Biofluid Mechanics
Biofluid Mechanics
Biofluid
Mechanics
The Human Circulation
7328_C000.fm Page ii Monday, October 16, 2006 11:03 AM
7328_C000.fm Page iii Monday, October 16, 2006 11:03 AM
Biofluid
Mechanics
The Human Circulation
Krishan B. Chandran
Ajit P. Yoganathan
Stanley E. Rittgers
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources. Reprinted material
is quoted with permission, and sources are indicated. A wide variety of references are listed. Reasonable
efforts have been made to publish reliable data and information, but the author and the publisher cannot
assume responsibility for the validity of all materials or for the consequences of their use.
No part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic,
mechanical, or other means, now known or hereafter invented, including photocopying, microfilming,
and recording, or in any information storage or retrieval system, without written permission from the
publishers.
For permission to photocopy or use material electronically from this work, please access www.copyright.
com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC) 222 Rosewood
Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and
registration for a variety of users. For organizations that have been granted a photocopy license by the
CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation without intent to infringe.
Chandran, K. B.
Biofluid mechanics : the human circulation / Krishnan B. Chandran, Ajit P.
Yoganathan, and Stanley E. Rittgers.
p. ; cm.
Includes bibliographical references and index.
ISBN-13: 978-0-8493-7328-2 (hardcover : alk. paper)
ISBN-10: 0-8493-7328-X (hardcover : alk. paper)
1. Blood flow. 2. Fluid mechanics. I. Yoganathan, A. P. (Ajit Prithiviraj), 1951- II.
Rittgers, Stanley E. III. Title.
[DNLM: 1. Blood Circulation--physiology. 2. Cardiovascular Physiology. 3. Heart
Valves--physiology. WG 103 C456b 2007]
QP105.C43 2007
612.1’181--dc22 2006031653
Dedication
To my wife, Vanaja, and daughters, Aruna and Anjana, for their loving support.
KBC
To my wife, Tripti, and daughter, Anila, for their love, encouragement and support.
APY
To my wife, Eva, and sons, David and Andrew, who have taught me so much.
SER
7328_C000.fm Page vi Monday, October 16, 2006 11:03 AM
7328_C000.fm Page vii Monday, October 16, 2006 11:03 AM
Preface
The field of biomedical engineering has seen a rapid growth in the past
decade and numerous new undergraduate and graduate programs have
been established in many universities. With the increasing number of stu-
dents enrolled in these programs, there is also an increase in demand for
suitable text books. This work is an attempt to provide such a text book for
a course in the application of fluid mechanics to the study of the human
circulatory system. This book is intended as a first course on fluid mechanics
in the human circulation, which is suitable for senior undergraduate or for
first-year graduate students in biomedical engineering. The topics contained
in the various chapters were organized based on the experience gained by
the three authors in teaching courses on cardiovascular fluid mechanics in
their respective programs.
The book is organized into three parts. Part I consists of introductory
review material on fluid and solid mechanics and also a review of cardio-
vascular physiology pertinent to the topics covered in the subsequent chap-
ters. The first two chapters are an introduction to foundational material in
fluid and solid mechanics for those students not having had prior exposure
to these topics. In curricula where the students would have previously
enrolled in courses in fluid mechanics, mechanics of deformable bodies, and
in human physiology, this section may serve as a brief review.
Part II of the book discusses the fluid mechanics in the human circulation,
primarily applied to blood flow at the arterial level. The first chapter in this
section (Chapter 4) examines viscometry and the rheological behavior of
human blood. The solid mechanics of the arterial wall subject to transmural
pressure is also briefly discussed in this chapter, since the interaction
between blood and the arterial wall needs to be considered in unsteady
flow simulation. The application of steady flow models to derive some
useful diagnostic parameters, such as vascular resistance and Gorlin’s equa-
tions to describe time-averaged flow behavior past heart valves, is treated
next (Chapter 5). The third chapter in this section (Chapter 6) briefly
describes the Windkessel model for unsteady flow followed by a detailed
treatment of the Moens–Korteweg and Womersley models of pulsatile flow
in the human circulation. The relationship between flow-induced stresses
and the initiation and growth of atherosclerosis is also discussed with
qualitative treatment of flow in curved vessels, branches, and bifurcations,
as well as unsteady flow past stenoses and aneurysms. The final chapter in
this section (Chapter 7) examines the flow through native heart valves and
cardiac chambers.
7328_C000.fm Page viii Monday, October 16, 2006 11:03 AM
Krishnan B. Chandran
Ajit P. Yoganathan
Stanley E. Rittgers
7328_C000.fm Page ix Monday, October 16, 2006 11:03 AM
Contents
Part I
1
Fundamentals of Fluid Mechanics
CONTENTS
1.1 Introduction ....................................................................................................3
1.2 Intrinsic Fluid Properties..............................................................................4
1.2.1 Density ................................................................................................4
1.2.2 Viscosity ..............................................................................................4
1.2.3 Compressibility..................................................................................7
1.2.4 Surface Tension..................................................................................8
1.3 Hydrostatics....................................................................................................9
1.4 Macroscopic Balances of Mass and Momentum ....................................10
1.4.1 Conservation of Mass ..................................................................... 11
1.4.2 Conservation of Momentum .........................................................13
1.5 Microscopic Balance of Mass and Momentum.......................................15
1.5.1 Conservation of Mass .....................................................................15
1.5.2 Conservation of Momentum .........................................................18
1.5.3 Mathematical Solutions..................................................................22
1.6 The Bernoulli Equation...............................................................................26
1.7 Dimensional Analysis .................................................................................30
1.8 Fluid Mechanics in a Straight Tube..........................................................32
1.8.1 Flow Stability and Related Characteristics .................................33
1.8.1.1 Steady Laminar Flow .......................................................33
1.8.1.2 Turbulent Flow ..................................................................36
1.8.1.3 Flow Development ...........................................................38
1.8.1.4 Viscous and Turbulent Shear Stress...............................40
1.8.2 Effect of Flow Pulsatility................................................................41
1.9 Boundary Layer Separation .......................................................................44
1.10 Problems........................................................................................................45
1.1 Introduction
Before considering the mechanics of biological fluids in the circulation, it is
necessary to first consider some key definitions and specific properties. Once
established, these “pieces” will be used to construct important laws and
3
7328_C001.fm Page 4 Friday, October 6, 2006 3:48 PM
principles that are the foundation of fluid mechanics. To begin, we will define
what we mean by a fluid. In general, a material can be characterized as a
fluid if it deforms continuously under the action of a shear stress produced
by a force, which acts parallel to the line of motion. In other words, a fluid
is a material that cannot resist the action of a shear stress. Conversely, a fluid
at rest cannot sustain a shearing stress. For our applications, we will treat a
fluid as a continuum (i.e., it is a homogeneous material) even though both
liquids and gases are made up of individual molecules. On a macroscopic
scale, however, fluid properties, such as density, viscosity, etc., are reasonably
considered to be continuous.
1.2.1 Density
Density, commonly denoted by the symbol , is defined as the mass of a fluid
per unit volume and has units of mass/length3 [M/L3]. In the system meter/
kilogram/second (MKS) this would be represented by (kg/m3) or by (g/cm3)
in the CGS system (centimeter/gram/second) where 1 g/cm3 = 103 kg/m3.
Values of density for several common biofluids are:
1.2.2 Viscosity
As we said earlier, a fluid is defined as a material that deforms under the
action of a shear force. The viscosity of a fluid (or its “stickiness”), denoted
7328_C001.fm Page 5 Friday, October 6, 2006 3:48 PM
A
U Ft
h
y
FIGURE 1.1
Fluid subjected to simple shearing stress.
∂u
τ = −µ (1.1)
∂y
Power law
fluid
Yield
stress
⋅
Rate of shear-° (sec–1)
FIGURE 1.2
Shear stress vs. rate of shear plots for Newtonian and non-Newtonian fluids.
τ = K pl γ n (1.2)
where n is less than unity. Such fluids are classified as power law fluids.
Another class of fluids is known as Bingham plastics because they will
initially resist deformation to an applied shear stress until the shear stress
exceeds a yield stress, τy . Beyond that point, there will be a linear relationship
between shear stress and rate of shear. The constitutive relationship for a
7328_C001.fm Page 7 Friday, October 6, 2006 3:48 PM
τ = τy + µbγ (1.3)
where τy is the yield stress and µb is the plastic viscosity. Fluids that exhibit
a yield stress and also a nonlinear relationship between shear stress and rate
of shear may be classified as Casson fluids. The specific empirical relation-
ship for such fluids, which deviate from the ideal Bingham plastic behavior,
is known as Casson’s equation or
τ = τ y + k c γ (1.4)
1.2.3 Compressibility
The compressibility of a fluid is quantified by the pressure change required
to produce a certain increment in either the fluid’s volume or density. This
property, known as the Bulk Modulus (k), is defined as
∆p ∆p
k= d∨ = dρ (1.5)
∨ ρ
c = k/ρ (1.6)
7328_C001.fm Page 8 Friday, October 6, 2006 3:48 PM
πR2∆p
2πRσ
FIGURE 1.3
Pressure-surface, tension force balance for a hemispherical drop.
2 πRσ = ∆pπR 2
and, hence,
R∆p
σ= (1.7)
2
1.3 Hydrostatics
While most fluids in the body exist in a state of continuous motion, there
are important effects on the fluid, which are due to static forces. A fluid at
rest in a gravitational field, for example, is in hydrostatic equilibrium as
shown in Figure 1.4. Under these conditions, the weight of the fluid is exactly
offset by the net pressure force supporting the fluid. Here, the pressure at
the base of the fluid element is p while that at a distance dz above the base
is p plus the gradient of pressure in the z direction, dp/dz, times the incre-
mental elevation, dz, where p is the gauge pressure referenced to the atmo-
spheric pressure, pa (= 1.01 × 105 N/m2). Thus, if we sum the pressure and
gravitational forces acting on the element in the vertical direction, we get
dp
pdA − p + dz dA − ρgdAdz = 0 (1.8)
dz
dp
= −ρg (1.9)
dz
where g is the gravitational acceleration. The negative sign for the pressure
gradient indicates that p decreases as z increases. Integrating the above
dp
( p+ dz ) dA
dz
z
dA
rg dA dz
dz
pdA
FIGURE 1.4
Hydrostatic equilibrium of a fluid element.
7328_C001.fm Page 10 Friday, October 6, 2006 3:48 PM
equation between the base and the top of the element, z1 and z2, respectively,
with corresponding pressures p1 and p2 , yields
∫ ∫
p2 z2
dp = − ρgdz (1.10)
p1 z1
or
p2 − p1 = −ρg( z2 − z1 ) (1.11)
Thus,
∆p = ρgh (1.12)
Example 1.1:
For the case of a column of mercury where ρHg = 1.35 × 104 kg/m3, the
pressure increase over a height of 1 mm would be:
dA2
V2n
dA1
dV
A2
V1n
A1
FIGURE 1.5
Mass flux balance for a stream tube.
This theorem relates the time rate of change of each of these properties in a
System relative to corresponding changes that occur within and across a
Control Volume and is given by
dBsys ∂
dt
=
∂t ∫ bρd ∨ + ∫ bρVdA
CV CS
(1.13)
where B denotes the extensive (i.e., absolute) property, b denotes the amount
of that property per unit mass (i.e., b = B/m), CV denotes the Control Volume
and CS denotes the Control Surfaces of that volume.
The rate of mass carried across a surface is equal to the density times the
flow rate, ρQ, where Q is the volume flow rate, or ρVn A, and Vn is the com-
ponent of velocity normal to the cross section. Furthermore, the rate of
change of mass within the tube is given by
∂
∂t ∫ ρd ∨
CV
Rate of Mass In = ∫ ρV dA
A1
1n
and
where V1n and V2 n are the velocities normal to the differential areas at cross
sections 1 and 2, respectively.
Inserting these into the law of Conservation of Mass, gives us
∂
∫
− ρV2 ndA +
A2
∫ ρV dA = ∂t ∫ ρd ∨
A1
1n
∨
(1.14)
If the velocity varies over the cross section as a function of radius, r, then
the mean velocities must first be obtained by
∫ V (r)dA
1
V1 = 1
A1
A1
7328_C001.fm Page 13 Friday, October 6, 2006 3:48 PM
and
∫ V (r)dA
1
V2 = 2
A2
A2
∑ F = ma (1.16)
Rewriting a as dV
dt
and bringing m inside the differential (since it is con-
stant for a System), results in
dV d(mV )
∑ F = m = dt
dt
(1.17)
∑F ∫ Vρ d ∨
D
sys = (1.18)
Dt
sys
When the System and the Control Volume are coincident at an instant of
time, the forces acting on the System and the forces acting on the Control
Volume are identical or
∑F = ∑F sys CV (1.19)
∂
∫ Vρ d ∨ = ∂t ∫ Vρ d ∨ + ∫ VρV ⋅ n dA
D
(1.20)
Dt
sys CV CS
Therefore, for a Control Volume that is fixed (i.e., with respect to an inertial
reference) and nondeforming
∂
∑F CV =
∂t ∫ Vρ d ∨ + ∫ VρV ⋅ n dA
CV CS
(1.21)
∑F CV = ∫ ρ g d ∨ + ∫ t dA
CV CS
(1.22a)
where g is the body force per unit mass acting on the Control Volume con-
tents and t is the stress vector acting on the Control Volume surfaces.
When viscous effects are important, the surface area integral of the stress
vector is nonzero and must be determined empirically (i.e., from experi-
mental data) and given as friction factors or drag coefficients (i.e., constants
relating drag forces to other variables). However, if viscous effects are
negligible, then the stress vector is given by
t = − np
ˆ (1.22b)
where n̂ is the outward directed unit vector normal to the surface and p is
the pressure.
The previous two balances of mass and momentum are called macroscopic
or integral balances because they consider the Control Volume as a large,
discrete space and are written in terms of bulk-flow variables. In order to
derive more general forms of these equations, which provide spatial detail
throughout the flow field, we need to take a microscopic or differential
approach. Such an approach leads us to what are commonly referred to in
fluid mechanics literature as the Continuity and the Navier–Stokes
Equations, for Conservation of Mass and Momentum, respectively.
7328_C001.fm Page 15 Friday, October 6, 2006 3:48 PM
V = uiˆ + vjˆ + wkˆ (1.23)
1. The net rate of mass flux across the control surfaces (Figure 1.6b)
in the x direction:
y
y
ρv|y+∆y
Control
v volume ρw|z+∆z
dy
u
w O
ρu|x ρu|x+∆x
dz
dx ρw|z
x x
ρv|y
z z
(a) (b)
FIGURE 1.6
(A) Differential control volume in rectangular coordinates. (B) Mass flux across surfaces of
control volume.
7328_C001.fm Page 16 Friday, October 6, 2006 3:48 PM
y direction:
z direction:
∂
(ρ∆x∆y∆z)
∂t
∂(ρu)
(ρu|x −ρu|x+∆x ) = ⋅ ∆x
∂x
etc.
Since the volume within the control element is time invariant, we can divide
each term by ∆x∆y∆z . Then, in the limit as ∆x∆y∆z approaches zero, we
obtain
∂ ∂ ∂ ∂ρ
(ρu) + (ρv) + (ρw) + =0 (1.25)
∂x ∂y ∂z ∂t
which is equivalent to
∂ρ
∇ ⋅ ρV + =0 (1.26)
∂t
The above equation is known as the Continuity Equation where the “del”
vector operator, ∇, is defined as
∂ ∂ ∂
∇=i +j +k (1.27)
∂x ∂y ∂z
7328_C001.fm Page 17 Friday, October 6, 2006 3:48 PM
ρ∇ ⋅ V = 0 (1.28)
or
∇ ⋅V = 0 (1.29)
Note that this equation is valid for both steady and unsteady (including
pulsatile, three dimensional) flows of an incompressible fluid.
If we return to Equation 1.24 and differentiate each numerator, we obtain
∂ρ ∂ρ ∂ρ ∂ρ ∂u ∂v ∂w Dρ ∂u ∂v ∂w
+ u + v + w + ρ + + = + ρ + + =0
∂t ∂x ∂y ∂z ∂x ∂y ∂z Dt ∂x ∂y ∂z
(1.30)
or
Dρ
+ ρ∇ ⋅ V = 0 (1.31)
Dt
where
D ∂ ∂ ∂ ∂
= +u +v +w
Dt ∂t ∂x ∂y ∂z
DT ∂T ∂T ∂T ∂T
= +u +v +w (1.32)
Dt ∂t ∂x ∂y ∂z
∂T
where ∂t
is the local time rate of change of temperature and the terms
∂T ∂T ∂T
u +v +w
∂x ∂y ∂z
7328_C001.fm Page 18 Friday, October 6, 2006 3:48 PM
represent the rate of change of temperature due to fluid motion (also known
as Convection).
Finally, for many fluid dynamic situations, such as the consideration of
liquids or gases at low speeds, it is quite common to assume an incompress-
ible fluid where ρ = Constant. In this case, Equation 1.31 can be further
reduced to
∇ ⋅V = 0
∑F (
ρV V ⋅ n dA ) ∂ ρVd ∨
lim
∆ ∨→0 ∆x ∆y ∆z
= lim
∆ ∨→0 ∫∫ ∆x∆y∆z
+ lim
∆ ∨→0 ∂t ∫∫∫ ∆x∆y∆z
(1.33)
lim
∑F
= dF
∆ ∨→0 ∆x∆y∆z
7328_C001.fm Page 19 Friday, October 6, 2006 3:48 PM
ρV (V ⋅ n)dA ∂ ∂ ∂
lim
∆∨→0 ∫∫ ∆x∆y∆z
= (ρVu) + (ρVv) + (ρVw)
∂x ∂y ∂z
∂ ∂ ∂ ∂V ∂V ∂V
= V (ρu) + (ρv) + (ρw) + ρ u +v +w
∂x ∂y ∂z ∂x ∂z ∂z
∂ ∂ ∂ ∂ρ
(ρu) + (ρv) + (ρw) = − (1.34)
∂x ∂y ∂z ∂t
ρV (V ⋅ n)dA ∂ρ ∂V ∂V ∂V
lim
∆∨→0 ∫∫ ∆x∆y∆z
= −V
∂t
+ ρ u
∂x
+v
∂y
+w
∂z
Vρd ∨ ∂ ∂V ∂ρ
lim
∆∨→0 ∫∫∫ = (ρV ) = ρ
∆x∆y∆z ∂t ∂t
+V
∂t
∂V ∂V ∂V dV
dF = ρ u +v +w +ρ (1.35)
∂x ∂y ∂z dt
The surface forces acting on the control volume are those due to the normal,
σ, and the shear, τ, stresses. These stresses can be assumed to vary contin-
uously from their nominal value at the center of the Control Volume in each
7328_C001.fm Page 20 Friday, October 6, 2006 3:48 PM
y ∂yx dy
yx +
∂y 2
zx − ∂zx dz
∂z 2
σxx _ ∂σxx dx
∂x 2
σxx + ∂σxx dx
∂x 2
∂yx dy
yx _
∂y 2
zx + ∂zx dz
∂z 2
x
FIGURE 1.7
Normal and shear stresses along the x coordinate in the Control Volume.
of the coordinate directions. Figure 1.7 depicts the normal and shear stresses
acting on the Control Volume in the x direction alone. Similar figures can be
constructed for the normal and shear stresses acting in the y and z directions.
Thus, the net surface force acting in the x direction is given by
∂σ ∂τ yx ∂τ
FSx = xx ∆x∆y∆z + ∆x∆y∆z + zx ∆x∆y∆z (1.37)
∂x ∂y ∂z
The total force acting on the control volume in the x direction then becomes
∂σ xx ∂τ yx ∂τ zx
dFx = ρgx + + + (1.39a)
∂x ∂y ∂z
∂σ yy ∂τ xy ∂τ zy
dFy = ρg y + + + (1.39b)
∂y ∂x ∂z
∂σ zz ∂τ xz ∂τ yz
dFz = ρgz + + + (1.39c)
∂z ∂x ∂y
7328_C001.fm Page 21 Friday, October 6, 2006 3:48 PM
Substituting the results of the above expressions 1.39 a–c back into the
expression for Newton’s Second Law (Equation 1.17) yields
∂σ xx ∂τ yx ∂τ zx ∂u ∂u ∂u ∂u
ρgx + + + = ρ + u + v + w (1.40a)
∂x ∂y ∂z ∂t ∂x ∂y ∂z
∂τ xy ∂σ yy ∂τ zy ∂v ∂v ∂v ∂v
ρg y + + + = ρ + u + v + w (1.40b)
∂x ∂y ∂z ∂t ∂x ∂y ∂z
∂τ xz ∂τ yz ∂σ zz ∂w ∂w ∂w ∂w
ρgz + + + = ρ +u +v +w (1.40c)
∂x ∂y ∂z ∂t ∂x ∂y ∂z
In this form, we can see that the right-hand side of the above equa-
tions actually represents density (mass/volume) × acceleration, or force/
volume, where the acceleration terms can be separated into local accel-
eration (∂u/∂t, etc.) and convective acceleration (u ∂u/∂x, etc.) compo-
nents. The total acceleration can be expressed in terms of the substantive
derivative as
Du ∂u ∂u ∂u ∂u
= +u +v +w (1.41)
Dt ∂t ∂x ∂y ∂z
Equation 1.40a to Equation 1.40c represent the complete form of the differ-
ential Conservation of Momentum balances. These equations cannot be
solved, however, because there are more unknowns (i.e., dependent variables)
than equations. Thus, it is necessary to derive additional information in order
to provide those equations. In practice, we will consider the special, although
not uncommon, case of incompressible Newtonian fluids. Here, the normal
and shear stresses can be expressed as
∂u ∂u ∂v
σ xx = − p + 2µ τ xy = τ yx = µ + (1.42a)
∂x ∂y ∂x
∂v ∂w ∂v
σ yy = − p + 2µ τ yz = τ zy = µ + (1.42b)
∂y ∂y ∂z
∂w ∂u ∂w
σ zz = − p + 2µ τ zx = τ xz = µ + (1.42c)
∂z ∂z ∂x
7328_C001.fm Page 22 Friday, October 6, 2006 3:48 PM
∂p ∂2 u ∂2 u ∂2 u ∂u ∂u ∂u ∂u
ρgx − + µ 2 + 2 + 2 = ρ + u + v + w (1.43a)
∂x ∂x ∂y ∂z ∂t ∂x ∂y ∂z
∂p ∂2 v ∂2 v ∂2 v ∂v ∂v ∂v ∂v
ρg y − + µ 2 + 2 + 2 = ρ + u + v + w (1.43b)
∂y ∂x ∂y ∂z ∂t ∂x ∂y ∂z
∂p ∂2 w ∂2 w ∂2 w ∂w ∂w ∂w ∂w
ρgz − + µ 2 + 2 + 2 = ρ +u +v + w (1.43c)
∂z ∂x ∂y ∂z ∂t ∂x ∂y ∂z
TABLE 1.1
Continuity and Momentum Equations in Cartesian, Cylindrical (Polar),
and Spherical Coordinate Systems
(a) Cartesian Coordinate System:
∂u ∂v ∂w
+ + =0
∂x ∂y ∂z
∂u ∂u ∂u ∂u ∂p ∂2 u ∂2 u ∂2 u
ρ + u + v + w = ρBx − +µ 2 + 2 + 2
∂t ∂x ∂y ∂z ∂x ∂x ∂y ∂z
∂v ∂v ∂v ∂v ∂p ∂2 v ∂2 v ∂2 v
ρ + u + v + w = ρBy − +µ 2 + 2 + 2
∂t ∂x ∂y ∂z ∂y ∂x ∂y ∂z
and
∂w ∂w ∂w ∂w ∂p ∂2 w ∂2 w ∂2 w
ρ +u +v +w = ρBz − +µ 2 + 2 + 2
∂t ∂x ∂y ∂z ∂z ∂x ∂y ∂z
1 ∂ ( rVr ) 1 ∂V ∂V
θ
∇ ⋅V = + + z =0
r ∂r r ∂θ ∂z
∂V ∂V V ∂Vr Vθ2 ∂V
ρ r + Vr r + θ − + Vz r
∂t ∂r r ∂θ r ∂z
∂V ∂V V ∂V V V ∂V
ρ θ + Vr θ + θ θ + r θ + Vz θ
∂t ∂r r ∂θ r ∂z
and
∂V ∂V V ∂Vz ∂V
ρ z + Vr z + θ + Vz z
∂t ∂r r ∂θ ∂z
(continued)
7328_C001.fm Page 24 Friday, October 6, 2006 3:48 PM
and
∂u
=0
∂x
7328_C001.fm Page 25 Friday, October 6, 2006 3:48 PM
or
u = f ( y) + C
where C is the integration constant, and Equation 1.44 would reduce to
∂p ∂2 u
= µ 2
∂x ∂y
Here, it is now possible to solve for u( y) as an explicit function of p.
The other approach taken is to solve these equations numerically. This
approach is more complex, but provides the ability to solve problems with-
out making unrealistic simplifying assumptions. The basic technique is to
first sub-divide the flow into many small regions, or elements, over which
the governing equations are applied. Rather than use the differential form
of the equations, however, they are rewritten in algebraic form in terms of
changes that occur in variables due to incremental changes in position and
time. Solutions are then obtained locally at specific locations, or nodes, on
the finite elements across a mesh of elements (Figure 1.8). This set of solutions
1 3
FIGURE 1.8
CFD Mesh for a reconstructed human coronary artery with a stenosis segment. It can be
observed that a finer mesh is employed in the area of stenosis in order to obtain more accurate
results in this region of interest.
7328_C001.fm Page 26 Friday, October 6, 2006 3:48 PM
is then updated at subsequent time intervals over the entire mesh until some
acceptable level of accuracy, or tolerance, is achieved based upon conver-
gence between successive values of certain output variables. Obviously, this
can be a very detailed and time-consuming process depending upon the
complexity of the geometry being analyzed and the initial and boundary
conditions imposed. Furthermore, additional features are sometimes
included in the equations to allow for simulation of non-Newtonian fluids
and turbulent flow conditions, for example. While it is always important to
validate such results, these computational fluid dynamic (CFD) software
programs are increasingly being used to solve challenging biomedical flow
problems unapproachable by any other means.
∂V 1
+ (V ⋅∇)V = − ∇p + g (1.46)
∂t ρ
As can be seen, the shear stress terms have been set equal to zero since
there is no friction in the system. To further simplify the equation of motion,
we will focus on flow along a streamline (Figure 1.9). A streamline is an
ds
∂p dz ) dA
( p + ∂z
V
θ dz
p dA
dA
pg dA ds
FIGURE 1.9
Force balance for a fluid particle along a stream line.
7328_C001.fm Page 27 Friday, October 6, 2006 3:48 PM
imaginary line in the flow field drawn so that it is always tangential to the
velocity vectors. The sum of the forces acting on a fluid particle along the
streamline is given by
∂p
∑ F = pdA − p + ∂s ds dA − ρg(sin θ)dAds
S (1.47)
where
dz
sin θ =
ds
∂p
∑ F = − ∂s ds − ρgdz dA
s (1.48)
dV ∂V ∂V
= +V (1.49)
dt ∂t ∂s
tan g
Taking the mass of the particle as ρdAdS and inputting Equation 1.48 and
Equation 1.49 into Newton’s Second Law of Motion gives us
∂V ∂V ∂p
ρdAdS +V = − ∂s ds − ρgdz dA (1.50)
∂t ∂s
2
Since an equivalent mathematical expression for the term V ∂∂Vs is ∂∂s V2 , we
will now substitute this into Equation 1.49 and integrate between any two
arbitrary points 1 and 2 to obtain
(V 2 − V12 ) + (p − p ) + ρg(z − z ) = 0
2
∂V
ρ ∫
1
∂t
ds + ρ 2
2 2 1 2 1 (1.51)
deceleration between the two locations and for steady flow conditions
(where ∂∂t = 0 ), it then becomes zero. Furthermore, since the points 1 and 2
are arbitrary, we can write a general expression of the steady flow Bernoulli
equation as
V2
p+ρ + ρgz = H (1.52)
2
where p , V, and z are evaluated at any point along the streamline. Here, H
is a constant and is referred to as the “total head” or total energy per unit
volume of fluid. This terminology is used because, even though we derived
this equation based on Conservation of Momentum principles, the terms in
Equation 1.52 all have the units of F/L2, which equivalently can be written
as FL/L3 or Energy/Volume.
In the absence of frictional losses, then, the Bernoulli equation states that
the total mechanical energy per unit volume at any point remains constant.
The mechanical energy of a system consists of a pressure energy component
(due to static pressure), a kinetic energy component (due to motion), and a
potential energy component (due to height) assuming there are no thermal
energy changes. Thus, between any two points, the form of the mechanical
energy may change from kinetic to potential and vice versa, for example,
but the total mechanical energy does not change.
The Bernoulli equation also states that the pressure drop between two
points located along a streamline at similar heights is only a function of the
velocities at these points. Thus, for z1 = z2, Bernoulli’s equation simplifies to
1
p1 − p2 = ρ(V22 − V12 ) (1.53)
2
Finally, when the downstream velocity is much higher than the upstream
velocity (V2 V1 ) , ρV12 can be neglected and Equation 1.53 then can be fur-
ther simplified to
1 2
p1 − p2 = ρV (1.54)
2 2
In the specific case of blood flow where we can assume a density of 1.06 g/cm3,
Equation 1.55 becomes
p1 − p2 = 4V22 (1.55)
Example 1.2:
A patient has a stenotic aortic valve producing a pressure drop between the
left ventricle and the aorta. The mean velocity in the left ventricle proximal
(upstream) to the valve is 1 m/s, while the mean velocity in the aorta distal
(downstream) of the valve is 4 m/s. Applying Equation 1.55, the pressure
drop across the valve is:
p1 − p2 = 4V22
= 4(4 ms )
2
= 64 mmHg
By comparison, if we include the proximal velocity in the calculation, then
the pressure drop reduces to
p1 − p2 = 4 (V22 − V12 )
= 4 ( 4 m s ) − (1 m s )
2 2
= 60 mmHg
∂V
ρ + ρV ⋅ ∇V = −∇p + ρ g + µ∇ 2 V (1.56)
∂t
where ρ ∂∂Vt is the local acceleration, ρV ⋅ ∇V is the convective accelera-
tion, −∇p is the pressure
force per unit volume, ρg is the body force per unit
volume, and µ∇ 2 V is the viscous forces per unit volume.
If we represent
g = −∇φ (1.57)
where
φ = −( g x x + g y y + g z z ) (1.58)
then,
−∇p + ρ g = −∇p − ρ∇φ = −∇( p + ρφ) = −∇p (1.59)
or
∂V
ρ + ρV ⋅ ∇V = −∇p + µ∇ 2 V (1.60)
∂t
ρV 2
Now, the ratio of 0 L0 also has the dimension of ( F⋅L) ( L3 ⋅L) , so multiplying
L
every term by 0 ρV 2 would result in
0
V
∂
Vo V V p µ 2 2 V
+ ⋅ ( L ∇) = ( Lo∇) 2 + ( Lo∇ )
V Vo o Vo
(1.61)
ρVo ρVo Lo Vo
∂t o
Lo
V
= , dimensionless velocity
Vo
P
P = 2 , dimensionless pressure
ρVo
υ
Τ = t o , dimensionless time
Lo
ρV L
Re = o o , dimensionless parameter, called the Reynolds number
µ
∂ p+ 1 ∇
= −∇ 2
+ ⋅∇ (1.62)
∂T Re
⋅ = 0
∇ (1.63)
Examination of Equation 1.62 and Equation 1.63 shows that, for a given
value of Re, there is only one solution to these equations for each driving
. One consequence of this is that by matching values of Re for
function, ∇p
various systems, which are similar but of different scale, it is possible to
achieve dynamic similarity and obtain predictive information about proto-
types from scaled model measurements.
7328_C001.fm Page 32 Friday, October 6, 2006 3:48 PM
Example 1.3:
It is desired to study the fluid dynamics in a coronary artery (diameter = 3 mm)
using a laser Doppler velocimetry (see Chapter 10, section 10.7: Laser
Doppler Velocimetry). However, the resolution of the LDV system is 300 µm,
resulting in a maximum of 10 velocity points across the artery diameter and
low accuracy for obtaining wall shear stresses from the fitted velocity profile.
In order to improve the resolution by ×10, an in vitro model is constructed
with a diameter of 3 cm. What flow rate should be used to ensure dynamic
similarity between the model and the prototype? (Assume the working fluid
has the same kinematic viscosity as blood — ν = 0.035 cm2/s).
Since Dp = 3 mm, Dm = 30 mm, dynamic similarity can only be achieved if
Rem = Re p
or
VmDm VpDp
=
νm νp
For νm = ν p,
Dp
Vm = V = 0.1Vp
Dm p
πD2 π10Dp2
Qm = Vm m = 0.1Vp = 10Qp
4 4
dr
r τ(2πr dx)
dp
(p + dx) 2πr dr
dx
p(2πr dr) dx
FIGURE 1.10
Force balances for steady flow through a straight, horizontal, circular tube.
the viscous forces. This ratio is classically known as the Reynolds Number
(Re), which is dimensionless since both terms have units of force [F]. It is
defined as
where ρ [kg/m3] is the density of the fluid, V [m/s] is the average velocity
of the fluid over the cross section of the tube, d [m] is the tube diameter,
and µ [kg/m·s] is the dynamic viscosity of the fluid. Although inertial forces
obviously begin to dominate for Re > 1, it has been determined experimen-
tally that in a smooth-surfaced tube, flow is laminar for all conditions where
Re < 2100. Furthermore, if the tube is long enough to have stabilized any
entrance effects (see section 1.8.1.3: Flow Development, below), then the
velocity profile takes on a parabolic shape and the flow is called fully devel-
oped laminar flow.
In section 1.4: Macroscopic Balances of Mass and Momentum, we dis-
cussed the principle of Conservation of Momentum and derived the
Navier–Stokes Equations. For incompressible Newtonian flow, the equation
of motion in vector notation is given by Equation 1.45
∂V 1
+ (V ⋅∇)V = − ∇p + g + ν(∇ 2V ) (1.45)
∂t ρ
where V [m/s] is the velocity vector, p [Pa] is the pressure, g [m/s2] is the
gravitational acceleration, and ρ [N-s2/m4] and ν [m2/s] are the density and
the kinematic viscosity of the fluid, respectively.
If we now apply these equations (in cylindrical coordinates, [Table 1.1]) to
the case of steady flow in a straight, circular, horizontal tube (Figure 1.10),
then the momentum balance in the z (axial) direction reduces to
∂p 1 ∂ ∂Vz
− +µ r = 0 (1.65)
∂z r ∂r ∂r
since the time rate of change (i.e., ∂ ∂t ), secondary velocity (i.e., Vr and Vθ ),
∂V
and circumferential velocity gradient (i.e., ∂θz ) terms are zero. As a conse-
∂V
quence, the conservation of mass balance results in z ∂z also being zero.
Rearranging terms then yields
∂p 1 ∂ ∂Vz
= µ r (1.66)
∂z r ∂r ∂r
7328_C001.fm Page 35 Friday, October 6, 2006 3:48 PM
dp 1 d dVz
= µ r (1.67)
dz r dr dr
We can further observe that, for the two terms of the equation (i.e., L.H.S.
and R.H.S.) to be equal for all values of the independent variables r and z
(each of which is only present in one of the terms), each term must be
constant. Equation 1.67 can then be integrated twice to yield
dVz 1 dp c
= r+ 1
dr 2µ dz r
1 dp 2
Vz = r + c1 ln r + c2 (1.68)
4µ dz
The constant terms, c1 and c2, can be evaluated by applying known values
of axial velocity at specific boundary locations. For example, Vz = 0 at r = R
due to the “no-slip” condition at the tube wall. The value of Vz , however,
is not known at the tube center, r = 0, although we can assume that it is a
maximum at that point due to overall symmetry of the tube. Thus, the
dV
appropriate boundary condition here is that drz = 0, which requires that c1 = 0
and which also constrains all velocities to be finite.
Evaluating c2 and substituting it into Equation 1.68 results in
1 dp 2
Vz = [r − R 2 ] (1.69)
4µ dz
∆pR 2 r
2
Vz (r ) = 1 −
4µL R
or
r 2
Vz (r ) = Vmax 1 − (1.70)
R
7328_C001.fm Page 36 Friday, October 6, 2006 3:48 PM
where V [m/s] is the velocity of the fluid at a distance r [m] from the center
of the tube, Vmax [m/s] is the maximum (centerline) velocity, R [m] is the
radius of the tube, d [m] is the diameter of the tube, and ∆p [Pa] is the
pressure drop along a length L [m] of the tube. By integrating this velocity
profile over the tube’s cross section and dividing by the area, we can obtain
the average velocity,
Vmax
Vave = (1.71)
2
Since flow rate in the tube, Q [m3/s], is equal to the average velocity, Vave ,
times the cross-sectional area, we can write
Vmax
Q = Vave ( πR 2 ) = ( πR 2 )
2
(1.72)
∆pR 2 ∆pπR 4
= ( πR 2 ) =
8µL 8µL
∆pπ d 4
Q= (1.73)
128µ L
µLQ
∆p = 128 (1.74)
πd 4
1.0
n = 10
n=6
Laminar n=8
r
R 0.5
Turbulent
0
0 0.5 1.0
–
µ
Vc
FIGURE 1.11
Plots of laminar (Poiseuille) and turbulent velocity profiles for n = 6, 8, and 10.
that for laminar flow, although it is not considered uniform or “plug” flow
(Figure 1.11).
In order to quantitatively describe turbulent flow, we think of the instan-
taneous velocity, V (t), at a given radial location as the sum of a time-averaged
mean velocity, V [m/s], and a randomly fluctuating velocity component,
V ′ [m/s], or
V(t) = V + V ′ (1.75)
Expressing turbulent flow in this way allows us to separate out one com-
ponent of a given variable, which can be quantified and expressed as an
analytic function while also allowing us to modify each variable by super-
imposing a random, disorderly motion on top of it. In the case of velocity,
the mean velocity profile (i.e., the quantifiable component) of a fully devel-
oped turbulent flow at Re < 105 is described by
1
r n
V = V max 1 − (1.76)
R
7328_C001.fm Page 38 Friday, October 6, 2006 3:48 PM
Vrms
I= 100% (1.77)
V ave
where Vrms = V ′ 2 is the root mean square of the fluctuating velocity [m/s].
It should be emphasized that blood flow in the circulatory system is
normally laminar, although in the ascending aorta it can destabilize briefly
during the deceleration phase of late systole; however, this time period is
generally too short for flow to become fully turbulent. Certain disease con-
ditions, though, can alter this condition and produce turbulent blood flow,
particularly downstream of an arterial stenosis (a vessel narrowing, usually
due to atherosclerosis), distal to defective (i.e., stenotic or regurgitant) heart
valves, or distal to prosthetic heart valves.
r
U0 u D
x
FIGURE 1.12
Flow in the entrance region of a pipe.
1
Le = 0.693 ⋅ d ⋅ Re 4 for turbulent flow (1.79)
7328_C001.fm Page 40 Friday, October 6, 2006 3:48 PM
Example 1.4:
For flow in the proximal aorta with a diameter of ~ 2.5 cm (0.025 m) and an
average velocity of ~ 0.25 m/s, the Reynolds number would be:
ρVD
Re = = (1.07 g/cm3 × 25 cm/s × 2.5 cm)/(0.035 Poise) [in cgs system]
µ
~ 1,900
Example 1.5:
Flow in the intercostal arteries derives from the thoracic aorta (diameter of
~ 2.0 cm) as they branch off to perfuse the musculature of the chest wall. These
vessels have a diameter of ~ 2 mm with an average velocity of ~ 10 cm/s.
Therefore, the Reynolds number would be:
ρVD
Re = = (1.07 g/cm3 × 10 cm/s × 0.2 cm)/(0.035 Poise)
µ
~ 60
In this case, the entrance length would be on the order of 0.7 cm. As a
result, flow development occurs in a relatively short distance and the major-
ity of the intercostal artery (several centimeters long) experiences Poiseuille-
type flow (in the mean).
dV
τ = −µ (1.1)
dr
where V is the velocity [m/s] at radial position r [m] and µ is the dynamic
viscosity [N-s/m2] of the fluid. For turbulent flow, we said that the boundary
layer is very thin and the velocity profile is flatter than for laminar flow. The
combination of a flatter profile with the nonslip condition at the wall results
in much higher shear rates and, thus, wall shear stresses for turbulent flow
are large as compared to laminar flow. For either case (laminar or turbulent),
the wall shear stress can be determined from a force balance within a control
volume if the pressure drop, ∆p, is known along a length L of the tube (refer
again to Figure 1.10)
d ∆p
τ wall = − (1.80)
4 L
that blood flow in the heart and arteries is quite pulsatile, meaning that
the pressure and velocity profiles vary periodically (here, the period
denotes the duration of a cardiac cycle [s]) with time. When the heart
contracts during systole, a pressure originates from the left ventricle, which
then travels out as a wave due to the elasticity of the arteries. The most
immediate consequence of this is that the pressure in the left ventricle
(upstream) exceeds that in the aorta (downstream) resulting in opening of
the aortic valve and the blood being ejected from the heart. Due to the
compliant nature of the arteries and their finite thickness, the pressure
travels like a sound wave at a speed, which is much faster than the flow
velocity (~ 500:1). The pulsatile nature of the flow also affects the pressure
distribution out into the vessels and the velocity profiles within them as
well as the point of transition from a laminar to a turbulent regime. This
later effect is due to the fact that flow accelerates rapidly in early systole,
when, based on the instantaneous Reynolds number in the ascending aorta,
flow would be expected to be turbulent during a major part of systole.
Despite this, however, it has been observed that aortic blood flow remains
laminar and well streamlined under these conditions. The reason for this
is partly due to the stabilizing effect that systolic acceleration has on the
flow. One might anticipate, then, that the onset of turbulence would occur
during the later part of systole when flow decelerates. This has, in fact,
been shown by the presence of erratic fluctuations in velocity recordings
(Figure 1.13).
100
50
Velocity (cm s–1)
–20
FIGURE 1.13
Disturbances during flow deceleration phase in a pulsatile (periodic unsteady) flow velocity
patterns in the aorta. (Redrawn from Seed, W.A. and Wood, N.B., Cardiovascular Research
5: 319–330, 1971. With permission.)
7328_C001.fm Page 43 Friday, October 6, 2006 3:48 PM
The flow destabilization that occurs in late systole is mainly due to the
development of an adverse (i.e., against the flow) pressure gradient during
the flow deceleration, but that is not the entire reason. Additionally, flow
in the aorta doesn’t become turbulent as soon or to the degree expected
based on the instantaneous Re values because there is not sufficient time
available for flow to become turbulent. To understand this, think of tur-
bulence as a process that requires both a minimum amount of energy to
be present and a triggering or seeding mechanism to initiate it. In this
case, the triggering is provided by some “disturbance” to the flow (e.g.,
a surface roughness or an external vibration) at some site. For “fully
developed” turbulence to occur, however, this disturbance must propagate
throughout the entire flow field. With steady flow, these conditions are
always present, but with pulsatile flow, they only exist for a limited time
(up to the cycle period) and then they reverse. Therefore, in a healthy
artery, turbulence is generally absent because the flow destabilization time
interval (mid-to-end systole) is short (~150 ms) and is immediately followed
by diastole when the velocity and instantaneous Re are low. When the
subsequent systolic phase begins, a new acceleration phase restabilizes
the flow. It should be pointed out here that, although complex vortex-
containing flows are present in regions like bifurcations or branches, such
flow is not necessarily turbulent in the sense that it is either random or
unpredictable. Many of these flows contain very laminar flow behavior,
just not simply in linear directions.
Based on the above discussion, Re alone is clearly not enough to completely
characterize pulsatile flow. Another dimensionless parameter, the Womersley
number (denoted as α), is also used to characterize the periodic nature of
blood flow. The definition of α is:
d ρω
α= (1.81)
2 µ
Example 1.6:
The ascending aorta has a diameter, d = 2.5 cm and the heart rate is 70 bpm
(beats per minute). The density, ρ, and the viscosity of blood, µ, are 1.06 g/cm3
and 0.035 Poise, respectively. Based on these data, the Womersley parameter
would be:
d ρω
α=
2 µ
What this example illustrates is that flow in the major arteries is charac-
terized by both large mean flow and large oscillatory flow components due
to the intermittent nature of the heart pump. As we’ll see later in Chapter 6
(Unsteady Flow and Nonuniform Geometric Models), the Womersley param-
eter is an important index of the net amount of forward flow and the shape
of the local velocity profile.
throughout the body. For this redirection to occur, a pressure gradient must
exist. The presence of an adverse pressure gradient (i.e., a pressure gradient
acting against the flow direction and tending to decrease the fluid velocity)
may have a magnitude that is large enough to cause the fluid within the
boundary layer to stop and deflect away from the surface. When this hap-
pens, boundary layer separation is said to have occurred. Boundary layer
separation begins at a separation point on the boundary, which is defined
as a location where, not only is the flow velocity, u, equal to zero, but the
rate of change of u along a direction normal to the surface, n, is also zero,
or ∂u/∂n = 0. Beyond the separation point, there may exist a region of
reversed, turbulent, or disturbed flow where recirculating or vortical flows
are commonly seen. These characteristics are important biologically since
they can lead to reverse and/or oscillatory shear stresses and to long particle
residence times, both of which can have a wide variety of clinical implica-
tions, including the initiation of atherosclerosis (see Chapter 6, section 6.4:.
Wall Shear Stress and Its Effect on Endothelial Cells).
In pulsatile flows, flow separation can be generated by a geometric
adverse pressure gradient or by time-varying changes in the driving pres-
sure. Geometric adverse pressure gradients, for example, are present
behind all prosthetic heart valves because of their small orifice areas. As
the area downstream of a prosthetic heart valve increases, the mean flow
velocity must decrease in accordance with the continuity equation. This
decrease in velocity is also accompanied by an adverse pressure gradient
due to the energy balance. Similar effects are seen as a result of the con-
tractile nature of the heart in which blood flow experiences both accelera-
tion and deceleration during a cardiac cycle. An adverse pressure gradient
exists during the deceleration phase of a particular flow and may lead to
boundary layer separation. Consequently, flow separation is even more
likely in regions where both geometric and temporal adverse pressure
gradients exist (e.g., arterial stenoses, stenotic heart valves, prosthetic heart
valves, and aneurysms).
1.10 Problems
1.1 For Hagen–Poiseuille flow, show that the shear stress at the tube wall
( τ w = τ r=R ) can be written as
4µQ
τw =
πR 3
1.3 For Hagen–Poiseuille flow, show that the mean velocity over the cross
section is 1/2 of the peak velocity in the tube.
1.4 What pressure will be required to force 1 cc/s of blood serum through
an intravenous tube of radius 0.50 mm and length 3 cm and into an artery with
a mean pressure of 100 mmHg? (Assume: Blood serum viscosity, µ = 7 × 10−3
Poise.)
1.5 A patient has atherosclerosis, which produces a stenosis of his aorta of
16% diameter reduction.
