). The circled region shows additional anatomical model of the hand included in the 67-artery model (grey), containing vessels of the superficial palmar arch and digital arteries. D: flow waveform prescribed at the aortic root as a reflective boundary condition (black), and simulated flow waveform at the thoracic aorta (grey) for the 55-artery model. E: simulated pressure waveforms for the 55-artery model at the aortic root (black) and thoracic aorta (gray).
Fig. 2.
Fig. 2.
A–C: reduction of a nonlinear 1D model single vessel attached to a 3-element windkessel model (R1–C–R2; A) into a single 2-element windkessel model (Cnew–Rnew; C), via an intermediary stage (B) in which the 1D vessel is simplified into a 2-element windkessel model (Cv–Rv). The 1D vessel is characterized by a length l, a cross-sectional area A(x) and a pulse wave velocity c(x). D–F: reduction of a nonlinear 1D model single bifurcation coupled to 3-element windkessel models (R1–C–R2; D) into a nonlinear 1D model single vessel coupled to a 3-element windkessel model (Rnew,1–Cnew–Rnew,2; F). In the intermediary stage (E), the daughter vessels 1 and 2 and their outlet windkessel models are transformed into two 2-element windkessel models (CT–RT). qin(t): inflow; qout(t): outflow; pin(t): inflow pressure; pout: outflow pressure.
Fig. 3.
Fig. 3.
A–F: pressure (P) and flow (Q) waveforms at the aortic root and midpoint of the thoracic aorta of the normotensive 55-artery model (solid grey) and several reduced models (dashed black), 1 for each row. The number of segments and the arterial topology for each reduced model are given in the 1st 2 columns. The aortic-root pressure waveform calculated by a 2-element windkessel model of the whole systemic circulation is shown in F. Average (avg), systolic (sys) and diastolic (dias) relative errors are indicated in the top right corner of each plot.
Fig. 4.
Fig. 4.
Evolution of the average (A and D), systolic (B and E), and diastolic (C and F) relative errors in the pressure waveform at the aortic root of the 55-artery model with the number of arterial segments, under normotensive (A–C) and hypertensive (D–F) conditions. For each plot, the vertical dashed line corresponds to the 19-artery model: arterial networks on its left (≥19 vessels) include a full aorta, whereas those on its right (<19 vessels) only include portions of it.
Fig. 5.
Fig. 5.
A–F: pressure (P) and flow (Q) waveforms at the aortic root and midpoint of the thoracic aorta of the hypertensive 55-artery model (solid grey) and several reduced models (dashed black), 1 for each row. The number of segments and the arterial topology for each reduced model are given in the 1st 2 columns. The aortic-root pressure waveform calculated by a 2-element windkessel model of the whole systemic circulation is shown in F. Average (avg), systolic (sys), and diastolic (dias) relative errors are indicated in the top right corner of each plot.
Fig. 6.
Fig. 6.
A–F: pressure (P) and flow (Q) waveforms at the aortic root and midpoint of the digital artery of the 67-artery model (solid grey) and several reduced models (dashed black), one for each row. The number of segments and the arterial topology for each reduced model are given in the 1st 2 columns. Average (avg), systolic (sys), and diastolic (dias) relative errors are indicated in the top right corner of each plot.
Fig. 7.
Fig. 7.
Evolution of the average (□), systolic (+), and diastolic (●) relative errors in the pressure waveform at the aortic root (A) and digital artery (B) with the number of arterial segments in the 67-artery model. Vertical dashed lines correspond to the 35-artery model: arterial networks displayed on the left and on the line include a full aorta while those on the right only include portions of it.

References

    1. Alastruey J, Parker KH, Peiró J, Sherwin SJ. Can the modified Allen's test always detect sufficient collateral flow in the hand? A computational study. Comput Methods Biomech Biomed Engin 9: 353–361, 2006.
    1. Alastruey J, Parker K, Peiró J, Sherwin SJ. Lumped parameter outflow models for 1-D blood flow simulations: effect on pulse waves and parameter estimation. Commun Math Phys 4: 317–336, 2008.
    1. Alastruey J, Parker KH, Peiró J, Sherwin SJ. Analysing the pattern of pulse waves in arterial networks: a time-domain study. J Eng Math 64: 331–351, 2009.
    1. Alastruey J, Parker K, Sherwin S. Arterial pulse wave haemodynamics. In: 11th International Conference on Pressure Surges, edited by Anderson S. Lisbon, Portugal: Virtual PiE Led t/a BHR Group, 2012, chapt. 7, p. 401–442.
    1. Almond NE, Jones DP. Noninvasive measurement of the human peripheral circulation: relationship between laser Doppler flowmeter and photoplethysmograph signals from the finger. Angiology 39: 819–829, 1988.
    1. Armentano R, Graf S, Barra J, Velikovsky G, Baglivo H, Sanchez R, Simon A, Pichel RH, Levenson J. Carotid wall viscosity increase is related to intima-media thickening in hypertensive patients. Hypertension 31: 534–539, 1998.
