You are on page 1of 14

Revisiting the pressure-area relation for the flow in

elastic tubes: Application to arterial vessels


Rafik Absi

To cite this version:


Rafik Absi. Revisiting the pressure-area relation for the flow in elastic tubes: Application to arterial
vessels. Series on Biomechanics, Bulgarian Academy of Sciences, 2018, 32, pp.47-59. �hal-01807385�

HAL Id: hal-01807385


https://hal.archives-ouvertes.fr/hal-01807385
Submitted on 4 Jun 2018

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

Revisiting the pressure-area relation for the flow in elastic tubes:


Application to arterial vessels
Rafik Absi
EBI - Ecole de Biologie Industrielle, 49 avenue des Genottes, CS 90009, 95895 CERGY Cedex, FRANCE

Abstract
For the description of the flow behaviour in elastic tubes as arterial vessels, we need a relationship between the
transmural (internal minus external) pressure and the variation in the cross-sectional area A (or diameter), i.e.,
the pressure-area − constitutive relation. However, a literature review shows different relations. In this study,
the method based on the linear theory of elasticity is revisited. A new pressure-area − relation is proposed.
Results for the variation of cross-sectional area, arterial compliance and distensibility are presented. To define
a unique threshold value for the applicability of the former equations, all results are presented in dimensionless form
using the parameter = ℎ / (where E is Young’s Modulus, ℎ and are respectively the vessel wall
thickness and the internal radius at =0). Comparisons with the so-called linear and non-linear − equations
show that all results are similar for / < 0.05. Our results indicate that the former equations could be used with
an accepted gap until / =0.1. However, the inaccuracy increases with and at / =0.2, the difference
is of 26.7% and 24.6% respectively for the linear and non-linear relations. Proposed equations were applied to arterial
vessels with =150mmHg for radius from 0.8 to 6 mm. Results show an increase in the diameter of 4% for
=0.8mm while it is of 30% for =6mm.

Keywords: Pressure-area relation, Transmural pressure, Elastic tubes, Arterial vessels, Compliance

1. Introduction

Physiological and cardiovascular fluid mechanics provide an understanding of advanced concepts in fluid
mechanics to study blood flow in the cardiovascular system. The knowledge of the fluid mechanics of the
circulatory system is indispensable for well understanding of many cardiovascular diseases [1,2]. Principles
of conservation of mass and momentum provide the main equations of fluid flow which are non-linear,
partial differential equations and need numerical solutions. For some cases, simplification of these
equations allows analytical solutions [3,4]. Computational fluid dynamics (CFD) modelling provides
detailed pressure and flow fields [5] and the quantification of some parameters which cannot be obtained
experimentally as wall shear stress [6,7].
Fluid dynamics in elastic tubes is of high interest in different industrial and biological applications [8-10].
The description of the flow behaviour in elastic tubes as arterial vessels needs three independent variables
namely the pressure ( , ), the fluid velocity ( , ) (or equivalently the flow rate ( , )) and the cross-
sectional area ( , ). The main governing equations are the conservation of mass and momentum (i.e. the
continuity and the momentum equations). In this problem, we only have two equations and three variables,
namely, , , and . Therefore, we need a third relation which describes the deformation of the vessel walls
due to a variation in the pressure. A third equation could be obtained from the energy conservation which
is related to the interaction between the fluid and the tube wall or by analytical equations which provide a
relationship between the transmural (internal minus external) pressure and the variation in the cross-
sectional area (or diameter), i.e., the so-called state equation or pressure-area − constitutive relations.
A literature review shows different pressure-area relations [11-21]. In this study, we will consider the
1
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

relations which are derived from the linear theory of elasticity without considering the viscoelastic
behaviour. The more used equations which show interest in practical purposes are the so-called linear and
non-linear − relations, e.g., Rammos et al. (1998) [11], Olufsen et al. (1999) [12], Formaggia et al.
(2003) [13], Sherwin et al. (2003) [14] and Urquiza et al. (2006) [15]).
The aim of this study is to provide an answer about these different pressure-area relations. It is important
to understand why there are different relations, and which one should be used. To provide an appropriate
answer to these questions, it is important to re-examine the method which provides these relations.

In the following sections we will first review the linear and non-linear − relations, then present the
main assumptions of the study. The method of obtaining a − relation will be revisited and finally our
results will be compared to former relations and we will discuss the implications for arterial vessels.