2
Introduction to Solid Mechanics
CONTENTS
2.1 Introduction to Mechanics of Materials...................................................47
2.1.1 Elastic Behavior ...............................................................................47
2.1.2 Engineering and True Strain .........................................................50
2.1.3 Incremental Elastic Modulus.........................................................51
2.1.4 Poisson’s Ratio.................................................................................51
2.1.5 Shearing Stresses and Strains........................................................53
2.1.6 Generalized Hooke’s Law..............................................................53
2.1.7 Bulk Modulus ..................................................................................56
2.2 Analysis of Thin-Walled Cylindrical Tubes ............................................57
2.3 Analysis of Thick-Walled Cylindrical Tubes...........................................60
2.3.1 Equilibrium Equation .....................................................................60
2.3.2 Compatibility Condition ................................................................61
2.4 Viscoelasticity ...............................................................................................63
2.5 Problems........................................................................................................67
As blood flows in the arterial system in the human circulation, there is a constant
interaction between the flowing blood and the wall of the arteries. In order to
study the unsteady blood flow dynamics in the arteries, it becomes necessary
to consider the material behavior of the arterial wall. In this chapter, the basic
stress–strain relationship for an elastic material is reviewed and basic relation-
ships for stresses in thin-walled and thick-walled cylindrical tubes are derived
that will be useful in modeling unsteady blood flow in the arteries. Visco-elastic
behavior of materials is also included, as this has arterial wall significance.
47
7328_C002.fm Page 48 Friday, October 6, 2006 12:23 PM
Stress (σ)
ᐍo
ᐍ
d
P
Strain (⑀)
(a) (b)
FIGURE 2.1
Stress–strain behavior of Hookean elastic material.
− 0 δ
ε= = (2.1)
0 0
σ = Eε (2.2)
It is apparent that the elastic modulus will have the same units as stress.
A material is said to be homogeneous if the elastic modulus, E, is the same
at every point on the bar. This will result in the elongation being uniformly
distributed along the length of the bar. A material is defined as isotropic if
the stress–strain behavior does not change when the direction of loading is
changed. The stress–strain diagram will be linear until the proportionality
limit is reached and then the relationship will become nonlinear. Engineering
materials can be divided into two broad categories based on the stress–strain
relationship. Figure 2.2 shows the typical stress–strain relationship for a
ductile and for a brittle material. For a ductile material, the stress initially
increases linearly with strain until a critical value of the yield stress, σy , is
reached. If the applied stress exceeds the yield stress for the material, the
material will be subject to a large deformation with a small increase in load
exhibiting plastic deformation.
σ
σU
Rupture σB ×
×
Stress (σ)
Stress (σ)
σY
σB
FIGURE 2.2
Stress–strain diagrams for typical ductile (a) and brittle (b) material.
7328_C002.fm Page 50 Friday, October 6, 2006 12:23 PM
After the maximum value of stress, or the ultimate stress, σult, has been
reached, the diameter of a region of the specimen decreases significantly.
This phenomenon is referred to as necking. After necking occurs, a lower
load is sufficient to induce further deformation of the specimen until rup-
ture takes place. The load to induce rupture is known as the breaking
strength, σb, of the material. In the case of a brittle material, there is no
noticeable change in the rate of elongation before rupture takes place. In
such a material, the ultimate strength and the breaking strength are the
same. Any material for which the stress–strain curve does not follow the
linear relationship is classified as being nonlinear. We will observe later
that most biological materials generally do not exhibit a linear stress–strain
relationship.
− 0 δ
ε= =
0 0
This relationship is defined as the engineering strain where the ratio of the
elongation δ and the initial length of the specimen 0 are used. However,
strain can also be computed by using the successive values of the instanta-
neous length as they are recorded. Taking the value of the instantaneous
length and the corresponding increment in length, the true strain can be
defined as
∆
ε1 = ∑ ∆ε = ∑
Replacing the summation by integrals, the true strain can be expressed as
∫ = ln
d
ε1 = (2.3)
0
0
The relationship between the true strain and the engineering strain can be
obtained as
ε1 = ln(1 + ε) (2.4)
Stress (σ)
Einc = dσ/d⑀
dσ
d⑀
Strain (⑀)
FIGURE 2.3
Schematic of a nonlinear stress vs. strain relationship for biological soft tissue.
dσ
Einc = (2.5)
dε
The incremental elastic modulus (and the material stiffness) increases with
the increase in slope of the nonlinear stress–strain curve schematically shown
in Figure 2.3.
σx
εx = (2.6)
E
7328_C002.fm Page 52 Friday, October 6, 2006 12:23 PM
d
A
P
x
sx = P
A
FIGURE 2.4
An axially loaded slender bar.
In this equation, subscript x denotes the coordinate along the axial direc-
tion. It was also pointed out earlier that, when the bar elongates axially, there
would be a corresponding contraction in the transverse directions. In other
words, even if there is no applied stress in the transverse directions, there
will be strains present in these directions. The Poisson’s ratio is defined as
lateralstrain
ν=
axialstrain
or
−ε y −εz
ν= = (2.7)
εx εx
σx σx
εx = and ε y = ε z = − ν (2.8)
E E
7328_C002.fm Page 53 Friday, October 6, 2006 12:23 PM
y y π +γ
yx xy
2
xy xy
yx π _γ
xy
2
x x
FIGURE 2.5
An element subjected to shearing stresses.
τ xy = Gγ xy (2.9)
τ yz = Gγ yz and τ zx = Gγ zx (2.10)
σy
yz yx
zy xy
xz
zx
σx
σz
x
z
FIGURE 2.6
General stress conditions on a cubic element.
the Principle of Superposition can be applied, and we can write the gener-
alized Hooke’s law as
σ x νσ y νσ y
εx = − −
E E E
σ y νσ x νσ z
εy = − −
E E E
(2.11)
σ νσ νσ y
εz = z − x −
E E E
τ xy τ yz τ
γ xy = ; γ yz = ; γ zx = zx
G G G
Here, σx, σy , and σz are the normal stresses in the respective coordinate
directions and εx, εy , and εz, the corresponding normal strains. The normal
strain, ε, is a measure of the elongation or contraction of a line segment in
the body. Hence, normal strains cause a change in the volume of a rectangular
element shown in Figure 2.6. τxy , τyz, and τzx are the shear stresses and γxy , γyz,
and γzx are the corresponding shear strains. In these terms, the first subscript
indicates the plane normal to the coordinate direction on which the stress
acts and the second subscript indicates the direction of the stress. The shear
strain is a measure of the change in angle between two small line segments
that are originally perpendicular to each other and, thus, represents a change
in the shape of the rectangular element.
7328_C002.fm Page 55 Friday, October 6, 2006 12:23 PM
E
σx = (1 − ν)ε x + ν(ε y + ε z )
(1 + ν)(1 − 2 ν)
E
σy = (1 − ν)ε y + ν(ε z + ε x )
(1 + ν)(1 − 2 ν)
E (2.12)
σz = (1 − ν)ε z + ν(ε x + ε y )
(1 + ν)(1 − 2 ν)
τ xy = Gγ xy
τ yz = Gγ yz
τ zx = Gγ zx
E
σx = (ε + νε y )
(1 − ν2 ) x
(2.13)
E
σy = (ε + νε x )
(1 − ν2 ) y
In the case of blood vessels, a subject of interest in this textbook, the vessel
geometry can be approximated as a tube and, hence, the use of polar (r, θ, z)
coordinates will be appropriate. The Hooke’s law in polar coordinates can
be written as
σ r νσ 0 νσ z
εr = − −
E E E
σ 0 νσ z νσ r
ε0 = − − (2.14a)
E E E
σ z νσ r νσ θ
εz = − −
E E E
E
σr = [(1 − ν)ε r + ν(ε θ + ε z )]
(1 + ν)(1 − 2 ν)
E
σθ = [(1 − ν)εθ + ν(ε z + ε r )] (2.14b)
(1 + ν)(1 − 2 ν)
E
σz = [(1 − ν)ε z + ν(ε x + ε θ )]
(1 + ν)(1 − 2 ν)
7328_C002.fm Page 56 Friday, October 6, 2006 12:23 PM
τ rθ = G γ rθ
τθz = Gγ θz (2.14c)
τ zr = Gγ zr
E
σr = (ε + νεθ )
(1 − ν2 ) r
(2.15)
E
σθ = (ε + νε r )
(1 − ν2 ) θ
Because the strains are much smaller than unity, their products will be
even smaller and can be neglected and, thus, V can be written as
V = 1 + εx + εy + εz (2.17)
V − V0 = ∆V = ε x + ε y + ε z
The change in volume per unit volume, also referred to as the volumetric
strain or cubical dilatation, is given by
∆V
= εx + εy + εz (2.18)
Vθ
7328_C002.fm Page 57 Friday, October 6, 2006 12:23 PM
∆V (1 − 2 ν)
= (σ x + σ y + σ z ) (2.19)
V0 E
This expression shows that when the Poisson’s ratio is 0.5, the volumetric
strain will be equal to zero. A special case to be considered is when the body
is subjected to a uniform hydrostatic pressure p. Then, each of the stress
components is equal to – p and the above relationship can be rewritten as
∆V −3(1 − 2V )
= p (2.20)
V0 E
∆V p
=− (2.21)
V0 k
E
G= (2.22)
2(1 + ν)
σθ
t
σz
σz
σθ
FIGURE 2.7
Stresses on a thin-walled cylindrical vessel.
∆z
σdA t
t
σzdA
r
x r
pdA z
r
z
σdA
t
pdA
(a) (b)
FIGURE 2.8
Forces on segments of the cylindrical vessel.
7328_C002.fm Page 59 Friday, October 6, 2006 12:23 PM
the blood pressure is the transmural pressure load resulting in the deforma-
tion of the vessel.
Static force equilibrium in the x direction can be written as
pR
σθ = (2.24)
t
σ z (2 πRt) − p( πR 2 ) = 0 (2.25)
pR
σz = (2.26)
2t
σ θ = 2σ z (2.27)
In open-ended tubes, σ z = 0 and only the hoop stress will exist. Note that
the radius R used in this derivation is the internal radius of the tube. The
initial circumferential length is 2πR and, if the radius increases by a small
amount, ∆R, due to the applied internal pressure, then the circumferential
strain can be computed as
∆R
εθ = [2 π(R + ∆R) − 2 πR] 2 πR = (2.28)
R
σ θ pR 2
E= = (2.29)
εθ t∆R
7328_C002.fm Page 60 Friday, October 6, 2006 12:23 PM
dθ
(σ r + dσ r )(r + dr )dθdz − σ r rdθdz − 2 σ θdrdz sin =0 (2.30)
2
P2
dz σr + dσr
dr σ
R2
R1 σr
dθ σ
o z Pl
Pl
P2
(a) (b)
FIGURE 2.9
Stresses on a thick-walled, open-ended vessel.
7328_C002.fm Page 61 Friday, October 6, 2006 12:23 PM
Since Sin d2θ ≅ d2θ for small angles, after gathering terms, discarding higher
order terms, and dividing by rdrdθ dz, we obtain
dσ r ( σ r − σ θ )
+ =0 (2.31)
dr r
du
εr = (2.32)
dr
and
2 π(r + u) − 2 πr u
εθ = = (2.33)
2 πr r
E E du u
σr = (ε + νεθ ) = +ν (2.34)
1 − ν2 r 1 − ν2 dr r
and
E E u du
σθ = (ε + νε r ) = +ν (2.35)
1 − ν2 θ 1 − ν2 r dr
d 2 u 1 du u
+ − =0 (2.36)
dr 2 r dr r 2
d 1 d
(ur ) = 0 (2.37)
dr r dr
7328_C002.fm Page 62 Friday, October 6, 2006 12:23 PM
Following the two integrations, a function u(r) satisfying the above equation
is given as
c2
u = c1r + (2.38)
r
E 1− v
σr = (1 + v)c1 − 2 c2 (2.39)
1− v
2 r
and
E 1− v
σθ = (1 + v)c1 + r 2 c2 (2.40)
1 − v2
The constants c1 and c2 in Equation 2.39 and Equation 2.40 are evaluated
by applying the boundary conditions. From Figure 2.9, the boundary con-
ditions can be specified as σr = −p1 at r = R1 (inner surface) σr = −p2 and at
r = R2 (outer surface). The negative signs in the above equations indicate that
the pressures act into the surface as shown in Figure 2.9. Substituting for c1
and c2 in the above equations, the radial and circumferential stresses are
derived as
R2 R2 R2 R2
σ r = p1 2 1 2 1 − 22 − p2 2 2 2 1 − 21 (2.41)
R2 − R1 r R2 − R1 r
and
R2 R2 R2 R2
σ θ = p1 2 1 2 1 − 22 − p2 2 2 2 1 + 21 (2.42)
R2 − R1 r R2 − R1 r
The stress equations (Equation 2.41 and Equation 2.42) were first published
in 1833 and are known as the Lame relationships.
In the stress analysis of blood vessels, the external pressure p2 is generally
assumed to be zero and the force acting on the vessel wall is the transmural
blood pressure. The expression for the radial displacement in an artery
considered as a thick-walled tube is given by the expression
Bergel (1961a and 1961b) obtained an expression for the incremental mod-
ulus corresponding to an increase in outer radial dimension ∆R2 due to an
increase in internal pressure ∆p. By replacing p1 with ∆p, r with R2 and u
with ∆R2 in the above equation, the incremental modulus can be shown to be:
2(1 − ν2 )R12 R2 ∆p
Einc = (2.45)
(R22 − R12 )∆R2
R2
Ep = ∆ p (2.46)
∆R2
This relationship ignores the tube wall thickness and, thus, Ep represents a
structural modulus rather than the elastic property of the tube material.
2.4 Viscoelasticity
The relationships for material properties derived above are applicable for
purely elastic material in which a load applied to the material produces an
instantaneous response in the form of deformation. Upon removal of the
external load, the specimen will instantaneously return to the undeformed
state (assuming inertial effects are small). In the presence of viscous behavior,
however, recovery of deformation will not occur instantly. This is because
the applied shear stress in viscous material is a function of rate of shear,
7328_C002.fm Page 64 Friday, October 6, 2006 12:23 PM
σ σ
G
µ
1 1
⑀ η ⑀
G
⑀ ⑀
σ σ σ σ
(a) (b)
G µ
σ σ σ
σ
µ
(c) (d)
FIGURE 2.10
A typical viscoelastic material schematically represented by (a) a linear spring representing the
elastic component with the corresponding stress–strain relationship and (b) the viscous com-
ponent represented by a dashpot with the corresponding stress–strain rate relationship; The
combined viscoelastic behavior of a solid represented by (c) a Maxwell model with the spring
and dashpot in series, and (d) a Kelvin model with spring and dashpot in parallel.
σ = Ksε (2.47)
σ = µε (2.48)
or in parallel (the Kelvin or Voigt model). In the case of the Maxwell model
(Figure 2.10c), the constitutive relationship is given by
σ σ
+ = ε (2.49)
Ks µ
In the case of the Kelvin model (Figure 2.10d), the constitutive relationship
can be expressed as
σ = K s ε + µε (2.50)
These two-element models are not generally adequate to represent the vis-
coelastic behavior of real materials and, thus, models with three or more
elements are employed that can incorporate the flexibility needed to describe
the response of actual materials. Creep and stress relaxation experiments are
generally performed to describe the viscoelastic properties of materials. In
the case of creep tests (Figure 2.11), the viscoelastic specimen is subjected to an
instantaneous constant stress σ0 and the strain (creep response) is measured
⑀∝
Strain (ε)
⑀0
t=0 t=∝
Time
σ0
Stress (σ)
Time
FIGURE 2.11
Creep test for a viscoelastic material.
7328_C002.fm Page 66 Friday, October 6, 2006 12:23 PM
σ0
Stress (σ)
σ∝
t=0 t=∝
Time
⑀0
Strain (ε)
Time
FIGURE 2.12
Stress relaxation experiment for a viscoelastic material.
as a function of time. For the stress relaxation experiment (Figure 2.12), the
specimen is subjected to an instantaneous constant strain and the stress (relax-
ation) is measured as a function of time. Thus, in viscoelastic material (includ-
ing blood vessels), a finite time is required after the application of the load
before the complete deformation is attained with a static load. With such a
material, the radius of the tube for a step increase in pressure will take a finite
time to increase to its steady state value (Figure 2.13). The same phenomenon
Pressure
Φ
Radius
Time
FIGURE 2.13
Response to time-dependent pressure loading of a viscoelastic structure.
7328_C002.fm Page 67 Friday, October 6, 2006 12:23 PM
will also occur when the pressure is removed instantly. To determine the elastic
modulus under static conditions of applied internal pressure, adequate time
should be allowed for the vessel to expand to its final radius before the
measurement is made. With such experiments, only the elastic response of
the material is determined and no information is obtained about the viscous
properties.
To determine the viscoelastic behavior of the material, dynamic experiments
are performed with the application of sinusoidally varying pressure. A typ-
ical response to such loading is shown in Figure 2.13. Even though the
oscillations of the radius have the same frequency as that of the applied
pressure, the effect of viscosity is to delay the response by a phase angle, φ.
Moreover, it is also observed that the value of peak distension under sinu-
soidal loading is smaller than that for static deflection. This is due to the fact
that before the delayed response can attain its peak, the load has already
reached its down slope and, thus, reversing the process so that the amplitude
of the distension is smaller.
Due to the phase difference between the applied pressure and the radial
changes, the ratio of stress to resultant strain will be a complex quantity,
which can be expressed by the relationship
The modulus, Ec, is a function of the applied oscillatory frequency and its
values can be determined by applying sinusoidal loading at varying fre-
quencies. For the case of dynamic loading in viscoelastic material, the incre-
mental elastic modulus expression given in Equation 2.45 can be modified as
2(1 − v 2 )R12 R2 ∆ p iφ
Einc = e (2.52)
(R22 − R12 )∆R2
2.5 Problems
2.1 A circular aluminum tube 40 cm in length is subject to a tensile load
of 2 kN. The outside and inside diameters of the tube are 4.2 and 4.0 cm,
respectively. What is the amount of tensile stress on the bar? Aluminum has
an elastic modulus of 73.1 GPa. Determine the axial strain and the increase
7328_C002.fm Page 68 Friday, October 6, 2006 12:23 PM
in length of the bar for the given load. Assuming a typical elastic modulus
of about 0.1 MPa for a typical artery, what would be the corresponding tensile
load on an arterial specimen of the same dimension that would result in the
same axial strain?
2.2 A high-strength steel rod (E = 200 GPa and ν = 0.32) with a diameter
of 5 cm is being subjected to a compressive load of 10 kN. Determine the
increase in diameter of the tube after the load is applied.
2.3 A brass specimen 10 mm in diameter and a length of 50 mm is loaded
with a 20 kN force in tension. If the length increases by 0.12 mm, determine
the elastic modulus of the brass. If the diameter of the bar decreases by 0.0083
mm, calculate the Poisson’s ratio of the material.
2.4 An aluminum soda can (E = 73.1 GPa and ν = 0.35.) has a radius-to-
thickness ratio of 200:1 and holds soda under pressure. When the lid is
opened to release the pressure, the strain in the longitudinal direction is
measured as 170 µm/m. What was the internal pressure in the can? Express
your answer in the units of mm Hg. Compare this pressure magnitude with
the typical mean blood pressure in an artery.
7328_C003.fm Page 69 Monday, September 18, 2006 10:57 AM
3
Cardiovascular Physiology
CONTENTS
3.1 Introduction ..................................................................................................69
3.2 The Heart ......................................................................................................70
3.2.1 Overview ..........................................................................................70
3.2.2 Cardiac Structure.............................................................................73
3.2.3 Cardiac Conduction ........................................................................74
3.2.4 Cardiac Function .............................................................................77
3.3 Cardiac Valves..............................................................................................82
3.4 Systemic Circulation....................................................................................84
3.5 Coronary Circulation ..................................................................................91
3.6 Pulmonary Circulation and Gas Exchange in the Lungs .....................94
3.7 Cerebral and Renal Circulations ...............................................................98
3.7.1 Cerebral Circulation........................................................................98
3.7.2 Renal Circulation.............................................................................99
3.8 Microcirculation ...........................................................................................99
3.9 Regulation of the Circulation ..................................................................103
3.10 Atherosclerosis ...........................................................................................105
3.10.1 Morphology of Atherosclerosis...................................................106
3.10.2 Sequence of Growth of the Lesion .............................................107
3.10.3 Physiological Implications ...........................................................107
3.11 Problems...................................................................................................... 110
3.1 Introduction
This chapter contains a brief review of the cardiovascular physiology
relevant to fluid mechanics in the human circulation. For a detailed study,
the reader is referred to any textbook on Physiology (e.g., Guyton and Hall,
2000 or Silverthorn, 2001). The cardiovascular system includes the heart and
blood vessels of the systemic and pulmonary circulation. Fundamental
requirements of the circulatory system are to provide adequate blood flow
without interruption and to regulate blood flow according to the various
69
7328_C003.fm Page 70 Monday, September 18, 2006 10:57 AM
demands of the body (Rushmer, 1976). The contracting heart supplies the
energy required to maintain the blood flow through the vessels. The pressure
gradient developed between the arterial and the venous end of the circula-
tion is the driving force causing blood flow through the blood vessels. The
energy is dissipated in the form of heat due to the frictional resistance. Blood
picks up oxygen in the lungs and nutrients in the intestine and delivers them
to the cells in all parts of the body. The circulating blood also removes cellular
wastes and carbon dioxide from the cells for excretion through the kidneys
and the lung, respectively. The circulating blood also maintains the visceral
organs, such as the heart, the kidney, the liver, and the brain, at a constant
temperature by convecting the heat generated and dissipating the same
through transfer across the skin. In the following brief review of the cardio-
vascular system, we will concentrate on the mechanical aspects of circulation.
Cardiovascular Physiology 71
Superior Aorta
vena cava
Pulmonary
semilunar valve
Right pulmonary Left pulmonary
arteries arteries
Left pulmonary
veins
Right
atrium
Cusp of left AV
(bicuspid) valve
Chordae
Cusp of right AV tendineae
(bicuspid) valve
Papillary
muscles
Inter ventricular
Inferior vena septum
cava Left ventricle
Descending
Right ventricle aorta
FIGURE 3.1
A schematic diagram of the four chambers of the heart and the heart valves. The arrows indicate
the direction of blood flow. (Redrawn from Silverthorn, D.U. (2001) Human Physiology: An
Integrated Approach, 2nd ed., Prentice Hall, Upper Saddle River, NJ. With permission.)
Pulmonary
Left artery
atrium Aorta
Right
atrium
P
A
M
T
Fibrous
skeleton
Right Left
ventricle ventricle
FIGURE 3.2
The fibrous skeleton and the four chambers of the heart. (Redrawn from Rushmer, R.F. (1976)
Cardiovascular Dynamics, W.B. Saunders Company, Philadelphia. With permission.)
Intercalated disks
Myocardial muscle cell
FIGURE 3.3
The striated appearance of the cardiac muscle. (Redrawn from Silverthorn, D.U. (2001) Human
Physiology: An Integrated Approach, 2nd ed., Prentice Hall, Upper Saddle River, NJ. With permission.)
7328_C003.fm Page 73 Monday, September 18, 2006 10:57 AM
Cardiovascular Physiology 73
Right Left
ventricle ventricle
FIGURE 3.4
A schematic drawing depicting the left and right ventricles. The left ventricle has a more
rounded shape, whereas the right ventricle has a semilunar shape and wraps around the left
ventricle.
circulation where the resistance to flow is relatively small and high pressures
are not needed. However, if the pulmonary resistance increases due to dis-
ease, the right ventricular myocardium cannot produce the sustained high
pressures needed and, hence, failure of the right ventricle (clinically known
as congestive heart failure or CHF) will result. Similarly, if a large pumping
volume is demanded of the left ventricle, the left ventricular chamber dilates
to accommodate the increased volume.
Cardiovascular Physiology 75
SA node depolarizes.
SA
node
AV
node
Electrical activity
goes rapidly to
AV node via
internodal
(b) pathways.
(c)
SA node
Internodal Depolarization
pathways spreads more
slowly across atria.
Conduction
slows through
AV node.
AV node
Bundle of (d)
his Bundle Depolarization moves
branches Purkinje rapidly through
fibers ventricular conducting
(a) system to the apex
of the heart.
(e)
Depolarization wave
spreads upward from
the apex.
(f )
FIGURE 3.5
The electrical conduction of the heart. (Redrawn from Silverthorn, D.U. (2001) Human Physiology:
An Integrated Approach, 2nd ed., Prentice Hall, Upper Saddle River, NJ. With permission.)
in Figure 3.6b. In the action potential of the S-A node, each time the resting
membrane potential is reestablished, the potential gradually decays until it
reaches the threshold for self-excitation. At this point, depolarization takes
place and the action potential is established. The action potential of the S-A
node repolarizes within about 15 ms after depolarization, whereas 15 to 30 ms
is required for the other cardiac muscles. During this plateau region, or refrac-
tory period, only extremely high electrical stimulation can initiate a new spike
and the normal cardiac impulses cannot initiate a new spike.
7328_C003.fm Page 76 Monday, September 18, 2006 10:57 AM
Membrane potential
of autorhythmic cell
Electrical Membrane
current potential of
contractile cell
Contractile cells
Cells of
SA node
Intercalated disk
with gap junctions
(a)
Plateau
+20
0 (b)
–20
–40
–60
–80
Millivolts
0 1 2 3 4
Seconds
(b)
FIGURE 3.6
(A) Electrical conduction in myocardial cells resulting in depolarization. (Redrawn from
Silverthorn, D.U. (2001) Human Physiology: An Integrated Approach, 2nd ed., Prentice Hall, Upper
Saddle River, NJ. With permission.) (B) The rhythmic action potentials of the Purkinje fiber and
the ventricular muscle. (Redrawn from Guyton, A.C. and Hall, J.E. (2000) A Textbook of Medical
Physiology, 10th ed., W.B. Saunders Company, Philadelphia. With permission.)
Cardiovascular Physiology 77
1 sec.
+2
+1 R
Millivolts
P T P
0
Q
S
–1
FIGURE 3.7
A typical normal electrocardiogram.
also follows the order in which the action potential reaches the particular
region. Hence, the interventricular septum starts contracting first followed by
the endocardial surface of the apical region, and the contraction spreads
toward the base (Figure 3.5). From a mechanical point of view, this sequence
of contraction enables an efficient ejection of blood from the ventricles to the
respective blood vessels through an effective “squeezing”of the heart.
Electrical potentials generated by the heart also spread to the surrounding
tissues and this electrical activity can be recorded by placing electrodes on
the surface of the skin. Such a recording is known as the electrocardiogram
(ECG) and a typical ECG signal is shown in Figure 3.7. The P wave and the
QRS complex shown in the figure are depolarization waves. The electrical
currents cause the P wave as the atrial muscles depolarize followed by the
contraction of the atrial muscles. The QRS complex represents the depolar-
ization of the ventricular muscles, which is followed by the ventricular
muscular contraction. The T wave is generated as the ventricular muscles
repolarize and, hence, it is known as the repolarization wave. In Figure 3.7,
the time between consecutive R waves is 0.8 sec or, in other words, the heart
rate is 75 cycles or beats per minute (bpm). For a normal adult, the heart
rate is between 70 to 75 bpm under resting conditions.
n
r ia
Ve
l sy
stol
e
EDV = End-diastolic volume.
7328_C003.fm Page 78 Monday, September 18, 2006 10:57 AM
FIGURE 3.8
Schematic of the cardiac cycle with the atrial and ventricular systolic and diastolic phases. (Redrawn from Silverthorn, D.U. (2001) Human Physiology: An
Integrated Approach, 2nd ed., Prentice Hall, Upper Saddle River, NJ. With permission.)
Biofluid Mechanics: The Human Circulation
7328_C003.fm Page 79 Monday, September 18, 2006 10:57 AM
Cardiovascular Physiology 79
Stroke volume
120
D
Left ventricular pressure (mm Hg)
Aortic valve
Aortic valve Ventricular ejection opens
closes;
ESV
80 C
One Isovolumic
Isovolumic
cardiac contraction
relaxation
cycle
40
Mitral valve
Mitral valve closes; EDV
opens Ventricular diastole + filling
A B
0 65 100 135
Left ventricular volume (ml)
A B: Passive filling and atrial contraction
B C: Isovolumic contraction
C D: Ejection of blood into aorta
D A: Isovolumic relaxation
FIGURE 3.9
A ventricular pressure volume curve. (Redrawn from Silverthorn, D.U. (2001) Human Physiology:
An Integrated Approach, 2nd ed., Prentice Hall, Upper Saddle River, NJ. With permission.)
7328_C003.fm Page 80 Monday, September 18, 2006 10:57 AM
the left ventricle. Point A in the figure depicts the point when the ventricular
chamber pressure and volume are the minimum. The mitral valve will open
once the atrial pressure exceeds that of the ventricle and the AB portion of
the curve represents diastolic ventricular filling as the ventricular muscles
continue to relax. Point B represents the end diastolic volume (EDV) and is
the maximum ventricular volume during a cardiac cycle. With the beginning
of ventricular contraction, the chamber pressure rises resulting in the closure
of the mitral valve and the isovolumic contraction phase shown as the BC
portion of the curve. At point C, the ventricular pressure exceeds that of the
aorta and the aortic valve opens. The ventricular muscles continue to contract
increasing the chamber pressure while ejecting the blood into the aorta;
hence, the ventricular volume decreases. The amount of blood left in the
ventricle at the end of contraction (point D) is the end systolic volume (ESV)
and represents the minimum ventricular volume in a cardiac cycle. The
difference between the EDV and the ESV is called the stroke volume (SV)
and represents the volume of blood pumped out by the left ventricle into
the systemic circulation in a cardiac cycle (ml/beat). At the end of contrac-
tion, the ventricular muscles start to relax, the ventricular pressure falls
below that of the aorta and the aortic valve closes. The ventricular muscles
continue through the isovolumic relaxation (segment “DA” of the curve)
and the cardiac cycle starts again. The amount of blood that is pumped out
of the ventricles in 1 minute is referred to as the cardiac output (CO) and is
the product of the stroke volume and the heart rate (HR):
Cardiovascular Physiology 81
The heart muscles receive the energy to perform this work from the oxygen
in the blood. Computing the input and output energies, efficiency of the
heart can be computed. Typically, the work done by the heart amounts to
only 10 to 15% of the total input energy and the remainder of the energy is
dissipated as heat.
A combined representation of the electrical activity of the heart (ECG), the
pressure pulses in the cardiac chambers and the aorta, and the ventricular
blood volume is illustrated in the Wiggers diagram (Figure 3.10). Between the
Time (m sec)
0 100 200 300 400 500 600 700 800
QRS QRS
Electro- complex Cardiac cycle complex
cardiogram
(ECG) P T P
90 Aorta
Dicrotic notch
Left
ventricular
pressure
60
Mitral valve
Left atrial closes Mitral valve
30 pressure
open
Heart S1 S2
sounds
135 End diastolic
Left volume
ventricular
volume
(ml)
65 End systolic volume
Atrial Ventricular systole Ventricular diastole Atrial
systole systole
FIGURE 3.10
The Wiggers diagram depicting the relationship with the electrocardiogram, the ventricular
and the atrial pressures, the heart sounds, and the volume changes in the ventricular chamber.
(Redrawn from Silverthorn, D.U. (2001) Human Physiology: An Integrated Approach, 2nd ed.,
Prentice Hall, Upper Saddle River, NJ. With permission.)
7328_C003.fm Page 82 Monday, September 18, 2006 10:57 AM
pressure and volume curves in the Wiggers diagram, the heart sounds gener-
ated during the cardiac cycle are also illustrated. Ventricular systole begins at
the apex of the heart and the blood is squeezed toward the base of the heart
during the ventricular contraction. The increase in pressure in the ventricular
chambers and the blood pushing upward on the ventricular side of the atrio-
ventricular (mitral and tricuspid) valves keeps the leaflets closed so that blood
does not flow back into the atria. Vibration induced following the closure of
the valves generates the first heart sound (S1 in Figure 3.10) that can be heard
through a stethoscope. Similarly, at the end of contraction, the ventricles relax
with a rapid pressure reduction in the chambers. Once the ventricular pressure
falls below that of the arteries, blood starts to flow back toward the ventricles
and the semilunar (i.e., aortic and pulmonic) valves close. The second heart
sound (S2) is generated due to the induced vibrations with the closure of the
semilunar valves. Third and fourth heart sounds (not illustrated in the figure)
can also be recorded with sensitive electronic stethoscopes. The third heart
sound is generated by the turbulent or agitated flow of blood into the ventricles
during the filling phase. The fourth heart sound is associated with turbulence
created by the atrial contraction. Diseased states, such as stenosed (i.e., stiffer)
or incompetent (i.e., leaking) valve leaflets, will result in changes in character-
istics of the heart sounds, such as frequency and amplitude. Hence, heart
sounds can also be used in diagnosis of valvular diseases.
Aortic semilunar
Chordae tendinea
valve (open)
(tense)
Papillary
Cardiovascular Physiology
Pulmonary semilunar
muscles (tense)
valve (open)
Left ventricle
(contracted)
7328_C003.fm Page 83 Monday, September 18, 2006 10:57 AM
FIGURE 3.11
An illustration of the four heart valves during the systolic and diastolic phases of a cardiac cycle. The top panel depicts the ventricular contractile phase
with the atrio-ventricular valves in the closed position and the semilunar valves in the open position. The bottom panel represents the ventricular filling
phase with the atrio-ventricular valves in the open position and the semilunar valves in the closed position. (Redrawn from Silverthorn, D.U. (2001) Human
83
Physiology: An Integrated Approach, 2nd ed., Prentice Hall, Upper Saddle River, NJ. With permission.)
7328_C003.fm Page 84 Monday, September 18, 2006 10:57 AM
of the coronary arteries by the valve cusps in the fully open position. During
systole, the three cusps open to the full dimensions of the valve ring resulting
in axial flow of blood into the aorta. The A-V valves have two large opposing
cusps and small intermediary cusps at each end. For the mitral valve,
chordea tendinae originating from the edge of the two larger leaflets are
connected to two papillary muscle groups. The chordea tendinae from the
tricuspid valves are attached to three groups of papillary muscles. The mitral
valve consists of a large anteromedial cusp that hangs like a curtain at the
fully open position and a shorter posterolateral cusp. The combined surface
area of the two leaflets is almost twice the valve ring area. In the open
position, the upper portion of the mitral valve resembles that of a funnel.
During contraction of the ventricle, the papillary muscles also contract and
pull the valve leaflets toward the ventricle for efficient closure of the valves
at the beginning of systole. The anatomy and function of the tricuspid valve
is similar to that of the mitral valve. The detailed analysis of fluid mechanics
past the heart valves is included in Chapter 7.
∆p
R= (3.3)
Q
7328_C003.fm Page 85 Monday, September 18, 2006 10:57 AM
Cardiovascular Physiology 85
Pulmonary veins
Superior Ascending
vena cava arteries
Pulmonary
arteries Lungs
Right atrium
Left Aorta
atrium
Hepatic
vein Hepatic artery
FIGURE 3.12
Schematic of blood flow to the visceral organs of systemic and pulmonary circulation. (Redrawn
from Silverthorn, D.U. (2001) Human Physiology: An Integrated Approach, 2nd ed., Prentice Hall,
Upper Saddle River, NJ. With permission.)
where R is the resistance to the flow of blood in that segment. Under these
conditions, the resistance to blood flow is predominantly due to the viscous
shear stresses. The resistance computed using units of pressure in mm Hg
(millimeters of mercury) and flow rate in cm3/s in Equation 3.3 is referred to
as Peripheral Resistance Units (PRUs). For example, in the normal systemic
circulation, if the mean pressure difference between the systemic arteries and
the veins is about 100 mmHg and the flow rate through the systemic circulation
is about 100 cm3/s, the resistance is 1 PRU. In the pulmonary circulation, the
time-averaged pressure difference between the pulmonary artery and the left
atrium is normally about 10 mmHg. Since the pulmonary circulation will also
7328_C003.fm Page 86 Monday, September 18, 2006 10:57 AM
maintain the same flow rate as the systemic circulation over a period of time,
the resistance in the pulmonary circulation is about 0.1 PRU. Clinically, the
resistance is computed in ‘‘Woods Units’’ where mmHg is used for pressure
and l/min is used for the flow rate.
Again, in Chapter 1, the relationship between the flow rate and the pressure
gradient was derived for steady flow of a viscous fluid through a rigid cylin-
drical pipe. The Hagen–Poiseuille flow relationship was given by the equation
πR 4 ∆p
Q= (1.72)
8µL
∆p 8µL
R= = (3.4)
Q πR 4
From Equation 3.4, it can be observed that the resistance to blood flow is
inversely proportional to the fourth power of tube radius. Thus, small changes
in the radius of arteries and arterioles will significantly alter the vascular
resistance. This is important since one of the characteristics of blood vessel
walls is that they are very responsive to stimulation. With sympathetic stim-
ulation, the tone of the vessel walls increases, causing a decrease in diameter
and, thereby, a significant increase in resistance to the flow of blood. On the
other hand, with sympathetic inhibition, the tone of the vessels decreases
resulting in an increase in the vessel diameter and a corresponding decrease
in flow resistance. Thus, sympathetic stimulation or inhibition is effectively
used in controlling the flow rate of blood through any region of the body and
this control is affected at the arteriolar level. For example, an increase in vessel
diameter of just 5% would produce a reduction in resistance of over 21%.
The ratio of change in volume to a change in pressure gives a measure of
the distensibility of the blood vessel. In practice, the increase in volume
normalized to the initial volume is used to define the volumetric strain of
the vessel. The definition for vascular compliance, C, is given by
∆V ∆V
C= V
= (3.5)
∆p V ∆p
Cardiovascular Physiology 87
140
imulation
Pressure (mm Hg) 120
stem
Sympathetic st
ic Inhibition
80
Arterial sy
60
40 Sympathet n
atio
s t i mul
etic tem
20 path ous sys
Sym Ven hibition
Normal volume In
0
0 500 1000 1500 2000 2500 3000 3500
Volume (ml)
FIGURE 3.13
Pressure–volume curves for the arterial and venous system. (Redrawn from Guyton, A.C. and
Hall, J.E. (2000) A Textbook of Medical Physiology, W.B. Saunders Company, Philadelphia. With
permission.)
The aorta and other major arteries are high-pressure conduits transporting
blood to the various regions and visceral organs in the body. The normal
arterial wall is made up of three layers — the intima, media, and adventitia.
The intimal layer of the artery consists of a layer of endothelial cells attached
to a basement membrane. The medial layer consists of vascular smooth mus-
cle cells, collagen, and elastin fibers in an extracellular matrix. The adventitial
layer consists of connective tissue that is attached, or tethered, to the surround-
ing organs in the body. Since blood flows under high pressure through the
arteries, these vessels have strong walls. The arterial walls are also less com-
pliant compared to the venous vessels as shown in Figure 3.13. In the systemic
arteries, relatively small increases in arterial volume will result in large
increases in pressure. Therefore, the systemic arteries serve as a pressure res-
ervoir (Rushmer, 1976). The small arteries further branch into arterioles that
also have strong muscular walls where the vascular tone is controlled by
regulatory mechanisms. Thus, the blood flow distribution to various regions
of the body is controlled by changes in resistance offered by various arterioles.
The blood from the arterioles then enters into capillaries where the exchange
of nutrients between the blood and the interstitial space takes place. The
capillary walls are very thin and permeable to small molecular substances to
allow for an efficient exchange of nutrients. Blood from the capillaries is
collected into venules that coalesce into larger veins, which eventually return
it to the right atrium. Table 3.1 shows the total cross-sectional area, the time-
averaged velocity of blood, and the percentage of blood at any given time in
the various vessels in the systemic circulation (Milnor, 1989).
7328_C003.fm Page 88 Monday, September 18, 2006 10:57 AM
TABLE 3.1
Cross-Sectional Area of the Various Blood Vessels in the Systemic and Pulmonary
Circulation along with the Percentage Volume of Blood for a 20 kg Dog
Total Percentage
Mean Mean Cross Total of Total
Diameter Number Length Section Volume Blood
Vessels (mm) of Vessels (mm) (cm2) (ml) Volume
Systemic Circulation
Aorta (19–45) 1 (2.0–16.0) 60
Arteries 4.000 40 150.0 5.0 75
Arteries 1.300 500 45.0 6.6 30
Arteries 0.450 6000 13.5 9.5 13 11%
Arteries 0.150 110,000 4.0 19.4 8
Arterioles 0.050 2.8 × 106 1.2 55.0 7
Capillaries 0.008 2.7 × 109 0.65 1357.0 88 5%
Venules 0.100 1.0 × 107 1.6 785.4 126
Veins 0.280 660,000 4.8 406.4 196
Veins 0.700 40,000 13.5 154.0 208
Veins 1.800 2.100 45.0 53.4 240 67%
Veins 4.500 110 150.0 17.5 263
Venae cavae (5–14) 2 (0.2–1.5) 92
Subtotal 1406
Pulmonary Circulation
Main Artery 1.600 1 28.0 2.0 6
Arteries 4.000 20 10.0 2.5 25 3%
Arteries 1.000 1550 14.0 12.2 17
Arterioles 0.100 1.5 × 108 0.7 120.0 8
Capillaries 0.008 2.7 × 109 0.5 1357.0 68 4%
Venules 0.110 2.0 × 106 0.7 190.0 13
Veins 1.100 1650 14.0 15.7 22
Veins 4.200 25 100.0 35 5%
Main veins 8.000 4 30.0 6
Subtotal 200
Heart
Atria 2 30
Ventricles 2 54 5%
Subtotal 84
Total 1690 100%
Source: Milnor, W.R. (1989) Hemodynamics, 2nd ed., Williams and Wilkins, Baltimore. With
permission.
At the root of the aorta, the systolic pressure of a normal adult is about
120 mmHg and the diastolic pressure is about 80 mmHg. The difference
between the systolic and diastolic pressures is referred to as the pulse pres-
sure. A typical pressure pulse in the aorta as a function of time is shown in
Figure 3.14. The pressure pulse in systole starts with a steep rise, which
gradually decreases to a plateau at peak pressure. This is followed by a sharp
incisura, or dichrotic notch, due to the closing of the aortic valve. During
diastole, there is essentially an exponential decay of the pressure pulse.
7328_C003.fm Page 89 Monday, September 18, 2006 10:57 AM
Cardiovascular Physiology 89
Sharp
Slow rise incisura
to peak Exponential diastolic
decline
120 Sharp 16.0
upstroke
Pressure (k Pa)
Pressure (mm Hg)
13.3
80 10.6
6.65
40
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Seconds
FIGURE 3.14
Schematic of a normal pressure pulse in the aorta depicting the sharp pressure rise in the
beginning of systole and the exponential pressure decay during diastole. The incisura during
the pressure decay represents the closing of the aortic valve.
Since systole lasts for about 1/3 of the cardiac cycle followed by diastole for
about 2/3 of the cardiac cycle, the mean arterial pressure (MAP) over a
cardiac cycle can be computed as
( ps + 2 pd )
MAP = (3.6)
3
1. The blood vessels in the distal portion are stiffer and, hence, the
velocity of transmission increases.
2. Curvature and branching sites in the blood vessels reflect the pulse
wave and the reflected wave interacts with the waves in the forward
direction.
7328_C003.fm Page 90 Monday, September 18, 2006 10:57 AM
120
100
Pulmonary arteries
Pressure (mm Hg)
Pulmonary veins
Venae cavae
Large veins
Small veins
Capillaries
Capillaries
80
Arterioles
Venules
Venules
60
Large arteries
Small arteries
40
Arterioles
Aorta
20
0
Systemic Pulmonary
FIGURE 3.15
Pressure pulses in the systemic and pulmonary circulation. (Redrawn from Guyton, A.C. and
Hall, J.E. (2000) A Textbook of Medical Physiology, W.B. Saunders Company, Philadelphia. With
permission.)
3. The high pressure portion of the pressure pulse travels faster than
the other parts of the pulse and results in nonlinear interactions.
Cardiovascular Physiology 91
Aortic valve
Right ventricle
Right coronary a.
Pulmonary valve
Left coronary a.
Orifices of
coronary a.s
Left ventricle
Tricuspid
Mitral valve valve
Coronary sinus
Atrioventricular (A-V) node
(a)
Aorta
Right coronary a.
Circumflex a.
Pulmonary a.
Right atrial
appendage Anterior descending
branch
Posterior descending (interventricular)
branch
(interventricular)
(b)
FIGURE 3.16
The anatomical details of the left and right coronary arteries: (a) top view showing the left and
right coronary ostia and the origin of the coronary arteries, and (b) main branches of the
coronary vessels and its perfusion field. (Redrawn from Jacob, S.W. and Francone, C.A. (1974)
Structure and Function in Man, 3rd ed., W.B. Saunders Company, Philadelphia. With permission.)
7328_C003.fm Page 92 Monday, September 18, 2006 10:57 AM
arteries from the left and right coronary sinuses. The bottom panel illustrates
the principal branches of the left and right coronary arteries. The openings
for the coronary arteries, the coronary ostia, are situated in the right and left
sinuses of Valsalva. The left coronary artery originates from the left aortic
sinus and divides immediately into the left anterior descending and the left
circumflex branches. The left coronary artery supplies most of the left ven-
tricle, the anterior two-thirds of the interventricular septum, the anterior left
margin of the free wall of the right ventricle, the apex, the left atrium, and
the lower half of the interatrial septum. The right coronary artery, originating
from the right coronary sinus, supplies the anterior and posterior walls of
the right ventricle, the right atrium and the sinus node, the posterior one-
third of the interventricular septum, the atrio-ventricular node, the upper
half of the interatrial septum, and the posterior base of the left ventricle. In
approximately half of human beings, more blood flows through the right
coronary artery, while the left coronary flow is predominant in about 20%
of humans. The coronary flow from the large conduit arteries, which are on
the epicardium (i.e., surface of the heart), branch into arterioles and then
capillaries that supply the heart muscle. The blood from the capillaries
circulates through the venules and the large conduit veins back to the heart.
Of the returning venous blood, 85% drains through the coronary sinus. The
remaining 15% flows through the anterior superficial veins and the thesbian
veins and directly empties into the cardiac chambers (Marcus, 1983).
The coronary artery circulation has been extensively studied because of
the importance of the coronary blood supply in maintaining the proper
function of the heart. The resting coronary blood flow is about 225 ml/min,
or about 0.8 ml/min/g of the heart muscle. When the heart muscles work
harder (e.g., during exercise conditions), the coronary blood supply can
increase by 4 to 10 times the resting blood flow rate. The rate of blood flow
through the coronary arteries during a cardiac cycle is illustrated in
Figure 3.17. During systole, when the heart muscles are contracted, very little
flow occurs in the coronary arteries due to the constriction of the intramus-
cular branch vessels of the coronary arteries. During diastole, the heart
muscles are relaxed, the lumen of the intramuscular arteries open fully, and
the major portion of the coronary flow occurs. Measurements of coronary
blood flow with electromagnetic flow meters have indicated that the flow
through the arteries in systole constitutes from 7 to 45% of the total coronary
flow during a cardiac cycle (Rushmer, 1976).