    1. Avolio AP. Multi-branched model of the human arterial system. Med Biol Eng Comput 18: 709–718, 1980.
    1. Avolio A. The finger volume pulse and assessment of arterial properties. J Hypertens 20: 2341–2343, 2002.
    1. Baruch MC, Warburton DE, Bredin SS, Cote A, Gerdt DW, Adkins CM. Pulse decomposition analysis of the digital arterial pulse during hemorrhage simulation. Nonlinear Biomed Phys 5: 1–15, 2011.
    1. Baruch M, Kalantari K, Gerdt D, Adkins CM. Validation of the pulse decomposition analysis algorithm using central arterial blood pressure. Biomed Eng Online 13: 1–19, 2014.
    1. Blanco PJ, Watanabe SM, Dari EA, Passos MA, Feijóo RA. Blood flow distribution in an anatomically detailed arterial network model: criteria and algorithms. Biomech Model Mechanobiol 13: 1303–1330, 2014.
    1. Chowienczyk PJ, Kelly RP, Maccallum H, Millasseau SC, Andersson TL, Gosling RG, Ritter JM, Anggard EE. Photoplethysmographic assessment of pulse wave reflection. J Am Coll Cardiol 34: 2007–2014, 1999.
    1. Davies JE, Alastruey J, Francis DP, Hadjiloizou N, Whinnett ZI, Manisty CH, Aguado-Sierra J, Willson K, Foale AR, Malik IS, Hughes AD, Parker KH, Mayet J. Attenuation of wave reflection by wave entrapment creates a “horizon effect” in the human aorta. Hypertension 60: 778–785, 2012.
    1. Dick DE, Kendrick JE, Matson GL, Rideout VC. Measurement of nonlinearity in the arterial system of the dog by a new method. Circ Res 22: 101–111, 1968.
    1. Formaggia L, Nobile F. Multiscale modelling of the circulatory system: a preliminary analysis. Comput Visual Sci 2: 75–83, 1999.
    1. Formaggia L, Gerbeau JF, Nobile F, Quarteroni A. On the coupling of 3D and 1D Navier-Stokes equations for flow problems in compliant vessels. Comput Methods Appl Mech Eng 191: 561–582, 2001.
    1. Frank O. Die Grundform des arteriellen Pulses. Z Biol 37: 483–526, 1899.
    1. Imholz B, Wieling W, Montfrans G, Wesseling K. Fifteen years experience with finger arterial pressure monitoring: assessment of the technology. Cardiovasc Res 38: 605–616, 1998.
    1. Jaeger G, Westerhof N, Noordergraaf A. Oscillatory flow impedance in electrical analog of arterial system. Circ Res 16: 121–34, 1965.
    1. Kakar P, Lip GY. Towards understanding the aetiology and pathophysiology of human hypertension: where are we now? J Hum Hypertens 20: 833–836, 2006.
    1. Karamanoglu M, Gallagher D, Avolio A, O'Rourke M. Functional origin of reflected pressure waves in a multibranched model of the human arterial system. Am J Physiol Heart Circ Physiol 267: H1681–H1688, 1994.
    1. Kearney P, Whelton M, Reynolds K. Global burden of hypertension: analysis of worldwide data. Lancet 365: 217–223, 2005.
    1. Khir A, Parker K. Wave intensity in the ascending aorta: effects of arterial occlusion. J Biomech 38: 647–655, 2005.
    1. Kind T, Faes TJ, Lankhaar JW, Vonk-Noordegraaf A, Verhaegen M. Estimation of three- and four-element windkessel parameters using subspace model identification. IEEE Trans Biomed Eng 57: 1531–1538, 2010.
    1. Laurent S, Boutouyrie P, Asmar R, Gautier I, Laloux B, Guize L, Ducimetiere P, Benetos A. Aortic stiffness is an independent predictor of all-cause and cardiovascular mortality in hypertensive patients. Hypertension 37: 1236–1241, 2001.
    1. Leguy CA, Bosboom EM, Gelderblom H, Hoeks AP, van de Vosse FN. Estimation of distributed arterial mechanical properties using a wave propagation model in a reverse way. Med Eng Phys 32: 957–967, 2010.
    1. Mancia G; The Task Force for the Management of Arterial Hypertension of the European Society of Hypertension (ESH) and of the European Society of Cardiology (ESC). 2013 ESH/ESC guidelines for the management of arterial hypertension. Eur Heart J 34: 2159–2219, 2013.
    1. Milisic V, Quarteroni A. Analysis of lumped parameter models for blood flow simulations and their relation with 1D models. ESAIM 38: 613–632, 2004.
    1. Millasseau S, Kelly R, Ritter J, Chowienczyk P. Determination of age-related increases in large artery stiffness by digital pulse contour analysis. Clin Sci 377: 371–377, 2002.