2. Review of pressure-area relations

The vessel is represented as a cylindrical tube (Fig. 1) of length , wall thickness ℎ, inner (or internal)
radius = , outer (or external) radius and circular cross-sectional area ( , ) = . Pressure-
area equations provide relation between the transmural pressure and the variation in the cross-sectional
area (or the diameter). The transmural pressure is defined as = − , where = is the
internal fluid (blood) pressure and is the external pressure (from surrounding tissue). The variation in
2
the area is between the tube section = (at internal fluid pressure p) and the area 0 = 0 when
there is zero transmural pressure (i.e. = ), where is the radius at =0.
In this study, we will consider the more used − relations, i.e., linear and non-linear equations
(presented in Table 1).

Fig. 1: A diagram of a vessel represented as a cylindrical tube, = is the internal radius; h is the wall
thickness; L is the length of the tube; and σ is the circumferential stress.

2.1. The linear pressure-area relation

We write the linear − relation in the following form

(1) = ( − )

where is the proportionality factor which is a measure for the stiffness of the tube wall, the value from
Rammos equation is therefore = , where the coefficient of proportionality is related to Young’s
Modulus , the vessel wall thickness ℎ, =2 the diameter and the cross-sectional area when
=0.

2
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

Table 1
Linear and non-linear − relations used in the literature.

Author − Relation
Rammos et al. (1998) [11]
= 1+ ( − )

Olufsen et al. (1999) [12]
4 ℎ
− = 1−
3
Sherwin et al. (2003) [14] √ ℎ
= + (√ − √ )
(1 − )
Urquiza et al. (2006) [15]

= + −1

2.2. Non-linear pressure-area relations

We write the non-linear − relation in a single general form as

(2) = (√ − √ )

In this relation the coefficient of proportionality is given therefore by different relations (Olufsen et al.
(1999) [12], Sherwin et al. (2003) [14], Urquiza et al. (2006) [15]) which are summarized in Table (2).

Table 2
Coefficient for the non-linear − relations.

Author
Olufsen et al. (1999) [12] 4 ℎ 1
3 √
Sherwin et al. (2003) [14] √ ℎ
(1 − )
Urquiza et al. (2006) [15] ℎ 1

In the two first non-linear relations of Olufsen et al. (1999) [12] and Sherwin et al. (2003) [14] (Table 2)
the variation of the thickness ℎ (equal to ℎ at =0) due to the deformation is introduced through the
Poisson ratio . In Olufsen’s relation =0.5 and therefore 1 − 2 = 3/4 and it contains ℎ instead of ℎ .
In the last nonlinear elastic relation of Table (2), Urquiza et al. (2006) [15] didn’t consider the Poisson
ratio. It is important to understand why there are linear and non-linear relations and why different relations
for β, and which relation should be used? An answer requires a re-examination of the method related to
these relations.

3
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

3. Pressure-area relation revisited

3.1. Main assumptions of flow in elastic tubes (arterial vessels)

Several approaches can be taken to write the relation between the pressure and the cross-sectional area. The
arterial wall shows a time-delay in the response from a change in pressure to the corresponding change in
cross-sectional area i.e., the viscoelastic behaviour [22-24]. However, these viscoelastic effects seem to be
small within the physiological range of the flow and pressure [25]. Therefore, many studies use relations
derived from the linear theory of elasticity and disregard the viscoelastic behaviour [11-22]. Our study is
based therefore on the following assumptions:
• the flow is axisymmetric
• the arterial vessels walls are thin, i.e., ℎ<< , that the loading and deformation are axisymmetric
• the structural arterial properties are constant
• the vessel is tethered in the longitudinal direction
To obtain the relationship between the pressure and the cross-sectional area, we need to examine the
equilibrium of the internal and external forces acting on a unit element of the wall.

3.2. Pressure force

The elementary force due to the pressure differences is given by

(3) = − =( − )

For the half cylinder (Figure 1), the vertical component is

(4) =( − ) sin

After integrating from 0 to π, we obtain the vertical force due to the pressure differences

(5) = 2( − )

If the vessel is thin-walled (ℎ ≪ ), then ≈ = and

(6) = 2( − ) =2

3.3. LaPlace’́s Law

The aim is to link the transmural pressure to the tension in the walls related to the wall stress (force per unit
area) σ. The force pulling the half cylinder down is

(7) =2 ℎ

In equilibrium, Fσ is balanced by the vertical force due to the transmural pressure (Eq. 6)

(8) ℎ=

Equation (8) is an expression of LaPlace’́s law for a thin-walled cylinder. Note that for a given transmural
pressure, the wall tension ( = ℎ) per unit length increases as the radius increases and vice-versa. The
elasticity of the tube involves a relation between stress and strain (proportional deformation) as = ∆ / .
By considering the variation of the wall thickness through the Poisson ratio and Eq. (8), we write

4
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.