During resting conditions, 65% of the oxygen is extracted from the coro-
nary arterial blood as it passes through the heart. The rate of blood flow
through the coronary arteries is regulated by the myocardial oxygen
demand. Under severe exercise conditions, the cardiac output can increase
three to six times the resting value. To meet the increased oxygen demand
as the cardiac muscles are working harder to increase the cardiac output,
the coronary blood flow also increases by several fold. The fact that coronary
blood flow is regulated in response to the myocardial oxygen demand can
be illustrated by occluding the coronary arteries for a few seconds and then
7328_C003.fm Page 93 Monday, September 18, 2006 10:57 AM
Cardiovascular Physiology 93
100
0
Systole Diastole
FIGURE 3.17
The blood flow rate in the human left coronary vessels in systole and diastole extrapolated
from dog data. (Redrawn from Guyton, A.C. and Hall, J.E. (2000) A Textbook of Medical Physiology,
W.B. Saunders Company, Philadelphia. With permission.)
releasing the occlusion. The myocardium distal to the occlusion has a reduc-
tion in blood supply and oxygen during the occlusion. Hence, immediately
after the occlusion is released, the blood flow through the vessel increases
significantly and stays high for a length of time depending upon the length
of occlusion. This phenomenon, referred to as reactive hyperemia, is illus-
trated in Figure 3.18.
Chest pain, or angina pectoris, results if the myocardium receives inade-
quate oxygen supply. Angina pectoris can be initiated by physical exertion
or emotional stress and is usually relieved by rest or administration of nitrites
or velocity
Debt 15 sec
area
FIGURE 3.18
Reactive hyperemia caused by occlusion of the coronary vessels.
7328_C003.fm Page 94 Monday, September 18, 2006 10:57 AM
(drugs that dilate the arteries). Myocardial infarction (MI) is the result of
acute occlusion of coronary vessels. Coronary artery occlusion is due to the
presence of vascular lesions called atherosclerosis and will result in lack of
blood supply to the portion of the myocardium perfused by the artery
together with injury or death to the affected segment. A description of this
disease, its predilection, and attempts to relate abnormal flow-induced
stresses to the etiology of these lesions are discussed later in this chapter
and in subsequent chapters in this book.
Cardiovascular Physiology 95
the narrowing of the smaller bronchi and bronchioles due to the constriction
of smooth muscle cells in these smaller-sized passages. Histamine and slow
reactive substances of anaphylaxis, which are released in the lungs due to
allergic reactions (caused, for example, by pollen in the air), can induce
bronchiolar constriction. The respiratory passages are kept moist by a layer
of mucus that coats the entire luminal surface. The mucus also traps small
particles from the inspired air and keeps most of the particles from reaching
the alveoli. The respiratory passages are also lined with epithelial cells con-
taining about 200 celia, or flagella, in each cell. These celia beat continually
at a rate of 10 to 20 times per second, inducing motion of the mucus towards
the pharynx which, together with the entrapped particles, is either swal-
lowed or coughed out to the exterior.
Blood flow to the lungs is provided by the contraction of the right ventricle
into the pulmonary artery that extends about 4 cm beyond the pulmonary
valve before branching into the right and left pulmonary arteries that feed
the two lungs. The pressure pulse curve for the right ventricle is compared
to that of the aorta in Figure 3.19a and the pressures in the blood vessels of
the lung are shown in Figure 3.19b. The systolic and diastolic right ventric-
ular pressures are about 22 and 0 mmHg, respectively. In the pulmonary
artery, the corresponding pressures are 22 and 8 mmHg with a pulse pressure
of 14 mmHg. By conservation of mass, the rate of blood flow from the right
ventricle into the lungs averaged over time must be the same as the rate at
which blood is pumped by the left ventricle into the systemic circulation.
Thus, the flow rate through the lungs is very high compared to that of other
organs because the entire cardiac output must flow through the lungs to
become oxygenated. This is possible fluid dynamically because the distance
traversed by the blood in the pulmonary circulation is relatively short com-
pared to that in the systemic circulation. Additionally, the pulmonary arte-
rioles are larger in size and more distensible than their systemic counterparts,
causing the resistance in the pulmonary circulation to be low. Therefore, the
lower pressures generated by the right ventricle are still sufficient to maintain
the same cardiac output through the pulmonary circulation and back into
the left atrium (see Equation 1.74). The lungs themselves are supplied with
oxygenated blood for nutrition through the bronchial arteries, which origi-
nate from the systemic circulation. The flow in this circuit amounts to only
1 to 2% of the cardiac output. The bronchial circulation supplies oxygen to
supporting tissues of the lungs, including connective tissue as well as the
bronchi. The blood from the bronchial circulation empties into the pulmo-
nary veins and returns to the left atrium, thus bypassing the lung capillaries
(i.e., alveoli).
Because of the low pulmonic pressures, distribution of blood flow in the
lung capillaries is noticeably affected by gravity. More specifically, when a
person is in the standing position, pressure in the base (inferior) region is
higher than in the apical (superior) region of the lung. Another important
phenomenon is that the capillaries in the alveolar walls are distended by
blood pressure, but are also compressed by air pressure in the alveoli. Hence,
7328_C003.fm Page 96 Monday, September 18, 2006 10:57 AM
8
0
0 1 2
Seconds
(a)
25 S
mm Hg
15 M
8 D
7
2
0
Pulmonary Pulmonary Left
artery capillaries atrium
(b)
FIGURE 3.19
(a) Typical pressure pulses in the right ventricle, pulmonary artery, and the aorta; (b) distribution
of pressures in the blood vessels in the lung. (Redrawn from Guyton, A. C. and Hall, J. E. (2000)
A Textbook of Medical Physiology, W.B. Saunders Company, Philadelphia. With permission.)
in the regions where the alveolar air pressure becomes larger than that of
the capillary blood pressure, the capillaries collapse and prevent blood flow.
Therefore, in the basal regions of the lungs, where the blood pressure is
always higher than that of the alveolar pressure, there is continuous blood
flow. In the apical zones, however, the blood pressure becomes higher than
the alveolar pressure only during the systolic phase and, thus, results in an
intermittent flow through the capillaries. During exercise when there is an
increase in cardiac output, the flow through the lungs may increase by
fourfold up to sevenfold. This additional flow is accomplished by increasing
7328_C003.fm Page 97 Monday, September 18, 2006 10:57 AM
Cardiovascular Physiology 97
dCi ∆Ci
Ni = − Di = − Di (3.7)
dz δ
Di K Di A
w=− ∆pi = − DLi ∆pi (3.8)
δ
In this equation, w is the total amount of gas transported across the respi-
ratory membrane cross-sectional area, A. Since A and δ are very difficult to
determine accurately, all the parameters in the equation are lumped together
as the diffusing capacity, DL. Under resting conditions, the diffusing capacity
of O2 is 21 ml min− mmHg and that of CO2 is 400 to 450 ml min− mmHg. The time-
averaged, partial pressure difference across the respiratory membrane under
resting conditions is about 11 mmHg and the total amount of oxygen trans-
ferred across the respiratory membrane is about 252 ml/min. The total amount
of CO2 transferred is about 200 ml/min. Since the solubility of oxygen in aque-
ous solution is rather poor, oxygen diffuses across the red blood cell membrane
in the capillary and then reacts with hemoglobin in the intracellular fluid
volume so that it is transported principally as oxy-hemoglobin.
7328_C003.fm Page 98 Monday, September 18, 2006 10:57 AM
Anterior Anterior
communicating cerebral
artery artery
Internal
carotid Anterior
artery choroidal
artery
Middle Posterior
cerebral communicating
artery artery
Posterior
cerebral
artery
Superior
Pontine cerebellar
arteries artery
Basilar artery
FIGURE 3.20
Schematic of blood flow in the human brain showing the Circle of Willis connecting flow from
the vertebral and carotid arteries. (Redrawn from DeArmond, S.J., Fusco, M.M., and Dewey,
M.M. (1974) Structure of the Human Brain, Oxford University Press, London. With permission.)
7328_C003.fm Page 99 Monday, September 18, 2006 10:57 AM
Cardiovascular Physiology 99
3.8 Microcirculation
In the previous sections, we reviewed blood flow in both the systemic and
pulmonary circulations. We started with the contractile function of the heart
in generating a sufficient pressure gradient in order to induce flow along the
circulatory system. We also discussed the arterial system, which acts as the
7328_C003.fm Page 100 Monday, September 18, 2006 10:57 AM
Renal artery
Renal vein
(a)
Glomerulus
Efferent (capillaries)
arteriole Peritubular
capillaries
Juxtaglomerular
apparatus
Afferent
arteriole
(b)
FIGURE 3.21
Schematic of the blood supply to the kidneys: (a) renal artery and vein, and (b) details of the
afferent and efferent arterioles along with the Bowman’s capsule used in the filtration of blood
of waste products. (Redrawn from Silverthorn, D.U. (2001) Human Physiology: An Integrated
Approach, 2nd ed., Prentice Hall, Upper Saddle River, NJ. With permission.)
7328_C003.fm Page 101 Monday, September 18, 2006 10:57 AM
Arteries Vein
Venule
Precapillary
sphincters
Arteriole
Metarteriole
Capillaries
Small venule
Arteriovenous
bypass
Precapillary
sphincters
FIGURE 3.22
Schematic of the blood vessels in a typical microcirculation. (Redrawn from Silverthorn, D.U.
(2001) Human Physiology: An Integreted Approach, 2nd ed., Prentice Hall, Upper Saddle River,
NJ. With permission.
main conduits transporting blood to the various visceral organs and the
peripheries of the human body. In large blood vessels, the typical size (i.e.,
diameter) of the blood vessel is several magnitudes larger than the formed
elements of blood, such as the red blood cells and the white blood cells, so
that blood may be treated as a homogeneous fluid. These large vessels mainly
function as conduits for supplying blood to the capillary network.
The exchange of nutrients and cellular excreta between the tissues and the
flowing blood occurs at the level of capillaries. The fluid mechanics within
the arterioles and capillaries are characteristically different from those in
large arteries. Hence, flow of blood in the system of blood vessels consisting
of arterioles, capillaries, and venules is termed the microcirculation. A sche-
matic of a typical micro-vascular bed is shown in Figure 3.22. The arteries
divide into arterioles and the diameters of arterioles can range up to 100µm.
The walls of the arterioles are highly muscular and have the ability to alter the
vessel diameter several fold and, hence, vary the resistance to flow across the
micro-vascular bed. The arterioles divide into metarterioles with diameters
of about 20 µm. The musculature in the metarterioles is not continuous, as
7328_C003.fm Page 102 Monday, September 18, 2006 10:57 AM
smooth muscle fibers are found intermittently surrounding the wall. The
branching pattern of the capillary network varies from tissue to tissue as
well as within any one bed. Some of the capillary vessels (referred to as “true
capillaries”) that branch from the metarterioles have precapillary sphincters
at their origin. The sphincters consist of one or two smooth muscle cells
wrapped around the arteriolar–capillary junction. The diameter of the true
capillary ranges from 5 to 8µm, barely large enough for red blood cells and
other elements to squeeze through. Some of the capillaries are large and do
not have the precapillary sphincters and are referred to as the preferential
channels. If the precapillary sphincters constrict, the blood is shunted
through the low-resistance preferential channels and, hence, the micro-
vascular bed unit is bypassed.
The concentration of oxygen in the tissue supplied by the unit regulates
the degree of opening and closing of the precapillary sphincters and amount
of blood flowing through the true capillaries. The wall of the true capillaries
consists of a unicellular layer of endothelial cells surrounded by a thin
basement membrane of about 0.5 µm in thickness. Pores or passageways,
far apart from one another with a fraction of the lumen surface area, are
present in the capillary membrane through which water and water-soluble
substances pass between the lumen of the capillary and the interstitial fluid.
Several of these water-soluble substances, such as sodium and chloride ions
and glucose, are not lipid-soluble and, therefore, can only pass through the
pores. Transport of fluid between the capillary lumen and the interstitial
fluid depends on the hydrostatic and colloid osmotic pressure difference
between the capillary blood and the interstitial fluid. The rate at which water-
soluble substances transport across the pores depends upon the permeability
of the capillary pores. The diameter of the pores is about 80 to 90 Α . Water
molecules are much smaller than the pore diameter and, hence, have a larger
permeability compared to that of the plasma protein molecules. Lipid soluble
substances, such as oxygen and carbon dioxide, are transported by means
of diffusion across the cell membrane into the capillary. The amount of
diffusion is directly proportional to the concentration difference across the
two sides of the membrane. At the distal end, the capillaries merge into post
capillaries and then into venules. In the systemic circulation, venules merge
into larger veins terminating in the superior and inferior vena cavae, which
drain the blood into the right atrium. The lymphatic circulation drains the
excess fluid from the interstitial locations. Lymphatic capillaries are blind-
ended sacs present in the interstitial space around the collecting venules and
which merge into collecting lymphatics. These become lymphatic vessels,
which run parallel to the veins and which drain via the thoracic duct into
the subclavian vein in the neck.
Even the largest arterioles in the micro-vascular bed have a diameter that
is only about 15 times larger than that of the red blood cells. The true capillary
diameter is about the same size or smaller than that of the red blood cell
and, thus, the cells must squeeze through the capillaries. In such flow situ-
ations, the blood cannot be considered as a continuum and the interaction
7328_C003.fm Page 103 Monday, September 18, 2006 10:57 AM
between particulates (i.e., cells) and the plasma must be taken into account.
Due to the small diameters and low velocities in the microcirculation, viscous
effects dominate over inertial effects and the magnitude of the Reynolds
number ranges between 0.005 and 0.5. However, the velocity gradient at the
wall or the wall shear rates in the microcirculation is of the order of 1000 s-1
and are relatively large compared to that of those in the arteries as we will
observe in subsequent chapters. The study of microcirculatory dynamics is
beyond the scope of this textbook, but can be found in treatises such as Caro
et al. (1978) and Fung (1984).
efferent arterioles in the kidneys to reduce the renal blood flow and the
glomerular filtrate. In the brain, the concentration of carbon dioxide and
hydrogen ions also influences the blood flow rate to the organ in addition
to the effect of changes in tissue oxygen concentration.
Long-term control of local blood flow (i.e., effecting changes over a period
of weeks or months) in proportion to the needs of the tissue is the result of
increase or decrease of the physical size and number of blood vessels sup-
plying the tissue. If the metabolic rate in a given tissue is increased over a
prolonged period, the vascularity is also increased. Growth of new blood
vessels to increase vascularity results from the presence of various peptides,
such as vascular endothelial growth factor (VEGF), fibroblast growth factor
(FGF), and angiogenin, among others. Another mechanism for increasing
the vascularity is by development of a collateral circulation. When a blood
vessel is blocked, collateral vessels develop in the region around the blocked
vessel over a period of time in order to compensate for the decrease in blood
supply. For example, collateral blood supply is an important mechanism for
keeping the cardiac muscles viable and possibly preventing a heart attack
after thrombosis of a coronary vessel.
The global nervous control of circulation and cardiac function is performed
by the autonomic sympathetic and parasympathetic nervous systems. The
control of both these nervous systems is affected by the vasomotor center in
the brain. The sympathetic vasomotor nerve fibers pass through the spinal
cord and innervate the visceral organs and the heart as well as the vascula-
ture in the peripheral region. Parasympathetic nerve fibers do not play a
major role in the control of circulation, but control the heart by the nerve
fibers contained in the vagus nerve. The sympathetic nerve fibers carry a
large number of vasoconstrictor fibers that secrete norepinephrine. Constric-
tion of the blood vessels by norepinephrine increases the resistance to blood
flow (due to a reduction in the radius of the vessels), resulting in an increase
in the blood pressure. The large veins are also constricted and, hence, blood
from these highly distensible vessels is displaced toward the heart. Thus,
the volume of ventricular filling increases, resulting in increased cardiac
output through the Frank–Starling mechanism. With increase in the filling
volume, the heart muscles contract with increased force and, hence, the
stroke volume (and the cardiac output) increases together with an increase
in blood pressure. The sympathetic nervous system also directly acts on the
heart by increasing both the heart rate and the force of contraction. Parasym-
pathetic nerve fibers, on the other hand, decrease the heart rate and cause a
slight decrease in heart contractility resulting in a reduction in cardiac output
and blood pressure.
Baroreceptors, or pressoreceptors, are stretch receptors in the nervous
control system that help to control the arterial blood pressure. These recep-
tors are located particularly in the wall of the aortic arch and in the walls of
the internal carotid artery slightly above the bifurcation in the carotid sinus.
The baroreceptors respond rapidly to changes in blood pressure (rather than
to the mean arterial pressure). Should the blood pressure rapidly rise, the
7328_C003.fm Page 105 Monday, September 18, 2006 10:57 AM
signals from the baroreceptors inhibit the vasoconstrictor nerves and excite
the vagal parasympathetic nerves. The net result is the dilation of the periph-
eral blood vessels, while the heart rate and force of contraction of the cardiac
muscles is reduced. Should the blood pressure rapidly decrease, the opposite
effects will take place resulting in an increase in blood pressure. The nervous
control system can act to rapidly control the blood pressure for a short time
on the order of minutes.
The kidneys play an important role in the long-term control of blood
pressure and in the development of hypertension (i.e., high blood pressure).
When the body contains too much extracellular fluid, the blood volume and
mean arterial pressure rises. The kidneys react to the increased blood pres-
sure by excreting excess body fluid and return the blood pressure to its
normal value. This phenomenon of increasing urinary output to control the
blood pressure is known as pressure diuresis. Aside from using changes in
the extracellular fluid in the kidneys to control blood pressure, the
Renin–Angiotensin system in the kidneys also plays an important role in
blood pressure control. With a decrease in the blood pressure, renin is
released in the kidneys. This enzyme reacts chemically with plasma proteins
to release Angiotensin I and II. Intense vasoconstriction caused by the effect
of Angiotensin II in the artery results in an increase in blood pressure. Mild
constriction of the veins displaces additional blood to the heart resulting in
increase in ventricular filling and cardiac output and an increase in blood
pressure as well. When the mean arterial pressure rises to magnitudes above
110 mmHg (along with corresponding increases in systolic and diastolic
blood pressures), this condition is referred to as hypertension. Hypertension
can be a result of release of large quantities of renin in the kidneys due to
malfunction of the renin-secreting juxtaglomerular cells. Chronic high blood
pressure will result in excess workload and eventual failure of the heart,
rupture of blood vessels, particularly in the brain, renal hemorrhage, and
renal failure.
3.10 Atherosclerosis
Diseases in the blood vessels and in the heart, such as heart attack and stroke,
are responsible for nearly half of the deaths in the U.S. The underlying cause
for these events is the formation of lesions, known as atherosclerotic plaques,
in the large- and medium-sized arteries in the human circulation. These
plaques can grow and occlude the artery and, hence, prevent blood supply
to its distal bed. Plaques containing calcium can also rupture and initiate the
formation of blood clots (thrombus). The clots can form emboli, which are
carried in the bloodstream and occlude smaller vessels, resulting in interrup-
tion of blood supply to the distal bed. Plaques formed in coronary arteries
can lead to heart attacks, while those formed in the cerebral circulation can
7328_C003.fm Page 106 Monday, September 18, 2006 10:57 AM
result in stroke. There are a number of risk factors for the presence of athero-
sclerotic lesions. Those for which we have no control include gender, age,
and familial history of early cardiovascular disease. Risk factors that can be
controlled to minimize the occurrence of atherosclerosis include smoking,
obesity, sedentary lifestyle, elevated serum cholesterol and triglycerides,
untreated high blood pressure, and diabetes mellitus.
FIGURE 3.23
Classical light micrograph of a cross section of a coronary artery that contains a large atheroscle-
rotic lesion. (Redrawn from Schlant, RC. and Alexander, R.W., Eds. (1994) Hurst’s The Heart, Arteries
and Veins, 8th ed., McGraw-Hill, Health Professions Division, New York. With permission.)
7328_C003.fm Page 107 Monday, September 18, 2006 10:57 AM
cells from monocytes and mast cells, and their continued accumulation in
the intima leads to an atherosclerotic lesion, or fatty streak. With the con-
tinuing effect of hypercholesterolemia, endothelial cell injury, and abnormal
hemodynamics, the fatty streak will grow into a fibro-fatty lesion containing
multiple layers of SMC, connective tissue, macrophages, and T-lymphocytes.
With further growth, a fibrous cap will form to create an advanced lesion.
The role of elevated blood cholesterol levels in the development of athero-
sclerosis is well established. Cholesterol is a lipid that is absorbed in the
digestive tract and is not soluble in aqueous solutions. Cholesterol combines
with a lipoprotein molecule so that it can dissolve in the plasma. Cholesterol
that combines with high density lipoproteins (HDL) is associated with
lower risk of heart attacks, whereas elevated levels of low density lipoprotein
(LDL) cholesterol is associated with arterial disease. The LDL carrier is
necessary for cholesterol uptake into all of the body’s cells. The LDL portion
of the complex combines with an LDL receptor in the cell membrane and
the receptor–LDL–cholesterol complex enters into the cell by endocytosis.
However, if LDL-cholesterol is not absorbed into the cells and remains in
the blood, the excess LDL-cholesterol will move into the walls of the arteries.
Endothelial cells
Elastic lamella
Smooth
(a) muscle cells
Monocytes
Platelets
(b)
Lipid arootets
(c)
Monocytes smooth
muscle cells
(d)
Cholesterol crystals
Blood vessel
(e)
FIGURE 3.24
Postulated sequence of events in the pathogenesis of atherosclerosis. (From White, R.A. and
Cavaye, D.M. (1993), in A Text and Atlas of Arterial Imaging – Modern and Developing Technology,
Cavaye, D.M. and White, R.A., Eds., W.W. Lippincott, Philadelphia. With permission.)
for normal blood flow. Once the remodeling limit is reached, however, the
lesion protrudes into the lumen forming a stenosis. In the advanced stages
of the lesion, the cells in the plaque may die and form calcified scar tissue.
The deposits of calcium create the hardening of the arteries, which reduces
the distensibility in the region of the atherosclerotic plaques. If the plaques
have a rough surface or if they rupture, platelets will stick to the damaged
area and to each other to form a thrombus on the arterial wall. Figure 3.26
7328_C003.fm Page 109 Tuesday, September 19, 2006 1:53 PM
Lipoprotein
with oxidized Foam cells
cholesterol Macrophage
FIGURE 3.25
Schematic of the components involved in the development of atherosclerosis. (From Kleinstreu-
er et al. (2001). A computational analysis of particle-hemodynamics and prediction of the onset
of arterial disease, in Cardiovascular Techniques, Leondes, C., Ed., CRC Press, Boca Raton, FL.
With permission.)
shows the progression of initial fatty lesions to calcified plaques in the intimal
layer of a blood vessel and the gradual decrease in the lumen cross section
available for blood flow.
The common sites of predilection of atherosclerosis in the human circulation
are shown in Figure 3.27. These include the coronary arteries, the branching
of the subclavian and common carotids in the aortic arch, the bifurcation of
the common carotid to the internal and external carotids (especially in the
carotid sinus region distal to the bifurcation), the renal arterial branching in
the descending aorta, and in the ileo-femoral bifurcations of the descending
aorta. The common feature for the development of the lesion is the presence
of curvature, branching, and bifurcation present in these sites. Besides the
geometry, the fluid dynamics at these sites can be anticipated to be consid-
erably different from that at other segments of arteries that are relatively
straight and devoid of any branching segments. Hence, a number of inves-
tigators have attempted to link fluid dynamically induced stresses with the
formation of atherosclerotic lesions in the human circulation. Hemodynamic
theories for the genesis of atheromatous plaques and detailed descriptions
of the fluid dynamics at corresponding sites will be the topic of discussion
in subsequent chapters, where steady and unsteady flow models in the
human circulation will be considered.
7328_C003.fm Page 110 Monday, September 18, 2006 10:57 AM
Liceration Normal
Medial fibrosis
Medial Intimal or
Thrombus
muscle fibers medial fibrosis
FIGURE 3.26
Schematic of the cross section of a blood vessel illustrating the growth of atherosclerotic lesions
progressively reducing the lumen area available for blood flow to the distal beds supplied by the
vessel. (Redrawn from Engeler et al., (1991) Intravascular sonography in the detection of arterioscle-
rosis and evaluation vascular interventional procedures, in AJR 156: 1087–1090. With permission.)
3.11 Problems
3.1 Under normal conditions, the mean systemic arterial pressure at the aortic
root is 95 mmHg and the corresponding central venous pressure is 0 mmHg.
The corresponding pressures in the pulmonary artery and the left atrium are
13 mmHg and 4 mmHg, respectively. If the cardiac output is 5.2 l/min, compute
the peripheral resistance in the systemic and pulmonary circulation. Express
your answers in PRUs as well as in SI units. In the case of moderate exercise,
the mean aortic pressure rose to 150 mmHg and the cardiac output increased
by three times the normal value. Compute the percentage change in the sys-
temic peripheral resistance. (Conversion factor: 1 mmHg = 133.3 Pa.)
3.2 In a clinical situation, the stroke volume for a patient was measured at
45 ml/stroke and the heart rate was recorded to be 85 beats per minute.
Compute the cardiac output for the patient. Cardiac Index is expressed as
the ratio of cardiac output over the body surface area. Assuming that the
7328_C003.fm Page 111 Monday, September 18, 2006 10:57 AM
I II
Proximal
Mid-proximal
Distal
IV
III
FIGURE 3.27
Common sites for the presence of atherosclerotic plaques in the human circulation. (Redrawn
from DeBakey et al., (1985) Patterns of artherosclerosis and their surgical significance, in Ann.
Surg. 201: 115–131. With permission.)
This estimate is smaller than the actual work done by the heart. Why? How
far could one raise a 100-lb weight with this amount of work?
3.6 The amount of work done per cycle computed in Problem 3.5 is the left
ventricular work output. For a heart rate of 70 bpm, the left ventricular
output per minute can be computed. The amount of oxygen consumed by
the myocardial muscle per minute computed in Problem 3.3 can be converted
as work input to the heart (1 ml of O2 consumed = 20 Joules = 14.75 lbf-ft).
Compute the efficiency of the heart as the ratio of work output to the work
input.
7328_S002.fm Page 113 Friday, July 28, 2006 4:36 PM
Part II
4
Rheology of Blood and Blood Vessel
Mechanics
CONTENTS
4.1 Rheology of Blood ..................................................................................... 116
4.1.1 Viscometry and Theory for Capillary, Coaxial
Cylindrical, and Cone and Plate Viscometers.......................... 116
4.1.1.1 Capillary Viscometer ...................................................... 116
4.1.1.2 Coaxial Cylinder Viscometer ........................................ 119
4.1.1.3 Cone and Plate Viscometer ...........................................122
4.1.2 Physical Properties of Blood .......................................................123
4.1.3 Viscous Behavior of Blood...........................................................124
4.1.3.1 Plasma...............................................................................124
4.1.3.2 Whole Blood ....................................................................126
4.1.3.3 Effect of Hematocrit .......................................................130
4.1.3.4 Effect of Temperature.....................................................130
4.1.3.5 Effect of Protein Content in Plasma ............................131
4.1.3.6 Yield Stress for Blood.....................................................131
4.1.3.7 Effect of Tube Diameter .................................................135
4.1.4 Pressure-Flow Relationship for Non-Newtonian Fluids ........139
4.1.4.1 Power Law Fluid ............................................................140
4.1.4.2 Bingham Plastic...............................................................141
4.1.4.3 Casson’s Fluid .................................................................144
4.1.5 Hemolysis and Platelet Activation with Fluid
Dynamically Induced Stresses ....................................................150
4.2 Blood Vessel Mechanics............................................................................152
4.2.1 Structural Components of the Blood Vessel Wall
and their Contribution to Material Behavior ...........................152
4.2.2 Material Behavior of Blood Vessels............................................155
4.2.2.1 Assumptions in the Analysis ........................................158
4.2.2.2 Experimental Results......................................................158
4.2.3 Residual Stress on Blood Vessels................................................162
4.2.4 Material Characterization of Cardiac Muscle...........................164
4.3 Summary .....................................................................................................165
4.4 Problems......................................................................................................165
115
7328_C004.fm Page 116 Friday, October 6, 2006 12:25 PM
P1 > P2
R
P1 P2
r Direction
of flow
FIGURE 4.1
Steady flow of Newtonian fluid through a straight tube of circular cross section.
Equation 1.43c). There, the relationship between the flow rate and the pres-
sure gradient was found to be:
Vmax
Q = Vave ( πR 2 ) = ( πR 2 )
2
(1.72)
∆pR 2 ∆pπR 4
= ( πR 2 ) =
8µL 8µL
Keep in mind that one of the fundamental assumptions behind this equa-
tion is that the flow is fully developed or, L >> 2R. The nature of flow devel-
oping in the entrance region of a tube is discussed in Chapter 1 (section 1.8:
Fluid Mechanics in a Straight Tube) where an approximate relationship for
the entry length prior to fully developed flow is derived. One way to achieve
this condition is to use a very narrow, or “capillary,” tube.
Thus, in a capillary viscometer of radius R and length L, if the flow rate
and pressure drop of a fluid can be accurately measured, the viscosity coef-
ficient can be determined as
∆pπ R 4
µ= (4.1)
8QL
Referring to Figure 4.1, the net force pushing the fluid to the right is:
Fp = p1π r 2 − p2 π r 2 = ( p1 − p2 )π r 2
while the shear force, which retards the motion acts on the circumferential
surface of the fluid element, is given by
Fτ = τ 2 π rL
7328_C004.fm Page 118 Friday, October 6, 2006 12:25 PM
In fully developed, steady flow, these two forces balance each other and
we obtain
( p1 − p2 )π r 2 = τ(2 π rL)
( p1 − p2 ) r ∆pr
τ= 2L =−
2L
From this expression for the shear stress, the magnitude of the shear stress
at the wall will be:
( ∆p)r ∆pR
τw = = (4.2)
2L r=R 2L
du
τ = −µ (1.1)
dr
where u is the velocity of the fluid. Hence, the shear rate at the wall is given
by
∆pR
γ w =
2µL
Using the relationship for ∆p from Equation 1.74, wall shear rate becomes
4Q
γ w = (4.3)
πR 3
Thus, having obtained values for wall shear stress and wall shear rate
(Equation 4.2 and Equation 4.3), respectively, one can calculate fluid viscosity
from their ratio.
Capillary viscometers include devices that measure the absolute viscosity
coefficient as well as the viscosity relative to fluids of known viscosity. The
flow through the capillary may be induced by an externally applied pressure
gradient or by gravitational forces. There are several precautions that need
to be observed in using these instruments. The derivation of the relationship
given in Equation 4.3 included several assumptions, such as laminar
7328_C004.fm Page 119 Friday, October 6, 2006 12:25 PM
(streamlined), steady, fully developed flow, and that the fluid is Newtonian.
To avoid the effects of turbulent flow in deviating from the derived relation-
ship, experiments must be performed at lower Reynolds numbers where
streamlined flow will be assured. The assumptions of steady and fully devel-
oped flow imply that the velocity profile within the capillary is parabolic and
no acceleration of the fluid is occurring within the viscometer. In most types
of capillary viscometers, part of the applied pressure is used to induce kinetic
energy to the fluid. Moreover, a small amount of energy is also expended in
overcoming the viscous forces at the converging and diverging streamlines
at the entrance and exit to the capillary. Kinetic energy and Couette correc-
tions, respectively, need to be taken into account to correct for the errors due
to these two effects and are detailed in Dinsdale and Moore (1962).
In viscometers with externally applied pressure, the effective pressure
gradient includes the externally applied pressure gradient as well as the
hydrostatic head of the fluid in the viscometer. Hydrostatic corrections are
also needed to account for the constantly decreasing hydrostatic head (Dins-
dale and Moore, 1962). Use of this device on Newtonian fluids is not only
required for the Poiseuille relation to hold, but it is also essential due to the
variable shear rate within the tube. Differentiation of the parabolic velocity
profile shows that the rate of shear is not constant, but is a function of radial
position, r. Thus, a fluid particle near the wall would be exposed to a high
shear rate while one at the tube center would experience zero shear rate. If
the fluid viscosity were shear rate dependent, i.e., µ( γ ) , then a range of
viscosities would be seen and no meaningful value would be obtained. The
relationships between pressure gradient and flow rate for non-Newtonian
fluids in straight circular tubes are discussed later in the chapter.
Torsion wire
Graduated scale
bob
Cup
h R
R2
FIGURE 4.2
Schematic of a coaxial (Couette) cylinder viscometer.
T = Ft r
or
T = τ r (2 πrh)r
The torque, T, measured at the surface of the inner cylinder must be the
same as the torque at any arbitrary radius r since the motion is steady.
Rewriting in terms of the shear stress, we get
T
τr = (4.4)
2 πr 2 h
It can be observed from the above relationship that the shearing stress is
inversely proportional to the square of the distance from the axis of rotation.
It is interesting to note, however, that by using an annular gap that is small
compared to the cylinder radii (i.e., dr << R1, R2), the shearing stress on the
fluid will be almost constant throughout the volume of the testing fluid and,
thus, approaches the case of Couette flow. (Couette flow is a classic config-
uration of flow between two infinitely large flat plates where either one plate
is in motion [as in this device] or a pressure gradient is present or both.)
7328_C004.fm Page 121 Friday, October 6, 2006 12:25 PM
V = rω
and, thus, the gradient of velocity (or rate of shear) will be given by
dV dω
=r +ω
dr dr
In this expression, the first term on the right-hand side represents the radial
velocity gradient and the second term is the angular velocity due to rigid
body rotation. Since we are interested in the tangential velocity gradient in
the fluid, the second term with the constant angular velocity can be
neglected. For a Newtonian fluid, τ r = µγ , or
dV dω
τ r = µγ = µ = µr
dr dr
T dω
= µr
2 πr 2 h dr
Consequently,
dω T
= (4.5)
dr 2 πµhr 3
4 πhR 2 R 2
T = 2 1 22 µΩ (4.6)
( R2 − R1 )
s
ψ h
r
FIGURE 4.3
Schematic of a cone and plate viscometer.
V = Ωr
The rate of shear will be given by the ratio of the linear velocity and the
gap between the cone and plate h at that radius. From the figure, it can be
observed that h = r tan ψ.
Ωr Ω
Thus, the shear rate γ = r tan ψ = ψ , provided that the cone angle ψ is small.
Furthermore, the shearing stress is also independent of the radius in this
configuration.
The total torque, T, can be obtained by the expression,
T= ∫ rτ dA
A
r
∫
T = τ r 2 π r 2 dr
0
and, thus,
3T
τr =
2 πR 3
7328_C004.fm Page 123 Friday, October 6, 2006 12:25 PM
τr 3T ψ
µ= = (4.7)
γ 2 πR 3Ω
8.5 µ
1.0 µ 2.5 µ
(a)
(b)
FIGURE 4.4
(a) Schematic of the biconcave shape of the red blood cell; (b) a scanning electron microscope
image of a human red blood cell magnified × 13,000.
TABLE 4.1
Summary of Selected Studies on Blood Viscosity Measurement Reported
in Literature of the 1960s
Viscometer Blood Anticoag. HCT Temp. Comments
Coaxial cylinder Human ACD 0–40% 20–40°C 6–200 sec1
Capillary Human Heparin 12.5–70% 25–32°C 12.8–335 sec1
power law
Cone and plate Dog Heparin 37°C 11.45–230 sec1
Cone and plate Human ACD Normal 37°C 2–100,000 sec1
Plate–coaxial Casson’s Eqn.
Coaxial cylinder Human Heparin 0–95% 37°C 0.052–52
Coaxial Human Heparin 43–48% 37°C 23–230 sec1
Cone and plate Human None/Heparin 33–57% 37°C 10–250 sec1
Oxalate
Coaxial cylinder Human ACD 42.5% 25–37°C 0.1–20 sec1
Cone and plate Human Heparin 0.80% 22–37°C 2.12 sec1
7328_C004.fm Page 127 Friday, October 6, 2006 12:25 PM
800
600
400
200
0
50 100 150 200 250
◊
Rate of shear γ (sec–1)
FIGURE 4.5
Shear stress normalized to the plasma viscosity vs. rate of shear plot from experimental data.
(Redrawn from Whitmore, R.L. (1968) Rheology of Circulation, Pergamon Press, New York. With
permission.)
the logarithmic shear stress vs. rate of shear plot (Figure 4.6). Even though
the data points exhibit a linear behavior over small ranges of shear rates,
a curvilinear relationship is observed over the entire range of shear rates
for which data have been obtained. Tests conducted with rotational vis-
cometers for shear rates ranging from 5 to 200 sec−1 follow the Power Law
1000
Shear stress/plasma viscosity µ (sec–1)
p
100
n = 0.75
10
n = 1.0
(Newtonian)
1
0.01 0.1 1 10 100 1000
◊
Rate of shear g (sec–1)
FIGURE 4.6
Shear stress vs. rate of shear plot in logarithmic scale. (Redrawn from Whitmore, R.L. (1968)
Rheology of Circulation, Pergamon Press, New York. With permission.)
7328_C004.fm Page 128 Friday, October 6, 2006 12:25 PM
30
(sec–1/2)
µp
20
Shear stress/plasma viscosity
10
0 5 10 15
Rate of shear γ◊ (sec –1/2)
FIGURE 4.7
Square root of shear stress adjusted to plasma viscosity vs. square root of shear rate from
viscometric data. (Redrawn from Whitmore, R. L. (1968) Rheology of Circulation, Pergamon Press,
New York. With permission.)
reasonably well for values of the exponent n between 0.68 and 0.8. Values
of n exceeding 0.9 have been reported from results obtained with capillary
viscometers.
Inspection of Figure 4.5 once again reveals that even though blood exhibits
a yield stress (y– intercept on the plot), the data does not suggest a linear
relationship between shear stress and rate of shear particularly at low shear
rates, a relationship indicative of an ideal Bingham plastic (τ = τ y + Kb γ ). As
an alternative, we could consider the constitutive relationship for the
Casson’s fluid ( τ = τ y + K c γ ) where there is a linear relationship between
the square root of shear stress vs. the square root of rate of shear. A plot of
such a curve together with the data obtained for whole blood is shown in
Figure 4.7.
A least square fit of the data yields the relationship
τ
= 1.53 γ + 2.0 (4.8)
µp
The apparent viscosity coefficient for whole blood at any rate of shear,
thus, can be computed from the above relationship. For example, at a rate
of shear of 230 sec−1, the apparent viscosity of whole blood is about 3.3 cP
using a value of 1.2 cP for plasma viscosity. This value compares favorably
7328_C004.fm Page 129 Friday, October 6, 2006 12:25 PM
0.6
0.4
0.2
Experimental data
Asymptotic
Casson values
equation
(4.8)
0
0.1 1 10 100 1000
◊
Rate of shear γ (sec–1)
FIGURE 4.8
A plot of apparent viscosity as a function of rate of shear. (Redrawn from Whitmore, R.L. (1968)
Rheology of Circulation, Pergamon Press, New York. With permission.)
with the experimentally determined values ranging from 3.01 to 5.53 cP.
Apparent viscosity data from experiments by various investigators for a
range of shear rates is shown in Figure 4.8. A plot of the constitutive rela-
tionship in Equation 4.8 is also included in the figure. As can be observed,
the apparent viscosity increases to relatively large magnitudes at low rates
of shear. This is because the red blood cells tend to aggregate into Rouleaux
formations and, thus, exhibit an increase in viscosity at low rates of shear.
Aggregates of red blood cells are shown in Figure 4.9. As the rate of shear
increases, the aggregates gradually break up if the rate of shear exceeds about
50 sec−1, and the viscosity coefficient approaches an asymptotic value of about
3.5 cP. Thus, when specifying the viscosity coefficient of blood, it is important
to specify the rate of shear at which the measurements were obtained.
As discussed earlier, apart from the fact that the rate of shear has a pro-
found effect on the measured apparent viscosity of whole blood, the vari-
ability of samples, the effect of the addition of anticoagulants, and the
experimental conditions must also be taken into account. Let us now look
into the effect of other variables on the viscosity coefficient of blood.
7328_C004.fm Page 130 Friday, October 6, 2006 12:25 PM
FIGURE 4.9
A scanning electron microscope view of aggregation of red blood cells magnified × 2000.
10
Spheres
8
52 sec–1
42 sec–1
6 64 sec–1
×
212 sec–1
×
Relative viscosity µr
3
×
2
×
×
1
0 20 40 60 80
Hematocrit H
(a)
FIGURE 4.10
(a) Effect of hematocrit on the relative viscosity of blood; (b) Effect of hardening of the red
blood cells on the relative viscosity. (Redrawn from Whitmore, R.L. (1968) Rheology of Circulation,
Pergamon Press, New York. With permission.)
10
Cells
Spheres hardened
8 normal
Relative viscosity µr 6
1
0 20 40 60 80
Hematocrit H
(b)
have shown that the yield stress depends strongly on fibrinogen concentra-
tion and is also dependent on the hematocrit (Figure 4.12 a and Figure 4.12b).
Several empirical formulae relating the yield stress to hematocrit and
fibrinogen concentration in blood have been reported in the literature
(Cooney, 1976). An empirical relationship for the yield stress as a function
of hematocrit can be given by
1 A ( H − Hm )
τy3 = (4.9)
100
40
Temp = 37°C
Hematocrit = 49%
Viscosity, cP 30
20
14
12 Globulin 2.2 wt%
10
8 Whole blood
6 Albumin 3.5 wt%
0 10 20 40 60 80 100 120
Shear rate, sec–1
FIGURE 4.11
Effect of plasma proteins on the viscosity of whole blood. (Redrawn from Copley, A.L. and
Stainsby, G., Eds. (1960) Flow Properties of Blood, Pergamon Press, Oxford. With permission.)
0.5
Hct
25%
30%
0.4 35%
40%
(dyn/cm2)
× 45% ×
××
0.3 ××
× ×
×
×
0.2
τy,
0.1
0
0 0.1 0.2 0.3 0.4 0.5
Fibrinogen conc., g/100 ml
(a)
0.3
(dyn/cm2)
0.2
0.1
τy,
0
25 35 45
Hematocrit (%)
(b)
FIGURE 4.12
Effect of (a) fibrinogen concentration and (b) hematocrit on the yield stress for blood. (Redrawn
from Merrill et al. (1965), The Casson equation and rheology of blood near zero shear, in
Proceedings of the Fourth International Congress on Rheology, Copley, A.L. (Ed.). With permission.)
zero at the center and a maximum at the wall. To initiate flow, the imposed
pressure drop must be large enough so that the wall shear stress,
R∆p
τw =
2L
will exceed the yield stress for blood. Equating the wall shear stress to the
value for the yield stress for blood (0.07 dynes/cm2), the pressure gradient
7328_C004.fm Page 135 Friday, October 6, 2006 12:25 PM
4.5
4.0
Apparent viscosity, cP
3.5
3.0
2.5
0 0.5 1.0 1.5 2.0 2.5
Tube radius, mm
FIGURE 4.13
Plot of apparent viscosity as a function of tube radius.
Axis of tube
1
V1 Bernoulli force
V2
2
Tube wall
Velocity profile
FIGURE 4.14
Schematic describing the rotation of red blood cells under the influence of a velocity gradient.
∆p 1 d du
− = µ r c 0≤r ≤R−δ
L r dr c dr
(4.11)
∆p 1 d dup
− = µ r R−δ ≤r ≤R (4.12)
L r dr p dr
where µ c and µ p represent the viscosities of the core and periphery, respec-
tively, while uc and up represent the velocities of the core and periphery,
respectively. The boundary conditions used to obtain a solution for the two
differential equations are (1) the velocity gradient is zero at the center of the
a a–δ
Core
region
FIGURE 4.15
Schematic of a cell-free zone layer model.
7328_C004.fm Page 137 Friday, October 6, 2006 12:25 PM
tube, (2) no slip occurs at the tube wall, and (3) the velocity and the shear
stress is continuous at the interface between the two zones. The equations
can be integrated with the above boundary conditions to solve for the veloc-
ity components and to obtain an expression for the flow rate as given below:
π R 4 ∆p δ µ
4
Q= 1 − 1 − 1 − p (4.13)
8µ p L R µc
By comparison, the expression for the flow rate for a homogeneous fluid
(Poiseuille’s Equation) was derived earlier as
π R 4 ∆p
Q=
8µ L
µ p
µ app =
)( )
(4.14)
(
1 − 1 −
δ 4
R 1−
µp
µc
µc
µ app =
(4.15)
1 + 4 Rδ µµc −1
p
∫
Q = 2 π u(r )rdr
0
7328_C004.fm Page 138 Friday, October 6, 2006 12:25 PM
Umax
Umax
FIGURE 4.16
Concentric layer model for flow in small tubes.
can be rewritten as
R R
∫ ∫
du
Q = π d[u(r )r ] − π r 2
2 dr
dr
0 0
Applying the “no slip” condition at the wall, the first integral on the right
side will be identically equal to zero. From the parabolic velocity profile
characteristic of Poiseuille flow, the expression for the velocity gradient will
be given by
du ∆pr
=−
dr 2µ L
R
π ∆p 3
Q=
2µ L
r dr ∫
0
(4.16)
N
π ∆p
Q=
2µ L ∑ (nε) ε
n= 1
3 (4.17)
7328_C004.fm Page 139 Friday, October 6, 2006 12:25 PM
N
N 2 ( N + 1)2
∑n
n= 1
3 =
4
N N
N 2 ( N + 1)2
∑n= 1
(nε)3ε = ε 4 ∑n
n= 1
3 = ε4
4
2
π ∆pR 4 ε
Q= 1+ (4.18)
8µ L R
µ
µ app = (4.19)
ε 2
1+ R
εµ p
δ= (4.20)
2(µ c − µ p )
( ∆p)r ∆pR
τw = = (4.2)
2L r=R 2L
7328_C004.fm Page 140 Friday, October 6, 2006 12:25 PM
The expression for the velocity profile and the flow rate for steady flow of
a Newtonian fluid through a tube was derived in Chapter 1, section 1.8: Fluid
Mechanics in a Straight Tube, Equation 1.70 and Equation 1.72, respectively,
as well as in the description of the capillary viscometer in this chapter. We
will now derive the corresponding relationships for fluids that follow the
Power Law, Bingham Plastic, and Casson’s constitutive relationships.
n
du
τ = K pl γ n = K pl (4.21)
dr
n
du r ∆p
K pl = −
dr 2L
Rewriting
1
∆p n
∫ ∫r
1
du = − n dr
2 K pl L
1
∆p n
n n+ 1 n
u = − r +C
2 K pl L n+1
1
∆p n
n n+ 1
C= R n
2 K pl L n+1
Substituting this value for C, the expression for the velocity profile becomes
1
n ∆p n
R n+ 1 n − r n+ 1 n
u=
n + 1 2 K pl L
(4.22)
7328_C004.fm Page 141 Friday, October 6, 2006 12:25 PM
Q= ∫ 2πrudr
0
n ∆p
1
n R n+ 1 R
= 2π
n + 1 2 K pl L
0
∫
R nrdr − r n+1 nrdr
0
∫
n ∆p
1
n R n+ 1 a
= 2π
n + 1 2 K pl L
0
∫
R nrdr − r 2 n+1 ndr
0
∫
1
n R 3 n+ 1 R
= 2π
n ∆p R n+ 1 n r 2 − r n
n + 1 2K L 2 3 n+ 1
n
pl 0 0
1
n ∆p n R 3 n+ 1 n R 3 n+ 1 n
= 2π − 3 n+ 1
n + 1 2 K pl L 2 n
1
n+ 1 R 3 n+ 1 n
n ∆p n
n
= 2π
n + 1 2 K pl L 2 3 n+1
n
1
nπ R 3 R∆p n
Q=
3n + 1 2 K pl L
(4.23)
The flow distribution in the tube given by Equation 4.23 depends upon
the value for n. The value for n for human whole blood can be approximated
to ~ 0.75 over the range of rates of shear from 10 to 1000 1/s.
du
τ = τ y + Kb γ = τ y − Kb (4.24)
dr
7328_C004.fm Page 142 Friday, October 6, 2006 12:25 PM
where τ y is the yield stress and Kb is the viscosity coefficient for the Bingham
plastic. Again, substituting τ into Equation 4.2, we get
du r ∆p τ y
=− + (4.25)
dr 2 Kb L Kb
or
∆p τy
du = − rdr + dr
2 Kb L Kb
Integrating, we obtain
∆p r 2 τ y
u=− + r+C
2 Kb L 2 Kb
∆p R 2 τ y
C= − R
2 Kb L 2 Kb
∆p r 2 τ y r ∆pR 2 τ y R
u=− + + −
2 K b L 2 K b 4K b L µ b
Rearranging terms
∆p τy
u= (R 2 − r 2 ) − (R − r ) (4.26)
4K b L Kb
From the expression for the wall shear stress at the inner surface of the
tube (r = R) given by
R∆p
τw =
2L
until the shear stress at the wall, τ w , exceeds the Bingham Plastic yield stress,
there is no flow possible through the tube. However, once the wall shear
7328_C004.fm Page 143 Friday, October 6, 2006 12:25 PM
stress exceeds the yield stress, flow is induced through the tube. We can also
determine a radius Rc at which the shear stress equals the yield stress
2 Lτ y
Rc =
∆p
The fluid will have a constant velocity uc between the center of the tube
and Rc and the velocity profile will follow Equation 4.26 from Rc to R. Using
Equation 4.26, the core velocity can be written as
∆p τy
uc =
4K b L
( R 2 − Rc 2 ) −
Kb
(R − Rc )
Rc ∆p
Substituting τ y = 2L , the core velocity is:
∆p
uc =
4K b L
( R 2 − Rc 2 ) − R2c ∆Lp K1 (R − Rc ) (4.27)
b
The flow rate through the tube is equal to the sum of the flow through the
core and the peripheral regions:
Q = Qcore + Qper
where
Qcore = uc π Rc 2
Thus,
∆pπ ( R 2 − Rc 2 ) Rc 2 Rc 3
Qcore = − (R − Rc ) (4.28)
Kb L 4 2
Qper = ∫ 2πrudr
Rc
7328_C004.fm Page 144 Friday, October 6, 2006 12:25 PM
∆p ( R 2 − r 2 ) Rc
uper = − ( R − r )
Kb L 4 2
( R 2 − r 2 ) Rc
R
∆p
Qper =
Kb L ∫
Rc
4
− ( R − r ) 2πrdr
2
Integrating
π∆p R 4 R 2 Rc 2 5 4 Rc R 3 RRc 3
Qper = − − R − +
Kb L 8 2
(4.29)
4 24 c 6
π∆p R 4 Rc 4 Rc R 3
Q= + −
Kb L 8 6
(4.30)
24
2τ y
Rc = ∆p
L
4
πR 4 ∆p 4 2 τ y 1 2 τ y
Q= 1 − ∆p + ∆p (4.31)
8K b L 3 R ( L ) 3 R ( L )
If the yield stress becomes zero, the flow relationship will reduce to that
of Poiseuille flow for a Newtonian fluid. Equation 4.31 is known as the
Buckingham’s equation for an idealized Bingham plastic.