    1. Mynard JP, Nithiarasu P. A 1D arterial blood flow model incorporating ventricular pressure, aortic valve and regional coronary flow using the locally conservative Galerkin (LCG) method. Commun Numer Meth Eng 24: 367–417, 2008.
    1. Olufsen M, Nadim A. On deriving lumped models for blood flow and pressure in the systemic arteries. Math Biosci Eng 1: 61–80, 2004.
    1. Olufsen M. Structured tree outflow condition for blood flow in larger systemic arteries. Am J Physiol Heart Circ Physiol 276: H257–H268, 1999.
    1. Perdikaris P, Grinberg L, Karniadakis GE. An effective fractal-tree closure model for simulating blood flow in large arterial networks. Ann Biomed Eng 2014. December 16 [Epub ahead of print].
    1. Quarteroni A, Formaggia L. Mathematical Modelling and Numerical Simulation of the Cardiovascular System. Lausanne, Switzerland: École Polytechnique Fédérale de Lausanne, 2002.
    1. Quick CM, Young WL, Noordergraaf A. Infinite number of solutions to the hemodynamic inverse problem. Am J Physiol Heart Circ Physiol 280: H1472–H1479, 2001.
    1. Quick CM, Berger DS, Stewart RH, Laine a G, Hartley CJ, Noordergraaf A. Resolving the hemodynamic inverse problem. IEEE Trans Biomed Eng 53: 361–368, 2006.
    1. Raines JK, Jaffrin MY, Shapiro AH. A computer simulation of arterial dynamics in the human leg. J Biomech 7: 77–91, 1974.
    1. Reymond P, Merenda F, Perren F, Rufenacht D, Stergiopulos N. Validation of a one-dimensional model of the systemic arterial tree. Am J Physiol Heart Circ Physiol 297: H208–H222, 2009.
    1. Segers P, Rietzschel ER, De Buyzere ML, Stergiopulos N, Westerhof N, Van Bortel LM, Gillebert T, Verdonck PR. Three- and four-element Windkessel models: assessment of their fitting performance in a large cohort of healthy middle-aged individuals. Proc Inst Mech Eng H 222: 417–428, 2008.
    1. Sherwin S, Franke V, Peiró J, Parker K. One-dimensional modelling of a vascular network in space-time variables. J Eng Math 47: 217–250, 2003.
    1. Smith NP, Pullan AJ, Hunter PJ. An anatomically based model of transient coronary blood flow in the heart. SIAM J Appl Math 62: 990–1018, 2002.
    1. Stergiopulos N, Young DF, Rogge TR. Computer simulation of arterial flow with applications to arterial and aortic stenoses. J Biomech 25: 1477–1488, 1992.
    1. Stergiopulos N, Westerhof BE, Westerhof N. Total arterial inertance as the fourth element of the windkessel model. Am J Physiol Heart Circ Physiol 276: H81–H88, 1999.
    1. Valdez-Jasso D, Bia D, Zócalo Y, Armentano RL, Haider AM, Olufsen MS. Linear and nonlinear viscoelastic modeling of aorta and carotid pressure-area dynamics under in vivo and ex vivo conditions. Ann Biomed Eng 39: 1438–1456, 2011.
    1. van den Bos G, Westerhof N, Elzinga G, Sipkema P. Reflection in the systemic arterial system: effects of aortic and carotid occlusion. Cardiovasc Res 10: 565–573, 1976.
    1. Wang JJ, O'Brien AB, Shrive NG, Parker KH, Tyberg JV. Time-domain representation of ventricular-arterial coupling as a windkessel and wave system. Am J Physiol Heart Circ Physiol 284: H1358–H1368, 2003.
    1. Wang TJ, Gona P, Larson MG, Levy D, Benjamin EJ, Tofler GH, Jacques PF, Meigs JB, Rifai N, Selhub J, Robins SJ, Newton-Cheh C, Vasan RS. Multiple biomarkers and the risk of incident hypertension. Hypertension 49: 432–438, 2007.
    1. Westerhof N, Lankhaar JW, Westerhof BE. The arterial windkessel. Med Biol Eng Comput 47: 131–141, 2009.
    1. Willemet M, Lacroix V, Marchandise E. Inlet boundary conditions for blood flow simulations in truncated arterial networks. J Biomech 44: 897–903, 2011.
    1. Willemet M, Lacroix V, Marchandise E. Validation of a 1D patient-specific model of the arterial hemodynamics in bypassed lower-limbs: simulations against in vivo measurements. Med Eng Phys 35: 1573–1583, 2013.
    1. Willemet M, Alastruey J. Arterial pressure and flow wave analysis using time-domain 1-D hemodynamics. Ann Biomed Eng 43: 190–206, 2015.
    1. Xiao N, Alastruey J, Alberto Figueroa C. A systematic comparison between 1-D and 3-D hemodynamics in compliant arterial models. Int J Numer Method Biomed Eng 30: 204–231, 2013.

Source: PubMed

3
订阅