(9) = = ( )

If the radius varies from the initial value related to =0 (zero transmural pressure) to a given value
, the strain is equal to ( − )/ , Eq. (9) becomes

(10) = ( )
−1

The former linear − relation (Eq. 1) is based therefore on the assumption / = / which is not
realistic and could be valid only for small deformations. The accurate relation is / =√ / and
therefore involves a non-linear − relation.

(11) = ( )
(√ − √ )

With Eq. (2), the coefficient is therefore equal to

(12) = ( ) √

With = 0.5 in Eq. (12), we obtain the same coefficient of Olufsen et al. (1999) [12] (Table 2).
However, in Eq. (12), ℎ and are not constants since they change with . Therefore, we need a relation
for with ℎ and instead of ℎ and . To write this equation, we need to know the variation of the wall
thickness. Based on the assumption related to this variation, we present the following formulations.

4. A new − relation

4.1. A first approximation

If we assume that the variation of the wall thickness is negligible (ℎ = ℎ and = 0) Eq. (12) becomes
= and we write Eq. (2) in the following dimensionless form

(13) = 1− where =

Coefficient has the unit of pressure, it could be interpreted as the initial value (at =0) of the
coefficient of proportionality between and the strain. From Eq. (13), we have the following explicit
equation for ( )

(14) =

In Eq. (14), / ≠1, this equation will be used for < .

4.2. Proposed − relation

If we consider the variation of the wall thickness ℎ through the Poisson ratio , it is possible to write a
relation between ℎ and ℎ based on the area conservation equation 2 ℎ = 2 ℎ, Eq. (12) becomes

= ( )
and we write Eq. (2) as

5
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

(√ ) √
(15) = where = ( )

In this equation, is known through input data , ℎ and . However, it is possible to write Eq. (15) in
dimensionless form using and , since = /(1 − ), as

(16) = ( )
1−

In equation (16), the dimensionless coefficient of proportionality ( )


is related to the assumption
based on the area conservation equation. When the variation of the wall thickness is negligible ℎ = ℎ ,
this coefficient becomes equal to 1 since for this case = 0 and = in this dimensionless coefficient
and Eq. (16) reverts to Eq. (13) of the first approximation (Appendix 1).

We write equations (13) and (16) in a single form as:

(17) = 1−

1 =0
where =
( )
≠0

However unlike in the first approximation, Eq. (16) doesn’t allow one to write an explicit equation for
( ). In all − relations (Table 3), the different coefficients of proportionality ( , , or ) play
an important role which impacts strongly the results. This coefficient depends on the biomechanical
behaviour of the blood vessels which are determined by the physical properties of the individual wall
constituents (mostly elastin, collagen, and smooth muscle), and their relative content [20].

Table 3
Summary of former and proposed − relations.

Coefficient /
Linear [11] ( − ) ℎ 1
= + −1
2
Non-linear [15] (√ − √ ) ℎ
= +
−1

1sr √ − ℎ
Approximation =
√ 1−
1−
Proposed (√ − ) √ ℎ
= 1
(1 − ) 1−
(1 − )

6
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

Table 4
Summary of arterial compliance or capacitance and distensibility from former and proposed relations.

Linear [11] 1 2
2
Non-linear [15] 2√
2 2

1st / /
2
Approximation 2 2

Proposed

( − 0.5 √ ) (1 − ) (1 − )
1 − 0.5 1 − 0.5

4.2. Arterial compliance or capacitance and the distensibility

The compliance or capacitance describes how volume changes in response to a change in pressure [26], it
is inversely proportional to elasticity. The arterial capacitance per unit length or cross-sectional compliance
[27] may be calculated assuming that vessel length does not vary with transmural pressure.

(18) =

The capacitance is a key parameter involved in many calculations such as:


• The flow rate is related to the cross-sectional compliance through =
• The distensibility is defined by = =
Our two proposed relations for compliance and distensibility are summarized in table (4) and presented
together with the former linear and non-linear equations.