τ = τ y + K c γ (4.32)
7328_C004.fm Page 145 Friday, October 6, 2006 12:25 PM
Replacing γ by − du
dr and the expression for the yield stress, we obtain
du 1 r ∆p τy
− = −
dr K c 2L Kc
du 1 r ∆p τ y 2 r τ y ∆p
− = 2 + 2− 2
dr K c 2 L K c Kc 2L
or
1 ∆p τy 2 τ y ∆p
du = − rdr − 2 dr + 2 rdr
Kc2 2 L Kc Kc 2L
Integrating
3
1 ∆p r 2 τ y 2 τ y ∆p r 2
u=− − 2 r+ 2 +C
2
Kc 2 L 2 Kc Kc 2L 3 2
3
1 ∆p a2 τ y 2 τ y ∆p R 2
C= + 2 R+ 2
2
Kc 2 L 2 Kc Kc 2L 3 2
and substituting for C and rearranging terms, we obtain the relationship for
the velocity profile as
u=
1 ∆p 2 2
Kc 2 4L
τy
(R − r ) + 2 (R − r ) −
Kc
4
3K c 2 2L (
τ y ∆p 3 3
R 2 −r 2 ) (4.33)
2 Lτ y
Rc =
∆p
7328_C004.fm Page 146 Friday, October 6, 2006 12:25 PM
Substituting for the yield stress from above in Equation 4.33, we get the
core velocity as
uc =
1 ∆p 2
Kc 2 4L
( R − Rc 2 ) + K1 2 R2c ∆Lp ( R − r ) − 3K4 2
c c 2L 2L (
Rc ∆p ∆p 3 2 3
R − Rc 2 )
Rewriting the equation, we get
uc =
( ) R
∆p
L
−
8 3 R2
Rc R 2 + 2 Rc R − c
4K c 2
2 (4.34)
3 3
For r > Rc , the expression for the velocity profile will be:
uper =
( ) R
∆p
L 2 − r 2 + 2 Rc R − 2 Rc R −
8 3
Rc R 2 +
8 3
4K c 2
Rc R 2 (4.35)
3 3
Q = Qcore + Qper
and the expression for the flow in the core region can be simplified as
Qcore = π
( ) R R
∆p
L
−
8 52 32 R4
Rc R + 2 Rc 3R − c
4K c
2 2 (4.36)
2 c
3 3
Qper = ∫u
Rc
per 2πrdr
( )
∆p R
3
4K 2 ∫
L 8 3 8
= 2π R 2 − r 2 + 2R R − 2R r − R R 2+ R r 2 rdr
c c 3 c 3 c
c Rc
7328_C004.fm Page 147 Friday, October 6, 2006 12:25 PM
Qper = π
( ) R
∆p 4
L
− R 2 Rc 2 +
13 4 2
R + R R 3 − 2 RRc 3 −
8 7 8 3 5
Rc R 2 + R 2 Rc 2
4K c 2 2 42 c 3 c 7 3
(4.37)
Adding Equation 4.36 and Equation 4.37, we get the expression for the
flow through the tube as
Q=
( ∆p L ) π R 4 − Rc 4 + 2 R R3 − 8 7
4K c 2 2 c Rc R 2
42 3 7
Substituting for Rc in terms of the yield stress, the flow rate is given by
1
4 2 τ y 16 2 τ y
4
1 2τ y
2
πR 4 ∆p
Q= 1 − + − (4.38)
8K c 2 L 21 ( ∆p L ) R 3 ( L ) 7 ( ∆p L ) R
∆p
Assuming steady flow, one can compute the flow dynamics though a
typical artery using the relationships derived above for Newtonian and non-
Newtonian fluids. For a typical human peripheral vessel, the pressure gra-
dient ∆p/L can be computed as 6 dynes/cm3 from the Poiseuille relationship
assuming a flow rate of 140 ml/min and a radius of 0.43 cm. Using the
Poiseuille relationship, we are assuming a Newtonian behavior for human
whole blood with a viscosity coefficient of 3.5 cP. The wall shear stress can
be computed to be 1.29 dynes/cm2 and the corresponding wall shear rate is
36.86/sec as shown below.
and
γ = τ w µ = 36.9/sec.
Applying the same wall shear stress and the shear rate in the constitutive
relationships for the other fluids, the consistency indices and the correspond-
ing flow rate through the artery can be computed with the magnitudes
shown in Table 4.2. The computed velocity profiles for the four fluids
described above are included in Figure 4.17.
7328_C004.fm Page 148 Friday, October 6, 2006 12:25 PM
TABLE 4.2
Computed Consistency Index Values and the Flow Rates through an
Artery Assuming a Constant Pressure Drop for the Various Fluids
Consistency Index Flow Rate Q
Newtonian fluid µ = 3.5cP 140 ml/sec
= 0.0862
Bingham plastic (τ w − τ y ) 136
µb =
τ y = 0.07 γ
(dynes/cm2 ) = 0.0331
Casson’s fluid τw − τ y
117
k c=
τ y = 0.07 γ
(dynes/cm2 ) = 0.1435
6 ×× × × Cassons fluid
××
××
5 ××
××
××
4 ×
××
3 ××
×
×
2 ×
×
×
1 ×
×
×
0 ×
0 0.1 0.2 0.3 0.4 0.5
Radius (cm)
FIGURE 4.17
Velocity profiles computed with the radius and consistency index values shown in Table 4.2
for the various fluids. A constant pressure drop was assumed for each case.
7328_C004.fm Page 149 Friday, October 6, 2006 12:25 PM
TABLE 4.3
The Pressure Gradient and the Wall Shear Stress Computed for the
Various Fluids with the Assumption of a Constant Flow Rate
Bingham
Newtonian Power Law Plastic Casson’s Fluid
Consistency 0.035 0.086 0.033 0.144
Index
6.0 6.37 6.10 6.71
∆p
dynes/cm3
1.29 1.37 1.31 1.44
τw
dynes/cm 2
computed pressure gradients and the corresponding wall shear stresses for
each of the fluids are shown in Table 4.3, when the flow rate through the
vessel is assumed to be the same for each case. The corresponding velocity
profiles are plotted in Figure 4.18.
It can be observed that there is an increase in pressure gradient of about
12% between the Newtonian fluid and the Casson’s fluid with the flow rate
and tube radius maintained constant and the wall shear stress increased by
about 10%. In simulation studies on blood flow in an artery, it is a common
practice to assume that blood is Newtonian since the shear rates are
expected to be larger than 50 sec−1 for most of the cardiac cycle. The above
6 ×× × Cassons
××
5 ××
××
4 ××
××
3 ×
×
2
×
1 ×
×
×
0 ×
0 0.1 0.2 0.3 0.4 0.5
Radius (cm)
FIGURE 4.18
Velocity profiles for the various fluids assuming a constant flow rate and the computed pressure
drops (and wall shear stress) shown in Table 4.3.
7328_C004.fm Page 150 Friday, October 6, 2006 12:25 PM
106
RBC destruction
Platelet destruction
104
102
Exposure time (s)
Surface effects
10–4 dominate
10–6
10–8
100 101 102 103 104 105 106
Shear stress (dynes/cm2)
FIGURE 4.19
Shear stress — exposure time plot for threshold hemolysis of red blood cells and destruction
of platelets.
fluid, thus ignoring the significant fraction of formed elements (red and
white blood cells and platelets) in it. The flow-induced stresses, predomi-
nantly shear stresses, can activate or even destroy the blood cells and platelets.
Many studies have been performed on the effect of high shear stresses on
blood cells. The magnitudes of the shear stresses and the time of exposure
of the cells to stresses are important determinants of the destruction or
activation of the blood cells. A shear stress vs. time of exposure plot for red
blood cells hemolysis threshold and for activation or destruction of platelets
is shown in Figure 4.19. The data included in the plot are a compilation of
experimental data in which the blood was subjected to a range of shear
stresses and exposure times collected by various investigators. As illustrated
in the figure, hemolysis and platelet activation can result even with relatively
lower magnitudes of shear stresses if the formed elements are exposed to
foreign surfaces during flow.
Hemolysis of red blood cells will result in the release of hemoglobin into
the plasma. Hemoglobin within the red blood cells is the principal material
for transport of oxygen from the lungs to the body tissues. Therefore,
hemolysis at an elevated rate will result in anemia so that the blood cannot
7328_C004.fm Page 152 Friday, October 6, 2006 12:25 PM
Composite reinforced by
coilagen fibers arranged
in helical structures
Collagenfibrils
Adve
ntitia
FIGURE 4.20
Schematic of the components of a healthy artery. (From Holzapfel, Gasser, and Ogden (2000)
A new constitutive framework for arterial wall mechanics and a comparative study of material
models, in J. Elasticity 61: 1–48. With permission.)
with a degree of slackness and, thus, their full stiffness is not apparent until
after the vessels are stretched to the extent where the slackness is removed.
Even though the individual components behave like linear elastic materi-
als, the combination of taut elastin and relaxed collagen fibers results in a
nonlinear behavior as illustrated in Figure 4.22. A schematic illustration of
the taut elastin and unstretched collagen fibers is shown in Figure 4.22a. A
schematic of the force-deformation test for a single elastin fiber is shown in
Figure 4.22b whose elastic modulus is about 106 dynes/cm2. A similar sche-
matic for the collagen fiber behavior with an elastic modulus of about 109
dynes/cm2 is shown in Figure 4.22c. For a combination of elastin and col-
lagen fibers, the effect of increasing distension where the collagen fiber
becomes taut and reacts to the load will result in a bilinear curve as shown
7328_C004.fm Page 154 Friday, October 6, 2006 12:25 PM
End.
Ela.
Mus.
Fib.
End.
Ela. > 200 µ
Mus. > 45 µ
Fib.
FIGURE 4.21
The diameter, wall thickness, and components of the wall of the various blood vessels in the
circulatory system. (Redrawn from Burton, A.C. (1971) Physiology and Biophysics of the Circulation.
Year Book Medical Publishers, Chicago. With permission.)
σ
Collagen alone
Combined
Elastin Collagen
(a) Elastin alone
σ ε
Collagen Elastin
Elastin
(d)
ε
Single elastin fiber σ
Recruitment of 3rd
collagen fiber
(b)
Recruitment of 2nd
σ collagen fiber
Collagen
Recruitment of 1st
collagen fiber
ε
ε Collagen Elastin
Single collagen fiber
(c) (e)
FIGURE 4.22
(a) Schematic of the elastin–collagen complex in an unstretched arterial wall; (b) stress–strain
relationship for an elastin; and (c) for a collagen fiber; (d) stress–strain behavior of an elastin
and collagen fiber combination; and (e) behavior with additional collagen fiber recruitment
with increasing load.
FIGURE 4.23
Images of a pig’s thoracic aorta being subjected to a uniaxial test (bottom row), and the
corresponding nonlinear stress–strain relationship.
unloading, the area under the loop will decrease rapidly at first and reach
a steady state after a number of cycles. For this reason, data on the load-
deformation behavior for biological tissue are obtained only after precondi-
tioning the tissue by loading and unloading for several cycles. Figure 4.23
shows a photograph of a strip of a pig aorta gripped between clamps and
attached to a testing apparatus. The corresponding measured stress–strain
behavior for the specimen is also shown in the figure. Such tests will yield
information about the directional dependence of the force-deformation
behavior, if any, of the arterial wall. These in vitro tests should be performed
in a physiological environment with controlled ionic content, moisture, tem-
perature, and other variables. Figure 4.24 represents a schematic comparison
of the stress–strain relationship for soft tissue, such as skin, and blood vessels
with dry bone and steel used in engineering applications.
More useful information on the actual behavior of blood vessels can be
obtained by performing the tests with the vessels in their natural geometry.
Segments of blood vessels in their original configuration can be used in in
vitro or in situ testing under an applied transmural pressure. For in vitro
testing, relatively long segments need to be obtained, so that making the
measurements sufficiently away from the supports can eliminate “end
7328_C004.fm Page 157 Friday, October 6, 2006 12:25 PM
σ Steel
Dry bone
Blood vessel
FIGURE 4.24
A comparison of stress–strain relationship for steel, bone, blood vessel, and skin.
pr ∆r d ∆r
σθ = =E + Eν (4.39)
t R2 dt R2
7328_C004.fm Page 159 Friday, October 6, 2006 12:25 PM
25
20
Einc (dynes/cm2 × 106)
15
10
0
20 40 60 80 100 120 140 160 180 200 220 240
Pressure (mm Hg)
FIGURE 4.25
Incremental static elastic modulus as a function of transmural pressure for dog’s arteries ∆
thoracic aorta, – abdominal aorta, – femoral artery, and O – carotid artery. (Redrawn from
Bergel, D.H. (1961a) J. Physiol. 156: 445–457. With permission.)
7328_C004.fm Page 160 Friday, October 6, 2006 12:25 PM
TABLE 4.4
The Static Elastic Properties of the Arterial Wall
Pressure Thoracic Abdominal Femoral Carotid
(mmHg) Aorta Aorta Artery Artery
40 1.2 + 0.1 (6) 1.6 + 0.4 (4) 1.2 + 0.2 (6) 1.0 + 0.2 (7)
100 4.3 + 0.4 (12) 8.9 + 3.5 (8) 6.9 + 1.0 (9) 6.4 + 1.0 (12)
160 9.9 + 0.5 (6) 12.4 + 2.2 (4) 12.1 + 2.4 (6) 12.2 + 2.7 (7)
220 18.2 + 2.8 (5) 18.0 + 5.5 (3) 20.4 + 4.4 (6) 12.2 + 1.5 (7)
Note: Figures in parentheses refer to the number of specimens in the experiments.
Source: Bergel, D.H. (1961a) J. Physiol., 156: 445–457. With permission.
TABLE 4.5
The Dynamic Elastic Properties of the Arterial Wall
Vessel/Frequency 0 Hz 2 Hz 5 Hz 18 Hz
Thoracic aorta 4.4 + 0.40 (10) 4.7 + 0.42 (10) 4.9 + 0.45 (10) 5.3 + 0.80 (4)
Abdominal aorta 9.2 + 0.94 (7) 10.9 + 0.88 (7) 11.0 + 0.82 (7) 12.2 + 0.46 (4)
Femoral artery 9.0 + 1.15 (5) 12.0 + 0.81 (5) 12.0 + 0.82 (5) 10.6 + 1.39 (5)
Carotid artery 6.9 + 0.48 (6) 11.0 + 1.00 (6) 11.3 + 0.99 (6) 11.5 + 1.03 (6)
Note: Figures in parenthesis show the number of specimens used in the experiments.
Source: Bergel, D.H. (1961a) J. Physiol., 156: 458–469. With permission.
7328_C004.fm Page 161 Friday, October 6, 2006 12:25 PM
Radial cut
FIGURE 4.26
Schematic of a ring-shaped arterial segment and the open sector (horseshoe-shaped), stress-
free state of the segment after a radial cut.
7328_C004.fm Page 163 Friday, October 6, 2006 12:25 PM
FIGURE 4.27
A photo of an excised, ring-shaped, pig’s femoral arterial segment and the open-sector-shaped,
stress-free state of the artery with an additional radial cut.
computed and have been shown to be compressive in the intimal layer and
tensile in the adventitial layer. The computed residual stress with an assumed
elastic modulus demonstrates that the magnitudes of the residual stress are
about 10% of the estimated circumferential stress in the arterial wall under
physiological transmural load. Under normal physiological loading neglecting
the residual stress, it has been shown that circumferential stress is generally
larger in the intimal border and decreases towards the adventitial border.
Thus, the inclusion of a residual stress that is compressive in the intimal
border and tensile in the adventitial border will result in a circumferential
stress distribution across the arterial wall thickness, which is more uniform
under physiological loading. The presence of residual strain and stress has
7328_C004.fm Page 164 Friday, October 6, 2006 12:25 PM
been implicated in the remodeling of the arterial wall. Finite element studies
indicate that the distribution of the circumferential stress throughout the wall
thickness is more uniform if the residual strain effect is included in the
analysis.
4.3 Summary
In this chapter, the rheological behavior of blood was considered. The basic
relationships for viscometry were developed and the application of viscometry
to measure the viscosity of blood and plasma was discussed. The importance
of understanding the rheological behavior of blood and its changes with
diseased states was briefly discussed. A study of the arterial wall material
behavior is important for a basic understanding of the physiology of blood
vessels. The arteries serve as a pressure reservoir in the circulation. As blood
is ejected from the left ventricle during systole, the vessel walls distend as
the arterial pressure rises. Later, during diastole, the passive arterial wall
tension maintains a force to drive the blood through the peripheral capillaries.
Assessment of mechanical properties of the blood vessels in various
segments of the peripheral circulation is important in the understanding of
the propagation of pressure pulses and the pressure-flow relationship in the
circulation under normal and diseased states. It is also important to under-
stand the interaction between blood flowing through the lumen and the
vessel wall. Knowledge of the behavior of the normal artery will also be
helpful in early detection of any diseases of the blood vessels, such as
atherosclerosis. Study of the material characterization of the cardiac muscles
is also important in our understanding of normal physiological behavior and
various alterations, which occur with the onset of diseases.
4.4 Problems
4.1 Derive the relationship for the torque developed in a Couette vis-
cometer (Equation 4.6). A typical Couette viscometer with a 4.0 cm
inner diameter, 4.2 cm outer diameter and 6 cm height is used to
measure the viscosity of blood. If the angular velocity is 50 rpm,
compute the torque required to rotate the outer cylinder. Also plot
the distribution of the rate of shear between the inner and the outer
cylinder. Use a whole blood viscosity of 3.5 cP.
4.2 For a typical red blood cell, the intracellular fluid volume is 87 µm3
and the surface area of the cell is 163 µm3 . If the red blood cell had
a spherical shape with the same intracellular fluid volume, compute
the surface area for the same. What is the advantage of the biconcave
shape for the red blood cell?
7328_C004.fm Page 166 Friday, October 6, 2006 12:25 PM
4.3 Plot the apparent viscosity as a function of shear rate for the rela-
tionship given by Equation 4.8 for a range of shear rates from 1 to
500 sec−1. Compute the apparent viscosity of blood at a rate of shear
of 230 sec−1. Use a plasma viscosity of 1.2 cP.
4.4 For a normal hematocrit of 43% and fibrinogen concentration of 0.3
g/100 ml, compute the yield stress for blood employing the empir-
ical formulae given in Equation 4.10. In a capillary with a diameter
of 5 µm and a length of 1 mm, estimate the minimal pressure gradient
required for the blood to overcome the computed yield stress and
induce flow through the vessel.
4.5 Derive expressions for the flow rate and the apparent viscosity for
the cell-free boundary layer model described in the text.
4.6 For human blood, assuming values of plasma viscosity of 1.2 cP and
whole blood viscosity of 3.3 cP at 37°C, compute the apparent vis-
cosity of blood flowing through a tube of 100 µm diameter using
both the cell free marginal layer and the Sigma effect theories. Assume
a cell free layer thickness of 3 µm and a mean thickness of unsheared
laminae of 15 µm . Compare the values from both theories.
4.7 Assume that whole blood with a hematocrit of 45% flows through
a small diameter tube. The total flow rate is 16 cc/sec, although in
the core region it is 12 cc/sec and in the peripheral region it is 4 cc/sec.
The blood cells accumulate in the core region with a volume of
10 mm3 and there are no blood cells present in the cell-free peripheral
region which has a volume of 6 mm3. By performing a mass balance
on the red blood cells, determine the hematocrit in the core region
and the average hematocrit in the whole tube. What is the effect on
apparent viscosity of the blood in the tube?
4.8 An artery of an animal has an internal diameter of 1 cm and a wall
thickness of 0.75 mm at an end diastolic pressure of 85 mm Hg. An
8 percent increase in the diameter was measured for a pulse pressure
of 45 mm Hg. Compute the circumferential stress in the wall of the
artery as well as the elastic modulus of the vessel at the mean arterial
pressure assuming that the arterial wall is thin and made of linear
isotropic elastic material.
4.9 In an experiment with a canine femoral artery, the following stress-
strain relationship was obtained. Using the data, compute the
incremental elastic modulus at a strain of 0.15.
Strain 0.00 0.03 0.05 0.07 0.09 0.10 0.12 0.15 0.17 0.19 0.21 0.23 0.25 0.27 0.28
Stress (× 105 0 0.3 1 2 5 10 15 31 48 66 83 102 121 119 102
dynes/cm2)
7328_C005.fm Page 167 Monday, September 18, 2006 12:20 PM
5
Static and Steady Flow Models
CONTENTS
5.1 Introduction ................................................................................................167
5.2 Hydrostatics in the Circulation ...............................................................168
5.3 Applications of the Bernoulli Equation .................................................169
5.3.1 Total vs. Hydrostatic Pressure Measurement ...........................169
5.3.2 Arterial Stenoses and Aneurysms ..............................................170
5.3.3 Cardiac Valve Stenoses.................................................................172
5.4 Rigid Tube Flow Models ..........................................................................178
5.4.1 Vascular Resistance .......................................................................180
5.5 Estimation of Entrance Length and Its Effect
on Flow Development in Arteries ..........................................................182
5.6 Flow in Collapsible Vessels......................................................................185
5.7 Summary .....................................................................................................189
5.8 Problems......................................................................................................189
5.1 Introduction
In this chapter, we will discuss some applications of hydrostatics and steady
flow models to describe blood flow in arteries. Even though the flow in the
human circulatory system is unsteady, particularly at the precapillary level,
steady flow models do provide some insight into the aspects of flow through
the arteries and some useful applications can be found using the steady
flow models. As would be expected, steady flow models are also simpler
to use because of the absence of time variations in the governing equations
(cf. the Equations of Motion, Equation 1.43a to Equation1.43c). Steady flow
models also avoid the complexity of having a moving interface between the
blood and the vessel wall as the artery distends in response to pulsatile
pressure.
167
7328_C005.fm Page 168 Monday, September 18, 2006 12:20 PM
∆p = ρgh (1.12)
Using the above equation and knowing the density of blood, the pressures
in the blood vessels can be estimated as shown in the figure. Since arteries
are relatively stiff vessels, the increase in pressure due to hydrostatic effects
will only cause minimal alterations in the blood volume since the vessel
cross section remains nearly constant. However, in the veins, the blood
volume will be greatly affected because they are much thinner and can
expand significantly (see Chapter 3, section 3.4: Systemic Circulation). In
fact, if the venous tone is low and a person suddenly stands up, they may
actually faint because of the increased pooling of blood in the lower extrem-
ities that reduces blood flow back to the heart and, thus, flow to the brain.
Moreover, due to the corresponding decrease of pressure in the head due to
further increases in elevation, the veins in that region may be in a partially
Venous
39
51
(–44 mm Hg)
100 2
95 100 95
Venous pressure (+88 mm Hg)
5 2 5 183
FIGURE 5.1
Hydrostatic pressure differences in the circulation. (Redrawn from Burton, A. C. (1971) Physi-
ology and Biophysics of the Circulation. Year Book Medical Publishers, Chicago. With permission.)
7328_C005.fm Page 169 Monday, September 18, 2006 12:20 PM
V2
p+ρ + ρgz = H (1.52)
2
ρV 2
pe − p1 = pk = (5.1)
2
7328_C005.fm Page 170 Monday, September 18, 2006 12:20 PM
To pressure To pressure
transducer transducer
FIGURE 5.2
Schematic for the measurement of lateral and end pressure.
A1V1 = A2V2
2
a1
P1 a2 P2
V1 V2
FIGURE 5.3
Application of Bernoulli equation to an arterial stenosis.
7328_C005.fm Page 171 Monday, September 18, 2006 12:20 PM
ρV12 ρV 2
p1 + = p2 + 2 (5.2)
2 2
Since p1, V1, and V2 are known, the pressure at the site of constriction, p2,
can be computed. To express the kinetic energy per unit volume (also in
terms of mmHg), the conversion factor 1 g/cm-s2 = 7.5 × 10−4 mmHg can be
used.
Example:
Consider a case where there is a focal stenosis of a 6 mm diameter femoral
artery in which its cross-sectional diameter is reduced to one-third of normal.
Then, the velocity at the stenosis, V2, will be nine times the upstream velocity,
V1 . Furthermore, if the flow rate through the artery were 50 cc/min, then
the velocities, V1 and V2, are:
Therefore, if the pressure at the upstream was 100 mmHg, then the pres-
sure at the stenosis will be reduced by
in the blood, an enzyme, which breaks down the elastin in the artery wall),
it is commonly seen in the distal, or abdominal, aorta where it is referred to
as an Abdominal Aortic Aneurysm (AAA). Analysis performed on the
energy present in the flow similar to that above shows that the flow velocity
is reduced at the cross section of the aneurysm and that some of the kinetic
energy will be converted into pressure at that site. Although the increase in
pressure may not be significant in the resting state (< 5 mmHg), the corre-
sponding conversion from kinetic energy to pressure may be substantial
under exercise conditions when aortic velocity increases several fold. The
increase in pressure will lead to a corresponding increase in wall stress
(see Chapter 2, section 2.2: Analysis of Thin-Walled Cylindrical Tubes), which
may be sufficient to cause further expansion of the aneurysm cross section.
This sequence of events acts as a positive feedback loop in that further
enlargement of the artery reduces velocity and, again, increases static pres-
sure. Ultimately, this process may result in the bursting of the vessel at that
site. Obviously, this is never a desirable outcome, but when it occurs in the
major outlet vessel from the heart, it is particularly critical, being fatal in
~ 50% of cases.
Flow D1 D0 V2 D2
V1
CV
1 2
FIGURE 5.4
Schematic of the flow through a nozzle.
7328_C005.fm Page 173 Monday, September 18, 2006 12:20 PM
ρ 2 ρV 2 V 2
p1 − p2 = (V2 − V12 ) = 2 1 − 12
2 2 V2
2 2
V1 A2
V = A
2 1
ρV22 A2
2
p1 − p2 = 1 −
2 A1
Solving the above equation, we can express the ideal velocity, V2ideal, as
2( p1 − p2 )
V2 ideal =
2
ρ 1 − AA2
1
A2
Cc =
A0
2( p1 − p2 )
V2 ideal =
2
ρ 1 − Cc2 AA0
1
7328_C005.fm Page 174 Monday, September 18, 2006 12:20 PM
Due to frictional losses, the velocity at position 2 will be less than the ideal
value given by the above expression. Hence, a velocity coefficient, Cv , is
introduced such that
V2 actual
CV =
V2 ideal
or
2( p1 − p2 )
Q = A0CcCV
2
ρ 1 − Cc2 AA0
1
2
ρQ 1 − C 2 A0
A1
2 c
p1 − p2 =
2 A0 Cc CV
2 2 2
or
ρQ 2 1
p1 − p2 =
2 A02Cd2
where
CcCV
Cd =
2
1 − CV2 AA0
1
Qm ρ
EOA = A0 = (5.3)
Cd 2 ∆p
7328_C005.fm Page 175 Monday, September 18, 2006 12:20 PM
MSF
AVA = (5.4)
44.5 ∆pm
where AVA is the aortic valve area, MSF is the mean systolic flow rate, and
∆pm is the mean pressure drop across the valve.
The constant, 44.5 (cm/s)/((mmHg)1/2) takes into account the conversion
of units between the mean systolic flow rate (cm3/s) and the mean systolic
pressure drop (mmHg) and it also includes an assumed discharge coefficient
to compute the orifice area in cm2. The corresponding Gorlin equation for
the mitral valve is given by
MDF
MVA = (5.5)
31.0 ∆pm
where the mean diastolic flow rate (cm3/s) and the mean diastolic pressure
gradient (mmHg) are used.
Measurements of the flow rate and the pressure drop across the valves
can be used to predict the effective valve orifice areas of the natural valves
suspected of being stenotic. Such information is useful for cardiac surgeons
in deciding when to replace diseased valves. For normal aortic and mitral
valves, a discharge coefficient close to unity is assumed. In the case of
natural heart valves with centralized flow and no obstructions, the analogy
with the flow through a nozzle may be reasonable. This formula is also
extensively used to predict the effective valve orifice area of prosthetic
valves. However, due to obstruction of the occluders, especially with
mechanical valves, the discharge coefficient can be expected to be signifi-
cantly different from that of natural valves. The common practice is to
perform in vitro experiments in which the actual effective orifice area can
be measured and to determine the discharge coefficient for each type of
prosthetic valve.
The expression for the effective orifice area derived above is based on
steady flow across a nozzle and, thus, is applied assuming constant systolic
flow conditions. In reality, however, even during this period when the valve
is fully open, the blood flow goes through acceleration and deceleration
phases so that the flow is time-dependent. Therefore, a more rigorous
analysis should include the acceleration of the fluid as well as viscous
7328_C005.fm Page 176 Monday, September 18, 2006 12:20 PM
dQ
∆p = A + BQ 2 + CQ (5.6)
dt
where Q is the flow rate, dQ/dt represents the temporal acceleration, BQ2
represents the convective acceleration and CQ represents the viscous
dissipation.
The inertial term (time dependence) can be eliminated by averaging
Equation 5.6 over the forward flow interval (time during which the valve
is open), resulting in
or by taking the measurements at the time of peak flow when dQ/dt will be
identically equal to zero, which results in
In this relationship, the subscript p denotes that the values are measured
at the instance of peak flow through the orifice. Performing a dimensional
analysis on the important parameters for flow through an orifice — pressure p,
flow rate Q, orifice area A, density ρ, and viscosity µ — shows that the
constants B and C can be related to
ρ
B=
A2
and
µ
C= 3
A 2
respectively. Hence, the relationships for the peak and mean pressure drops
can be rewritten as
ρ 2 µ
∆pm = k1 Q +k Q (5.9)
A2 m 2 A 3 2 m
and
ρ 2 µ
∆pp = k1 Q + k 2 3 Qp (5.10)
A2 p A 2
7328_C005.fm Page 177 Monday, September 18, 2006 12:20 PM
ρ
EOA = KQrms (5.11)
∆pm
The difference between the above relationship and the Gorlin equation
derived earlier is the root mean square flow rate used here instead of the
mean flow rate used in the Gorlin equation. The relationship using the peak
flow will be given by
ρ
EOA = KQp (5.12)
∆pp
where Qp is the peak flow rate and ∆pp is the pressure drop across the valve
at the instant of peak flow rate. Use of the above two relationships rather than
the Gorlin equation provides more accurate values for the valve orifice area
from the rigorous fluid mechanical analysis discussed above.
A dimensional analysis of steady flow through an orifice has shown that
d
Cd = f , Re
D
where d is the orifice diameter, D is the diameter of the pipe, and Re is the
Reynolds number = ρVD µ ( )
. For a given valve geometry, the only variable in
the above relationship is the flow rate and, hence, it has been proposed that
kQ
EOA = (5.13)
Cd (Q) ∆p
where EOA represents the effective orifice area and Cd(Q), the discharge
coefficient as a function of flow rate Q. Using an in vitro experimental setup
in which the orifice areas of prosthetic valves of various geometries were
measured using planimetry over a range of flow rates, studies have demon-
strated that when the discharge coefficient was expressed as a linear function
of flow rate, such as
the prediction capability of the effective valve orifice areas was significantly
improved. Further experimental studies are necessary in order to be able to
afford a more accurate clinical prediction of the effective valve orifice areas
in order to improve diagnostic capabilities.
K ∆pD4
Q= (5.15)
L
πR 4 ( p1 − p2 )
Q= (5.16)
8µL
where µ is the coefficient of viscosity and a is the tube radius. From Equation
5.15 and Equation 5.16, it can be observed that K = π 128µ . Thus, the relation-
ship in Equation 5.2 is also referred to as the Hagen–Poiseuille law. We have
already derived the above relationship in Chapter 1 from the Navier–Stokes
equation (Equation 1.72). We also utilized this relationship considering the
forces acting on a volume element of the fluid in Chapter 4 in our discussions
on the principles of the capillary viscometer. Several assumptions were made
in deriving the relationship for the flow rate given in Equation 5.16 and we
should critically examine the validity of these assumptions in models
describing blood flow in arteries. The assumptions are:
the veins and the pulmonary arteries are more elliptical in shape. The
arteries also have a taper with their cross sections narrowing with
distance downstream. Thus, the general assumption of a circular cross
section without taper is a deviation from reality.
6. Rigid wall: As was discussed in the previous chapter (Chapter 3,
section 3.4: Systemic Circulation), the arterial walls are visco-elastic
and distend with the pulse pressure. The interaction between the
flowing blood and the distensible arterial wall is an important factor
in the description of the flow dynamics. Thus, the assumption of
rigid walls in the model is also not valid. However, for the special
case of steady flow models in the circulatory system, distensibility
of the vessels will not affect the solution.
7. Fully developed flow: The model described above also assumes that
the flow is fully developed, which implies that the velocity profile
remains the same at any cross section with distance downstream.
However, as the blood leaves the ventricle through the aortic valve,
the velocity profile in the aorta is relatively flat and a finite length
is needed before the flow may become fully developed (as
described in the next section). Similarly, even in the distal arteries,
flow passes through several branching points and curved arterial
sections. At each location where the cross-sectional geometry devi-
ates from a straight arterial segment, the flow will also be appro-
priately altered. Thus, the assumption of fully developed flow is
also not valid.
∆p
R= (3.3)
Q
7328_C005.fm Page 181 Monday, September 18, 2006 12:20 PM
R1 R2 R3
RT = R1 + R2 + R3
R1
R2
1/RT = 1/R1 + 1/R2 + 1/R3
R3
FIGURE 5.5
Total resistance with tubes in series or parallel configuration.
E
R= (5.17)
I
where I is the current and E the voltage across a segment of a circuit. If the
pressure drop is measured in terms of mmHg and the flow rate in terms of
cc/s, then the resistance is expressed as mmHg-s/cm3 or a peripheral resis-
tance unit (PRU) and it is used in physiological literature. From the Poiseuille
expression for flow rate through a tube, we obtained the relationship
8µL
R= (3.4)
πR 4
When the vessels are in series, the total resistance will be the sum of the
individual resistances. When the vessels are in parallel, the total resistance
across the vessels can be computed using the relationship shown in Figure 5.5.
This formula for the vascular resistance, derived from the steady, fully
developed flow relationship, can be used to estimate the resistance in seg-
ments of the vascular system. For example, the mean pressures at the aortic
root and at the terminal end of the vena cavae can be measured along with
the time averaged flow rate through the systemic circulation. These data
can be used to compute the resistance in the systemic circulation as a mea-
sure of the functioning of the circulatory system. Similarly, the pulmonary
vascular resistance can also be determined by measurement of the corre-
sponding pressures. The expression for the resistance (Equation 3.4) shows
that it is inversely proportional to the fourth power of the radius and, thus,
a small change in the radius of the vessel will considerably affect the resis-
tance to flow. The implication of this is that the autonomic nervous system
in the body controls the tension of the smooth muscles in the vessel wall in
7328_C005.fm Page 182 Monday, September 18, 2006 12:20 PM
the arterioles. Thus, with the alteration of the muscle tension, the arterioles
can be distended or contracted selectively to control the amount of blood
flow into the various segments of the body.
Even though the measurement of vascular resistance yields information
about the state of the circulatory system and, hence, can be used as a
diagnostic parameter, several limitations must be kept in mind on the
information provided: (1) It does not indicate which pathway between the
two points of measurement is constricted or dilated; (2) it does not indicate
the cause for the change (due to nervous stimulation, increased transmural
pressure or other causes); (3) it yields information only on net changes and
does not indicate local changes; and (4) it does not provide information to
distinguish between dilation of the vessels or the opening of new vessels
(Angiogenesis).
U
d
x
Boundary-layer edge
FIGURE 5.6
Concept of entry length before the flow becomes fully developed.
7328_C005.fm Page 183 Monday, September 18, 2006 12:20 PM
U
A2 du
u m
dy 2
δ du A1
m
dy 1 y2
y1
FIGURE 5.7
A force balance on a fluid element in the boundary layer.
d du
µ A( y2 − y1 )
dy dy
µU
A( y2 − y1 )
δ2
7328_C005.fm Page 184 Monday, September 18, 2006 12:20 PM
U2
ρ A( y2 − y1 )
X
U2 U
ρ = kµ 2
X δ
µX
δ∝
ρU
From the above relationship, we can see that the boundary layer grows in
proportion to the square root of the distance. Also, the boundary layer
thickness decreases at any given location as the flow rate through the tube
increases, producing a corresponding increase in the free stream velocity.
The boundary layer will fill the tube and the flow will become fully devel-
oped (i.e., with no further convective acceleration in the fluid) when the
boundary layer thickness equals the radius, or δ = D/2, where D is the
diameter of the tube. Then, the entrance length will be given by
U
X = kD2
ν
or
UD
X = kD (5.18)
ν
where the terms in the parenthesis represent the Reynolds number. In this
relationship, ν = µρ is referred to as the kinematic viscosity. Equation 5.18 can
7328_C005.fm Page 185 Monday, September 18, 2006 12:20 PM
be used to predict the entrance length for steady laminar flow through a
straight pipe of circular cross section. The magnitude for the constant k has
been experimentally determined to be approximately 0.06. Thus, if an artery
was assumed to be a straight cylindrical tube with steady and laminar flow,
we would be able to estimate the distance at which the flow would become
fully developed using the expression given in Equation 5.18. Note that the
relationship in Equation 5.18 is valid only for flows with Reynolds numbers
greater than 50. In cases where the Reynolds number is close to zero (e.g.,
in capillaries where the Reynolds number is ≤ 0.01), the entrance length
becomes a constant of 0.65D.
In arterial flow, flow development is affected by a number of factors. In
large arteries, for example, the entrance length is relatively long since it
depends not only on vessel diameter but also on the Reynolds number
(Equation 5.18), both of which are large. This produces a situation in which
a large portion of most major arteries is exposed to developing flow with
higher velocity gradients near the wall. Secondly, the axial velocity profile
becomes skewed at sites of curvature and bifurcation of the arteries
(see Chapter 6, section 6.6: Flow Through Curved Arteries and Bifurcations)
with higher velocity (and, thus, a higher velocity gradient) toward one wall
and a lower velocity toward the opposite wall. One of the implications of
this is that the intimal lining of arteries is exposed to higher shear forces
proximally and lower shear forces distally. Also, there is a general tendency
for higher shear forces on the outer wall of curvatures and branch inlets and
on the flow divider of bifurcations. In terms of the boundary layer, it will
be thinner in regions of high velocity compared to regions with a low veloc-
ity. The size of the boundary layer is also important in terms of mass trans-
port of molecules (i.e., gas and nutrient) between the blood and artery wall
since diffusive effects are more important than convective effects in large
boundary layer regions. This is the basis for one theory of atherosclerosis
(see Chapter 6, section 6.4: Hemodynamic Theories of Atherosclerosis) in
that certain molecules, such as low-density lipoproteins (LDL), may accu-
mulate within regions of thicker boundary layers and tend to stay in that
region for longer time, enhancing their diffusive transport to the subendo-
thelial region and initiating atherosclerotic lesions.
h P′
(cm) (mm
Hg) Q = 1.0
15 0
(9) (3)
10 4 1 –5
5 8
Q = 1.0
Q = 2.0
0 12 12 9 3 0
(12) (9) (3) (0)
FIGURE 5.8
Pressure and flow through rigid pipe segments. (Redrawn from Milnor, W.R. (1989) Hemo-
dynamics, Williams and Wilkins, Baltimore. With permission.)
system, however, the conduits are flexible and, in some instances, the trans-
mural (i.e., across the vessel wall) pressure can cause significant collapse of
the conduit. This is especially true downstream of a stenosis where the
pressure in the conduit can drop below that of the extramural (i.e., outside
the vessel wall) pressure. In such cases, the flow through the conduit is no
longer dependent upon the upstream, p1, and downstream, p2, pressure
difference, but rather on the difference between the upstream pressure and
the pressure surrounding the conduit, pe. When this happens, flow through
the conduit can remain relatively constant even though the downstream
pressure, p2, varies widely. This phenomenon is variously described as flow
limitation, Starling resistor phenomenon, sluice effect, and waterfall
effect. It has been observed in blood flow situations such as from extratho-
racic to intrathoracic veins, diseased coronary arteries, pulmonary blood
flow, flow through cerebral vessels, and in urine flow.
Let us consider a system of blood vessels as shown in Figure 5.8 where all
the vessels are assumed to be rigid. Pressure at the inlet of the channel is
assumed to be 12 mmHg and the pressure drop across the arteries, capillaries,
and veins is assumed to be 3 mmHg, 6 mmHg, and 3 mmHg, respectively.
These pressure drops are due to viscous dissipation as the fluid flows across
the vessel. A total flow rate of 2 ml/s is assumed to be evenly divided
between the two segments. The vertical distance between the vessels and
the corresponding hydrostatic pressure difference are also shown on the scale
on the left of the figure. The resulting pressures, which include the hydro-
static effect, are given inside the vessels and the pressures due to the effects
7328_C005.fm Page 187 Monday, September 18, 2006 12:20 PM
H P′
(cm) (mm
Hg)
15 0 Q = 0.43
(10.7) (8.1)
–6.7
10 4 2.7
0.1
5 8
Q = 1.43
0 12 12 9 3 Q
(12) (9) (3) (0)
Q = 1.0
FIGURE 5.9
Effect of vessel collapsibility on the pressure and flow. (Redrawn from Milnor, W.R. (1989)
Hemodynamics, Williams and Wilkins, Baltimore. With permission.)
of viscous losses alone are given at the outside. If the vessel in the upper
segment is not rigid, but is made of a thin elastic material, it will collapse
due to the low transmural pressure and, thus, the flow would cease in the
upper segment. However, as soon as the flow ceases, the pressure would
rise to 4 mmHg due to the static conditions and the vessel would once again
open. Thus, this sequence of events will be repeated continuously and has
been demonstrated in in vitro experiments. However, in vivo, thin-walled
vessels will tend to attain an equilibrium state as shown in Figure 5.9 (Milnor,
1989). Assume here that the capillary segments collapse to a narrow lumen
at a transmural pressure of 0.2 mmHg and completely collapse at 0 mmHg.
Pressure and flow through the upper channel will reach a state of equilibrium
at about 0.1 mmHg transmural pressure with a lower flow through the upper
segment, as shown in Figure 5.9.
The phenomenon described above can be used to explain flow through
the lung capillaries. When a person is standing, little flow occurs through
the capillaries in the apices of the lungs due to the collapse of the blood
vessels. As discussed earlier, the flow rate through a rigid tube is dependent
on the distal pressure, which is not necessarily true in a collapsed vessel as
discussed below. Figure 5.10 shows two rigid vessels connected by a collaps-
ible vessel. The collapsible vessel is enclosed in a box so that the pressure
outside the vessel can be independently set. This arrangement is also referred
to as Starling’s resistor. The collapsible segment is assumed to be fully open
at a transmural pressure of 0.2 mmHg and to be completely closed at 0 mmHg.
The pressures at various points in the system (in mmHg) are included in the
7328_C005.fm Page 188 Monday, September 18, 2006 12:20 PM
P′
h (mm
(cm) Hg) (Pa – Pv)(mm Hg)
15 0 20 15 10 5 0
10 4 15
Q 5 8
Flow (ml/min)
4 10
12 10 –6 0 12
a b c
5
–5 16
–8
d –10 20 0
–5 0 5 10
Pv (mm Hg)
FIGURE 5.10
Pressure and flow through a collapsible tube. (Redrawn from Milnor, W. R. (1989) Hemodynam-
ics, Williams and Wilkins, Baltimore. With permission.)
figure. The flow through the vessel will reach equilibrium when the pressure
inside the vessel is 4.1 mmHg representing a transmural pressure of 0.1
mmHg. Changing the distal pressure at point “c” by raising or lowering the
outflow beaker will not alter the flow through the vessel provided it is not
raised above 4.2 mmHg. However, changing the pressure within the box will
affect the flow. The interaction between the variables is represented in the
curve shown in the right half of the figure.