5. Results and discussion

5.1. Variation of the cross-sectional area and diameter

Figure (2) shows a comparison between − curves obtained from the methods presented in tables (2)
and (3).
For =0, fig. (2.a) presents comparisons between three explicit equations of namely, the linear
− relation = + with the coefficient of Rammos [11] (dashed line), the non-linear −
relation = + with the coefficient (Table 2) given by Urquiza et al. [15]) (dash-dotted
line) and our approximation given by Eq. (14) (solid line). Figure (2.a) shows that all equations present
practically same results for / <0.05. This could explain the applicability of the linear equation for
small values of / .

7
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

For =0.5, fig. (2.b) presents comparisons between two non-linear − relations = +
with coefficients (Table 2) given rby Sherwin et al. [14] (dashed line) and our result which is obtained
from the resolution of the ordinary differential equation (ODE) given by the cross-sectional compliance
(Table 4) (solid line). Figure (2.b) shows that Sherwin’s equation underestimates the area compared to our
equation. They present similar results for / <0.05.
All other equations remain quite similar to the result from our proposed equations until a value of /
equal to about 0.05. However, from / =0.1, the gap increases significantly with . Former linear
and non-linear relations underestimate the variation of cross-sectional area for a given transmural pressure.
At / =0.2, the difference is of 26.7% and 24.6% respectively for linear (Rammos) and non-linear
(Urquiza) relations.
0
A/A

Fig. 2: Comparison of − relations, / ( / ), (a) =0, from linear (Rammos, dashed line), non-
linear (Urquiza et al., dash-dotted line) equations and results from Eq. (14) solid line, (b) =0.5 from non-
linear Sherwin et al. (dashed line) and our solution (solid line) obtained from the resolution of ODE of
compliance (Table 4).

5.2. Variation of cross-sectional compliance and distensibility

Figure (3) shows a comparison between cross-sectional compliance obtained from the four equations
presented in table (4) namely, the linear with the coefficient of Rammos [11] (dashed line), the non-linear
with the coefficient of Urquiza et al. [15]) (dash-dotted line), Eq. (13) (thin solid line) and our proposed
Eq. (16) with =0.5 (thick solid line). At the opposite of figure (2) where results were similar for lower
values of / , results for compliance (figure 3) show more scatter. The difference increases with /
and therefore with (figure 3.b). For / =1.5, the gap of the non-linear − relation (with the
coefficient of Urquiza et al.) is of 43% with respect to our Eq. (16). Figure (4) presents results for

8
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

distensibility . Even if the difference between the four curves seems more important, the difference (of
the non-linear − relation) is the same as for cross-sectional compliance namely 43%.

In figures (3) and (4), all curves begin at a value of dimensionless cross-sectional compliance equal to 2
except our Eq. (16) which starts from a value of 1.5. This value is due to the term of the Poisson ratio, i.e.,
(1 − ) which is equal to 3/4 (for =0.5) and therefore at / =1 a value equal to 1.5. The
approximation of ≈ (3 )/(2 ℎ ) (for << ) [12] provides the same value / = 1.5
and confirms that this approximation is valid only for very small values of / . For larger values, the
dependency on (or ) should be considered (figure 3).

a)
/ A0
Cc1

b)
Fig. 3: Comparison of dimensionless cross-sectional compliance obtained by the four equations (Table 4).
9
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

a)

b)
Fig. 4: Comparison of dimensionless distensibility obtained by the four equations (Table 4).

5.3. Application to arterial vessels

In order to understand the implications for arterial vessels, we will consider the empirical relation for the
parameter β1 as a function of [12] namely

10
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

(18) 1
( 0) = 1 exp( 2 0) + 3

The constants =2.00 × 107 g.s−2.cm−1, =−22.53 cm−1 and =8.65 × 105 g.s−2.cm−1 were obtained by
fitting the data from Segers et al. (1998) [28], Stergiopulos et al. (1992) [29] and Westerhof et al. (1969)
[30]. Values of β1 obtained from Eq. (18) are therefore in g.s−2.cm−1.

Table 5
Parameters for =150mmHg (0.2×105Pa).