When the outlet pressure Pv is equal to the inlet pressure Pa, there isn’t
any flow through the vessel. Gradually decreasing the outlet pressure results
in a linear increase in the flow through the vessel until the outlet pressure
equals the pressure in the box surrounding the vessel. Further decrease in
the outlet pressure does not increase the flow through the vessel even though
the gradient Pa – Pv continues to increase. The model described above is
similar to that occurring in the lungs. Here, the pressure in the box would
depict the alveolar pressure and, due to the low pressure in the vessels in
the lung circulation combined with the hydrostatic effects, the capillary acts
like a sluice gate controlled by the arterial pressure and the alveolar pressure.
Such a phenomenon where the flow is independent of the downstream
pressure is also described as the “vascular waterfall.” The physics of this
phenomenon, in which the mechanism by which the flow becomes indepen-
dent of the variation in the downstream pressure, is explained by the
“inertial” and “frictional” mode of flow limitation. In the ‘inertial’ mode
explanation, the upstream and downstream pressure is assumed to be decou-
pled at a point in the converging (collapsed) conduit at which the flow has
accelerated to a velocity that exceeds the pressure-wave propagation velocity
of the system at that point. Pressure disturbances distal to the ‘‘choke point’’
7328_C005.fm Page 189 Monday, September 18, 2006 12:20 PM
5.7 Summary
In this chapter, we used simple steady flow models to derive some relation-
ships between the flow and pressure in the circulatory system and discussed
some clinical applications based on the simplified models. We will now
consider the effect of unsteady flow and the distensibility of the blood vessels
as we proceed on in following chapters to more realistic models to describe
blood flow dynamics in the arterial system.
5.8 Problems
5.1 For a typical human, the diameter at the root of the aorta is 2.5 cm, the
time averaged flow rate (Cardiac Output) is 5.5 lpm, and the peak flow rate
during systole is 20 lpm.
5.2 Assume that the blood is flowing through an aorta of 1.0 cm in diameter
at an average velocity of 50 cm/s. Let the mean pressure in the aorta be
100 mmHg. If the blood were to enter a region of stenosis where the diameter
of the aorta is only 0.5 cm, what would be the approximate pressure at the
site of narrowing? (Assume blood density to be 1.056 g/cc: 1 g/cm-s2 = 7.5 ×
10−4 mmHg.)
7328_C005.fm Page 190 Monday, September 18, 2006 12:20 PM
5.3 Reconsider the above problem where the blood enters an aneurysm
region with a diameter of 1.5 cm.
5.4 Estimate the pressure difference between the inlet and outlet of a capillary
that would be needed if blood is flowing through the capillary with a velocity
of 0.2 cm/sec. Assume that (1) the capillary diameter is 8 mm, (2) the capillary
length is 200 mm, and (3) the apparent viscosity of blood is 2.5 cP. The friction
factor for laminar flow in tubes can be expressed as f = 16/Re, where Re is the
Reynolds number and f = ρRV∆2pL .
5.5 Assume that blood entering the aorta from the left ventricle has a steady
state mean velocity of 30 cm/s. If the diameter of the aorta at the entry is
1.2 cm, estimate the distance in the aorta that the blood has to travel before
the flow becomes fully developed. What are the assumptions made in the
formula used in computing this entry length, which is not realistic in the
human circulation?
7328_C006.fm Page 191 Monday, October 16, 2006 10:51 AM
6
Unsteady Flow and Nonuniform
Geometric Models
CONTENTS
6.1 Introduction ................................................................................................191
6.2 Windkessel Models for the Human Circulation...................................192
6.3 Continuum Models for Pulsatile Flow Dynamics ...............................195
6.3.1 Wave Propagation in the Arterial System.................................195
6.3.1.1 Moens–Korteweg Relationship.....................................196
6.3.1.2 Womersley Model for Blood Flow:
Including Viscosity Effects ............................................205
6.3.1.3 Wave Propagation in an Elastic
Tube with Viscous Flow ................................................213
6.4 Hemodynamic Theories of Atherogenesis ............................................222
6.4.1 Low Pressure, Low and High Wall Shear
Stress Theories ...............................................................................223
6.4.2 Time Varying Wall Shear Stress, Oscillatory
Shear Index, and Wall Shear Stress Gradients .........................227
6.5 Wall Shear Stress and Its Effect on Endothelial Cells .........................228
6.6 Flow through Curved Arteries and Bifurcations .................................229
6.6.1 Curved Vessels...............................................................................231
6.6.2 Bifurcations and Branches ...........................................................236
6.7 Flow through Arterial Stenoses and Aneurysms .................................241
6.8 Summary .....................................................................................................254
6.9 Problems......................................................................................................254
6.1 Introduction
Even though the steady flow models considered in the previous chapter
provided us with some insight on the flow through arteries and we consid-
ered some applications using the same models, more realistic ones need to
take into account the unsteady nature of flow through the circulatory system.
191
7328_C006.fm Page 192 Monday, October 16, 2006 10:51 AM
dV
Di = (6.1)
dp
Inflow P Outflow
Q (t) V per. res.
Rs
FIGURE 6.1
Schematic representation of blood flow in the circulation based on the Windkessel theory.
7328_C006.fm Page 193 Monday, October 16, 2006 10:51 AM
Vdp V
k= =
dV Di
dV dV dp
=
dt dp dt
or
dV dp
= Di
dt dt
A mass balance can be written for the fluid in the elastic chamber as
i.e.,
p dp
Q (t ) − = Di
Rs dt
Q(t) = Q0 0 ≤ t ≤ ts
and
Q(t) = 0 ts ≤ t ≤ T
where ts is the time at end systole and T is the duration of the cardiac cycle
and Qo is a constant. Then the equation for systole can be written as
dp p Q
+ = 0
dt Rs Di Di
t
−
p(t) = RsQ0 − (RsQ0 − p0 )e RsDi (6.2)
dp p
=−
dt Rs Di
−( t − T )
p ( t ) = pT e RsDi (6.3)
Vs = Q0ts
I II
FIGURE 6.2
Pressure pulse plot from the Windkessel theory. II represents the time interval in which the
inflow stops and the pressure decays exponentially, given by the relationship in Equation 6.9.
An actual pressure recording is displayed in the insert. (Redrawn from Noordergraaf, A. (1978)
Circulatory Systems Dynamics (1978) Elsevier, Philadelphia. With permission.)
7328_C006.fm Page 195 Monday, October 16, 2006 10:51 AM
ts
∫
Vs = Q ( t )dt
0
The stroke volume determined using this theory could also be used to
estimate the cardiac output. Even though the Windkessel theory is able to
estimate the pressure changes in the arterial system, there are several draw-
backs in applying this theory for understanding the flow dynamics in
circulation. The Windkessel theory assumes that the pressure pulse wave
propagates instantly throughout the arterial system and neglects the finite
time needed for the pulse wave transmission, as we will see later. Several
improved models using the Windkessel theory, incorporating traveling
waves and wave reflections, have also been considered. However, even with
such models, details of the velocity profiles and mechanical forces due to
the blood flowing at various segments in the circulatory system cannot
be studied.
h
r u(r, z, t)
R
z Flow w(r, z, t)
FIGURE 6.3
Schematic diagram of a model for wave propagation in a thin-walled elastic tube.
1 ∂
r ∂r
( rVr ) + ∂∂Vzz = ∂∂Vrr + Vrr + ∂∂Vzz = 0 (6.4)
∂V ∂V ∂V
ρ r + Vr r + vz r
∂t ∂r ∂z
∂V ∂V ∂V
ρ z + vr z + v z z
∂t ∂r ∂z
terms in our analysis. We will write the estimated magnitude of each term
below the corresponding term in the continuity and momentum equations.
In the momentum equations, we will assume that the leading temporal
acceleration term has the predominant effect and compare the relative mag-
nitudes of the nonlinear convective acceleration terms with respect to the
leading term. We will scale the magnitude of the dependent variables in
order to determine the magnitude of each term as follows:
Vr → u, Vz → w , r → R , z → λ , and t→τ
If λ is the wavelength and τ is the period of the pulse wave through the
elastic tube, then the wave speed c will be given by
λ
c=
τ
We will also assume that the ratio of fluid velocity to the pulse wave
velocity is much smaller than unity, i.e., w c 1 .
The magnitudes of the terms in the continuity equation with Vφ neglected
will be:
∂Vr Vr ∂Vz
+ +
∂r r ∂z
u u w
~
R R λ
R
Multiplying each term with , we have
w
u u R
~
w w λ
We observe that the ratio u/w, which is the same order as that of R/λ, can
also be assumed to be much less than unity. Hence, the radial velocity
induced by the tube displacement in the radial direction can be expected to
be much smaller than the axial velocity magnitudes.
Let us consider the magnitude of the inertial terms in the r-direction
momentum equation
∂Vr ∂V ∂V
+ Vr r + Vz r
∂t ∂r ∂z
u u2 wu
~
τ R λ
7328_C006.fm Page 198 Monday, October 16, 2006 10:51 AM
τ
Multiplying each term with , we get
u
τu wτ
1
R λ
u w w
1 λ
w Rτ τ
u R R λ
Since ∼ and ∼ ∼ c , we have
w λ τ τ
Rw w
1
λ c c
∂Vz ∂V ∂V
+ Vr z + Vz z
∂t ∂r ∂z
w uw w2
~
τ R λ
τ
Multiplying each term with , we get
w
u w w
1 λ
w Rτ τ
and, hence,
Rw w
1
λ c c
∂u ∂p
ρ =− (6.5)
∂t ∂r
7328_C006.fm Page 199 Monday, October 16, 2006 10:51 AM
and
∂w ∂p
ρ =− (6.6)
∂t ∂z
If we assume that the pressure gradient in the axial direction ∂∂pz is the
dominant driving force inducing the flow, we can assess the magnitude of
the radial pressure gradient, ∂∂pr , with respect to the axial pressure gradient.
( )
In Equation 6.4, the left-hand side term is ∼ ρ uτ and the right-hand side
term is ∼ Rp . In the axial direction, the left-hand side term of Equation 6.5
is ∼ (ρ wτ ) and the right-hand side term is ∼ λp .
Taking the ratio of magnitudes of these two equations, we get
ρ uτ p
= p
R
ρ wτ λ
Since the axial pressure gradient is the most important term, we assign a
value of unity for that term and get
ρ uτ Rp
=
ρ wτ 1
or
p u R
= ∼
R w λ
∫
Aw ( z, t ) = πR 2 w ( z, t ) = 2 π wrdr
0
R R
∂w 1 ∂
∫0
∂z
2 πrdr + ∫ r ∂r (ur)2πrdr = 0
0
7328_C006.fm Page 200 Monday, October 16, 2006 10:51 AM
sθ sθ
dθ
FIGURE 6.4
An element of the tube wall with the forces acting on the same.
∂ 2
πR w( z, t) + 2 πur 0 = 0
R
∂z
R ∂w
uR ( z, t) = − (6.7)
2 ∂z
Thus, with the simplifying assumptions and the order of magnitude study,
the two reduced governing equations for the fluid motion are Equation 6.5
and Equation 5.6.
Let us now consider the equations of motion for the tube wall. We will
assume that the tube is thin walled and, hence, neglect the bending stresses
and the stress variation in the radial direction. Let η( z, t) be the displacement
of the tube in the radial direction. Figure 6.4 represents a segment of the
tube wall with the forces acting on the element shown. The length of the
segment along the axial direction is assumed to be unity.
The circumferential strain can be computed as
2 π ( R + η) − 2 πR η
εθ = =
2 πR R
7328_C006.fm Page 201 Monday, October 16, 2006 10:51 AM
The hoop stress (stress in the wall along the circumferential direction) will
be given by
η
σ θ = Eεθ = E
R
d2 η
ρt hRdθ = pRdθ − σ θ hdθ
dt 2
In this equation, the term on the left-hand side represents the inertial force
and the terms on the right-hand side represent the forces due to the internal
pressure and the hoop stress in the membrane, respectively and ρt is the
density of the tube wall material. As an initial approximation, if we neglect
the inertial forces using the same reasoning as that for neglecting ∂p/∂r, the
equation of motion for the tube reduces to
hEη
p ( z, t ) = (6.8)
R2
Finally, the boundary condition at the interface equates the fluid velocity
in the radial direction at the wall with the rate of the tube displacement in
the radial direction and is given by the equation
dη
= η = uR
dt
R ∂w
η = −
2 ∂z
R 2 ∂p R ∂w
η = =−
hE ∂t 2 ∂z
∂ w 2 R ∂p
− =
∂z hE ∂t
7328_C006.fm Page 202 Monday, October 16, 2006 10:51 AM
∂ ∂ w 2R ∂2 p
− =
∂t ∂z hE ∂t 2
Since the pressure is a constant over a cross section and, hence, indepen-
dent of r-coordinate, integrating Equation 6.5 over the cross section, we get
∂w 1 ∂p
=−
∂t ρ ∂z
2R ∂2 p ∂ ∂ w ∂ ∂ w ∂ 1 ∂p
= =− =
hE ∂t 2 ∂t ∂z ∂z ∂t ∂z ρ ∂z
and, hence,
∂2 p 1 ∂2 p 1 ∂2 p
= =
∂z 2 ( hE
2 Rρ ) ∂t 2 c02 ∂t 2
The above equation is the wave equation from which we obtain the rela-
tionship for the velocity of wave propagation, Co, as
hE
c02 = (6.9)
2 Rρ
Eh d2 η
p= 2
η + ρt h 2
R dt
7328_C006.fm Page 203 Monday, October 16, 2006 10:51 AM
where the inertial term has been retained in the last term. To obtain the
solution of this second-order differential equation, we will assume that the
pulse pressure is sinusoidal of the form
p = A sin( kz − ωt)
where k = λ1 , λ being the wave length, and ω the circular frequency of oscil-
lation. We will assume that the tube displacement will also follow the same
form and, hence,
η = B sin( kz − ωt)
Eh
A sin( kz − ωt) = 2 − ρt hω 2 B sin( kz − ωt)
R
or
Eh
p = 2 − ρt hω 2 η
R
∂p Eh ∂η a Eh ∂w
= 2 − ρt hω 2 = − 2 − ρt hω 2
∂t R ∂t 2 R ∂z
∂w 2 R ∂p
− =
∂z hE∗ ∂t
where
Eh
E∗ = E 2 − ρt hω 2
R
7328_C006.fm Page 204 Monday, October 16, 2006 10:51 AM
2R ∂2 p ∂ ∂w ∂ ∂w
=− =−
hE ∂t
∗ 2 ∂t ∂z ∂z ∂t
Thus,
∂2 p 1 ∂2 p
= hE∗
∂z 2 2 Rρ ∂t 2
The above relationship is the wave equation and, thus, the wave speed is
given by
hE ω 2ρt R 2
c02 ( ω ) = 1−
2 Rρ E
(6.10)
22
20
18
Experimental data
16
Phase velocity (m/sec)
14
Eqn. 6.13
12
10
0
0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 9.5 10
Frequency (/2p)
FIGURE 6.5
Pulse wave velocity plotted as a function of frequency from the theoretical models compared
with experimental data. (Experimental curve redrawn from Noordergraaf, A. (1978) Circulatory
Systems Dynamics, Elsevier, Philadelphia. With permission.)
7328_C006.fm Page 205 Monday, October 16, 2006 10:51 AM
∂Vr Vr ∂Vz
+ + =0
∂r r ∂z
∂Vz ∂2Vz
= =0
∂z ∂z 2
7328_C006.fm Page 206 Monday, October 16, 2006 10:51 AM
∂p
=0
∂r
dp
= Ae iωt
dz
Vz = w(r )e iωt
A iωt µ d 2 w 1 dw iωt
iωwe iωt = − e + 2 +
r dr
e
ρ ρ dr
d 2 w 1 dw iωρ A
+ − w=
dr 2 r dr µ µ
r′ = r R
7328_C006.fm Page 207 Monday, October 16, 2006 10:51 AM
d 2 w 1 dw A
+ − iα 2 w = R 2 (6.11)
dr ′ 2 r ′ dr ′ µ
where
ωρ
α 2 = R2
µ
or
d ωρ
α= (1.81)
2 µ
ωρ ρRv τ
α 2 = R2 ∝ µv
µ R
The numerator represents the inertial forces and the denominator, the
viscous forces, similar to the Reynolds number defined earlier. Table 6.1 gives
TABLE 6.1
Radius of the Vessel, Heart Rate, and the Womersley Parameter for the Various
Species
Species Weight (kg) Vessel Radius (cm) Heart Rate αa (min-1)
Mouse 0.017 Aorta 0.035 500 1.4
Rat 0.6 Aorta 0.13 350 4.3
Cat 3.0 Aorta 0.21 140 4.4
Rabbit 4.0 Aorta 0.23 280 6.8
Dog 20.0 Aorta 0.78 90 13.1
Man 75.0 Aorta 1.5 70 22.2
Ox 500.0 Aorta 2.0 52 25.6
Elephant 2000.0 Aorta 4.5 38 49.2
the typical values for the radius of the aorta as well as the femoral artery
for various species and the corresponding Womersley parameter values. As
can be observed, even though the radius and heart rate vary over a wide
range among the species, the variation in the Womersley parameter for the
various species is within one order of magnitude.
The differential equation given in Equation 6.10 is of the form
d2w dw
z2 2
+z + ( z 2 − γ 2 )w = 0
dz dz
and is known as the Bessel’s equation of the first kind and γth order. The
solution for the equation is the Bessel function of the first kind and γth order
is J γ . With the appropriate boundary conditions, the solution for the Equation
6.10 is:
J 0 αr′i 2
3
AR 2
w= 1 − 3
iµα 2 J 0 αi 2
and, hence,
J 0 αr′i 2
3
AR 2
Vz = 1 − 3 e iωt (6.12)
iµα 2 J 0 αi 2
1 2 J 1 i 2α
3
πR 2 Ae iωt
0
2
∫
Q = 2 πR wr ′dr ′ = 1−
iωρ i 3 2αJ 0 i 3 2α
(6.13)
The results given above for the axial velocity and the instantaneous flow
rate are complex functions and can be separated into the real and imaginary
parts. They can also be expressed in terms of the modulus and phase, and
the phase angle will represent the phase lag or lead with respect to the applied
harmonic pressure gradient. Womersley (1955a) expressed the complex
expressions in terms of modulus and phase angle by using the relationship
J i 3 2αr′
0
M0′ e iε0 = 1 − 3
J 0 i 2α
7328_C006.fm Page 209 Monday, October 16, 2006 10:51 AM
and
2 J i 3 2α
iε10 = 1 −
1
′
M10 e 3 2 3 2
i J 0 i α
Ae iωt
Vz = ( M0′ e iε0 )
i ωρ
and
πR 2 Ae iωt
Q= ′ e iε10 )
( M10
iωρ
A R 2 M0′
Vz = − sin(ωt − φ + ε 0 )
µ α2
and
A πR 4 M10
′
Q= sin(ωt − φ + ε′10 )
µ α 2
A πR 4 M10
′
Q= cos(ωt − φ + [90° − ε′10 ])
µ α2
Thus, for a given applied sinusoidal pressure gradient, the results shown
above describe the velocity profile over a cross section as well as the instan-
′
′ , M10
taneous flow rate. Figure 6.6 shows a plot of M10 α2
, and ε′10 as a function
′
of a. It can be observed from the figure that as α approaches 0, M10 α2
approaches 1/8, which is the constant for Poiseuille flow relationship in
steady flow (zero frequency) and ε′10 approaches 90. It can be observed that
7328_C006.fm Page 210 Monday, October 16, 2006 10:51 AM
1.0
M′10
0.10
M′10/a2
M′10
0.5
0.05
M′10/a2
0 0
a
90°
⑀′10
60°
30°
FIGURE 6.6
' , M /α2, and ε'
M10 10 10 as a function of α. (Redrawn from Milnor, W.R. (1989) Hemodynamics,
2nd ed., Williams and Wilkins, Baltimore. With permission.)
the phase lag is zero when ε′10 is 90. Thus, with α equal to 0, we have steady
flow with maximum flow rate and zero phase lag between the applied
pressure gradient and flow. With increasing frequency of oscillation, the flow
rate decreases and the phase lag between the pressure gradient and the flow
increases. Figure 6.7 shows a measured pressure gradient from the main
pulmonary artery of a human subject. This pressure gradient pulse, which
repeats itself for every cycle, can be resolved into various harmonics using
the Fourier series expansion. The middle panel shows the first harmonic of
the pressure gradient pulse and the bottom panel shows the instantaneous
flow rate computed. The modulus and amplitude of the first harmonic of
the pressure gradient and the computed flow rate are also shown in
Figure 6.7. It can also be observed that the flow lags behind the pressure
gradient. Since the applied pressure gradient and the computed flow rate
are sinusoidal, averaging the flow rate over a cycle will yield a zero flow rate.
Figure 6.8 shows the synthesis of a measured pressure gradient (panel D)
into four harmonics. Panel A illustrates the first two harmonics and Panels
B, C, and D show the consecutive addition of the other two harmonics to
obtain the resulting pressure gradient. The computation of the corresponding
instantaneous flow rate for each of the harmonics is shown superimposed
with the corresponding pressure gradient in Figure 6.9. Note the phase
7328_C006.fm Page 211 Monday, October 16, 2006 10:51 AM
0 p 2p
(a)
0
mm Hg/cm
A
0
50 (c)
Q
ml/sec
–50
0 0.3 1.0
Time(s)
FIGURE 6.7
Pressure gradient, p/z, measured from the main pulmonary artery of a human; (a) the first
harmonic of the pressure gradient, (b) and the first harmonic of the computed instantaneous
flow rate, and (c) curves. (Redrawn from Milnor, W.R. (1989) Hemodynamics, 2nd ed., Williams
and Wilkins, Baltimore. With permission.)
difference between the pressure gradient and the flow rate in each of the
harmonics.
As pointed out earlier, when the flow rate curves are time averaged over
a cardiac cycle, zero net flow will be the result. However, we know that the
arteries have a net positive flow (cardiac output) and this is incorporated by
the addition of a steady or d.c. component of flow as shown in Figure 6.10.
In this figure, the addition of the four harmonics of the flow is shown and
in the last panel, the x axis is shifted down to include a steady flow compo-
nent. The time average of this flow rate curve in a cycle will give the stroke
volume and multiplied by the heart rate will give the cardiac output.
Figure 6.11a shows a mean velocity curve computed from the theory. The
measured pressure gradient curve used in the computations is included in
Figure 6.11b. A comparison between the computed mean velocity curve and
the measured curve (McDonald, 1974) shows good agreement, in spite of
the assumptions used in deriving the expression for the theoretical flow rate.
Equation 6.11 gives the expression for the velocity profile in the lumen
using the above theory and, once again, the velocity profiles can be computed
for the various harmonics of the pressure gradient and plotted for various
times during the cardiac cycle. Figure 6.12 shows velocity profiles plotted
as a function of the nondimensional radius for the first four harmonics, the
7328_C006.fm Page 212 Monday, October 16, 2006 10:51 AM
6 6
dp/dz (mm Hg/cm)
0 0
–2 –2
Harmonic 1 Harmonic 3
–4 –4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Time (sec) Time (sec)
(a) (b)
6 6
dp/dz (mm Hg/cm)
0 0
–2 –2
Harmonic 4
–4 –4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Time (sec) Time (sec)
(c) (d)
FIGURE 6.8
The first four harmonics of the pressure gradient measured in the dog’s femoral artery. The
pressure gradient data were obtained from McDonald’s Blood Flow in Arteries. (Nichols, W.W.
and O’Rourke, M.F. (1990) McDonald’s Blood Flow in Arteries, 3rd ed., Lea and Febiger, Phila-
delphia.)
angle in the y axis representing the time in the cardiac cycle (ωt). Panel A
represents the fundamental frequency with α = 3.34 and B, C, and D represent
the next three harmonics with values of 4.72, 5.78, and 6.67, respectively.
It can be noted that the profiles are plotted only for one-half of the cardiac
cycle. Since the velocity varies as a sinusoidal function, the profiles for the
other half will be a mirror image of the first half. Several interesting features
can be noted from Figure 6.12. It can be observed that the profiles do not
become completely parabolic at any time during the cardiac cycle. Since the
velocity of the fluid at the wall is zero, the laminae near the wall are slowed
down first due to the viscous forces and the fluid motion near the wall is
slower. Thus, when the pressure gradient reverses, the fluid near the wall
tends to reverse direction easily. Also, some time must elapse before the
viscous diffusion occurs towards the core region. Hence, as the frequency
increases in the subsequent harmonics, the viscous effects are not felt in the
core region and the velocity profile is relatively flat during most of the
7328_C006.fm Page 213 Monday, October 16, 2006 10:51 AM
FIGURE 6.9
The first four harmonics of the pressure gradient are plotted along with the corresponding flow
rate computed using Equation 6.12. The radius of the femoral artery was assumed to be 0.15
cm and the first harmonic was assumed to be 3.34.
cardiac cycle. Figure 6.13 shows the computed velocity profile in the femoral
artery of a dog from the measured pressure gradient and with an addition
of a steady flow component (a parabolic velocity profile with an axial velocity
of 22.65 cm/s). Once again, it can be observed that the velocity profile
reversal occurs first near the wall.
6 6
Flow rate (cm3/sec)
0 0
–2 –2
–4 Harmonic 2 –4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Time (sec) Time (sec)
(a) (b)
6 6
4 1+2+3 4 1+2+3+4
2 Harmonic 4 2
0 0
–2 –2
–4 –4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Time (sec) Time (sec)
(c) (d)
FIGURE 6.10
Shown is the addition of the first four harmonics of the computed flow rate to yield the flow
rate through the femoral artery. The plot on the bottom right illustrates an offset in which a
steady flow component with a mean velocity of 22.65 cm/s has been added. The mean velocity
was computed based on the Poiseuille flow relationship with the steady flow component of
the measured pressure drop.
fluid and S represents the stresses within the tube membrane. Displacements
in the r, θ, z directions are respectively η, ζ, and ξ. ρt is the tube material
density and h is the thickness of the tube.
The equation of motion in the r direction will be:
∂2 η
ρt hRdθdz = σ rr Rdθdz − Sθ hdzdθ
∂t 2
∂2 η h
ρt h = σ rr − Sθ
∂t 2 R
7328_C006.fm Page 215 Monday, October 16, 2006 10:51 AM
100
80
40
20
–20
–40
4
dp/dz (mm Hg/cm)
–2
–4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Time (sec)
(b)
FIGURE 6.11
Plots of the computed flow rate and measured pressure gradient (steady flow component and
the first four harmonics) through the femoral artery of a dog. The agreement between the
computed and the measured (Nichols, W.W. and O’Rourke, M.F. (1990) McDonald’s Blood Flow
in Arteries, 3rd ed., Lea and Febiger, Philadelphia) flow rate profiles is very good.
∂2 ξ ∂S
ρt hRdθdz = − σ rz Rdθdz + z hdzRdθ
∂t 2 ∂z
7328_C006.fm Page 216 Monday, October 16, 2006 10:51 AM
180°
165°
150°
135°
120°
Axial velocity (cm/sec)
105°
90°
75°
60°
45°
30°
100 cm/sec
15°
0°
FIGURE 6.12
Velocity profiles plotted as a function of nondimensional radius in the femoral artery of a dog
computed using the relationship given in Equation 6.11. A, B, C, and D refer to the first four
harmonics with a = 3.34, 4.72, 5.78, and 6.67, respectively.
∂2 ξ ∂S
ρt h = − σ rz + z
∂t 2 ∂z
2 π(R + η) − 2 πR η
εθ = =
2 πR R
and
∂ξ
εz =
∂z
From the generalized Hooke’s law, we can derive the expressions for the
tube stresses as follows:
Eε z = Sz − νSθ
7328_C006.fm Page 217 Monday, October 16, 2006 10:51 AM
360°
165° 345°
150° 330°
135° 315°
120° 300°
Axial velocity (cm/sec)
105° 285°
90° 270°
75° 255°
60° 240°
45° 225°
30° 210°
15° 195° 100 cm/sec
0° 180°
FIGURE 6.13
Computed velocity profiles plotted as a function of the nondimensional radius during a cardiac
cycle in the femoral artery of a dog. A parabolic velocity profile with a mean axial velocity of
22.65 cm/s has been superposed on the first four harmonics in the plots.
σrr, η
Sθ
δSz
Sz + d
δz z
Szθ
σrz, ξ σrθ, ζ
Sθz
Sz
Sθ
h
FIGURE 6.14
Stresses on an element of the tube.
7328_C006.fm Page 218 Monday, October 16, 2006 10:51 AM
and
Eεθ = Sθ − νSz
(εθ + νε z )
Sθ = E
(1 − ν2 )
and
(ε z + νεθ )
Sz = E
(1 − ν2 )
Thus,
η ∂ξ
+ν
R ∂z
Sθ = E
(1 − ν )
2
and
∂ξ η
+ν
∂z R
Sz = E
(1 − ν )
2
or
∂2 ξ ν ∂η
+
∂Sz ∂z2 R ∂z
=E
∂z (1 − ν2 )
1 ∂Vθ Vr ∂V 1 ∂Vz
σ θθ = − p + 2µ + σ θz = µ θ +
r ∂θ r ∂z r ∂θ
∂Vz ∂V ∂V
σ zz = − p + 2µ σ rz = µ z + r
∂z ∂r ∂z
7328_C006.fm Page 219 Monday, October 16, 2006 10:51 AM
∂2 η ∂V hE η ν ∂ξ
ρt h = − p + 2µ r − +
∂t 2 ∂r r=R
(1 − ν2 ) R 2 R ∂z
and
∂2 ξ ∂Vz ∂Vr hE ∂2 ξ ν ∂η
ρt h = − µ + + +
∂t 2
∂r ∂z r = R (1 − ν2 ) ∂z 2 R ∂z
The governing equations of motion for the fluid, along with the above two
equations must be solved simultaneously to obtain the solutions for the
dependent variables (i.e., the velocity components, pressure, and the tube
displacements). The boundary conditions at the interface between the fluid
and the tube are specified as
∂η
= u r=R
∂t
and
∂ξ
= w r=R
∂t
p = Pe i ( kz−ωt )
Eh
1 1
ω 1 − 1 − ν + ν 1 µ
2 2 2
C = = ± (6.14)
k1 2 Rρ
4 R 2ρω
7328_C006.fm Page 220 Monday, October 16, 2006 10:51 AM
and
1 1
2 Rρ 2 µ 2
1 ν2
k2 = ±ω 1− ν +
Eh 2ρω
R 4
1 1
Eh 2 ωρ 2
1
C = ± R
2 Rρ µ (5 − 4 ν)
and
1 1
2 Rρ 2 1 µ 2
k2 = ± (5 − 4 ν)
Eh R ωρ
The plot of the wave velocity for the low viscosity case as a function of
frequency is also shown in Figure 6.5. The theoretical results predict an
increase in phase velocity with increase in frequency in contrast to the exper-
imental results. Even though the theoretical models presented above appear
to yield reasonable results for blood velocity profiles, further improvements
are necessary in order to simulate the propagation of pulse waves in the
arterial wall as briefly described below. In deriving the above relationships
for flow of incompressible viscous fluid through an elastic tube, we made
several simplifying assumptions and following is a discussion on the validity
of those assumptions.
reduces in the distal portions even though the tapering is very gradual.
For a single conduit, the reduction in cross section can be approxi-
mated by an exponential form (Li, 1987). For the aorta, the change
in cross section can be expressed as
kz
−
A( z) = A0e r
r A
k = ln 0
z A
A = A0e − kz
where the magnitudes for k were computed for each of the vessels
from in vivo measurements.
4. Entrance effects: The derivations given above assume a fully devel-
oped flow, which is a condition that does not exist in the circulation.
In order to distribute blood throughout the body, blood vessels very
often branch out and their geometry is very irregular and includes
some curvature. Hence, fully developed flow cannot be expected in
the blood vessels and this assumption is also invalid.
5. Reflected waves: The elastic tube was assumed to be infinitely long
without the presence of any discontinuities, such as curvature or
branching where wave reflections can occur. The effect of wave
reflections has not been taken into account in our formulation.
6. Linearization of equations: Through an order of magnitude study, the
equations were linearized with the assumption that the contribu-
tions from the nonlinear terms are relatively small. Since the equa-
tions were linearized, it was also possible for us to superimpose the
solutions for the higher harmonics to obtain the solution for physi-
ological pulsatile flow. Hence, we have neglected the contributions
of the nonlinear terms and have also ignored the nonlinear interac-
tions between the harmonics. Womersley (1957a) has shown that
neglecting the nonlinear terms in the governing equations for the
fluid will result in an error of about 5% in the computed velocity
magnitudes.
7328_C006.fm Page 222 Monday, October 16, 2006 10:51 AM
distal to the bifurcation; the renal arterial branching in the descending aorta;
and in the ileo-femoral bifurcations of the descending aorta. The common
feature in the location for the development of the lesion is the presence of
curvature, branching, and bifurcation present in these sites. The fluid dynam-
ics at these sites can be anticipated to be vastly different from other segments
of the arteries that are relatively straight and devoid of any branching
segments. Therefore, several investigators have attempted to link the fluid
dynamically induced stresses with the formation of atherosclerotic lesions
in the human circulation. The hemodynamic theories for the genesis of
atheromatous plaques, and the detailed description of the fluid dynamics
at corresponding sites in the human circulation will be considered in the
following sections.
It has been observed that, coincident with the geometric features, there is
a deviation of blood flow from simple antegrade, axial path lines to more
complex patterns (Figure 6.15). What all of these changes have in common,
fluid dynamically, is the presence of low, and potentially reversed, flow
velocities associated with locally altered pressures and shear stresses.
6.4.1 Low Pressure, Low and High Wall Shear Stress Theories
One of the earliest theories on the cause of atherosclerosis development was
the Low Pressure theory (Texon, 1957). This publication was one of the first
attempts to link fluid dynamics with atherogenesis. It was based on the
hypothesis that at sites of arterial curvature, such as the human aorta, as the
blood is forced to turn around the curve, pressure will increase on the outer
wall and correspondingly decrease at the inner wall of curvature. It was
proposed that the pressure at the inner wall would decrease significantly to
become negative and would literally pull the endothelium into the lumen,
initiating the disease process. It was later shown that the radial pressure
gradient is very small and that a significant decrease in pressure will not
occur; hence, the effect proposed in this theory is not probable. Nonetheless,
this publication was important in that it did contribute toward the thought
process on hemodynamic effects on the endothelial cells and initiation of
atheroma.
Some investigators have also suggested a role for elevated shear stress as
an initiating factor in atherosclerosis because it would produce actual phys-
ical damage to the endothelial cells and also impair the normal mass trans-
port process. In 1968, Fry published a significant study on the effect of wall
shear stresses on the endothelial cells, which formed the basis for the High
Shear Stress theory. It consisted of in vivo canine studies in which he intro-
duced a plug in the descending aorta with known dimensions of an orifice
in order to increase the flow rate and wall shear stresses to which the
endothelial cells would be exposed. The opening in the plug was eccentric
so that part of the luminal circumference was exposed to the endothelial
cells. By the analysis of the equations of motion, he justified neglecting the
7328_C006.fm Page 224 Monday, October 16, 2006 10:51 AM
Q0
A
3.3 mm
1.3
Q1 3.2 Intermediate
0.49 branch
Q2
0.05
Left main coronary artery
3.7 Left anterior
B 3.9
descending branch
3.4 mm
Q0
S P
C 3.9 3.7
44 Q1
87 dynes/cm2
0.50
766
766
mm/sec 3.4
Q0 82
23
12
766
20
Left circumflex branch Q1
Aortic sinus 0.50
FIGURE 6.15
Steady flow path lines in models of human left coronary arteries. (Note the relative thickness
of the artery wall at each location.) (From Askura, T. and Karino, T. (1990) Flow patterns and
spatial distributions of atherosclerotic lesions in human coronary arteries, in Circ. Res. 66: 1054.
With permission.)
R ∂p
τ=
2 ∂z
7328_C006.fm Page 225 Monday, October 16, 2006 10:51 AM
1 NO Low arterial
shear stress 1 tPA
Anticoagulant shear stress
(τs > 15 dyne/cm2)
antithrombotic state (τs – ± 0 – 4 dyne/cm2) 1 PGl2
Procoagulant 1 NO
prothrombotic state 1 tPA
Slesdy 1 MCP–1
high flow Ose IVCAM
slow
τs
7328_C006.fm Page 226 Monday, October 16, 2006 10:51 AM
EC quiescence
1 Proliferation
1 Apoptosis Prooxidant state
Paracrine quiescent state 1 COX-1, 2
1 TGF-β 1 PDGF-B 1 Mn SOD
1 NO/eNOS 1 ACE 1 Cu/Zn SOD
1 PGl2/PGl2 Synthase 1 ET–1 Paracrine proliferative state
1 Adrenomedulin 1 ECE 1 ET–1 1 NO/aNOS
1 CNP 1 ACE 1 Adrenomedulin
High EC antioxidant activity
1 COX-1, 2 1 PDGF-B 1 PGl2/PGl2 Synthase
VCAM molecule Fibroblast 1 ECE 1 TGF-β
1 Mn SOD
1 Cu/zn SOD Activated monocyte Platelet aggregate 1 CNP
Smooth muscle cell Fibrin plug
FIGURE 6.16
Effect of shear stress on endothelial cells (EC) and smooth muscle cells (SMC). (From Malek, A.M., Alper, S.L., and Izumo, S., (1999) Hemodynamic shear
stress and its role in atherosclerosis, in JAMA 282: 2035. With permission.)
Biofluid Mechanics: The Human Circulation
7328_C006.fm Page 227 Monday, October 16, 2006 10:51 AM
Studies were extended next to unsteady flow analysis for detailed analysis
of time-dependent velocity profiles and wall shear stress distribution.
With the development of more sophisticated measurement techniques, such
as laser Doppler and particle image velocimetry techniques, more detailed
data analysis became possible. Simultaneously, with computational fluid
dynamic (CFD) algorithm development and the advent of high-speed com-
puters, detailed CFD models were also developed. In addition, advent of
sophisticated imaging modalities, such as MRI, computed tomographic (CT),
and ultrasound imaging, morphologically realistic 3-D geometrical reconstruc-
tion of the segments of interest also became available. Studies on arterial sites
of interest resulted in additional theories as well. As a result of this work, the
high shear stress theory became less widely supported than the low shear
stress theory since the levels of wall stress necessary for cellular erosion to
occur were determined to be rarely seen under normal flow conditions.
∂τ w
TWSSG =
∂t
∫ TWSSG dt
1
TWSSG =
T
0
T
∫ τ w dt
OSI = 0.5 1 − T
0
∫ w
0
τ dt
the OSI is greater than zero as long as there is some directional change during
the flow cycle. The maximum value of OSI is 0.5 in the case of pure oscillatory
flow without any net forward flow.
The spatial wall shear stress gradient (SWSSG) has been evaluated to
determine if local changes in forces acting on the endothelial cells could
induce gaps in the intercellular junction where mass transport would pro-
ceed in an uncontrolled manner. Since the wall shear stress tensor has two
components, one in the mean wall shear stress direction “m” and one normal
“n” to the same, the SWSSG is defined as
2 2
∂τ ∂τ
SWSSG = m + n
∂m ∂n
∫ SWSSG dt
1
SWSSG =
T
0
PGl2 ?
PGG2
MAG + AA
PA
Go PLC
PKC Gq AC Gi GC Gi
rho DAG Grb2 Src ATP GTP
Sos cAMP cGMP K+ Ca++ Cytosol
PIP2 IP3
Ca++
sequestering ras Protein
compartment ?
Ca++ MEKK kinases
raf
p50 p50
B p65 c-fos Cytoskeleton NO
p85
changes
NF-kB
Cytoskeleton Transcription
factors Protein
changes +/– Coding region
DNA AUUUA
IEG Cis-elemen
Paxillin t
protein mRNA
Nucleus
rho
FAK Focal
src adhesions
Cell membrane
Extracellular matrix
FIGURE 6.17
Schematic representation of the multiple “shear–stress responsive” signaling pathways, their
crosstalk, and integration. (From Papadaki, M. and Eskin, S. G., (1997) Effects of fluid shear
stress on gene regulation of vascular cells, in Biotechnol. Prog. 13: 212. With permission.)
carrying blood from the heart to other organs, they impose major changes
on the patterns of flow through them, which, in turn, affect the dynamics
of the circulation and the forces acting on the inner lining of the vessels. As
discussed below, this latter effect is the subject of much research into the
possible causes of arterial disease. (See section 6.4: Hemodynamic Theories
of Atherosclerosis).
7328_C006.fm Page 231 Monday, October 16, 2006 10:51 AM
(a) (b)
(c) (d)
FIGURE 6.18
Phase contrast micrographs of BAEC (bovine aortic endothelial cells) monolayers subjected to
various flow conditions for 24 hours. (From Helminger et al., (1991) Effects of pulsatile flow in
cultured vascular endothelial cell morphology, in J. Biomech. Eng. 113: 128. With permission.)
Facial
External carotid
Carotid bifurcation
Common carotid
Vertebral
Thoracic aorta
Diaphragm
Iliac
Level of inguinal
Sacral ligament
Femoral
Saphenous
Tibial
FIGURE 6.19
Anatomy of major systemic arteries (canine). (From Caro, C. G., Pedley, T. J., Schroter, R. C., and
Seed, W. A. (1978) The Mechanics of the Circulation, Edward Arnold, London. With permission.)
7328_C006.fm Page 233 Monday, October 16, 2006 10:51 AM
Outer
wall
Axial velocity
profile Secondary
motions
(a) (b)
FIGURE 6.20
Schematic of the flow development in a curved vessel with secondary flow.
hence, the axial velocity profile is skewed toward the inner wall of curvature
and no secondary flow develops. With viscous fluid, the two opposing forces
do not balance each other and a secondary flow develops resulting in two
vortices, called “Dean Vortices.” The axial velocity profile becomes skewed
towards the outer wall of curvature, as shown in Figure 6.20 where higher
wall shear stresses are then found.
As the curvature continues, however, those high velocity regions eventu-
ally impact with the outer wall and are deflected in a circumferential direc-
tion. The result is that blood with a high axial velocity now acquires a
circumferential component as well and is convected around the outer surface
of the aorta toward the inner curvature of the vessel. The velocity profile at
this point takes on a reverse image of the outwardly skewed profile seen
more proximally. The axial locations where these transitions occur vary
depending upon the exact geometry of the vessel and the flow rate (i.e., the
cardiac output). Superimposed on this is the fact that flow into the aortic
arch is a developing flow since the velocity profile at the aortic root past the
aortic valve is relatively flat. In the developing flow, the viscous effect is
confined to the boundary layer region initially and the fluid in the core region
acts as inviscid fluid. Hence, the velocity profile is initially skewed toward
the inner wall of the tube and slowly moves toward the outer wall in the
downstream region. Furthermore, the aortic arch is not only curved in
the lateral plane (i.e., left to right), but it also has a degree of curvature in
the anterio–posterior plane (i.e., front to back). Arterial branches from the
arch directed toward the head (e.g., the innominate, carotid, and vertebral
arteries) also affect these flow patterns as they siphon off about 25% of the
cardiac output and the cross section in the descending aorta is reduced by
about 40% from that in the ascending aorta.
The classic approach to analyzing steady flows in this type of geometry
is to apply the Bernoulli equation (see Chapter 1, section 1.6: The Bernoulli
7328_C006.fm Page 234 Monday, October 16, 2006 10:51 AM
Equation) in a form that also allows for energy losses to occur. This
“modified” Bernoulli equation
V2
p+ρ + ρgz = H (1.52)
2
includes the term, H, called the “head loss” (in units of pressure), which
accounts for the fact that energy is dissipated through frictional and nonaxial
flow motion effects. The value of H is obtained by applying an experimental
2
coefficient, KL, to the term ρ V2 , where V is the mean velocity in the tube. For
a 180-degree bend, KL is approximately equal to 0.2.
Further complicating these geometric factors is the high degree of pulsatil-
ity of flow in the aorta ( α ≈ 20 ), which leads to rapid changes in inertia,
limited boundary layer development, and more stable (i.e., less turbulent)
flow. Analysis of purely oscillatory flow in a curved tube has shown that the
viscous effects are confined to the boundary layer region, thus producing four
Dean vortices in a cross section where the vortices in the boundary layer
region are opposite in direction to the vortices in the core region. During
systole, the flow is accelerated forward along the curved tube, and secondary
flow leads to skewing of the axial velocity profile. During early diastole, flow
moving along the inner wall of curvature reverses direction; hence, the wall
shear stress along the inner wall is reversed, whereas slight forward flow
persists along the outer wall of curvature. Thus, the axial velocity profile is
initially skewed toward the outer wall, the maximum axial velocity then
moves toward the inner wall, and subsequently moves toward the outer wall
again. As a result, the inner wall region is subjected to an oscillating shear
stress, which is forward in systole but reverse in diastole (Figure 6.21a/b).
The secondary flow consists of a single vortex in the form of a corkscrew
that persists below the diaphragm in the descending aorta and this secondary
flow affects the flow distribution in the renal arteries as well as in the iliac
bifurcation. The secondary flow also reverses direction during diastole.
Because of the importance of vessel curvature upon flow development, a
nondimensional parameter known as the Dean number, ND, has been
defined as
RV
ND = Re
RC
and can be used for evaluating flow patterns in the aortic arch. This param-
eter is based on the more familiar Reynolds number, Re, which has been
modified by the inclusion of the relative size of the vessel, Rv, and its radius
of curvature, Rc. As can be seen, a high Dean number indicates significant
inertial effects and/or a small radius of curvature resulting in enhanced
secondary flow motions. A low Dean number, on the other hand, would
indicate relatively unimportant secondary flow effects. In the human aorta,
7328_C006.fm Page 235 Monday, October 16, 2006 10:51 AM
50
30
20
10
Inner 0 Outer
wall –1.0 0 +1.0 wall
Nondimensional radius
(a)
40
L/a 3.4
L/a 10.2
30
L/a 16.9
20
Axial velocity (cm/s)
10
0 Outer
0 +1.0 wall
–10
–20
Inner
wall –30
Nondimensional radius
(b)
FIGURE 6.21
Axial velocity profiles in (a) systole and (b) diastole at various cross sections in a curved tube.
(From Chandran, K.B., (1993) Flow dynamics in the human aorta, in J. Biomech. Eng. 115: 611.
With permission.)
7328_C006.fm Page 236 Monday, October 16, 2006 10:51 AM
FIGURE 6.22
Forward and backward directed helices in the aorta. (From Frazin, L. et al., (1990) Functional
chiral asymmetry in descending thoracic aorta, in Circulation 82: 1992. With permission.)
the average Dean number is high (~ 1500) due to a large Re and small Rc.
The increased magnitude of the Reynolds number during systole (~ 8000)
also leads to a higher Dean number and, thus, stronger secondary motions.
The combination of strong secondary motions and rapid changes in flow
rate gives rise to reversals of the helical pattern in the aortic arch as flow
first accelerates (systole) and then decelerates (diastole) (Figure 6.22).
Other vascular sites where curvature effects can be important are in the
coronary bifurcation (whose geometry varies dynamically with the beating
heart) and the aorto-iliac bifurcation, which has a slight curvature out of the
anterio-posterior (AP) plane.
f
so
d iu ure
t
Ra r va
cu
Velocity profile
Parent vessel
Angle of bifurcation
Flow divider
Boundary layer
FIGURE 6.23
Schematic of the aorto-iliac bifurcation geometry and the flow development in the branch
vessels.
At the divider walls of bifurcations, the boundary layers are relatively thin
with maximum axial velocity outside the boundary layer. As the blood enters
the daughter vessel with a finite radius of curvature, the faster moving fluid
is found towards the flow divider due to secondary flow development. If
the corners of the outer wall are sharp, flow separation may be present along
the outer wall of the bifurcation (lateral iliac arteries).
High wall
shear stress
Flow divider
Low wall
shear stress High wall
shear stress
A′
Low wall
shear stress
FIGURE 6.24
Schematic of an arterial branching site.
7328_C006.fm Page 238 Monday, October 16, 2006 10:51 AM
With an asymmetric bifurcation, the flow in the daughter vessels will not
be equal and flow separation may exist in the branch with reduced flow
(Q2 > Q1 ) as shown in Figure 6.25. A similar flow distribution would be
anticipated in a branch vessel, such as a T-junction, roughly corresponding
to renal arterial branches off the descending aorta. With an unequal flow
distribution, flow separation and low wall shear stresses are found along
the outer walls.
(a)
uAvg
(b)
FIGURE 6.25
Axial velocity profiles in the left anterior descending and left circumflex coronary artery during
(a) systole and (b) diastole. (From He, X. and Ku, D.J. (1996) Pulsatile flow in the human left
coronary artery bifurcation: average conditions. Biomech. Eng., 118: 74. With permission.)
7328_C006.fm Page 239 Monday, October 16, 2006 10:51 AM
(a) (b)
FIGURE 6.26
Stream lines in a model of the carotid bifurcation. (From Berger, S.A. and Jou, L.A., (2000) Flows
in stenotic vessels, in Ann. Rev. Fluid Mech., 32: 247. With permission.)
Other important arterial bifurcations occur in the carotid (Figure 6.26) and
coronary (Figure 6.27) circulations where similar flow effects are seen.
The common carotid artery originates at the aortic arch and bifurcates into
the external and internal carotid arteries at the neck level. The external
carotid supplies blood to the facial area while the internal carotid provides
blood to the brain. A unique feature of the internal carotid artery is that it
has a bulge at the bifurcation region, referred to as the carotid sinus. Presence
of the sinus acts to further accentuate the tendency for flow separation along
the outer wall, resulting in a large separation zone, which has been widely
observed in healthy volunteers. Over time, however, these features have
been implicated in the initiation of atherosclerosis due to the resultant low
and reversing shear stresses along the outer vessel walls. (See section 6.4:
Hemodynamic Theories of Atherogenesis.)
Arterial branches are characterized by the fact that they have a large
discrepancy between the cross-sectional areas of the side branch and the
main artery (Figure 6.24). Typical ratios for the total area of the outlet
branches to the area of the inlet artery can range from 0.8 to 1.3. There is
also a wide variation of take-off angles for the branches with many of these
being large and some approaching 90 degress, such as the intercostal arteries
(branches off the thoracic aorta) and the renal arteries (branches off the
abdominal aorta). Due to the large asymmetry at a branch, flow in the branch
vessel is characterized by (1) velocity profiles that are highly skewed toward
the flow divider, (2) secondary flow motions, and (3) a tendency for flow
separation along the outer walls. Distal to the branch point, flow in the main
vessel (e.g., the abdominal aorta distal to the mesenteric artery) becomes
skewed toward the branch, contains secondary flow, and may demonstrate
separated flow along the wall opposite to the branch orifice (Figure 6.28).
As a result of these features, wall shear stresses are generally higher along
7328_C006.fm Page 240 Monday, October 16, 2006 10:51 AM
X1 X2 X3 X4 X5
Flow
FIGURE 6.27
Composite laser Doppler velocity profiles in flow model at five axial locations and at four times
during the flow pulsatile wave. Peak flow rates estimated from slope of velocity profile and
the corresponding wall shear rates are shown on second figure from top. (From Lutz et al.
(1983), Comparison of steady and pulsatile flow in a double branching arterial model, in J. Biomech.
9: 753–766. With permission.)
(a) (b)
FIGURE 6.28
Arteriograms of a left common femoral artery. (From Cho, K.J. and Williams, D.M. (1991) Current
Therapy in Vascular Surgery, 2nd ed., C.B. Ernst and J.C. Stanley, Eds., B.C. Decker, Mosby
Yearbook, Elsevier, Philadelphia. With permission.)
7328_C006.fm Page 241 Monday, October 16, 2006 10:51 AM
the flow divider and lower along both the outer wall of the branch and the
wall opposing the branch. Again, these latter features coincide closely with
the observed sites of atherosclerosis development in patients.
It is possible to estimate the changes in energy distribution through a
branch section if we apply the “modified” Bernoulli equation to this geom-
etry. Here, the appropriate value of KL would depend on the degree of corner
roundness, the branch angle, etc., and would typically have a value of
approximately 1.0.
V1 A1 = V2 A2
or
A
V2 = 1 V1
A2
7328_C006.fm Page 242 Monday, October 16, 2006 10:51 AM
V2
A2
V1
A1
FIGURE 6.29
Schematic of flow through a stenosis.
where v1,2 are the mean velocities and A1,2 are the areas of cross sections 1
and 2, respectively.
Since A1 is always greater than A2 (site of stenosis), V2 will be larger than
V1. This increased velocity is accompanied by formation of a jet through the
stenosis (Figure 6.29).
Due to the Bernoulli effect, the increased velocity present at the throat of
the stenosis also produces a reduced pressure. With increasing degree of
stenosis, the pressure at the throat can decrease significantly and may even
cause the vessel to collapse. Distal to the stenosis, vortices may be present
and the flow may become turbulent with more severe degrees of stenosis.
In these cases of advanced stenosis, autopsy results have shown that the
plaques actually rupture, producing thrombus deposition at this site, which
induces a sudden occlusion of the vessel. Research studies are also being
7328_C006.fm Page 243 Monday, October 16, 2006 10:51 AM
V2
∆p = K L ρ
2
0.02
0.015
Y
0
0.05 0.055 0.06 0.065 0.07
X
(a)
2.56322
0.01 2.02383
1.48444
Inlet velocity = 0.2 m/s 0.945055
Reynold’s number = 483 0.405666
40% area stenosis –0.133724
0.005
0
0.04 0.045 0.05 0.055 0.06 0.065
X
(b)
FIGURE 6.30
Flow simulation in varying degrees of symmetric stenoses displaying the stream lines and shear
stress distribution.
7328_C006.fm Page 245 Monday, October 16, 2006 10:51 AM
0.02
0.015
Y
0.01
Inlet velocity = 0.2 m/s
Reynold’s number = 483
60% area stenosis
0.005 Reattachment point: x = 0.06868 m
0
0.05 0.055 0.06 0.065 0.07
X
(c)
5.95422
0.01 4.65545
3.35669
Inlet velocity = 0.2 m/s 2.05792
Reynold’s number = 483 0.75915
60% area stenosis –0.539618
0.005
0
0.04 0.045 0.05 0.055 0.06 0.065
X
(d)
0.02
0.015
Y
0.01
Inlet velocity = 0.2 m/s
Reynold’s number = 483
80% area stenosis Reattachment point: x = 0.13267 m
0.005 (Beyond this image)
0
0.05 0.055 0.06 0.065 0.07
X
(e)
19.6614
0.01 15.1846
Inlet velocity = 0.2 m/s 10.7077
6.23081
Reynold’s number = 483 1.75394
80% area stenosis –2.72294
0.005
0
0.04 0.045 0.05 0.055 0.06 0.065
X
(f )
FIGURES 6.31
Flow patterns through symmetric stenoses of 20% diameter reduction. (From Rittgers, S.E. and
Shu, M.C.S. (1990) Doppler color-flow images from a stenosed arterial model: Interpretation of
flow patterns, in J. Vasc. Surg. 12: 511–22. With permission.)
With further stenosis (80% case), the jet shortens and becomes turbulent.
In all cases, a separation zone is seen outside the jet in which flow recircu-
lation occurs across a region proportional in size to the degree of stenosis.
It can further be seen that these patterns fluctuate during the flow cycle,
FIGURES 6.32
Flow patterns through symmetric stenoses of 40% diameter reduction. (From Rittgers, S.E. and
Shu, M.C.S. (1990) Doppler color-flow images from a stenosed arterial model: Interpretation of
flow patterns, in J. Vasc. Surg. 12 : 511–22. With permission.)
7328_C006.fm Page 248 Monday, October 16, 2006 10:51 AM
FIGURES 6.33
Flow patterns through symmetric stenoses of 60% diameter reduction. (From Rittgers, S.E. and
Shu, M.C.S. (1990) Doppler color-flow images from a stenosed arterial model: Interpretation of
flow patterns, in J. Vasc. Surg. 12 : 511–22. With permission.)
FIGURES 6.34
Flow patterns through symmetric stenoses of 80% diameter reduction. (From Rittgers, S.E. and
Shu, M.C.S. (1990) Doppler color-flow images from a stenosed arterial model: Interpretation of
flow patterns, in J. Vasc. Surg. 12 : 511–22. With permission.)
7328_C006.fm Page 249 Monday, October 16, 2006 10:51 AM
Wss
2.45699
2.29319
Asymmetric stenosis 1 Asymmetric stenosis 1 2.12939
1.96559
1.80179
0.015 0.015 1.63799
1.47419
1.31039
1.14659
0.982794
0.01 0.01 0.818995
0.655196
0.491397
Y
Y
0.327598
0.163799
0.005 0.005
0 0
0.045 0.05 0.055 0.06 0.065 0.07 0.045 0.05 0.055 0.06 0.065 0.07
Z Z
(a) (b)
Wss
20 2.45699
2.29319
Asymmetric stenosis 2 Asymmetric stenosis 2 2.12939
20 1.96559
1.80179
15 1.63799
1.47419
1.31039
15 1.14659
0.982794
0.818995
10 0.655196
Y
10 0.491397
Y
0.327598
0.163799
5 5
0 0
50 60 50 60 70
Z Z
(c) (d)
FIGURE 6.35
Flow simulation displaying the stream lines and shear stress distribution with varying degrees
of asymmetric stenoses.
with the jet being most pronounced at either peak systole or during the
deceleration phase. Flow separation zones are generally larger and more
stable during the late diastolic phase. The presence of a severe stenosis
(Figure 6.34) may cause a reduced flow rate through the artery and then a
lack of oxygen (ischemia) and nutrients to tissues downstream. Mild to
moderate stenoses, however, often do not produce a significant reduction in
blood flow because of the vascular bed dilatation that occurs downstream,
compensating for the increased resistance upstream.
Stenosis eccentricity produces a further elevation of the wall shear stresses,
especially along the crown, or cap, of the plaque as well as enlarging the
size of the separation zone (Figure 6.35 and Figure 6.36).
It is easy to see from these data how wall shear stresses might cause plaque
cap rupture in which the connective tissue covering the atherosclerotic
7328_C006.fm Page 250 Monday, October 16, 2006 10:51 AM
Vmax
(a)
(b)
(c)
(d)
FIGURE 6.36
Flow simulation in a 95% eccentric arterial stenosis.
by studying the high shear levels in the jet and the long residence times in
the recirculation zones. If the flow is turbulent, it may be severe enough to
cause hemolysis of red blood cells or activation of platelets. This latter effect
is an important consequence since, once activated, platelets may enter the
recirculation zones and reside long enough to attach to the vessel wall and
stimulate other platelets to become activated. If the resultant platelet aggre-
gation proceeds to a critical mass, a thrombus may form and grow to occlude
the entire vessel.
Another important clinical problem involves the abnormal dilation of
major arteries, called aneurysm formation. Abdominal aortic aneurysms
(AAAs) are by far the most common sites, while aneurysms of the descend-
ing thoracic aorta, transverse aortic arch, thoraco-abdominal aorta, and
popliteal artery are also seen, often in conjunction with AAAs (Figure 6.37).
It is thought that an aneurysm forms due to a weakening of the vessel
wall, possibly because of an excess of a degradation enzyme known as
elastase, which acts to break down elastin, a normal load-bearing fiber in
the artery wall. There are several biomechanical implications of this disease.
One is that the vessel will experience increased wall stresses, which, if
allowed to progress, will eventually rupture the vessel. The level of wall
stresses can be approximated by using LaPlace’s Equation for a symmetric,
thin-walled cylinder (Figure 6.3). Summing the forces acting on the upper
half of the vessel and setting them equal to zero gives
pR
σθ = (2.24)
t
where σ is the circumferential wall stress [N/m2], t is the wall thickness [m],
p is the internal pressure [N/m2], and R is the cylinder radius [m]. Thus, the
stress in the wall will increase directly with its diameter, assuming that the
pressure and wall thickness remain constant (in practice, the wall also thins
due to conservation of mass and continued enzymatic activity). This, in
turn, further increases the vessel diameter until it reaches a size and wall
stress beyond which rupture is imminent. From clinical experience, an
abdominal aorta of 5 cm diameter (compared to a normal diameter of ≤ 2 cm)
has a rupture risk of about 5% per year and is, therefore, used as a value for
initiating treatment (usually placement of a bypass graft). It is critical to
diagnose the condition by this time since patients with aneurysms that
rupture have a mortality rate of about 50%.
From a fluid mechanics viewpoint, this large expansion of the vessel pro-
duces unusual flow characteristics that include large vortices near the vessel
walls and a central stream whose dimensions depend on the size and eccen-
tricity of the AAAs. Vortices form in the bulge region of an aneurysm and
due to the Bernoulli effect, the pressure at the bulged cross section increases.
Due to stagnant blood in the bulge, thrombus deposition and growth can
also occur in the bulge region. Figure 6.38 shows examples of flow simulation
7328_C006.fm Page 252 Monday, October 16, 2006 10:51 AM
Descending
thoracic aorta
Transverse arch
of aorta
Ascending
aorta
Abdominal
aorta
Thoracoabdominal
aorta
Descending
aorta
FIGURE 6.37
Patterns of aneurysmal vascular disease. (From Gotto, A.M. et al. (1997) Atherosclerosis, in
Surgical Management of Atherosclerotic Vascular Disease, DeBakey, M.E. and McCollum, M.E., III,
Eds., Pfizer Company, New York. With permission.)
in asymmetric aneurysms including the stream lines, wall shear stress, and
pressure distribution in the bulge region. The increase in transmural pressure
can induce additional forces on the wall and, hence, the bulge can grow.
Conservation of Mass requires that the average velocity through the aneu-
rysm must decrease and since blood has a tendency to clot when moving
7328_C006.fm Page 253 Monday, October 16, 2006 10:51 AM
Wss
1.43277
Asymmetric aneurysm 1 Asymmetric aneurysm 1 1.33382
0.025 0.025 1.23488
1.13593
1.03698
0.938038
0.02 0.02 0.839091
0.740145
0.641198
0.542252
0.015 0.015 0.443305
0.344359
Y
Y
0.245412
0.146466
0.01 0.01 0.047519
–0.0514275
–0.150374
0.005 0.005
0 0
0.216053
0.01 0.01 0.114324
0.012595
–0.089134
0.005 0.005
0 0
FIGURE 6.38
Simulation of stream lines, shear stress, and pressure distribution in an asymmetric abdominal
aortic aneurysm.
slowly or when static, most aneurysms develop thrombus within the dilated
pockets. Some may then form a pseudo flow channel similar in size to that
of the nondiseased (normal) vessel. This can be quite dangerous because the
aorta may appear normal on radiographic arteriography when, in fact, the
vessel wall is highly stretched and weakened. In some cases, thrombus from
an AAA may embolize to distal sites, such as the foot, producing an ischemic
condition known as a “blue toe” syndrome. Surgical treatment of AAA
involves exposure of the aorta, removal of the diseased segments, and place-
ment of a graft (typically Dacron) and this also has an associated risk to the
patient. Since not all AAAs larger than 5 cm rupture and some smaller
aneurysms do, biomechanical studies are ongoing in order to study their
fluid dynamics and effects upon the wall to better predict wall stress distri-
butions in the AAA and the possible potential for rupture.
7328_C006.fm Page 254 Monday, October 16, 2006 10:51 AM
6.8 Summary
In this chapter, some of the unsteady flow models to simulate blood flow in
arteries are presented. Even though the wave propagation through the elastic
tube is included in the models, more emphasis is put on the nature of blood
flow through the arteries. Description of the various hemodynamic theories
correlating flow dynamics in curved arterial segments, as well as in regions
of bifurcations and branches with the genesis and growth of atheroma in
these regions, has been included. The nature of unsteady flow dynamics in
arterial curvature and bifurcation sites, as well as in the region of stenosis
and aneurysm is also discussed.
6.9 Problems
6.1 Show that the Womersley parameter is analogous to the Reynolds num-
ber for unsteady flow.
6.2 For a typical human aorta, the diameter of the lumen is 30 mm and the
thickness of the wall is 4 mm. Assuming a blood density of 1.056 gm/cc and
a viscosity of 0.035 Poise, calculate the Womersley number if the heart rate
is 72 bpm. Diameters for the carotid and femoral arteries of a typical human
are about 0.8 cm and 0.5 cm, respectively. Compute the Womersley numbers
for each of these arterial segments for the same heart rate.
6.3 In an experiment on an exposed abdominal aorta of a dog, the pulse
wave speed is determined to be 1.5 m/s. The wall thickness of the artery
was measured to be 5% of the diameter. Estimate the Young’s modulus for
the aorta. In another segment of the same artery, the pulse wave speed was
double the previous value. What can you conclude from this?
6.4 List five assumptions made in deriving the Moens–Korteweg relation-
ship for wave propagation.
6.5 In deriving the relationship for the velocity profiles for oscillatory
motion in an elastic tube, Morgan and Kiely assumed that the pressure
gradient followed the relationship
the boundary conditions used in obtaining the solution. Here, u and w are
the velocity components, while η and ξ are the tube displacements in the
radial and axial directions, respectively.
6.7 The pressure gradient and flow rate measured in the femoral artery of
a dog are shown in Figure 6.11. Resolve them into their first four harmonics
and then plot them. Also, plot the sum of the four harmonics and compare
this with the measured signals.
7328_C006.fm Page 256 Monday, October 16, 2006 10:51 AM
7328_C007.fm Page 257 Monday, September 18, 2006 2:06 PM
7
Native Heart Valves
CONTENTS
7.1 Introduction ................................................................................................257
7.2 Aortic and Pulmonic Valves ....................................................................259
7.2.1 Mechanical Properties ..................................................................261
7.2.2 Valve Dynamics .............................................................................263
7.3 Mitral and Tricuspid Valves.....................................................................267
7.3.1 Mechanical Properties ..................................................................269
7.3.2 Valve Dynamics .............................................................................270
7.1 Introduction
The heart has four valves whose main function is to control the direction of
blood flow through the heart, permitting forward flow and preventing back
flow. On the right side of the heart, the tricuspid and pulmonic valves
regulate the flow of blood that is returned from the body to the lungs for
oxygenation. The mitral and aortic valves control the flow of oxygenated
blood from the left side of the lungs back to the body. The aortic and
pulmonic valves allow blood to be pumped from the ventricles into arteries
on the left and right side of the heart, respectively. Similarly, the mitral and
tricuspid valves lie between the atria and ventricles of the left and right sides
of the heart, respectively. The aortic and pulmonic valves open during systole
when the ventricles are contracting and close during diastole when the
ventricles are filling with blood, which enters through the open mitral and
tricuspid valves. During isovolumic contraction and relaxation, all four
valves are closed (Figure 7.1a). In Figure 7.1b, the pressure and volume
information for the left heart is shown to illustrate the timings of a normal
cardiac cycle. The details of the cardiac events and the opening of the valves
are included in Chapter 3, section 3.3: Cardiac Valves.
When closed, the pulmonic and tricuspid valves must withstand a pressure
of approximately 30 mmHg. The closing pressures on the left side of the
heart, however, are much higher. Specifically, the aortic valve withstands
257
7328_C007.fm Page 258 Monday, September 18, 2006 2:06 PM
Aorta
Left pulmonary
artery
Right pulmonary
artery
Pulmonary trunk
Superior vena Left atrium
cava Mitral valve
120 Aorta
Pressure (mm Hg)
Left
ventricle
20 Left atrim
100
Left
Volume (ml)
ventricle
60
Left atrim
20
FIGURE 7.1
(a) A schematic of the four chambers of the heart with the valves, (b) typical pressure and flow
curves for the left heart.
7328_C007.fm Page 259 Monday, September 18, 2006 2:06 PM
pressures of approximately 100 mmHg, while the mitral valve resists pres-
sures up to 150 mmHg. Since diseases of the valves on the left side of the
heart are more prevalent than those on the right side, most of this chapter
will focus on the aortic and mitral valves. Where pertinent, reference will be
made to the pulmonic and tricuspid valves, which tend to be impacted more
by congenital heart disease.
Sinuses
of valsalva
RCC
LCC
Aorta
left ventricle
NCC
FIGURE 7.2
Illustration of the aortic sinuses and valve in the closed position. The noncoronary cusp (NCC)
is in front. The left and right coronary cusps (LCC and RCC) are positioned as marked. The
aorta is above the closed valve in this orientation and the left ventricle is below the dashed
line. (From Yoganathan, A.P. et al. (2000) Heart valve dynamics, in The Biomedical Engineering
Handbook, 2nd ed., J.D. Bronzino (Ed.), CRC Press, Boca Raton, FL. With permission.)
7328_C007.fm Page 260 Monday, September 18, 2006 2:06 PM
Cusp free
edge Aorta S
V
Fibrosa
Spongiosa
Ventricularis
FIGURE 7.3
Schematic and histologic section of an aortic valve leaflet showing the fibrosa, spongiosa, and
ventricularis layers.
and is the major fibrous layer within the belly of the leaflet. The layer
covering the ventricular side of the valve is called the ventricularis and is
composed of both collagen and elastin. The ventricularis is thinner than the
fibrosa and presents a very smooth surface to the flow of blood. The central
portion of the valve, called the spongiosa, contains variable loose connective
tissue, proteins, and glycosaminoglycans (GAGs) and is normally not vas-
cularized. The collagen fibers within the fibrosa and ventricularis are not
organized in the unstressed state, but when a stress is applied, they become
oriented primarily in the circumferential direction with a lower concentra-
tion of elastin and collagen in the radial direction.
The fibrous annular ring of the aortic valve separates the aorta from the
left ventricle. Superior to this ring is a structure called the sinus of Valsalva,
or the aortic sinus. The sinus is comprised of three bulges at the root of the
aorta, with each bulge aligned with the belly or central part of the specific
valve leaflet. Each valve cusp and corresponding sinus is named according
to its anatomical location within the aorta. Two of these sinuses give rise
to coronary arteries that branch off the aorta, providing blood flow to the
heart itself. The right coronary artery is based at the right or right anterior
sinus, the left coronary artery exits the left or left posterior sinus, and the
third sinus is called the noncoronary or right posterior sinus. Figure 7.2
shows the configuration of the normal aortic sinuses and valve in the closed
position.
Because the length of the aortic valve cusps is greater than the annular
radius, a small overlap of tissue from each leaflet protrudes and forms a
coaptation surface within the aorta when the valve is closed. This overlapped
tissue, called the lunula, may help to ensure that the valve is sealed. When
the valve is in the open position, the leaflets extend to the upper edge of the
sinuses of Valsalva. The anatomy of the pulmonic valve is similar to that of
the aortic valve, but the surrounding structure is slightly different. The main
7328_C007.fm Page 261 Monday, September 18, 2006 2:06 PM
differences are that the sinuses are smaller in the pulmonary artery and the
pulmonic valve annulus is slightly larger than that of the aortic valve.
The dimensions of the aortic and pulmonic valves and their leaflets have
been measured in a number of ways. Before noninvasive measurement tech-
niques, such as echocardiography, became available, valve measurements
were recorded from autopsy specimens. An examination of 160 pathologic
specimens revealed the aortic valve diameter to be 23.2 ± 3.3 mm, whereas
the diameter of the pulmonic valve was measured at 24.3 ± 3.0 mm. However,
according to M-mode echocardiographic measurements, the aortic root
diameter at end systole was 35 ± 4.2 mm and 33.7 ± 4.4 mm at the end of
diastole. The differences in these measurements reflect the fact that the
autopsy measurements were not performed under physiologic pressure con-
ditions and that intrinsic differences exist in the measurement techniques.
On average, pulmonic leaflets are thinner than aortic leaflets: 0.49 mm vs.
0.67 mm, although the leaflets of the aortic valve show variable dimensions
depending on the respective leaflet. For example, the posterior leaflet tends
to be thicker, have a larger surface area, and weigh more than the right or
left leaflet, and the average width of the right aortic leaflet is greater than
that of the other two.
meshes, while elastin in the fibrosa consists of complex arrays of large tubes
that extend circumferentially across the leaflet. These tubes may surround
the large circumferential collagen bundles in the fibrosa. Mechanical testing
of elastin structures from the fibrosa and ventricularis have separately shown
that the purpose of elastin in the aortic valve leaflet is to maintain a specific
collagen fiber configuration and to return the fibers to that state during cyclic
loading. The valve's viscoelastic properties are actually dominated by the
elastic component (over the range of in vitro testing) so that the viscous
effects, which are largely responsible for energy losses, are small.
In addition to the collagen and elastin, clusters of lipids have been
observed in the central spongiosa of porcine aortic valves. Vesely et al. (1994)
have shown that the lipids tend to be concentrated at the base of the valve
leaflets, while the coaptation regions and free edges of the leaflets tend to
be devoid of these lipids. In addition, the spatial distribution of the lipids
within the spongiosal layer of the aortic leaflets corresponded to areas in
which calcification is commonly observed on bioprosthetic valves, suggesting
that these lipid clusters may be potential nucleation sites for calcification. In
contrast, pulmonic leaflets showed a substantially lower incidence of lipids.
The aortic valve leaflets have also been shown to be slightly stiffer than
pulmonary valve leaflets although the extensibilities and relaxation rates of
the two tissues are similar.
Using fluoroscopy, in which the aortic valve leaflets were surgically tagged
with radio-opaque markers and imaged with high speed x-rays, the leaflets
have been shown to change length during the cardiac cycle. The cusps are
longer during diastole than systole in both the radial and circumferential
direction. The variation in length is greatest in the radial direction, approx-
imately 20%, while the strain in the circumferential direction is about 10%
of the normal systolic length. The difference in strain is due to the presence
of the compliant elastin fibers aligned in this radial direction. The length
change in both directions results in an increased valve surface area during
diastole. During systole, the shortening of the valve leaflets helps to reduce
obstruction of the aorta during the systolic ejection of blood. It should be
noted that this change in area is by no means an active mechanism within
the aortic valve; the valve simply reacts to the stresses it encounters in a
passive manner.
In addition to this change in surface area, the aortic valve leaflets also
undergo bending in the circumferential direction during the cardiac cycle. In
diastole when the valve is closed, each leaflet is convex toward the ventricular
side. During systole when the valve is open, the curvature changes and each
leaflet is concave toward the ventricle. The bending is localized on the valve
cusp near the wall of the aorta. This location is often thicker than the rest of
the leaflet. The total diastolic stress in an aortic valve leaflet has been esti-
mated at 2.5 × 106 dynes/cm2 for a strain of 15%. The stress in the circumfer-
ential direction was found to be the primary load-bearing element in the
aortic valve. Due to the collagen fibers oriented circumferentially, the valve
is relatively stiff in this direction. The strain that does occur circumferentially
7328_C007.fm Page 263 Monday, September 18, 2006 2:06 PM
aortic axis. The aortic annulus moves downward toward the ventricle during
systole and then recoils back toward the aorta as the ventricle fills during
diastole. The annulus also experiences a slight side-to-side translation with
its magnitude approximately one-half the displacement along the aortic axis.
During systole, vortices develop in all three sinuses behind the leaflets of
the aortic valve. Leonardo da Vinci first described the function of these vor-
tices in 1513, and they have been researched extensively in this century
primarily through the use of in vitro models. It has been hypothesized that
the vortices help to close the aortic valve so that blood is prevented from
returning to the ventricle during the closing process. These vortices create a
transverse pressure difference that pushes the leaflets toward the center of
the aorta and each other at the end of systole, thus minimizing any possible
closing volume. However, as shown in vitro by Reul and Talukdar (1981), the
axial pressure difference alone is enough to close the valve. Without the vortices
in the sinuses, the valve still closes, but its closure is not as quick as when
the vortices are present. The adverse axial pressure difference within the aorta
causes the low inertia flow within the developing boundary layer along the
aortic wall to be the first to decelerate and reverse direction. This action forces
the belly of the leaflets away from the aortic wall and toward the closed
position. When this force is coupled with the vortices that push the leaflet
tips toward the closed position, a very efficient and fast closure is obtained.
Closing volumes have been estimated to be less than 5% of the forward flow.
The parameters that describe the normal blood flow through the aortic
valve are the velocity profile, time course of the blood velocity or flow, and
magnitude of the peak velocity. These are determined in part by the pressure
difference between the ventricle and aorta and by the geometry of the aortic
valve complex. As seen in Figure 7.4, the velocity profile at the level of the
aortic valve annulus is relatively flat. However there is usually a slight skew
toward the septal wall (less than 10% of the center-line velocity), which is
caused by the orientation of the aortic valve relative to the long axis of the
left ventricle. This skew in the velocity profile has been shown by many
experimental techniques, including hot film anemometry, Doppler ultra-
sound, and magnetic resonance imaging (MRI). In healthy individuals, blood
flows through the aortic valve at the beginning of systole and then rapidly
accelerates to its peak value of 1.35 ± 0.35 m/s; for children this value is
slightly higher at 1.5 ± 0.3 m/s. At the end of systole, there is a very short
period of reverse flow that can be measured with Doppler ultrasound. This
reverse flow is probably either a small closing volume or the velocity of the
valve leaflets as they move toward their closed position. In the presence of
aortic stenosis, peak velocities through the valve could be as high as 5 m/s
and large magnitudes of turbulent shear stresses that could induce hemolysis.
The flow patterns just downstream of the aortic valve are of particular
interest because of their complexity and relation to arterial disease. Highly
skewed velocity profiles and corresponding helical flow patterns have been
observed in the human aortic arch using magnetic resonance phase velocity
mapping.
7328_C007.fm Page 265 Monday, September 18, 2006 2:06 PM
50 cm/mm
L P L P L
12 m
m
A R A R A
P L P L P L
R A R A R A
P L P L P L
R A R A R A
P P L
P L L
R A R A
R A
FIGURE 7.4
Velocity profiles measured 2 cm downstream of the aortic valve with hot film anemometry in
dogs. The marker on the aortic flow curve shows the timing of the measurements during the
cardiac cycle. (From Paulsen, P. K. and Hansenkam, J. M. (1983). Three-dimensional visualiza-
tion of velocity profiles in the ascending aorta in dogs, measured with a hot-film anemometer,
in J. Biomech. 16: 201–210. With permission.)
Flow through the pulmonic valve behaves similarly to that of the aortic
valve, but the magnitude of the velocity is not as great. Typical peak velocities
for healthy adults are 0.75 ± 0.15 m/s; again these values are slightly higher
for children at 0.9 ± 0.2 m/s. As seen in Figure 7.5, a rotation of the peak
velocity can be observed in the pulmonary artery velocity profile. During
acceleration, the peak velocity is observed inferiorly with the peak rotating
counterclockwise throughout the remainder of the ejection phase. The mean
7328_C007.fm Page 266 Monday, September 18, 2006 2:06 PM
m/s
1.0
0.8
39 ms 70 ms 101 ms 0.6
Trans_neg
0.4
Left 0.2
Inf. 0.0
–0.2
Sup. –0.4
Right
(a) (b) (c)
m/s
132 ms 163 ms 256 ms 1.0
0.8
0.6
Trans_neg
0.4
0.2
0.0
–0.2
–0.4
(d) (e) (f )
m/s
1.0
0.8
349 ms 442 ms 721 ms 0.6
Trans_neg
0.4
0.2
0.0
–0.2
–0.4
FIGURE 7.5
Velocity profiles downstream of the human pulmonary valve obtained with magnetic resonance
phase velocity mapping. The timing of the measurements is shown by the marker on the flow
curve. (From Sloth, E. et al. (1994). Three-dimensional visualization of velocity profiles in the
human main pulmonary artery with magnetic resonance phase-velocity mapping, in Am. Heart
J. 128: 1130–1138. With permission.)
7328_C007.fm Page 267 Monday, September 18, 2006 2:06 PM
Annulus
Anterior leaflet
Posterior leaflet
Chordae
tendineae
Posteromedial
papillary muscle
Anterolateral
papillary muscle
FIGURE 7.6
Schematic of the mitral valve showing the valve leaflets, papillary muscles, and chordae ten-
dinae. (From Yoganathan, A.P. et al. (2000) Heart valve dynamics, in The Biomedical Engineering
Handbook, 2nd ed., J.D. Bronzino (Ed.), CRC Press, Boca Raton, FL. With permission.)
7328_C007.fm Page 268 Monday, September 18, 2006 2:06 PM
Posteromedial Anterolateral
commisure commisure
Medial Lateral
scallop Central scallop
scallop
Chordae Chordae
tendineae tendineae
Posteromedial Anterolateral
papillary muscle papillary muscle
FIGURE 7.7
Diagram of the mitral valve as a continuous piece of tissue. The posterior and anterior leaflets
are indicated, as are the scallops, chordae tendinae, and papillary muscles. (From Yoganathan,
A.P. et al. (2000) Heart valve dynamics, in The Biomedical Engineering Handbook, 2nd ed., J.D.
Bronzino (Ed.), CRC Press, Boca Raton, FL. With permission.)
leaflet is 1.3 cm, while the commisure height is less than 1.0 cm. The pos-
terior leaflet typically has indentations, called scallops, that divide the
leaflet into three regions — the medial, central, and lateral scallops.
The mitral leaflet tissue can be divided into both rough and clear zones.
The rough zone is the thicker part of the leaflet and is defined from the free
edge of the valve to the valve’s line of closure. The term “rough” is used to
denote the texture of the leaflet due to the insertion of the chordae tendinae
in this area. The clear zone is thinner and translucent and extends from the
line of closure to the annulus in the anterior leaflet and to the basal zone in
the posterior leaflet. Unlike the mitral valve, the tricuspid valve has three
leaflets — an anterior leaflet, a posterior leaflet with a variable number of
scallops, and a septal leaflet. The tricuspid valve is larger and structurally
more complicated than the mitral valve and the separation of the valve tissue
into distinct leaflets is less pronounced than with the mitral valve. The
surface of the leaflets is similar to that of the mitral valve; however, the basal
zone is present in all of the leaflets.
Chordae tendinae from both leaflets attach to each of the papillary muscles
(anterior and posterior). The chordae tendinae consist of an inner core of col-
lagen surrounded by loosely meshed elastin and collagen fibers with an outer
layer of endothelial cells. In the mitral complex structure, there are marginal
and basal chordae that insert into the mitral leaflets. From each papillary
muscle, several chordae originate and branch into the marginal and basal
chordae. The thinner marginal chordae insert into the leaflets’ free edge at
multiple insertion points, while the thicker basal chordae insert into the leaflets
at a higher level toward the annulus. The marginal chordae function is to keep
the leaflets stationary while the basal chordae seem to act more as supports.
The left side of the heart has two papillary muscles, called anterolateral
and posteromedial, that attach to the ventricular free wall and tether the
mitral valve in place via the chordae tendinae. This tethering prevents the
mitral valve from prolapsing into the left atrium during ventricular ejection.
On the right side of the heart, the tricuspid valve has three papillary muscles.
The largest one, the anterior papillary muscle, attaches to the valve at the
commissure between the anterior and posterior leaflets. The posterior pap-
illary muscle is located between the posterior and septal leaflets. The smallest
papillary muscle, called the septal muscle, is sometimes not even present.
Improper tethering of the leaflets will result in valve prolapse during ven-
tricular contraction, permitting the valve leaflets to extend into the atrium.
This incomplete apposition of the valve leaflets can cause regurgitation,
which is leaking of the blood being ejected back into the atrium.
and collagen density. Analysis of the leaflets under tension indicated that
the anterior leaflet would be more capable of supporting larger tensile loads
than the posterior leaflet. The differences between the mechanical properties
of the two leaflets may require different material selection for repair or
replacement of the individual leaflets.
Studies have also been done on the strength of the chordae tendinae. The
tension of chordae tendinae in dogs was monitored throughout the cardiac cycle
by Salisbury et al. (1963). They found that the tension only paralleled the left
ventricular pressure tracings during isovolumic contraction, indicating slack-
ness at other times in the cycle. Investigation of the tensile properties of the
chordae tendinae at different strain rates by Lim and Boughner (1975) found
that the chordae had a nonlinear stress–strain relationship. They found that the
size of the chordae had a more significant effect on the development of the
tension than did the strain rate. The smaller chordae with cross-sectional areas
of 0.001 to 0.006 cm2 had a modulus of 2 × 109 dynes/cm2, while larger chordae
of cross-sectional areas of 0.006 to 0.03 cm2 had a modulus of 1 × 109 dynes/cm2.
Ghista and Rao (1972) performed a theoretical study of the stresses sus-
tained by the mitral. This study determined that the stress level could reach
as high as 2.2 × 106 dynes/cm2 just prior to the opening of the aortic valve,
with the left ventricular pressure rising to 150 mmHg. A mathematical model
has also been created for the mechanics of the mitral valve. It incorporates
the relationship between chordae tendinae tension, left ventricular pressure,
and mitral valve geometry. This study examined the force balance on a closed
valve, and determined that the chordae tendinae force was always more than
half the force exerted on the mitral valve orifice by the transmitral pressure
gradient. During the past 10 years, computational models have been devel-
oped of mitral valve mechanics, with the most advanced modeling being
three-dimensional finite element models (FEM) of the complete mitral appa-
ratus. Kunzelman et al. (1993) has developed a model of the mitral complex
that includes the mitral leaflets, chordae tendinae, contracting annulus, and
contracting papillary muscles. From these studies, the maximum principal
stresses found at peak loading (120 mmHg) were 5.7 × 106 dynes/cm2 in the
annular region, while the stresses in the anterior leaflet ranged from 2 x ×106
to 4 × 106 dynes/cm2. This model has also been used to evaluate mitral valve
disease, repair in chordal rupture, and valvular annuloplasty.
Aortic valve L
A Posterior mitral
leaflet
P R
Anterior
mitral leaflet
Base of the heart
Velocity (cm/s)
Velocity (cm/s)
20.0
Vel Vel
0.0 oc ity oc ity
(cm (cm
/s) /s)
0.0 500.0
Time (ms) 60 60
40 40
20 20
0 0
–20 –20
670 ms –40 Mean –40
A L A L
(e) P R (f ) P R
FIGURE 7.8
Two-dimensional transmitral velocity profiles recorded at the level of the mitral annulus in a
pig — (a) systole, (b) peak E-wave, (c) deceleration phase of early diastole, (d) mid-diastolic
period (diastasis), (e) peak A-wave, and (f) time-averaged diastolic cross-sectional mitral
velocity profile. (From Kim et al. (1994). Two-dimensional mitral flow velocity profiles in pig
models using epicardial doppler echocardiography, in JACC 24: 532–545. With permission.)
7328_C007.fm Page 272 Monday, September 18, 2006 2:06 PM
The initial filling is enhanced by the active relaxation of the ventricle, main-
taining a positive transmitral pressure. The mitral velocity flow curve shows
a peak in the flow curve, called the E-wave, which occurs during the early
filling phase. Normal peak E-wave velocities in healthy individuals range from
50 to 80 cm/s. Following active ventricular relaxation, the fluid begins to
decelerate and the mitral valve undergoes partial closure. Then, the atrium
contracts and the blood accelerates through the valve again to a secondary
peak, termed the A-wave. The atrium contraction plays an important role in
additional filling of the ventricle during late diastole. In healthy individuals,
velocities during the A-wave are typically lower than those of the E-wave,
with a normal E-wave/A-wave velocity ratio ranging from 1.5 to 1.7. Thus,
normal diastolic filling of the left ventricle shows two distinct peaks in the
flow curve with no flow leaking back through the valve during systole.
The tricuspid flow profile is similar to that of the mitral valve although the
velocities in the tricuspid valve are lower because it has a larger valve orifice.
In addition, the timing of the valve opening is slightly different. Since the peak
pressure in the right ventricle is less than that of the left ventricle, the time
for right ventricular pressure to fall below the right atrial pressure is less than
the corresponding time period for the left side of the heart. This leads to a
shorter right ventricular isovolumic relaxation and, thus, an earlier tricuspid
opening. Tricuspid closure occurs after the mitral valve closes, since the
activation of the left ventricle precedes that of the right ventricle.
A primary focus in explaining the fluid mechanics of mitral valve function
has been understanding the closing mechanism of the valve. Bellhouse and
Bellhouse (1968) first suggested that the vortices generated by ventricular filling
were important for the partial closure of the mitral valve following early diastole.
Their in vitro experiments suggested that without the strong outflow tract vor-
tices, the valve would remain open at the onset of ventricular contraction, thus
resulting in a significant amount of mitral regurgitation before complete closure.
Later in vitro experiments by Reul et al. (1981) in a left ventricle model made
from silicone suggested that an adverse pressure differential in mid-diastole
could explain both the flow deceleration and the partial valve closure, even in
the absence of a ventricular vortex. Thus, the studies by Reul suggest that the
vortices may provide additional closing effects at the initial stage; however, the
pressure forces are the dominant effect in valve closure.
A more unified theory of valve closure put forth by Yellin et al. (1981)
includes the importance of chordal tension, flow deceleration and ventricular
vortices, with chordal tension being a necessary condition for the other two.
The animal studies by Yellin indicated that competent valve closure could
occur even in the absence of vortices and flow deceleration. Recent studies
using magnetic resonance imaging to visualize the three-dimensional flow
field in the left ventricle showed that in normal individuals a large anterior
vortex is present at initial partial closure of the valve, as well as following
atrial contraction. Studies conducted in our laboratory using magnetic
resonance imaging of healthy individuals clearly show vortices in the left
ventricle, which may be an indication of normal diastolic function.
7328_C007.fm Page 273 Monday, September 18, 2006 2:06 PM
Another area of interest has been the motion of the mitral valve complex.
The heart itself moves throughout the cardiac cycle and, similarly, the mitral
apparatus also moves and changes shape. Recent studies have been conducted
that examined the three-dimensional dynamics of the mitral annulus during
the cardiac cycle. These studies have shown that, during systole, the annular
circumference decreases from the diastolic value due to the contraction of the
ventricle and this reduction in area ranges from 10 to 25%. This result agrees
with an animal study by Tsakiris et al. (1971) that looked at the difference in
the size, shape, and position of the mitral annulus at different stages in the
cardiac cycle. They noted an eccentric narrowing of the annulus during both
atrial and ventricular contractions that reduced the mitral valve area by 10 to
36% from its peak diastolic area. This reduction in the annular area during
systole is significant, resulting in a smaller orifice area for the larger leaflet
area to cover. The annulus not only changes size, but it also translates during
the cardiac cycle. The movement of the annulus toward the atrium has been
suggested to play a role in ventricular filling, possibly increasing the efficiency
of blood flow into the ventricle. During ventricular contraction, there is a
shortening of the left ventricular chamber along its longitudinal axis, and also
the mitral and tricuspid annuli move toward the apex.
The movement of the papillary muscles is also important in maintaining
proper mitral valve function. The papillary muscles play an important role
in keeping the mitral valve in position during ventricular contraction.
Abnormal strain on the papillary muscles could cause the chordae to rupture,
resulting in mitral regurgitation. It is necessary for the papillary muscles to
contract and shorten during systole to prevent mitral prolapse; therefore, the
distance between the apex of the heart to the mitral apparatus is important.
The distance from the papillary muscle tips to the annulus was measured
in normal individuals during systole and was found to remain constant. In
patients with mitral valve prolapse, however, this distance decreased, cor-
responding to a superior displacement of the papillary muscle toward the
annulus.
The normal function of the mitral valve requires a balanced interplay
between all of the components of the mitral apparatus as well as the inter-
action of the atrium and ventricle. The in vitro engineering studies of mitral
valve function have provided some insight into its mechanical properties
and dynamics. Further fundamental and detailed studies are needed to aid
surgeons in the repair of diseased mitral valves and in understanding the
changes in function due to mitral valve pathologies. In addition, these stud-
ies are crucial for improving the design of prosthetic valves that more closely
replicate native valve function. (See Chapter 8, Prosthetic Heart Valve Fluid
Dynamics.)
7328_C007.fm Page 274 Monday, September 18, 2006 2:06 PM
7328_S003.fm Page 275 Friday, July 28, 2006 4:36 PM
Part III
Cardiovascular Implants
and Biomechanical
Measurements
7328_S003.fm Page 276 Friday, July 28, 2006 4:36 PM
7328_C008.fm Page 277 Tuesday, October 10, 2006 11:04 AM
8
Prosthetic Heart Valve Fluid Dynamics
CONTENTS
8.1 Introduction ................................................................................................278
8.2 A Brief History of Heart Valve Prostheses ............................................279
8.2.1 Mechanical Valves.........................................................................279
8.2.2 Tissue Valves ..................................................................................284
8.2.3 Summary of Mechanical Versus Tissue Valves ........................288
8.2.4 Current Types of Prostheses........................................................288
8.3 Hemodynamic Assessment of Prosthetic Heart Valves ......................289
8.3.1 Pressure Drop ................................................................................290
8.3.2 Effective Orifice Area (EOA) .......................................................292
8.3.3 Regurgitation..................................................................................294
8.3.3.1 Turbulent Jet Theory ......................................................295
8.3.3.2 Proximal Flow Convergence.........................................295
8.3.4 Flow Patterns and Shear Stresses ...............................................296
8.3.4.1 Caged Ball Valve .............................................................298
8.3.4.2 Tilting Disc Valve............................................................299
8.3.4.3 Bileaflet Valve ..................................................................299
8.3.4.4 Porcine Valve ...................................................................302
8.3.4.5 Pericardial Valve .............................................................303
8.3.5 Leakage Flow through Heart Valve Prostheses .......................304
8.3.6 Cavitation and HITS.....................................................................306
8.4 In Vitro Studies of Coagulation Potential and Blood Damage ..........308
8.4.1 Implications for Thrombus Deposition .....................................308
8.5 Durability of Prosthetic Heart Valves ....................................................309
8.5.1 Wear.................................................................................................310
8.5.2 Fatigue.............................................................................................310
8.5.3 Mineralization................................................................................ 311
8.6 Current Trends in Valve Design..............................................................312
8.7 Conclusions.................................................................................................312
8.8 Problem .......................................................................................................313
277
7328_C008.fm Page 278 Tuesday, October 10, 2006 11:04 AM
8.1 Introduction
Heart valve disease is one of the main afflictions of the cardiovascular system
and can be caused by rheumatic fever, ischemic heart disease, bacterial or
fungal infection, connective tissue disorders, trauma, and malignant carci-
noid. In advanced form, it leads to various disabilities and, ultimately, to
death. The valves that are most commonly affected are the mitral, aortic, and
tricuspid valves. Malfunction of the valves affects their hemodynamic (i.e.,
fluid dynamic) performance in two primary ways:
Both of these conditions reduce the efficiency of the heart and place addi-
tional stresses and strains upon it.