(cm) ×105(Pa) /
0.08 4.16 0.048
0.2 1.08 0.18
0.6 0.86 0.23

The application of the proposed equation to an arterial vessel with =150mmHg (Table 5) show an
increase in the diameter of 4 % for =0.8mm while it is of 30 % for =6mm. Figure (5) presents the
variation of the dimensionless cross-sectional area and diameter for different arterial vessels with initial
radius equal to 0.8, 2 and 6 mm. Results show that at =150mmHg, the dimensionless cross-sectional
area is more than 1.6 for =6mm, about 1.4 for =2mm and very small 1.08 for =0.8mm.

Fig. 5: Variation of dimensionless arterial vessel area and diameter for from 0 to 150 mmHg.

5. Conclusions

In this study, the pressure-area relation was revisited and a new relation was proposed. For the variation of
the cross-sectional area, the proposed result was obtained from the resolution of the ordinary differential
equation (ODE) given by the cross-sectional compliance or capacitance . The parameter = ℎ /
allowed us to write all results in dimensionless form and therefore to define a unique threshold value for
the applicability of the former equations. Comparisons with the so-called linear and non-linear −
equations show that results of cross-sectional area are similar until a value of dimensionless transmural
pressure / equal to about 0.05. Former equations could be used with an accepted gap until
/ =0.1. However, the inaccuracy increases with and at / =0.2, the difference is of 26.7%
and 24.6% respectively for the linear and non-linear relations. Results for the variation of arterial

11
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

compliance and distensibility were presented. These results showed non-negligible differences
between proposed and former equations. The proposed equation was applied to arterial vessels with
=150mmHg for radius from 0.8 to 6 mm. Results show an increase in the diameter of 4 % for
=0.8mm while the increase is of 30 % for =6mm.
This study was about the linear theory of elasticity. In future studies, for better understanding of the flow
behaviour in arterial vessels, we will consider fluid–structure interaction phenomena [31] and the influence
of transmural pressures on the change in tube shapes [32].

References

[1] Korsakova, N., Penkovskij, V., Shilova, A., Shevchenko, V., 2016. Model of blood circulation in the cerebral
cortex on the theory of fluid flow in heterogeneous medium. Series on Biomechanics 30(2), 5-14.
[2] Liepsch, D., Sindeev, S., Balasso, A., Frolov, S., 2017. Study of wall shear stress in an idealized 90-bifurcation
with Newtonian and non-Newtonian fluid models. Series on Biomechanics 31(1), 3-11.
[3] Wang, C. Y., 2017. Exact solution for laminar flow in partially collapsed tubes. Mechanics Research
Communications 83, 65-67.
[4] Absi, R., 2008. Analytical solutions for the modeled k equation, ASME Journal of Applied Mechanics 75(4),
044501-4.
[5] El Gharbi, N., Absi, R., Benzaoui, A., Bennacer, R., 2011, An improved near-wall treatment for turbulent channel
flows. International Journal of Computational Fluid Dynamics 25(1), 41-46.
[6] Morris, P.D., Narracott, A., von Tengg-Kobligk, H., Silva Soto, D.A., Hsiao, S., Lungu, A., Evans, P., Bressloff,
N.W., Lawford, P.V., Hose, D.R., Gunn, J.P., 2016. Computational fluid dynamics modelling in cardiovascular
medicine. Heart 102:18-28.
[7] Antonova, N., Dong, X., Tosheva, P., Velcheva, I., 2013. Numerical analysis of 3D blood flow and common
carotid artery hemodynamics in the carotid artery bifurcation without stenoses, Series on Biomechanics 28(3-4).
[8] Hoekstra, A.G., van’t Hoff, J., Artoli, A. M., Sloot, P. M. A., 2004. Unsteady flow in a 2D elastic tube with the
LBGK method, Future Generation Computer Systems 20(6), 917-924.
[9] Riahi, D.N., 2016. Modeling unsteady two-phase blood flow in catheterized elastic artery with stenosis,
Engineering Science and Technology, an International Journal 19(3), 1233-1243.
[10] Ghigo, A.R, Fullana, J.-M., Lagrée, P.-Y., 2017. A 2D nonlinear multiring model for blood flow in large elastic
arteries, Journal of Computational Physics 350, 136-165.
[11] Rammos, K.S., Koullias, G.J., Papou, T.J., Bakas, A.J., Panagopoulos, P.G., Tsangaris, S.G., 1998. A computer
model for the prediction of left epicardial coronary blood flow in normal, stenotic and bypassed coronary arteries, by
single or sequential grafting. Cardiovascular Surgery 6(6), 635-648.
[12] Olufsen, M.S., 1999. Structured tree outflow condition for blood flow in larger systemic arteries. American
Journal of Physiology 276(1), H257-268.
[13] Formaggia, L., Lamponi, D., Quarteroni, 2003. A. One-dimensional models for blood flow in arteries. Journal of
Engineering Mathematics 47, 251-276.
[14] Sherwin, S.J., Franke, V., Peiro, J., Parker, K., 2003. One-dimensional modelling of a vascular network in space-
time variables. Journal of Engineering Mathematics 47(3), 217-250.
[15] Urquiza, S.A., Blanco, P.J., Venere, M.J., Feijoo, R.A., 2006, Multidimensional modelling for the carotid artery
blood flow. Computer Methods in Applied Mechanics and Engineering 195, 4002-4017.
[16] Canic, S., 2002. Blood flow through compliant vessels after endovascular repair: wall deformations induced by
the discontinuous wall properties. Computing and Visualization in Science 4(3),147-155.
[17] Smith, N.P., Pullan, A.J., Hunter, P.J., 2002. An anatomically based model of transient coronary blood flow in
the heart. SIAM Journal on Applied Mathematics 62(3), 990-1018.
[18] Mynard, J.P., Davidson, M.R., Penny, D.J., Smolich, J.J., 2010. A numerical model of neonatal pulmonary atresia
with intact ventricular septum and RV-dependent coronary flow. Int. J. Numer. Meth. Biomed. Engng. 26(7), 843-
861.
[19] Valdez-Jasso, D., 2010. Modeling and Identification of Vascular Biomechanical Properties in Large Arteries,
PhD, North Carolina State University Raleigh, North Carolina.
[20] Mynard, J.P., Davidson, M.R., Penny, D.J., Smolich, J.J., 2012. A simple versatile valve model for use in lumped
parameter and one-dimensional cardiovascular models Int. J. Numer. Meth. Biomed. Engng. 28, 626-641.
[21] Sochi, T., 2014. The flow of Newtonian and power law fluids inelastic tubes. Int. J. Non-Linear Mech. 67, 245-
250.
[22] Caro, C. G., Pedley, T. J., Schroter, R. C., Seed W. A., 1978. The Mechanics of the Circulation. Oxford, UK:
Oxford Univ. Press.
12
Absi R., Series on Biomechanics, vol.32, 1, 2018, 47 - 59.

[23] McDonald, D.A., 1974. Blood Flow in Arteries (2nd ed.). London, UK: Arnold.
[24] Rockwell, R. L., Anliker, M., Elsner, J., 1974. Model studies of the pressure and flow pulses in a viscoelastic
arterial conduit. J. Franklin Inst. 297, 405-427.
[25] Tardy, Y., Meister, J., Perret, F., Brunner, H., Arditi, M., 1991. Non-invasive estimates of the mechanical
properties of peripheral arteries from ultrasonic and photoplethysmographic measurements. Clin. Phys. Physiol. Meas.
12, 39–54.
[26] Spencer, M.P., Denison, A.B.Jr., 1963. Pulsatile blood flow in the vascular system. In: Handbook of Physiology.
Circulation. Washington, DC: Am. Physiol. Soc., vol. II, chapt. 25, 842.
[27] Reneman, R.S., van Merode, T., Hick, P., Muytjens, A.M., Hoeks, A.P., 1986. Age-related changes in carotid
artery wall properties in men. Ultrasound Med Biol. 12, 465-471.
[28] Segers, P., Dubois, F., De Wachter, D., Verdonck, P., 1998. Role and relevancy of a cardiovascular simulator. J.
Cardiovasc. Eng. 3, 48–56.
[29] Stergiopulos, N., Young, D.F., Rogge, T.R., 1992. Computer simulation of arterial flow with applications to
arterial and aortic stenosis. J. Biomech. 25, 1477-1488.
[30] Westerhof, N., Bosman, F., DeVries, C.J., Noordergraaf, A., 1969. Analog studies of the human systemic arterial
tree. J. Biomech. 2, 121-143.
[31] Pielhop, K., Klaas, M., Schröder, W., 2015. Experimental analysis of the fluid–structure interaction in finite-
length straight elastic vessels. European Journal of Mechanics B/Fluids 50, 71–88.
[32] Nahar, S., Jeelani, S.A.K., Windhab, E.J., 2012. Influence of elastic tube deformation on flow behavior of a shear
thinning fluid. Chemical Engineering Science 75, 445–455.

13

You might also like