The decision to perform corrective surgery on the natural valve or to
replace it with a prosthetic valve is often made on the basis of an evaluation
of the functional impairment of the natural valve. A classification for such
an evaluation as proposed by the New York Heart Association is utilized
and surgery is usually limited to patients belonging to Classes III and IV.
The first clinical use of a cardiac valvular prosthesis took place in 1952
when Dr. Charles Hufnagel implanted an artificial caged-ball valve to treat
aortic insufficiency. The Plexiglas cage contained a ball occluder, and was
inserted into the descending aorta without the need for cardiopulmonary
bypass. It did not cure the underlying disease, but it did relieve regurgitation
from the lower two-thirds of the body.
The first implant of a replacement valve in the actual anatomic position
took place in 1960 along with the advent of cardiopulmonary bypass. Since
that time, the achievements in valve design and the success of artificial heart
valves as replacements have been remarkable with more than 50 different
cardiac valves having been introduced over the past 35 years. Unfortunately,
after many years of experience and success, not all problems associated
with heart valve prostheses have been eliminated. The most serious current
problems and complications are:
(a)
(b)
(c)
FIGURE 8.1
Photographs of various mechanical heart valves: (a) Starr–Edwards caged-ball valve, (b) caged-
disc heart valve, (c) Björk–Shiley tilting disc valve, (d) Medtronic Hall tilting disc valve, (e) St.
Jude Medical bileaflet valve, and (f) CarboMedics bileaflet valve. (From Yoganathan, A.P. (2000)
Cardiac valve prostheses, in The Biomedical Engineering Handbook, Bronzino, J.D. (Ed.), CRC
Press, Boca Raton, FL. With permission.)
7328_C008.fm Page 281 Tuesday, October 10, 2006 11:04 AM
(d)
(e)
(f)
Due to the high profile design characteristics of the ball valves, low
profile caged-disc valves were developed in the mid-1960s, especially for
the mitral position. Examples of the caged-disc designs are the Kay–Shiley
and Beall prostheses, which were introduced in 1965 and 1967, respectively
(Figure 8.1b). These valves were used exclusively in the atrioventricular
position, but due to their inferior hemodynamic characteristics, caged-disc
valves are rarely used today.
It is interesting to note that even after 35 years of valve development; the
ball-and-cage format remains the valve of choice for some surgeons. How-
ever, it is no longer the most popular mechanical valve, having been super-
seded, to a large extent, by tilting-disc and bileaflet valve designs. These
designs overcome two major drawbacks of the ball valve — namely, high
profile configurations and excessive occluder-induced turbulence in the flow
through and distal to the valve.
The most significant developments in mechanical valve design
occurred in 1969 and 1970 with the introduction of the Björk–Shiley and
Lillehei–Kaster tilting-disc valves (Figure 8.1c). Both prostheses involve
the concept of a ‘‘free’’ floating disc, which, in the open position, tilts to
an angle that depends on the design of the disc-retaining struts. In the
original Björk–Shiley valve, the angle of the tilt was 60 degrees for the aortic
and 50 degrees for the mitral model. The Lillehei–Kaster valve has a
greater angle of tilt of 80 degrees, but, in the closed position, is preinclined
to the valve orifice plane by an angle of 18 degrees. In both cases the
closed valve configuration permits the occluder to fit into the circumfer-
ence of the inflow ring with virtually no overlap, thus reducing mechan-
ical damage to erythrocytes. A small amount of regurgitation backflow
induces a ‘‘washing out’’ effect of debris and platelets and theoretically
reduces the incidence of thromboemboli.
The main advantage of the tilting-disc valve is that in the open position
it acts like an aerofoil in the blood flow where induced flow disturbances
are substantially less than those obtained with a ball occluder. Although the
original Björk–Shiley valve employed a Delrin occluder, all present-day
tilting-disc valves use pyrolytic carbon for these components. It should also
be noted that the ‘‘free’’ floating disc can rotate during normal function, thus
preventing excessive contact wear from the retaining components on one
particular part of the disc. Various improvements to this form of mechanical
valve design have been developed, but have tended to concentrate on alter-
ations either to the disc geometry, as in the Björk-–Shiley convexo-concave
design, or to the disc retaining system, as with the Medtronic Hall and
Omniscience valve designs (Figure 8.1d).
The Medtronic Hall prosthesis was introduced in 1977 and is characterized
by a central, disc-control strut, with a mitral opening angle of 70 degrees
and an aortic opening angle of 75 degrees. An aperture in the flat, pyrolytic,
carbon-coated disc affixes it to the central guide strut. This strut not only
retains the disc, but also controls its opening angle and allows it to move
downstream 1.5 to 2.0 mm; this movement is termed disc ‘‘translation’’ and
7328_C008.fm Page 283 Tuesday, October 10, 2006 11:04 AM
improves flow velocity between the orifice ring and the rim of the disc. The
ring and strut combination is machined from a single piece of titanium for
durability. All projections into the orifice (pivot points, guide struts, and disc
stop) are open-ended, streamlined, and are placed in the region of highest
velocity to prevent the retention of thrombi by valve components. The sew-
ing ring is of knitted Teflon®. The housing is rotatable within the sewing
ring for optimal orientation of the valve within the tissue annulus.
Most recent mechanical valve designs are based on the bileaflet, all-
pyrolytic carbon valve designed by St. Jude Medical, Inc. and introduced in
1978 (Figure 8.1e). This design incorporates two semicircular, hinged, pyro-
lytic carbon occluders (leaflets), which, in the open position, are intended to
provide minimal disturbance to flow. The leaflets pivot within grooves made
in the valve orifice housing. In the fully open position, the flat leaflets are
designed to open to an angle of 85 degrees.
The CarboMedics bileaflet prosthesis gained FDA approval in 1993
(Figure 8.1f). It is also made of Pyrolite® which is known for its durability
and thrombo-resistance. The valve has a recessed pivot design and is rotat-
able within the sewing ring. The two leaflets are semicircular, radiopaque,
and open to an angle of 78 degrees. A titanium-stiffening ring is used to
lessen the risk of leaflet dislodgment or impingement. The CarboMedics Top
Hat — an aortic version of this prosthesis — was designed for implantation
in the supraannular position. This allows for an increase in the size of the
implanted valve, thus improving forward flow hemodynamics of the valve.
Improving bulk flow hemodynamics of the bileaflet valve type has been
a recent design goal for St. Jude Medical as well. The St. Jude Medical HP
valve, introduced in 1993, features a decrease in thickness from the sewing
cuff design of the original St. Jude Medical valve. This modification allows
the placement of a proportionally larger housing within the cuff for a given
tissue annulus diameter, thus increasing the orifice area available for flow.
The St. Jude Medical Regent valve offers a further improvement in bulk
flow hemodynamics over the original St. Jude Medical valve. This improve-
ment is created by structurally reinforcing the housing of the St. Jude Medical
HP, allowing a decrease in its thickness, which further increases the internal
orifice area available for flow.
The Sorin Bicarbon valve, marketed by Sorin Biomedica of Italy, is popular
outside the U.S. The curvature of the leaflets of the Bicarbon valve reduces
pressure loss caused by the small central orifice of the original bileaflet valve
design. The pivots of this valve are made up of two spherical surfaces of
different radii of curvature. This design may reduce wear because the leaflet
projection rolls across the housing, instead of sliding against it, when the
valve opens and closes.
The ATS Open Pivot valve also features a change in the pivot design.
Traditional bileaflet valve pivots are made up of an extension of the leaflet
fitting into a recess in the housing where thrombus then has a tendency to
form. The Open Pivot design inverts this traditional pivot mechanism, expos-
ing the pivot to bulk forward flow. Exposure to high velocity flow may then
7328_C008.fm Page 284 Tuesday, October 10, 2006 11:04 AM
remove the protein and cell deposition on the pivot surface that leads to
thrombus formation.
The On-X valve, marketed by the Medical Carbon Research Institute, is
the most recent design introduction to the U.S. Some of the innovative
features of this valve are a length-to-diameter ratio close to that of native
heart valves, a smoothed pivot recess that allows the leaflets to open at an
angle of 90 degrees relative to the valve housing, and a two-point landing
mechanism during valve closure. The first two features result in improved
bulk flow hemodynamics, while the latter may result in a more even distri-
bution of stresses during valve closure, a quiet closure event, and a reduced
cavitation potential.
The majority of valve replacements today utilize mechanical valves and
the most popular design is the bileaflet, with approximately 80% of the
mechanical valves implanted today being bileaflet prostheses.
(a)
(b)
FIGURE 8.2
Photographs of biological leaflet valves: (a) a Hancock porcine bioprosthetic valve and (b) a
Carpentier–Edwards pericardial valve prosthesis. (From Yoganathan, A.P. (1995) Cardiac valve
prostheses, in The Biomedical Engineering Handbook, Bronzino, J.D. (Ed.), CRC Press, Boca Raton,
FL. With permission.)
7328_C008.fm Page 287 Tuesday, October 10, 2006 11:04 AM
approved by the FDA in 1997. Hemodynamic tests of this valve showed very
little resting pressure gradient and clinical results of resistance to wear are
very promising. Two other stentless aortic valve designs have since been
brought to the market: the Medtronic Freestyle and Edwards Prima valves.
The Prima valve allows full root or subcoronary implantation and removes
the coronary arteries from the root, while the Freestyle offers implantation
by a variety of surgical techniques and offers ligated coronary arteries.
Stentless bioprostheses are currently only approved for aortic valve
replacements in the U.S. Stentless mitral valves have been under develop-
ment for a number of years, as clinical trials with these valves were first
reported in 1992. Clinical and in vitro studies of stentless mitral valves have
shown that they have superior hemodynamics, but the questionable dura-
bility of these valves and complexity of the implantation technique have
delayed their approval for the U.S. market.
• Caged ball
• Tilting disc
• Bileaflet
• Stented bioprosthesis
7328_C008.fm Page 289 Tuesday, October 10, 2006 11:04 AM
In terms of specific issues related to heart valve design, the basic engineer-
ing considerations are:
• Flow dynamics
• Durability (structural mechanics and materials)
• Biological response to the prosthetic implant
1
p1 − p2 = ρ(V22 − V12 ) (1.53)
2
p1 − p2 = 4V22 (1.55)
where the constant has units of (mmHg)/(m2/s2). Thus, if the velocity (m/s)
is measured with continuous-wave Doppler ultrasound velocimeter (see
Chapter 10, section 10.6: Ultrasound Doppler Velocimetry), the pressure drop
(mmHg) can be obtained noninvasively using Equation 1.55.
When measuring pressure distal to prosthetic valves, it is important to
note whether or not the recovered pressure is measured. This is especially
important when using continuous-wave Doppler ultrasound to evaluate
prosthetic heart valves. Because of its large sample volume size, continu-
ous-wave Doppler ultrasound measures the highest flow velocity, which
occurs at the vena contracta. Consequently, the Doppler-derived pressure
drops are always based on the pressure at the vena contracta, which is at a
site proximal to where pressure recovery occurs. Thus, one is likely to
overestimate the transvalvular pressure drop with this method because
pressure recovery is not considered. However, even though Doppler ultra-
sound may overestimate transvalvular pressure drops, they are extremely
7328_C008.fm Page 291 Tuesday, October 10, 2006 11:04 AM
30
SJM regent
Medtronic hall
Pressure difference (mm Hg)
Hancock II
CE pericardial
20
Medtronic freestyle
10
0
200 300 400 200 300 400 200 300 400
Flow rate (cm3/s) Flow rate (cm3/s) Flow rate (cm3/s)
390 µm
(a) (b) (c)
FIGURE 8.3
Comparison of transvalvular pressure drop as a function of root mean square flow rate across
various mechanical and biological valve prostheses with tissue annulus diameter of (a) 19/20 mm,
(b) 25 mm, and (c) 27 mm. (Note that the pressure drop decreases with larger tissue annulus
diameters due to decrease in resistance of flow across the valve.)
7328_C008.fm Page 292 Tuesday, October 10, 2006 11:04 AM
∫
V1 (t)A1
A2 = dt (8.1)
V2 (t)
T
A2 is the area of the vena contracta or the effective valve area and A1 is the
area at the upstream face of the control volume, measured by echocardio-
graphy. The integral is taken with respect to time over the systolic flow
period, V1 is the velocity upstream of the aortic valve, and V2 is the velocity
at the vena contracta. Inherent to Equation 8.1 is the assumption that the
spatial velocity profile is uniform over the entire orifice area, but because
the velocity profile is nonuniform in mechanical prostheses, it is difficult to
apply this equation accurately in that case.
By extending the control volume from the left ventricular outflow tract to
encompass the entire left ventricle, it is possible to use this technique to
determine the mean mitral valve area. The time-velocity integral of both the
mitral and the left ventricular outflow tract positions must be calculated over
the entire cardiac cycle. It is only necessary to integrate the velocity at the
mitral valve during diastole and the velocity of the aortic valve during systole
because these are the only times in which there is flow through the valves.
If regurgitation is not present in either valve, Equation 8.1 can be applied and
the effective area of the mitral valve can be obtained. When aortic regurgita-
tion is present, the velocity and area at the right ventricular outflow tract can
be used instead for calculation of effective mitral orifice area.
Another method of determining valve area can be derived using the
Bernoulli equation and Conservation of Mass. A contraction coefficient (Cd)
can be defined as the area of the vena contracta divided by the area of the
valve orifice (Ao). The continuity equation for flow through an orifice, includ-
ing the contraction coefficient appears in Equation 8.2, where Q is the flow
rate and v is the velocity at the vena contracta
Q = A0VCd (8.2)
Solving for V2 in Equation 1.54, substituting into Equation 8.2 and rearrang-
ing, yields an equation that can be used to determine the effective orifice
area, EOA, of a cardiac valve
Q rms
EOA(cm 2 ) = (8.3)
51.6 ∆p
7328_C008.fm Page 293 Tuesday, October 10, 2006 11:04 AM
TABLE 8.1
Test Data for 27 mm Tissue Annulus Diameter-Size Valves with
Measurements Obtained Using an In Vitro Pulse Duplicator with a Heart
Rate of 70 bpm and a Cardiac Output of 5.0 l/min
Peak Turb.
SS+
Reg. Vol. (dynes/
Valve Type EOA*(cm2) PI (cm3/beat) cm2) ∆P mmHg
Caged ball 1.75 0.30 5.5 1850 13.6
Tilting disc# 3.49 0.61 9.4 1800 3.4
Bileaflet# 3.92 0.68 9.15 1940 2.7
# Values reported are mean values from several valve models of the same type.
where Qrms is the root mean square systolic/diastolic flow rate (cm3/s),
and ∆p is the mean systolic/diastolic pressure drop (mmHg). The nondi-
mensional constant, Cd, was determined by using the density of blood ( ρ =
1050 kg/m3) and converting the units to those convenient for biomedical
use. The effective area will be in cm2 if the root mean square systolic or
diastolic flow rate (Qrms) is in cm3/s and the mean systolic or diastolic
pressure drop ( ∆p) is measured in mmHg. Similar relationships were
derived earlier in Chapter 5 (Equation 5.4 and Equation 5.5) through the
application of Bernoulli’s equation where the magnitude of the constants
employed depends upon the discharge coefficient for a particular valve.
Table 8.1 lists EOAs obtained in vitro, for mechanical and tissue valve
designs in clinical use today. These results illustrate the fact that the newer
mechanical valve designs have better pressure gradient characteristics than
porcine bioprostheses in current clinical use.
Another method that has been developed to estimate the orifice area
involves measuring the pressure half-time of an orifice. The pressure half-time
( p t ) is the time required for left ventricular pressure to fall to one-half of
2
its peak value. An empirical relation has been established between the area
of the mitral valve ( AreaMV ) and p t as
2
220
AreaMV = pt (8.4)
2
7328_C008.fm Page 294 Tuesday, October 10, 2006 11:04 AM
8.3.3 Regurgitation
Because mechanical prosthetic valves are fairly rigid, they are unable to form
tight seals between the occluder and supporting ring when closed. Conse-
quently, regurgitant jets are present in the resultant gaps in the prosthetic
valves under normal conditions, although the amount of regurgitation is
usually small. Normal regurgitant flow is characterized by having a closing
volume during valve closure and leakage after closure. These parameters
are illustrated in Figure 8.4. Regurgitant flow is often characterized by the
regurgitant volume. The regurgitant volume is the total volume of fluid
through the valve per beat due to the retrograde flow and is equal to the
sum of the closing volume and the leakage volume. The closing volume is
the volume of fluid flowing retrograde through the valve during valve clo-
sure. Any fluid volume accumulation after valve closure is due to leakage
and is referred to as the leakage volume.
Closing regurgitation is related to the valve shape and closing dynamics,
and the percentage of stroke volume that succumbs to this effect is in the
range of 2.0 to 7.5% for mechanical valves and is typically less for tissue
valves, ranging from 0.1 to 1.5%. Leakage, which depends upon how well
the orifices are ‘‘sealed’’ upon closure, has a reported incidence of 0 to 10%
in mechanical valves and 0.2 to 3% in bioprosthetic valves. The overall
Forward flow
FIGURE 8.4
A schematic illustrating the closing and leakage volume across a valve prosthesis. (From
Yoganathan, A.P. (1995) Cardiac valve prostheses, in The Biomedical Engineering Handbook,
Bronzino, J.D. (Ed.), CRC Press, Boca Raton, FL. With permission.)
7328_C008.fm Page 295 Tuesday, October 10, 2006 11:04 AM
tendency is for regurgitation to be less for the bioprosthetic heart valves than
for mechanical valve designs (see Table 8.1). To distinguish normal from
abnormal valve function, it is important to differentiate between normal
regurgitant volume and additional regurgitation due to disease. Fluid
mechanical analysis of this problem has provided two different techniques —
Turbulent Jet Theory and Proximal Flow Convergence — that at least
partially fulfill this requirement.
Q0 = ( 4 πr 2 )Vr (8.5)
7328_C008.fm Page 296 Tuesday, October 10, 2006 11:04 AM
Q0 is the flow rate at the regurgitant orifice and the term within parentheses
is the surface area of the hemispherical surface. Vr is the velocity measured
with the color Doppler echocardiography and r is the radial distance at which
the Velocity is measured.
Because of its simplicity and ease of application, this technique has
received a great deal of attention, especially for mitral regurgitation. The
effects of regurgitant orifice motion, orifice geometry variation, and ultra-
sound machine settings have all been addressed with both in vitro and in
vivo investigations. In addition, its application to prosthetic valve regurgita-
tion has also been considered. Recent studies have found that isovelocity
contours from regurgitant orifices can be described better by a hemi-elliptical
shape than a hemispherical shape, yielding more accurate estimations of
regurgitant flow rate in vitro. Unfortunately, rigorous in vivo validation of
the proximal flow convergence technique is difficult.
are measured from the valve sewing ring. Table 8.1 lists peak turbulent shear
stress levels measured downstream of the common 27-mm valve prostheses.
200.0 2000.0
100.0 1000.0
0.0 0.0
(a) (b)
FIGURE 8.5
(a) Velocity and (b) turbulent shear stress profiles 30 mm downstream of a caged ball valve at
peak systole (from Yoganathan, A.P. Cardiac Valve Prostheses, in The Biomedical Engineering
Handbook, Bronzino, J.D. (Ed.), 1995. With permission from CRC Press, Boca Raton, Florida).
7328_C008.fm Page 299 Tuesday, October 10, 2006 11:04 AM
200.0
200.0 200.0
100.0 100.0
100.0
0.0
0.0
0.0
(a) (b)
2000.0
2000.0 2000.0
1000.0 1000.0
1000.0
0.0
0.0 0.0
(c) (d)
FIGURE 8.6
(a) Velocity profile 15 mm downstream, (b) velocity profiles 13 mm downstream, (c) turbulent
shear stress profile 15 mm downstream, and (d) turbulent shear stress profiles 13 mm down-
stream of a tilting disc valve at peak systole (from Yoganathan, A.P. Cardiac Valve Prostheses,
in The Biomedical Engineering Handbook, Bronzino, J.D. (Ed.), 1995. With permission from CRC
Press, Boca Raton, Florida).
is more evenly distributed across the flow chamber during the deceleration
phase than during the acceleration phase. Regions of flow separation are
observed around the jets adjacent to the flow channel wall as the flow
separates from the orifice ring. The measurements across the central orifice
show that the maximum velocity in the central orifice is 220 cm/s. Small
regions of low-velocity reverse flow are observed adjacent to the pivot/hinge
mechanism of the valve. More flow emerges from the central orifice during
the deceleration phase than during the acceleration phase.
High turbulent shear stresses occur at locations of high velocity gradients
and at locations immediately distal to the valve leaflets. The flow along the
7328_C008.fm Page 301 Tuesday, October 10, 2006 11:04 AM
200.0
200.0
100.0
100.0
0.0
0.0
(a) (b)
2000.0
2000.0
1000.0
1000.0
0.0
0.0
(c) (d)
FIGURE 8.7
Velocity profile 13 mm downstream of a bileaflet valve; (a) across the leaflets, (b) across the
central orifice and turbulent shear stress profiles 13 mm downstream of a bileaflet valve, (c)
across the leaflets, and (d) across the central orifice at peak systole (from Yoganathan, A.P.
Cardiac Valve Prostheses, in The Biomedical Engineering Handbook, Bronzino, J.D. (Ed.), 1995.
With permission from CRC Press, Boca Raton, Florida).
400.0 5000.0
200.0 2500.0
0.0 0.0
(a) (b)
FIGURE 8.8
(a) Velocity profile and (b) turbulent shear stress profiles 15 mm downstream of a porcine
bioprosthesis at peak systole (from Yoganathan, A.P. Cardiac Valve Prostheses, in The Biomedical
Engineering Handbook, Bronzino, J.D. (Ed.), 1995. With permission from CRC Press, Boca Raton,
Florida).
200.0 2000.0
100.0 1000.0
0.0 0.0
(a) (b)
FIGURE 8.9
a) Velocity profile and (b) turbulent shear stress profiles 17 mm downstream of a pericardial
tissue valve prosthesis at peak systole (from Yoganathan, A.P. Cardiac Valve Prostheses, in The
Biomedical Engineering Handbook, Bronzino, J.D. (Ed.) 1995. With permission from CRC Press,
Boca Raton, Florida).
7328_C008.fm Page 304 Tuesday, October 10, 2006 11:04 AM
Hinge
Leaflet
Thumbnail
Housing
FIGURE 8.10
Schematic of the hinge for one bileaflet valve model. (From Leo et al. (2002) Microflow fields
in the hinge region of the CarboMedics bileaflet mechanical heart valve design, in J. Thorac.
Cardiovasc. Surg., 124: 561–574. With permission.)
Leakage can also exit the narrow spaces between the occluder and housing
just upstream of mechanical valves in the form of high speed jets. The pattern
of these leakage jets is particular to the valve design. Figure 8.12 shows the
results of a laser Doppler velocimetry experiment designed to measure
velocity and turbulence within half the leakage jets of a bileaflet valve in
vitro under aortic flow conditions. The lengths and directions of the arrows
in this figure are representative of the jet velocity magnitude and direction.
The valve diagram in the upper right hand corner of this figure shows the
relative positions of the leakage jets with respect to the valve. Two of these
jets exit the valve from pivots, while the remaining jet appears to exit from
the intersection of the two leaflets and the valve housing. The largest velocity
magnitude measured in these jets was 0.6 m/s, and the largest turbulent
shear stress magnitude measured was 28.9 N/m2.
A glance at Figure 8.12 shows that the jets exiting from the valve are
asymmetric, though the valve itself is symmetric. This is likely due to the
Leaflets
FIGURE 8.11
A comparison of hinge geometry for two commercially available bileaflet valve designs. (From
Leo et al. (2002) Microflow fields in the hinge region of the CarboMedics bileaflet mechanical
heart valve design, in J. Thorac. Cardiovasc. Surg., 124: 561–574. With permission.)
7328_C008.fm Page 306 Tuesday, October 10, 2006 11:04 AM
1
2
1 2
3 Velocity
magnitude (m/s)
0.6
0.0
FIGURE 8.12
Laser Doppler measurement of a leakage jet of a bileaflet valve. (From Travis, B.R. et al. (2002)
An analysis of turbulent shear stress in leakage flow through bileaflet mechanical prostheses,
in J. Biomech Eng., 2002, 124(2):155–165.)
portion of systole and diastole, a level that could lead to damage to blood
elements. In the case of mechanical prostheses, the chances for blood cell
damage are increased due to the presence of foreign surfaces. Furthermore,
the regions of flow stagnation and/or flow separation that occur adjacent to
the superstructures of these valve designs could promote the deposition of
damaged blood elements, leading to thrombus formation on the prosthesis.
The clinical performance of the Medtronic Parallel valve is an example of
the effects of these flow patterns on likelihood for thromboembolic compli-
cations. This valve showed superior forward flow hemodynamics in an in
vivo porcine model, suggesting a good clinical performance. However, the
Parallel valve performed very poorly in clinical trials; approximately 20% of
patients in these trials developed thrombosis. Patient and material factors
were statistically eliminated from the potential reasons for the poor perfor-
mance. Subsequent in vitro analysis of flow through the valve pivot revealed
regions of highly disturbed vortical flow within this area during the leakage
phase. Thrombi from explant valves were localized within the pivots to these
disturbed flow regions.
8.5.1 Wear
Abrasive wear of valve parts has been, and continues to be, a serious issue
in the design of mechanical prosthetic valves. Various parts of these valves
come in contact repeatedly with each other for hundreds of millions of cycles
over the lifetime of the device. A breakthrough occurred with the introduc-
tion of pyrolytic carbon (PYC) as a valve material because it has relatively
good blood compatibility characteristics and wear performance. It has been
shown that while wear of PYC upon PYC and PYC upon metals is relatively
low, wear of PYC by metals is considerably greater. One example of this is
a PYC disc mounted on a metallic orifice/hinge combination. The most
durable wear couple is PYC–PYC; therefore, PYC-coated components are
very attractive. The first valve to employ a PYC–PYC couple was the St. Jude
Medical valve, which has fixed pivots for the leaflets. Tests indicate that it
would take 200 years to wear halfway through the PYC coating on a leaflet
pivot. By creating designs that allow wear surfaces to be distributed rather
than focal (e.g., the Omnicarbon valve, which has a PYC-coated disc that is
free to rotate), it is possible to reduce wear even further.
Despite these advances, wear on prosthetic heart valve surfaces can still
be problematic, as can be seen by the mechanical failures of the Edwards–
Duromedics valve. The Edwards–Duromedics valve design was introduced
in 1982, but withdrawn from the market after one of the leaflets in a number
of these valves fractured and embolized within the vascular system of
patients. Scanning electron microscopy examinations of the fractured leaflets
revealed regions of pitting on the leaflet and housing surfaces where the
leaflet contacts the housing during closure and within the pivot area as well
as several regions of concentrated micro pores in the carbon surfaces near
where the fractures occurred. A subsequent study showed that these micro
pores existed only in the Edwards–Duromedics design, and hypothesized
that they were created as a result of the shaping of carbon components with
a mandrel, a process that was unique to the Edwards–Duromedics valve.
This study linked pitting (indicative of cavitation damage) to the presence
of these micro pores. Other studies have disputed the connection between
micro pores and pitting damage. The fact that the cause of leaflet emboliza-
tion in these valves remains unknown means that wear remains a potential
problem in the design of mechanical prostheses.
8.5.2 Fatigue
Metals are prone to fatigue failure. Their polycrystalline nature contains
structural characteristics that may produce dislocations under mechanical
loading. These dislocations can migrate when subjected to repeated loading
cycles and can accumulate at intercrystalline boundaries with the end result
7328_C008.fm Page 311 Tuesday, October 10, 2006 11:04 AM
being tiny cracks. These tiny cracks are sites of stress concentration in which
the fissures may worsen until fracture occurs. The Haynes 25 Stellite
Björk–Shiley valve, which used a chromium–cobalt alloy, experienced the
most severe fatigue problem for a mechanical valve. While previous inves-
tigations had suggested that fatigue was not a problem for PYC, recent data
contradicts this and suggests that cyclic fatigue-crack growth occurs in
graphite/pyrolytic carbon composite materials. This work suggests a fatigue
threshold that is as low as 50% of the fracture toughness and views cyclic
fatigue as an essential consideration in the design and life prediction of heart
valves constructed form PYC. The FDA now requires detailed characteriza-
tion of PYC materials used in different valve designs.
8.5.3 Mineralization
The major cause of both porcine aortic and pericardial bioprosthetic valve
failure is calcification, a process which stiffens leaflets and frequently causes
cuspal tears. Calcific deposits occur most commonly at the commissures and
basal attachments and are most extensively deep in (intrinsic to) the cusps
in the spongiosa layer. Ultrastructurally, calcific deposits are associated with
cuspal connective tissue cells and collagen. Degenerative cuspal calcific
deposits are composed of calcium phosphates that are chemically and struc-
turally related to physiologic bone mineral (hydroxyapatite). The flexing
bladders in cardiac assist devices, flexing polymeric heart valves, and vas-
cular grafts have also been found to be vulnerable to calcific deposits. In
such cases, calcification is usually related to inflammatory cells adjacent to
the blood-contacting surface, rather than to the implanted material itself.
The mechanisms of calcification and the methods of preventing calcifica-
tion are an active area of current research. The most common methods of
studying calcification involve valve tissue implanted either subcutaneously
in 3-week-old weanling rats or valves implanted as mitral replacements in
young sheep or calves. Results of both types of studies show that biopros-
thetic tissue calcifies in a fashion similar to clinical implants, but at a greatly
accelerated rate. The subcutaneous implantation mode is a well accepted,
technically convenient, economical, and quantifiable model for investigating
mineralization issues. It is also very useful for determining the potential of
new antimineralization treatments.
Host, implant, and biomechanical factors impact the calcification of tissue
valves. Patients who are young or have renal failure are vulnerable to valve
mineralization, but immunological factors seem to be unimportant. Pretreat-
ment of valve tissue with an aldehyde cross-linking agent has been found
to cause calcification in rat subcutaneous implants; non-preserved cusps, on
the other hand, do not mineralize. Calcification of bioprosthetic valves is
greatest at the cuspal commissures and bases where leaflet flexion is the
greatest and deformations are maximal. Most data suggest that the basic
mechanisms of tissue valve mineralization result from aldehyde pretreat-
ment, which changes the tissue microstructure.
7328_C008.fm Page 312 Tuesday, October 10, 2006 11:04 AM
8.7 Conclusions
Direct comparison of the ‘‘total’’ performance of artificial heart valves is
difficult, if not impossible. The precise definition of criteria used to bench-
mark valve performance varies from study to study. To study long-term
performance, large numbers of patients and lengthy observation periods are
required. During these periods, there may be an evolution in valve materials
7328_C008.fm Page 313 Tuesday, October 10, 2006 11:04 AM
While the current status of artificial valves leaves room for further
improvement, the superior prognosis for the patient with a replacement
heart valve is dramatic and convincing.
8.8 Problem
8.1 A 27-mm bileaflet valve was tested in a pulse duplicator at a mean
systolic flow rate of 15 lpm. The corresponding mean pressure drop across
the valve was measured to be 5.2 mmHg. Assuming a discharge coefficient
of 0.61 for the bileaflet valve, compute the effective valve orifice area and
the performance index for the valve.
7328_C008.fm Page 314 Tuesday, October 10, 2006 11:04 AM
7328_C009.fm Page 315 Monday, October 16, 2006 11:05 AM
9
Vascular Therapeutic Techniques
CONTENTS
9.1 Vascular Graft Implants............................................................................315
9.2 Arteriovenous Fistulas..............................................................................317
9.3 Types of Vascular Graft Materials Used ................................................319
9.4 Clinical Experience with Vascular Grafts ..............................................322
9.5 Biomechanics and Anastomotic Intimal Hyperplasia .........................324
9.6 Angioplasty, Stent, and Endoluminal Graft Implants.........................333
9.7 Biomechanics of Stent Implants ..............................................................338
9.8 Problems......................................................................................................342
315
7328_C009.fm Page 316 Monday, October 16, 2006 11:05 AM
LMCA LAD
CIRC
OM
FIGURE 9.1
Angiogram of the left coronary system of a patent with severe coronary artery disease. (From
Gotto, A.M. Jr. et al. (1977) Atherosclerosis, Upjohn, Pfizer, Inc., New York. With permission.)
FIGURE 9.2
Coronary arteriograms, including 7-year follow-up of patient who underwent coronary artery
bypass to left anterior descending coronary artery. (From Gotto, A.M. Jr. et al. (1977) Atheroscle-
rosis, Upjohn, Pfizer, Inc., New York. With permission.)
7328_C009.fm Page 317 Monday, October 16, 2006 11:05 AM
In this surgical procedure, the affected vessel is exposed and the bypass
graft is placed alongside it and sewn to arterial sites well upstream and
downstream of the stenosis. The junctions between the artery and the graft
are called anastomoses and are generally constructed in an end-to-side
fashion. The primary exception to this is with grafts used for repair of an
abdominal aortic aneurysm where the anastomoses are placed in an end-to-
end fashion. Once the graft is in place, blood flow is then reinstituted around
the diseased area, often with some residual flow still being carried through
the native artery. A variety of surgical procedures are used to perform cor-
onary artery bypass grafting with internal mammary arteries, ranging from
the traditional open-chest procedure to newer approaches using either a
thoracotomy (OPCAB) or only a small incision in the ribcage (MIDCAB),
both of which avoid the need for simultaneous heart–lung bypass support
during the procedure.
FIGURE 9.3
A brachiobasilic bridge (loop) graft. (From Ernst, C.B. and Stanley, J.C. (1991) Current Therapy
in Vascular Surgery, 2nd ed, B.C. Decker, Philadelphia. With permission.)
FIGURE 9.4
Flow pattern as seen in the high volume flow graft. (From Villemarette, P.T. et al. (1989) Use of
color flow Doppler to evaluate vascular access graft function, in J. Vasc. Tech. 13: 164–170. With
permission.)
7328_C009.fm Page 319 Monday, October 16, 2006 11:05 AM
(a) (b)
FIGURE 9.5
Technique for resection and replacement of infrarenal aneurysm of abdominal aorta. (From
Ernst, C.B. and Stanley, J.C. (1991) Current Therapy in Vascular Surgery, 2nd ed, B.C. Decker,
Philadelphia. With permission.)
(a) (b)
FIGURE 9.6
Arteriogram in a patient with patent femoroposterior tibial ePTFE graft 10 years after placement
of the graft: (a) proximal graft, (b) distal anastamosis. (From Ernst, C.B. and Stanley, J.C. (1991)
Current Therapy in Vascular Surgery, 2nd ed, B.C. Decker, Philadelphia. With permission.)
FIGURE 9.7
Scanning electron micrograph illustrating results of the prime electrostatic endothelial cell
transplantation onto ePTFE (GORE-TEX®). (From Fields, C. et al. (2001) The persistence of
electrostatically seeded endothelial cells lining a small diameter expanded polytetrafluoroeth-
ylene vascular graft, in J. Biomaterials Applications 16:157-173. With permission.)
7328_C009.fm Page 322 Monday, October 16, 2006 11:05 AM
response, but also may actively reduce the potential for thrombosis and
intimal hyperplasia by the production of agents, such as prostaglandins
and nitric oxide.
100
440
90
375
80
291 Secondary patency
201
70 120
81
57 27
60 21
% Patency
11 6 3 2
50 51%
Primary patency
40 44%
30
20
10
0
0 6 12 18 24 30 36 42 48 54 60 66 72
Months after operation
FIGURE 9.8
Cumulative life table patency rates for 440 PTFE femoropopliteal bypasses. The lower line
illustrates primary patency rates, while the upper line shows secondary (achieved after throm-
bectomy) patency rates. (From Sawyer, P. N. (1987) Modern Vascular Grafts, McGraw-Hill, New
York. With permission.)
7328_C009.fm Page 323 Monday, October 16, 2006 11:05 AM
Flow
POS flow Flow
Proximal or donor end
Graft heel
Bypass graft
Hood
Artery
floor
Suture line
Graft toe
Dos flow
Distal or receptor end
Flow
FIGURE 9.9
Sites of intimal hyperplasia formation in a bypass graft distal anastomosis (POS — proximal
outflow segment, DOS — distal outflow segment). (From Li, X.-M. (1998) Evaluation of hemo-
dynamic factors at the distal end-to-side anastomosis of a bypass graft with different POS:DOS
ratios, Ph.D. Dissertation Dept. of Biomedical Engineering, The University of Akron, OH.)
(see Chapter 6, section 6.5: Wall Shear Stress and its Effect on Endothelial
Cells). Since these variables are particularly affected by the geometric
configuration of the anastomosis, extensive work has been performed to
examine the effect of the choice of graft caliber and the anastomotic angle,
in addition to the flow rate, upon eventual patency of the graft.
(a)
(b)
FIGURE 9.10
Three-dimensional sketch of the path line of the high inertia fluid seen under steady flow
conditions at Re = 950 and 1300, as well as for unsteady flow at around peak flow: (a) side
view, showing development of the double helix; (b) top view. (From Ojha, M. et al. (1990)
Influence of angle on wall shear stress distribution for an end-to-side anastomosis model, in
J. Vasc. Surg. 12: 747–53. With permission.)
7328_C009.fm Page 325 Monday, October 16, 2006 11:05 AM
residence times for blood cells and other components to interact with the
vessel wall. A compounding factor related to this is the stiffness, or lack of
compliance (= Volume/Pressure), of the graft material since changes in
vessel diameter under pulsatile flow conditions will further alter the local
geometry. A mismatch in compliance between the graft and native artery
could also lead to significant changes in local fluid and solid wall stresses,
which might then elicit various tissue responses. We have mentioned earlier,
for example, the relatively high stiffness of ePTFE material, which when
interposed with more elastic native vessels may produce elevated wall
stresses and impedance mismatches. This problem may persist, however,
even when using more compliant biological materials, such as veins, since
they may also become very stiff when exposed to arterial pressures. It is not
a surprising observation, then, that intimal hyperplasia commonly develops
at the toe and heel regions as well as along the artery floor of the anastomosis
(Figure 9.11).
Because these sites coincide closely with areas of altered biomechanical
variables, a number of studies have been performed to look at the effect of
geometric and material parameters upon the local fluid dynamics and sub-
sequent vascular responses. Specific variables that have been investigated
include: (1) material stiffness differences between the two vessels, (2) the
angle of the anastomosis, (3) the diameter mismatch between the graft and
the host artery, and (4) the flow distribution through the artery (i.e., the
FIGURE 9.11
Intimal hyperplasia has developed at the anastomosis of this Dacron aortofemoral graft with
the profunda femoris artery. At the time of patch angioplasty, a smooth, glistening, white plaque
was found. (From Gupta, S.K. (1993) Vascular Graft Monitoring, R.G. Landes Co., Austin, TX.)
7328_C009.fm Page 326 Monday, October 16, 2006 11:05 AM
Compliance vs % Patency
R = 0.873
100
75
% Patency
50
Arterial autografts
25 Vein autografts
Dacron
P.T.F.E.
0
1 2 3 4 5 6
Compliance (%/mm Hg × 10–2)
FIGURE 9.12
Linear regression analysis of compliance vs. graft patency. (From Abbott W.M. and Bouchier-
Hayes D.J. The Role of Mechanical Properties in Graft Design in (1978) Graft Materials in Vascular
Surgery, Dardik, H., Ed. Year Book Medical Publishers, Chicago. With permission.)
FIGURE 9.13
Flow patterns in anastamoses of 30, 45, and 60 degrees at Re = 205 (Note: Images were obtained
using H2 bubble flow visualization.) (From Keynton R.S. et al. (1991) The effect of graft caliber
upon wall shear within in vivo distal vascular anastomosis, in J. Biomech. Eng. 113: 458–463.
With permission.)
Re = 100
30° Inner wall
Outer wall
Re = 205
30° Inner wall
Outer wall
Re = 100
45° Inner wall
Outer wall
Re = 205
45° Inner wall
Outer wall
Re = 100
60° Inner wall
Outer wall
Re = 205
60° Inner wall
Outer wall
FIGURE 9.14
Normalized axial shear rate in anastomoses of 30, 45, and 60 degrees at Re = 100 and 205. (Note:
Wall shear rates were derived from laser Doppler anemometry velocity measurements.) (From
Keynton R.S. et al. (1991) The effect of graft caliber upon wall shear within in vivo distal vascular
anastomosis, in J. Biomech. Eng. 113: 458–463. With permission.)
7328_C009.fm Page 329 Monday, October 16, 2006 11:05 AM
60
20
Graft inlet
0 4 8 12 (mm)
Artery outlet
–12 –8 –4 0 4 8 12 (mm)
60
Mean velocity (cm/s)
40
20
Artery floor
–20
60
Toe region
Mean velocity (cm/s)
Radial position
0.35 mm
0.70 mm 40
1.05 mm
20
Graft inlet
0
0 4 8 12 (mm)
Artery outlet
–12 –8 –4 0 4 8 12 (mm)
60
Mean velocity (cm/s)
40
20
Artery floor
–20
FIGURE 9.15
Typical plots of mean axial velocities at three radial positions along the artery toe and floor
regions for diameter ratio (DR) = 1.0 (top) and DR = 1.5 (bottom). (Note: Velocity measurements
made using pulse Doppler ultrasound.) (From Keynton, R.S. et al. (1999) The effect of graft
caliber upon wall shear within in vivo distal vascular anastomosis, in J. Biomech. Eng. 121: 79–88.
With permission.)
7328_C009.fm Page 330 Monday, October 16, 2006 11:05 AM
t = 40 ms t = 40 ms
t = 150 ms t = 150 ms
t = 210 ms t = 210 ms
S
R
t = 300 ms t = 300 ms
t = 625 ms t = 625 ms
FIGURE 9.16
Vector plots in the ETS anastomosis for flow splits (ratio of distal to proximal arterial outflow)
of 50:50 and 75:25 at various times during the flow cycle. Flow separation at S and reattachment
at R occur on the graft hood during the systolic deceleration phase (t = 210 and 300 ms) with
a flow split of 50:50. (From How T.V. et al. (2000) Interposition vein cuff anastomosis alters wall
shear stress distribution in the recipient artery, in J. Vasc. Surg. 31: 1008–17. With permission.)
Y Symmetry plane
X
Z
Mean stress
Host 32
Sutures
artery 28
24
20
Dacron 16
graft 12
8
4
Symmetry plane
0
FIGURE 9.17
Mean stress distribution at the end-to-end Dacron graft–artery anastamosis projected onto a
three-dimensional image of the geometry. (From Ballyk P.D. et al. (1998) Compliance mismatch
may promote graft-artery intimal hyperplasia by altering suture-line stresses, in J. Biomech. 31:
229–37. With permission.)
Z
Y X
all
ew Mean stress
Sid
Sutures 32
28
Toe 24
Host 20
artery 16
12
Heel Dacron
graft 8
4
Plane of 0
Symmetry
FIGURE 9.18
Mean stress distribution at the end-to-side Dacron graft–artery anastomosis projected onto a
three-dimensional image of the geometry. (From Ballyk P.D. et al. (1998) Compliance mismatch
may promote graft-artery intimal hyperplasia by altering suture-line stresses, in J. Biomech. 31:
229–237. With permission.)
7328_C009.fm Page 332 Monday, October 16, 2006 11:05 AM
1200
800
600
400
200
–200
–500 –250 0 250 500 750 1000 1250 1500
Mean wall shear rate (1/s)
(a)
1200
1000
Intimal hyperplasia (µm)
800
600
400
200
–200
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Oscillatory shear index
(b)
FIGURE 9.19
Nonlinear regression of (top) intimal hyperplasia (IH) vs. mean wall shear rate (WSR) (r = – 0.483)
and (bottom) IH vs. oscillatory shear index (OSI) (r = 0.600). (From Keynton R.S. et al.
(2001) Intimal hyperplasia and wall shear in arterial bypass graft distal anastomoses, in
J. Biomech. Eng. 123: 464–73. With permission.)
development of IH with the level of wall shear rate (WSR) at the time of
implantation (Figure 9.19). These data show a modest inverse relationship
(r = – 0.483) between IH and WSR, where IH is greatest at low and reversing
WSRs and reaches a near-zero asymptote around 1000 s−1 — a level compa-
rable to ~ 2 times the normal mean WSR seen in major arteries. Furthermore,
there is a direct correlation (r = 0.6) of IH with the OSI (oscillatory shear
7328_C009.fm Page 333 Monday, October 16, 2006 11:05 AM
C.P.
Dye
Guiding cath.
FIGURE 9.20
Schematic of a percutaneous balloon angioplasty device showing guidewire (Wire), catheter,
and ports for balloon pressure (Balloon P.), dye injection (Dye), and catheter pressure (C.P.).
FIGURE 9.21
Digital subtraction angiograms showing excellent and long-term immediate results of percuta-
neous transluminal angioplasty in a patient with severe ulcerated internal carotid stenosis.
(From Brown, M.M. (1996) Balloon angioplasty for extracranial carotid disease, in Advances in
Vascular Surgery, Whittemore, A.D., Ed. (1996) Advances in Vascular Surgery, vol. 4, Elsevier,
Philadelphia. With permission.)
PTA with a separate catheter. Once in place, the stent provides a mechanical
scaffold that can help sustain the vessel opening over time (Figure 9.23).
Other variations on this approach involve compressing a coiled wire mesh
inside a small plastic sleeve and then deploying it within the stenosis by
pulling back the sleeve and allowing the stent to spring open (Figure 9.24).
Another approach is to use a unique metal, Nitinol, which has “thermal
memory” properties. This type of stent is first shaped into a coil at warm
temperatures and then straightened into a wire at room temperature (~ 22°C).
Once it is reheated by the body’s blood (~37°C) at the site of stenosis, the
device then returns to its original coiled shape.
FIGURE 9.22
Balloon-expandable coronary artery stent. (From Didisheim, P. and J.T. Watson, Cardiovascular
Applications. Ratner, B.D., Ed. (1996) Biomaterials Science: An Introduction to Materials in Medicine,
Elsevier, Philadelphia. With permission.)
7328_C009.fm Page 336 Monday, October 16, 2006 11:05 AM
FIGURE 9.23
Digital subtraction angiogram immediately after the percutaneous transluminal insertion of a
stent across a carotid stenosis at the bifurcation. (From Brown, M.M. (1966) Balloon angioplasty
for extracranial carotid diseases, in Advances in Vascular Surgery Vol. 4, 1996, A.D. Whittemore,
Ed., Elsevier, Philadelphia. With permission.)
FIGURE 9.24
A photo of a Wallstent. The tubular, highly permeable braided mesh has considerable longitudinal
flexibility and radial strength. (From Gwertzman, G. (1995) Clinically tested intravascular stents
and endovascular stent grafts, in Endovascular Stented Grafts for the Treatment of Vascular Diseases,
Marin, M.L., Veith, F.J., and Levine, B.A., Eds. R.G. Landes Co., Austin, TX. With permission.)
for deployment. The catheter used for this procedure is a heavier design
than that used for intra-arterial stent delivery since it carries a larger object
and also because it only passes through the larger arteries of the upper leg
and abdomen. In constructing an endoluminal graft for AAAs, which also
extend down into the iliac arteries from the aorta (and, thus, require distal
anastomoses to each iliac artery), a modular device is used (Figure 9.26).
Here, the main (Y-shaped) component is first inserted between the distal
aorta and one of the iliac arteries. Then, a second (straight) graft segment is
placed in the contra-lateral (opposite) iliac artery and connected to the main
section with additional stents.
Endoluminal grafts have proven to be relatively effective devices, espe-
cially in patients who would not have successfully tolerated surgery. How-
ever, there are several technical difficulties associated with this procedure
that include the need to pass this large object through a potentially diseased
femoral or iliac artery and the need to achieve proper placement of the graft
such that it doesn’t interfere with flow to the renal arteries. The primary
long-term complication experienced with this device has been the develop-
ment of perivascular leaks. These occur because of an incomplete seal
between the graft and the artery, resulting in blood at relatively high pressure
entering the gap between the two vessels. This buildup of blood pressure
could eventually cause a rupture of the aorta very similar to that of an
untreated aneurysm.
7328_C009.fm Page 338 Monday, October 16, 2006 11:05 AM
FIGURE 9.25
The MinTec system for aortic reconstruction. (From Gwertzman, G. (1995) Clinically tested
intravascular stents and endovascular stent grafts, in Endovascular Stented Grafts for the Treatment
of Vascular Diseases, Marin, M.L. Veith F.J. and Levine, B.A. Eds., R.G. Landes Co., Austin, Texas.
With permission.)
FIGURE 9. 26
Device used in patients with aortoiliac aneurysms with no distal neck. (From Gwertzman, G.
(1995) Clinically tested intravascular stents and endovascular stent grafts, in Endovascular Stented
Grafts for the Treatment of Vascular Diseases, Marin, M.L. Veith F.J. and Levine, B.A. Eds., R.G.
Landes Co., Austin, Texas. With permission.)
FIGURE 9.27
Photomicrograph of histological hematoxylin and eosin light microscopic section of atheroscle-
rotic abdominal aorta of a rabbit 8 weeks after a Palmaz stent implantation. (From Becker, G.F.
(1991) Intravascular stents: General principles and status of lower-extremity arterial applica-
tions, in Circulation (Supp. I), 83(2): With permission.)
–1E-5 1E-5
(a)
8 cm/s
–0.0001 0
(b)
FIGURE 9.28
Stream lines (m2/s) (for L/D = 3.53, resting conditions, 4 mm vessel) at (a) midsystole and
(b) peak diastolic forward flow. (From Berry, J.L. et al. (2000) Ann. Biomed. Eng. 28: 386–388.
With permission.)
7328_C009.fm Page 341 Monday, October 16, 2006 11:05 AM
FIGURE 9.29
Illustration of the CMS. The ends of the stent are gradually more flexible toward the ends,
providing a smooth transition in compliance when deployed in an artery. (From Moore, J.E.,
Jr. and Berry, J. L. (2002) Ann. Biomed. Eng. 30: 498–508. With permission.)
0.194 MPa
0.186
0.178
0.170
0.162
0.154
02
0.147
0.139
(a)
8
Circumferential stress (MPa)
CMS
Palmaz
4
0
0 10 20 30 40
–4
Axial position (mm)
(b)
FIGURE 9.30
Color-encoded maximum principal stress in an artery into which the Palmaz stent (top) or the
CMS (bottom) has been deployed. (From Moore, J.E., Jr. and Berry, J.L. (2002) Ann. Biomed. Eng.
30: 498–508. With permission.)
7328_C009.fm Page 342 Monday, October 16, 2006 11:05 AM
9.8 Problems
9.1 A segment of the femoral artery of a dog is replaced with an arterial
graft. The change in radius in response to a change in pulse pressure was
measured in the graft as well as in the femoral artery on the other side
(“contra-lateral”) to yield the following data:
For each of these vessels, compute the compliance and the incremental
elastic modulus at the mean pressure.
9.2 A catheter-tip balloon is designed to expand diseased regions of arteries,
which have a normal (nondiseased) internal diameter of 6 mm. The balloon
has a wall thickness of 0.1 mm and becomes fully distended at 6 mm. It can
withstand up to 6 atm of internal pressure. If the artery can only tolerate a
20% increase over its normal diameter before rupturing, what should the
elastic modulus of the balloon material be to provide full dilatation at
maximum pressure? Is this a maximum or minimum value?
7328_C010.fm Page 343 Monday, October 16, 2006 2:27 PM
10
Fluid Dynamic Measurement Techiques
CONTENTS
10.1 Introduction ..............................................................................................344
10.2 Pressure Measurement............................................................................344
10.3 Blood Flow Measurement ......................................................................347
10.4 Impedance Measurements......................................................................352
10.5 Flow Visualization ...................................................................................356
10.6 Ultrasound Doppler Velocimetry..........................................................359
10.7 Laser Doppler Velocimetry ....................................................................374
10.7.1 General Features ........................................................................375
10.7.2 Probe Volume Specifications ....................................................376
10.7.2.1 Calculation of Probe Volume Dimensions ............376
10.7.2.2 Fringe Patterns...........................................................377
10.7.3 Photodetectors ............................................................................379
10.7.4 Signal Processing .......................................................................379
10.7.5 Phase Window Averaging of LDV Data in Pulsatile
Flow..............................................................................................380
10.8 Magnetic Resonance Imaging and Velocity Mapping
Techniques.................................................................................................381
10.8.1 Slice Excitation ...........................................................................383
10.8.2 Spatial Encoding ........................................................................384
10.8.3 Imaging Procedure and Pulse Sequences ..............................385
10.8.3.1 Spin-Echo....................................................................386
10.8.3.2 Gradient-Echo ............................................................387
10.8.4 Magnetic Resonance Phase Velocity Mapping .....................388
10.9 Computational Fluid Dynamics ............................................................390
10.9.1 Governing Equations ................................................................390
10.9.2 Grid Generation .........................................................................393
10.9.3 Discretization Techniques.........................................................396
10.9.3.1 Finite Difference Method .........................................396
10.9.3.2 Temporal Integration ................................................398
10.9.4 Computational Modeling of Blood Flows .............................399
10.9.4.1 Body Motion ..............................................................399
10.9.4.2 Scale Disparity ...........................................................400
10.9.4.3 Blood Flow Properties ..............................................400
343
7328_C010.fm Page 344 Monday, October 16, 2006 2:27 PM
10.1 Introduction
We have now used models for steady (Chapter 5) and unsteady (Chapter 6)
flow through rigid or elastic tubes to understand the flow dynamics in the
blood vessels. Analysis of the governing equations resulted in expressions
for pressure pulse, flow rate, and velocity profiles. To verify the accuracy of
our models, data on these variables must be obtained from in vitro or, pre-
ferably, from in vivo measurements. Moreover, measurements of pressure,
flow rate, and resistance within specific segments of the circulatory system
are also important diagnostically to the physician. Measurement of detailed
velocity profiles and other flow parameters will also be helpful in under-
standing the factors involved in the initiation of disease processes, such as
thrombus formation and atherosclerosis (see Chapter 6).
∆R
e0 = E
R
e0 = FEQ (10.1)
ECG II
1 mv
–175 mm Hg
Strain-gauge –100 mm Hg
6 Fr catheter
–175 mm Hg
Catheter-tip –100 mm Hg
transducer
No bubble in transducer
1 second
(a)
ECG II
1 mv
–175 mm Hg
Strain-gauge –100 mm Hg
42 milliseconds
6 Fr catheter
–175 mm Hg
–100 mm Hg
Catheter-tip
transducer
Bubble in transducer
1 second
(b)
FIGURE 10.1
Comparison of pressure signals recorded with fluid-filled catheter transducer and catheter-
tipped transducer. Note the similarity of the signals between the two transducers in the absence
of air bubble in the catheter. Changes in pulse shape and a time delay are observed with the
presence of the air bubble with the fluid-filled catheter. (From Geddes, L.A. (1984) Cardiovascular
Devices and Applications, John Wiley & Sons, New York. With permission.)
7328_C010.fm Page 347 Monday, October 16, 2006 2:27 PM
(a)
(b)
FIGURE 10.2
Distorted pressure signals in a fluid-filled catheter due to catheter whip: (a) good quality
pressure recording and (b) the distorted signal due to catheter whip. (From Geddes, L. A. (1984)
Cardiovascular Devices and Applications, John Wiley & Sons, New York. With permission.)
E f = BV
where B is the flux density of the magnetic field (Webers/m2), is the spacing
between electrodes (m) and V is the mean flow velocity (m/s). Since the flow
rate through the vessel, Q (m3/s), is related to the area of cross section A
(m2) and the mean velocity by
Q = VA
A
Q= E (10.2)
Bl f
A
The ratio B
is a constant specified for a given flow meter system.
7328_C010.fm Page 348 Monday, October 16, 2006 2:27 PM
Flow meter
Range
Zero
Signal voltage
Magnet current
Electromagnet
Magnetic field
Flow
FIGURE 10.3
Schematic of an electromagnetic flow probe used for blood flow measurement.
FIGURE 10.4
Typical electromagnetic flow probes: (a) Extracorporeal. (Courtesy of Carolina Medical Electronics
Inc., King, NC.)
only for a relatively shorter time and the voltage-sensing circuits must be
gated for a short time. On the other hand, with square wave excitation, the
time varying component is effectively zero, except during the short periods
in which the signals are switched. The signal-to-noise ratio is proportional
to the peak-to-peak value of excitation, and for the same signal-to-noise ratio,
the square wave excitation requires twice as much power as the sine wave
excitation. Hence, the meters for the square wave devices are larger and
must operate at higher temperatures. Shown in Figure 10.6 is a typical flow
rate signal from an in vitro pulse duplicator distal to a heart valve. Measure-
ments were made with a Carolina square wave electromagnetic flow meter
and an in vivo metric extracorporeal flow probe. The time-averaged flow rate
(cardiac output) and other information, such as the amount of back flow,
can be determined from these curves.
Even though the electromagnetic flow meter is an ideal device for in vitro
testing, there are several drawbacks in using it for in vivo measurements.
One is that this is an invasive technique in which the vessel must be exposed
in order to fit the flow probe around the vessel. Secondly, if left in the body
for an extended length of time, problems of contact of the electrodes with
the arterial wall, such as protein or thrombus coating and subsequent dete-
rioration of the signals, may occur. Moreover, this device actually measures
the mean flow velocity and, thus, the area of the probe lumen must be accu-
rately known to determine the flow rate. The EMF theory also assumes that
the fluid has a flat velocity profile at the mean velocity magnitude. Since in
7328_C010.fm Page 350 Monday, October 16, 2006 2:27 PM
FIGURE 10.5
Typical electromagnetic flow probes: (b) Intracorporeal. (Courtesy of Carolina Medical Electronics
Inc., King, NC.)
blood vessels the lumen is not an ideal circular cross section and the flow is
asymmetric, some errors will occur in measuring the flow rate using this
technique.
A more recent flow-measuring device is the Transit Time Flow Meter.
This instrument is based on the fact that sound waves travel through a
fluid at slightly greater or lower velocities depending upon whether the
flow is forward or reverse, respectively. The transit time flow meter utilizes
ultrasonic energy in the kilohertz range (as opposed to ultrasound imaging
7328_C010.fm Page 351 Monday, October 16, 2006 2:27 PM
400
300
200
Flow rate (ml/sec)
100
–100
–200
–300
0 0.2 0.4 0.6 0.8 1
Time (sec)
FIGURE 10.6
A typical flow rate signal distal to a heart valve from an electromagnetic flow meter in a pulse
duplicator.
and Doppler devices that operate in the megahertz range), which is trans-
mitted through a vessel along the axis of the flow in alternate directions
(Figure 10.7). The device then detects small differences (in the order of
nanoseconds) in transit times between the two signals, tf and tr, as
tf =
(c + u)
tr =
(c − u)
T, R R, T
u d
θ
FIGURE 10.7
Transit-time flow probe with two ultrasonic crystals alternately transmitting and receiving
signals passing through a fluid and reflecting off of a backplate.
7328_C010.fm Page 352 Monday, October 16, 2006 2:27 PM
where
= 2d (sin θ)
c 2 ∆t
u=
(2 d cos θ)
where ∆t = tr − t f .
The volume flow rate is determined by multiplying this velocity by the
known vessel’s cross-sectional area. While this device is also invasive
(i.e., the probe must be placed around the vessel), it does not depend upon
any special features of the vessel (e.g., electrical conductivity) other than the
presence of reflectors in the form of red blood cells. Since the flow rate is
proportional to the transit time of the sound signals, it is a very linear and
easily calibrated device, which is very stable (i.e., not susceptible to other
electrical signals) and which can be left in place for long-term (i.e., days to
weeks) experiments.
∆p
R=
Q
When measuring the resistance across the systemic circulation, the mean
pressures at the aorta and vena cava are used along with the mean flow rate
(cardiac output). Using a familiar electrical analogy, the resistance in DC
circuits is given by R = E/I, where E is the voltage and I is the current in a
branch of the circuit. In AC circuits, the corresponding measure is the elec-
trical impedance. Thus, impedance to flow in the circulation under unsteady
flow can be defined as similar to resistance under steady flow conditions.
Milnor (1989) defines three different impedance relationships as
Longitudinal impedance:
− dp
ZL = dz
(10.3)
Q
7328_C010.fm Page 353 Monday, October 16, 2006 2:27 PM
Input impedance:
p
Zz = (10.4)
Q
Transverse impedance:
p
ZW = − dQ (10.5)
dz
pn
Zz n = (10.6a)
Qn
7328_C010.fm Page 354 Monday, October 16, 2006 2:27 PM
and
θn = φ n − ξn (10.6b)
where φn is the phase angle of the pressure and ξn is the phase angle of the
flow. Frequency analysis, or decomposition, is usually performed by Fourier
analysis, where the phase and amplitude of each frequency component is
obtained. A basic assumption of this technique is that the system is linear
so that there is no interaction between signals of different frequency. A
limitation of the characteristic impedance is that it cannot be directly mea-
sured in vivo. However, in larger arteries, the input impedance is essentially
independent of the frequency except for a drop in the impedance at low
frequencies. Thus, the characteristic impedance, Zo, is computed by averag-
ing the modulus of the input impedance across the frequency spectrum,
neglecting the low frequencies where the steep drop in impedance occurs.
The input impedance in the ascending aorta for a dog and a human is
given in Figure 10.8, where it is seen to fall steeply from 0 to 2 Hz and then
fluctuate with frequency thereafter. Such an input impedance spectrum usu-
ally shows a minimum of the impedance modulus between 2 and 8 Hz. The
phase angle of the input impedance is usually zero at the first minimum of
the modulus. Oscillations in the modulus of the impedance depend upon
reflections of the waves in the peripheral arteries as they encounter discon-
tinuities (i.e., branches, end-organs, etc.), and vasoconstriction will increase
1400
1.8
Modulus (103 dyn sec cm–5)
1.6 1200
1.4
1.2 300
1.0
0.8 200
0.6
0.4 100
0.2 z0
0 0
2 4 6 8 10 12 2 4 6 8 10
Frequency (Hz) Frequency (Hz)
2
Phase (rad)
Phase (rad)
0 0
–1
–2
(a) (b)
FIGURE 10.8
Input impedance in the aorta of dog (A) and human (B). (Redrawn from Milnor, W.R. (1989)
Hemodynamics, Williams and Wilkins, Baltimore, MD. With permission.)
7328_C010.fm Page 355 Monday, October 16, 2006 2:27 PM
400
200
100
F (Hz)
0
Phase (rad)
0.5 2 4 6 8 10
–0.5
FIGURE 10.9
Input impedance in the main pulmonary circulation of a dog (Redrawn from Milnor, W.R.
(1989) Hemodynamics, Williams and Wilkins. Baltimore, MD. With permission.)
the reflections and, hence, increase the oscillations in the impedance. The
ratio of Zz/Z0 and also the ratio of
Zz (max) − Zz (min)
Z0
are useful measures of the oscillations in the impedance. In this ratio, the
first maximum and first minimum of the modulus are used. Under normal
resting conditions, this ratio is about 0.6 to 0.85 in man and dogs.
The input impedance in the main pulmonary circulation is shown in
Figure 10.9. As can be observed, the shape of the profile is similar to that for
the aorta even though the magnitude is smaller. The first minimum of the
modulus occurs between 2 and 4 Hz and the next maximum occurs at 6 to
8 Hz. The characteristic impedance in pulmonary circulation is obtained
by averaging the impedance modulus at higher frequencies and it is about
190 dyne-sec/cm5 in dogs and about 23 dyne-sec/cm5 in man.
As mentioned earlier, the impedance will change due to alterations in the
vascular tone as well as changes in the elastic properties of the distal vessels.
Thus, measurements of vascular impedance can be used for diagnostic pur-
poses to determine changes in the vasculature in the distal vessels. For
example, pulmonary vascular hypertension increases the vascular imped-
ance to as much as three times the normal value and, as a consequence, the
7328_C010.fm Page 356 Monday, October 16, 2006 2:27 PM
FIGURE 10.10
Schematic of the experimental setup for flow visualization using colored dye. LoGerfo et al.
(1985) Structural details of boundary layer separation in a model human carotid bifurcation
under steady and pulsatile flow conditions, in J. Vasc. Surg., 2: 263–269.
dissolves in the fluid medium (usually an aqueous solution). The dye can then
be injected at a specific location, generally using a fine needle, and then its path
tracked using photographs or video recordings. One disadvantage of dye flow
visualization is the eventual buildup of the dye in the fluid, which then clouds
the visual field. Another is that it is not useful for either pulsatile or turbulent
conditions. Solid particles (e.g., aluminum flakes or microspheres) are often used
to obtain pathlines, especially if either the light source or the camera shutter is
strobed (Figure 10.11). The particles must be chosen so that they are good
reflectors of light, have densities similar to that of the fluid medium, and are
small in comparison to the finest flow structure being imaged.
FIGURE 10.11
An example of flow visualization using solid particles. (From Lutz, R.J. et al. (1983), Comparison
of steady and pulsatile flow in a double branching arterial model, in J. Biomech. 16: 753–766,
With permission.)
7328_C010.fm Page 358 Monday, October 16, 2006 2:27 PM
Doppler
Nitrogen laser frequency
analyzer
Pulsatile pump
Test section
35 mm
camera
Microcomputer
(a)
Video terminal
From doppler
system
FIGURE 10.12
Schematic of the experimental setup for flow visualization using a photochromic dye technique.
(From Poots, K. et al., (1986), A new pulsatile flow visualization method using a photochromatic
dye with applications to Doppler ultrasound, in Ann. Biomed. Eng. 14: 203–218, With permission.)
7328_C010.fm Page 359 Monday, October 16, 2006 2:27 PM
(a) S R
(b) S R VR
(c) S R VR
FIGURE 10.13
Doppler effect caused by a moving receiver. Atkinson, P. and Woodcock, J. (1982) Doppler
Ultrasound and Its Use in Clinical Measurement, Academic Press, London.
Nr = Ns + ∆N (10.7)
Since the number of wave peaks in a given time interval, ∆t, would be
equal to the corresponding frequency times ∆t and the additional peaks
would be the distance traveled divided by the wavelength, λ,
Vr ( ∆t)
fr ( ∆t) = fs ( ∆t) + (10.8)
λs
7328_C010.fm Page 361 Monday, October 16, 2006 2:27 PM
Therefore,
Vr
fr = fs + (10.9)
λs
Rewriting the wavelength, λs, in terms of its frequency, fs, and the speed
of sound in the medium, c (= 1540 m/s for water and 1560 m/s for blood)
c
λs = (10.10)
fs
Vr Vr
fd = fr − fs = = f (10.11)
λ s c s
The same analysis can be carried out for the case of a moving source,
yielding
V
fd = fs (10.12)
c
Both of these equations show that the amount of frequency shift caused
by the motion between the source and receiver is directly proportional to
the relative velocity between them.
In practice, most fluid velocity measurements are made using a sound
wave, which is transmitted to an object and then echoed back to the source
where it is received. Examples of this would be a particle reflector in the
fluid medium (in vitro) or a red blood cell in the blood (in vivo). In this
configuration, the frequency-shift, ∆f, becomes
2Vfs
∆f = (10.13)
c
2Vfs
∆f = cos θ (10.14)
c
7328_C010.fm Page 362 Monday, October 16, 2006 2:27 PM
To obtain the axial flow velocity, the Doppler shift equation is simply
rewritten in terms of V
∆f c
V= (10.15)
fs 2 cos θ
or
V ∆f
= (10.16)
c fs 2 cos θ
In this last expression, we can see the direct proportionality between the
ratio of reflector velocity to sound wave speed and the ratio of frequency
shift to source frequency. Also note in Equation 10.14 that a forward velocity
will produce an up-shifted frequency, while a reverse velocity will produce
a down-shifted frequency; thus, the device has directional capability.
In order to obtain the flow velocity from the Doppler shift waveform, the
received signal must be processed for its frequency content and, in particular,
for the frequency shift information. This outcome is accomplished in three
stages. The first step, called Phase Quadrature Demodulation (Figure 10.14),
extracts the frequency shift of the received signal from the frequency of the
transmitted signal by mixing the two. This process produces a combination of
signals at both summed and differenced frequencies. Those that are summed
are all in the range of fs , while those that are differenced are in the range of ∆f.
Receiving
transducer
Direct channel Quadrature channel
output output
Vf
Vr
FIGURE 10.14
Phase-quadrature demodulation. Atkinson, P. and Woodcock, J. (1982) Doppler Ultrasound and
Its Use in Clinical Measurement, Academic Press, London.
7328_C010.fm Page 363 Monday, October 16, 2006 2:27 PM
D Q D Q D Q
FIGURE 10.15
Directional processing in the time domain. Atkinson, P. and Woodcock, J. (1982) Doppler
Ultrasound and Its Use in Clinical Measurement, Academic Press, London.
By running all signals through a low-pass filter (e.g., one with a cut-off
frequency ≥2∆f ), only those signals with Doppler-shifted frequencies will
remain. The second step, called Phase Domain Demodulation (Figure 10.15),
separates out those signals that were up-shifted in frequency (i.e., forward
velocity) from those that were down-shifted in frequency (i.e., reverse velocity).
The final step involves actually quantifying the frequencies of those analog
signals. This can be done using one of two common signal-processing methods —
either zero crossing detection or Fourier transformation.
Zero-crossing detection [ZCD] or counting [ZCC] simply counts the num-
ber of times an oscillatory signal crosses the zero baseline (Figure 10.16).
Thus, a signal with high frequency has a larger count than a signal with low
frequency over a fixed period of time. This is a simple and inexpensive
approach, which can be built into any ultrasound Doppler device. However,
it doesn’t give a true mean frequency, as its output is actually proportional
to the Root Mean Square (rms) value of the signal. An alternate approach to
the ZCD/ZCC is to perform a fast Fourier transformation [FFT] on the signal
(Figure 10.17). This process converts the signal from a time domain to a
frequency domain representation in terms of its harmonics. By doing this, the
relative magnitude of each harmonic is determined and a resultant spectrum
of frequencies can be obtained (Figure 10.18). The result is that the frequency
shift (α particle velocity) and the power of the signal (α number of particle
reflectors present) can be obtained over time.
While all ultrasound Doppler velocimeters are based on the same funda-
mental principles and utilize virtually the same signal processing techniques,
7328_C010.fm Page 364 Monday, October 16, 2006 2:27 PM
Output
pulses
Set Reset S R S R S R
Time
FIGURE 10.16
SET–RESET zero-crossing counter. Atkinson, P. and Woodcock, J. (1982) Doppler Ultrasound and
Its Use in Clinical Measurement, Academic Press, London.
Internal timing
and logic unit
Synchronous
ADC
p( fi)
Input buffer Processor Display Spectrum
memory memory generator Display
fi outputs
FIGURE 10.17
The fast Fourier transformation (FFT) analyzer. Atkinson, P. and Woodcock, J. (1982) Doppler
Ultrasound and Its Use in Clinical Measurement, Academic Press, London.
7328_C010.fm Page 365 Monday, October 16, 2006 2:27 PM
Power
uency
Freq
Tim
e
FIGURE 10.18
Three-dimensional plot of time-variable spectrum. Hatte, L. and Angelsen, B. Doppler Ultrasound
in Cardiology: Physical Principles and Clinical Applications. (1982) Lea & Febiger, Phileadelphia.
Transmitting Master
amplifier oscillator
Transmitting
transducer Receiving
Demodulator
amplifier
Receiving
Ultrasonic transducer
beam
Doppler difference
signal
Moving targets
(blood) Blood vessel
FIGURE 10.19
The continuous-wave flow meter. Atkinson, P. and Woodcock, J. (1982) Doppler Ultrasound
and Its Use in Clinical Measurement, Academic Press, London.
7328_C010.fm Page 366 Monday, October 16, 2006 2:27 PM
2 –m/s
FIGURE 10.20
Continuous-wave spectrum of patient with mitral valve stenosis. This patient has a peak
mitral valve ejection velocity of 2 m/s, which is quickly and easily obtained by continuous-
wave Doppler. The velocity then can be converted into a pressure gradient of 16 mm Hg
via the simplified Bernoulli equation to provide an estimate of the severity of the stenosis.
Traditionally, such a pressure gradient would have been measured painstakingly by cathe-
terization. Hatte, L. and Angelsen, B. Doppler Ultrasound in Cardiology Physical Principles and
Clinical Applications. (1982) Lea & Febiger, Philadelphia.
the past decade. The only limitation to such a modality is the lack of range
resolution, which actually is an advantage for the assessment of valvular
stenosis and, because of the continuous nature of ultrasound emission in
this modality, there is no maximum velocity limit.
The pulse Doppler (Figure 10.21), on the other hand, uses only a single
crystal that transmits a signal for a brief time (i.e., a “burst”) and then pauses
to listen for returning echoes from previous transmissions. In this way, the
device cannot only acquire the frequency shift (i.e., velocity) information,
but it can also determine the location of the moving reflector by recording
the time required for the sound waves to travel out and back from the
reflector. This capability is known as range resolution and is very useful
when it is important to be able to detect velocities at specific locations. The
resolution of the measurement along the beam axis is dependent upon the
length of the signal burst sent out (Figure 10.22). Typically, the number of
cycles transmitted in a pulse Doppler burst, n, is of the order of four to six,
so that there is sufficient information in the echo for Doppler-shift frequen-
cies to be detected. Thus, the sample volume length along the beam axis and
the device resolution is nλ. Consequently, high spatial revolution is achieved
by using a high frequency system. Situations where this feature is useful
would be in obtaining a velocity profile across an artery or in determining
the maximum blood velocity through a stenosed heart valve.
As mentioned above, the relationship used to convert Doppler velocities
to stenotic pressure gradients is generally the simplified Bernoulli equation
p1 − p2 = 4V22 (1.55)
7328_C010.fm Page 368 Monday, October 16, 2006 2:27 PM
PRF
generator
Receiving Range
Demodulator
amplifier gate
Transducer Sample
and hold
Delay gate
generator
Selected
range
Audio
filters
Sample
volume
Output
FIGURE 10.21
Pulse-Doppler layout. Atkinson, P. and Woodcock, J. (1982) Doppler Ultrasound and Its Use in
Clinical Measurement, Academic Press, London.
Transducer
(a) Transmit
Pulse
Scatter
(c) Sample
Echo
Time
FIGURE 10.22
Pulse-Doppler basic principles. Atkinson, P. and Woodcock, J. (1982) Doppler Ultrasound and Its
Use in Clinical Measurement, Academic Press, London.
7328_C010.fm Page 369 Monday, October 16, 2006 2:27 PM
1
p1 − p2 = ρ(V22 − V12 ) (1.53)
2
where V2 is the maximal distal velocity and V1 is the proximal velocity. The
distance between points 1 and 2 is the distance over which the pressure drop
occurs.
Since the distal velocity V2 in a stenotic jet usually significantly exceeds
any proximal velocity (and this is especially true for the values squared), the
proximal velocity is neglected to enable ease of application in the form of
Equation 8.1 with continuous-wave Doppler. However, in some cases such
as combined valvular insufficiency and stenosis for the same valve, blood
cell velocities are already elevated beyond normal at a location proximal to
the stenotic valve. These situations can be identified with 2-D echocardio-
graphy and, in such cases, measurement of the proximal velocity is necessary
using pulse Doppler techniques to correct, in effect, Equation 1.55.
Another useful in vivo application of pulse Doppler velocimetry is in
determining the velocity profile across a vessel. This can be important in
evaluating complex flow patterns, such as exist in the aortic arch (Figure 10.23),
branches and bifurcations, and in key arteries, such as the coronaries and
carotids (see Chapter 6, section 6.5: Flow Through Curved Arteries and
Bifurcations). It is also helpful in pathologic situations, such as arterial
stenoses and aneurysms, where the flow patterns are greatly altered by
sudden changes in geometry. Once a velocity profile is obtained (either in
real time or in the mean), additional parameters may be derived from it,
such as the wall shear stress (including the OSI (oscillatory shear stress) and
shear stress gradients) and also local residence times.
While range resolution is a valuable feature, the act of pulsing the trans-
mitted signal imposes certain limitations on the operation of the system.
First, the maximum range of the device is directly restricted by the time
period between pulses since the distance traveled to the target is the product
c·∆t. The period between successive transmissions is determined by the
instrument’s pulse repetition frequency, PRF, and can be expressed as 1/PRF
(s/cycle). Thus, a pulse Doppler device’s maximum range, Zmax, becomes
c( ∆t) c
Zmax = = (10.17)
2 2(PRF)
7328_C010.fm Page 370 Monday, October 16, 2006 2:27 PM
4
4 5
1A 2
1A 2 5
3c 3c
1
1 3
3
2B
2B
90 ms 240 ms
(a) (b)
4 4
1A 2 5 5
1A 2
3c 3c
1 1
3 3
2B 2B
140 ms 320 ms
(c) (d)
4
5 4
1A 2 2
1A 2 5
3c
3c
1
1
3
3
2B
2B
170 ms 380 ms
(e) (f )
2
V1 4 5
50 cm.s–1 1
3
1A 3c
Vt 2B
FIGURE 10.23
Velocity profiles in the aorta of a dog using ultrasound Doppler velocity measurement tech-
nique: (a) 90 msec and (b) 320 msec after the opening of the valve. 1A, 2B, and 3C represent
profiles in perpendicular planes corresponding to the profiles shown in the plane of the figure.
(Redrawn from Farthing, S. and Peronneau, P. (1979) Flow in the thoracic aorta. Cardiovas Res.
13(11): 607–20. With permission.)
7328_C010.fm Page 371 Monday, October 16, 2006 2:27 PM
Secondly, the pulsed nature of the device also limits the amount of infor-
mation that can be obtained from a measurement since a returning echo is
only received once every 1/PRF seconds. Therefore, the frequency of a Dop-
pler-shifted signal, ∆fmax, must be within certain values in order for it to be
accurately determined. Exceeding this limit, also known as the Nyquist
sampling limit, will cause “aliasing” to occur where the frequency of the
signal appears to be lower than it really is. Since the Nyquist theory requires
a sampling rate, which is at least twice that of the signal frequency, the pulse
Doppler can only accurately collect data with frequency shifts of ∆fmax <
(PRF/2). This, in turn, imposes a limitation on the maximum velocity, Vmax,
obtainable through the Doppler-shift Equation 10.15
(PRF/2)c
Vmax < for θ = 0° (10.18)
2( fs )
c (PRF)c c2
Zmax (Vmax ) < < (10.19)
2(PRF) 4( fs ) 8( fs )
02: 53: 24 PM
V219 24
Depth = 180
ADLT HRT
PWR = 0dB
LV 49dB B/0/F
RA LA
(a) (b)
4 MAP
??
select Green
TAG
7Flow DIR
±, +, –
8 Playback
select
0 Color
ON/OFF
(c) (d)
FIGURE 10.24
Linear array transducer. Kremkau, F.W. (2002) Diagnostic Ultrasound Principles and Instruments,
6th ed., W.B. Saunders Co., Philadelphia.
Conventional
doppler signal processor
Velocity &
variance
Array calculator
Display
2-D B-mode
imager
FIGURE 10.25
Schematic of a color Doppler imaging system. Shung, K.K. (1992) Principles of Medical Imaging,
Academic Press, Inc. San Diego, CA.
In its present state, however, the full, 2-D CDFM display provides only
semiquantitative velocity information. This occurs because the system uti-
lizes a faster autocorrelation technique in place of direct spectral analysis
due to the greatly increased amount of data processing requirements of
CDFM compared to conventional Doppler. Also, pulse repetition frequencies
are set very low to avoid range ambiguity and to obtain greater depth
penetration (Equation 10.16) and, thus, Nyquist velocity limits also become
very low. With these low Nyquist velocities, aliasing frequently occurs with
physiologic flows, especially in high velocity situations, such as mitral regur-
gitation or aortic stenosis. The concept of aliasing in color flow results in a
‘‘wrap-around’’ of color to a value, which corresponds to flow in the opposite
direction. In these cases, multiple aliasing effects have restricted color image
quantification to a simple measurement of the color boundary location, since
quantitative distinction of velocities within the jet is virtually impossible for
high-velocity flows.
Fiber optic
2-D transceiver
blue and green
Table
FIGURE 10.26
Transceiver and receiver orientation around a flow model.
7328_C010.fm Page 376 Monday, October 16, 2006 2:27 PM
d
tan α = 2
(10.20)
f
Flow direction
Reference beam
d α
α
Bragg beam Probe volume
(exaggerated)
Transceiver Lens
Exit
df
dm
Entrance
lm
Particle
FIGURE 10.27
Probe volume formed by the intersection of a laser beam pair.
7328_C010.fm Page 377 Monday, October 16, 2006 2:27 PM
The dimensions of the ellipsoidal probe volume, dm and lm, formed by a pair
of intersecting beams, are given by
4λ f
dm = (10.21)
πDe E
dm
lm = (10.22)
tan α
where λ is the wavelength of the beam in air, f is the focal length of the
transceiver lens, De is the original beam diameter, and E is the beam expan-
sion ratio.
1
f s = fl + V ⋅ ( js − ji ) (10.23)
λ
Light intensity
Time
FIGURE 10.28
Schematic of a raw Doppler burst.
fd = fs − fi (10.24)
Therefore,
1
fd = V ⋅ ( js − ji )
λ
The distance between the fringes is fixed by the wavelength of the laser
beam and the angle between the intersecting beams. The light scattered by
the particle passing through the probe volume is modulated with a fre-
quency, which is related to the velocity of the particle by the equation
λ
V = fd (10.25)
2 sin α
V = fd d f (10.26)
7328_C010.fm Page 379 Monday, October 16, 2006 2:27 PM
and
λ
df = (10.27)
2 sin α
It can be seen from Equation 10.25 that the Doppler frequency is directly
proportional to the velocity. The constant of proportionality is a direct func-
tion of the wavelength of the laser and of the optical arrangement of the
experiment.
10.7.3 Photodetectors
Particles moving through the measuring volume scatter light of varying
intensity. However, the light signals are usually too weak to be analyzed
and, thus, photomultipliers (PM) are needed to amplify and convert the
input to a voltage output for analysis. A photomultiplier tube consists of a
photocathode, a multiplier chain, and an anode. When photons from the
scattered laser beam of a passing particle hit a PM tube, they are converted
into electrons by means of the photoelectric effect. These electron emissions
are amplified and accelerated by a string of successive electron absorbers
called dynodes to produce enhanced secondary emission. Finally, the anode
at the end of the tube is used to collect the resulting voltage that is large
enough to be picked up and analyzed. Some laser systems make use of
photodiodes for the signal amplification; however, the main advantage of
using photomultiplier tubes over photodiodes is the high gain and low noise
levels, which allow the system to operate with a light signal of low intensity.
The light signal picked up by the photodetectors contains the Doppler signal
as well as the frequency shift from the Bragg cell.
Voltage
Time
FIGURE 10.29
Down-mixed filtered Doppler signal.
The raw Doppler signal and a superimposed pedestal are created when-
ever a particle crosses the probe volume. The resultant signal thus consists
of (1) the true Doppler frequency of the particle, (2) the 40 MHz Bragg shift,
and (3) the frequency of the pedestal. The signal is passed through a pream-
plifier and through real-time signal analyzers (RSA). In the RSA, a high-pass
filter is applied to the signal to remove the low intensity pedestal. The signal
is then down-mixed, a process that removes the 40 MHz shift and leaves the
Doppler frequency fD. After down-mixing, the signal is passed through a
low-pass filter, which eliminates high frequency noise. The down-mixed,
filtered Doppler signal is illustrated in Figure 10.29.
Phase window
Velocity (m/s)
0 860
Cycle time (ms)
FIGURE 10.30
Phase window averaging.
follows. The mean and standard deviation of the measurements are calcu-
lated as
a n
∑ ∑ uijGij
i =1 j =1
mean = a n (10.28)
∑ ∑ Gij
i =1 j =1
a n
∑ ∑ (Uij − mean)2 Gij
i =1 j =1
S.D. = a n (10.29)
∑ ∑ Gij
i =1 j =1
Φ = γ B0 (10.30)
Precession
Bo Mz
M
M
Proton Mxy
FIGURE 10.31
The combination of the angular momentum and external magnetic field results in a complex
motion called precession.
7328_C010.fm Page 383 Monday, October 16, 2006 2:27 PM
magnetic field (Tesla). Because hydrogen with a single proton in its nucleus
is most commonly used as the target atom in clinical MRI, the term “nucleus”
will be replaced hereafter by “proton.”
Of the two possible orientations for the proton’s magnetic moment vector,
the parallel orientation prevails slightly over the antiparallel. It is this net
magnetic moment of the whole mass of protons directed parallel to the
magnetic moment of the static magnetic field that is actually responsible for
detecting a signal at the end of the acquisition procedure.
An advantage of MRI compared to other imaging techniques is that any
region in the body at any orientation can be imaged without limitations
related to acoustic windows or to the depth of the region of interest (in
contrast to ultrasound, for example). The imaging procedure consists of
(1) slice selection, or excitation, (2) encoding in the phase and frequency direc-
tions, or spatial encoding, (3) signal read-out, and (4) image reconstruction.
Excitation
z
Bo
y
Mz
B1
FIGURE 10.32
Proton excitation: By applying an RF pulse, B1, the magnetization vector is tipped from the
z-axis towards the x–y plane.
Bo z
y
Mz
FID
FIGURE 10.33
Free Induction Decay: After termination of the RF pulse, the excited protons emit their excessive
energy while returning to their original condition. The magnetization vector returns to the
z direction alignment. It is during this period that the image data is actually acquired.
to the phase of this precession from row to row. After phase encoding,
another magnetic field gradient is applied in the frequency encoding
direction. This creates a change in the precession frequency along each
row. Since each row already had an identity in space through the differ-
entiation in the phase of precession, the second encoding causes an iden-
tification of each separate voxel with a unique frequency and phase. This
is also the stage when the receiving coil reads the signal. This signal fills
the frequency k-space and, through 2-D Fourier transform, the final image
is reconstructed.
10.8.3.1 Spin-Echo
In this imaging modality, a 90° RF pulse is used to tip the net magnetization
vector M by 90°, moving it completely into the x–y plane (Figure 10.34).
Then, free induction decay occurs while the individual spins begin to precess
out-of-phase or dephase due to local magnetic field inhomogeneity. Since
dephasing reduces the amplitude of the received signal, rephasing of the
spins is desirable. This can be achieved by the application of an additional
180° RF pulse that forces the spins to rephase, producing an echo. The
emitted signal is read when this echo occurs. The components of the spin-
echo sequence are shown in Figure 10.35. It is widely used clinically, espe-
cially to obtain anatomical information anywhere in the body. In the heart,
it is used mainly to observe vessel walls that appear bright compared to
blood that appears dark (also referred to as MRI Angiography).
Y
90 deg pulse
Echo
FIGURE 10.34
Spin-echo sequence: Initially, the magnetization vector (bold upward arrow) is tipped 90°. Then,
spins start to dephase until another 180° pulse rephases them and an echo is generated.
7328_C010.fm Page 387 Monday, October 16, 2006 2:27 PM
90 180 Echo
RF pulses
and echo
Slice select
direction
Phase encoding
direction
Frequency
encoding direction
(read-out)
FIGURE 10.35
Diagram of a typical spin-echo pulse sequence.
10.8.3.2 Gradient-Echo
This is the most commonly used pulse sequence for cardiovascular imaging,
especially to obtain flow information, either qualitatively or quantitatively.
In gradient-echo, the RF pulse tips the magnetization vector M to an angle
less than 90°. This provides a faster acquisition procedure, since less time
is required for relaxation, which is important for cine cardiac imaging. After
excitation, the relaxation process and the dephasing of the individual spins
begin. Application of an additional magnetic field gradient in the frequency
encoding direction causes the dephasing to accelerate. Then, another gra-
dient of opposite polarity rephases the spins and an echo is formed and
read. The great advantage of the gradient-echo pulse sequence is its speed
compared to spin-echo, since the flip angle is usually much less than 90°.
A cine acquisition throughout a specific time period then can be achieved.
This is very important when imaging structures that are in motion, such as
the heart or arteries. Gradient-echo sequence components are shown in
Figure 10.36.
7328_C010.fm Page 388 Monday, October 16, 2006 2:27 PM
RF Echo
RF pulses
and echo
Slice select
direction
Phase encoding
direction
Frequency encoding
direction (read-out)
FIGURE 10.36
Diagram of a typical gradient-echo pulse sequence.
Φ= ∫ γ G(t)[r(t)]dt (10.31)
where Φ is the phase of the signal, γ is the gyromagnetic ratio, G(t) is the
vector of the magnetic field gradient (mTesla/m), and r(t) is the position
vector of the protons (in meters). For the case of motion in only a single
7328_C010.fm Page 389 Monday, October 16, 2006 2:27 PM
Time
FIGURE 10.37
A schematic of a Bipolar Gradient, Gz (t): The total area of the two lobes is zero, resulting in a
linear relationship between signal phase and proton velocity.
direction where acceleration and higher order motion terms are neglected,
the position vector is:
r(t) = Z0 + u ⋅ t (10.32)
where Z0 is the initial position of the proton (in meters), u is the velocity of
the proton in the z direction (m/s), and t is the time of the motion (in seconds).
Combining Equation 10.31 and Equation 10.32 results in
By applying a bipolar gradient Gz(t), as shown in Figure 10.37, the first term
on the right side of Equation 10.33 becomes zero. In addition, if the particle
velocity is assumed constant, then
∫
Φ = γ u Gz (t)tdt (10.34)
As seen from Equation 10.34, a linear relationship exists between the phase
of the signal and the velocity of protons. Thus, if the phase is obtained, the
velocity can be determined by
Φ
u= (10.35)
γ G (t)tdt
∫ z
7328_C010.fm Page 390 Monday, October 16, 2006 2:27 PM
The behavior of such a fluid was described in Chapter 1 with Equation 1.1.
Blood is a fluid that does not obey the relation of Equation 1.1 and is called
non-Newtonian because its viscosity changes with the applied shear forces.
However, non-Newtonian fluid behavior is extremely complex, and the
assumption of a Newtonian fluid is commonly made in order to predict flow
dynamics using numerical simulation tools. Another simplification com-
monly used is to consider the fluid as being incompressible. In this case, the
fluid density is assumed to be constant throughout the flow field and over
time. Despite the fact that in reality all fluids are compressible to some extent,
this idealization holds in most cases.
The equations that govern the motion of Newtonian fluids are called the
Navier–Stokes equations and are given by Equation 1.43a to Equation 1.43c.
These equations, derived from the conservation laws of mass, momentum,
and energy, are the so-called continuity equation, and momentum and
energy balances. The governing equations for an incompressible isothermal
flow of a constant density fluid r can be expressed in Cartesian coordinates,
using Einstein notations, as follows.
Continuity Equation
∂ui
=0 (10.36)
∂xi
Momentum Equation
where ui is the velocity in the ith direction, xj is the jth direction coordinate,
P is the pressure divided by the density ρ, υ is the fluid kinematic viscosity
(defined as υ = µ ρ ), and t denotes time.
Since the fluid density is assumed constant, the continuity and momen-
tum equations constitute a closed system of equations, i.e., there are two
equations for the two unknowns — velocity and pressure. In compressible
flow computations, the energy equation is decoupled from the momentum
and continuity equation and discarded. Therefore, the continuity and
momentum equations are the only equations solved to obtain velocity and
pressure fields. The energy equations are only solved if heat transfer is of
interest.
The remaining equations can be written in their noncompact forms as
Continuity Equation
∂u ∂v ∂w
+ + =0 (10.38)
∂x ∂y ∂z
7328_C010.fm Page 392 Monday, October 16, 2006 2:27 PM
Momentum Equations
∂u ∂ 2 P ∂u ∂ ∂u ∂ ∂u
+ u + − υ + uv − υ + uw − υ = 0
∂t ∂x ρ ∂x ∂y ∂y ∂z ∂z
(10.39a)
∂v ∂ ∂v ∂ P ∂v ∂ ∂v
+ uv − υ + v 2 + − υ + vw − υ = 0
∂t ∂x ∂x ∂y ρ ∂y ∂z ∂z
(10.39b)
∂w ∂ ∂w ∂ ∂w ∂ 2 P ∂w
+ uw − υ + vw − υ + w + − υ =0
∂t ∂x ∂x ∂y
∂y ∂z ρ ∂z
(10.39c)
where x, y, and z are the Cartesian coordinates; u, v, and w are the respective
velocity components; t is time; υ is the fluid kinematic viscosity; P is the
pressure; and ρ is the fluid density.
To simplify programming and ease algorithm implementation, the
Navier–Stokes equations are commonly written in vector form. To do so, all
governing equations are regrouped into a single equation for a vector con-
taining all flow variables (i.e., density, velocity components, pressure, etc.).
For instance, the incompressible Newtonian Navier–Stokes equations in vec-
tor form are given by
Γ = diag(0, 1, 1, 1) (10.41)
P
u
Q= (10.42)
v
w
7328_C010.fm Page 393 Monday, October 16, 2006 2:27 PM
and
u v w
u2 + P uv uw
E = ρ F=
v 2 + P G = vw
(10.43)
uv ρ P
w 2 +
uw vw ρ
0 0 0
∂
u
∂u ∂y ∂u
∂x ∂z
Eυ = ∂v Fυ = ∂v Gυ = ∂v (10.44)
∂x ∂y ∂z
∂w ∂w ∂w
∂x ∂z
∂
y
θ
y
x
(a) (b)
FIGURE 10.38
Grid generation for flow through a model of a straight blood vessel.
7328_C010.fm Page 394 Monday, October 16, 2006 2:27 PM
Geometry
FIGURE 10.39
Generalized coordinate grid for an aortic arch model.
will not guarantee the presence of grid nodes located near the surface of the
blood vessel. This will make the enforcement of the physical boundary
conditions challenging. For this rather simple problem, a more suitable coor-
dinate system is the cylindrical polar system shown in Figure 10.38b. The
coordinate system consists of two coordinate lines: (1) lines of constant θ
and perpendicular to the surface body, and (2) lines of constant r and parallel
to the surface body. As shown in Figure 10.38b, a coordinate line coincides
with the surface body and, therefore, the boundary conditions can be easily
applied to all nodes located along that line. Such a coordinate system is
called a body-fitted coordinate system.
For complex body shapes, the choice of a body-fitted coordinate system
is not always straightforward. For instance, consider the aortic arch geometry
depicted in Figure 10.39. The body-fitted coordinates do not resemble any
commonly used coordinate systems (i.e., cylindrical, spherical, etc.). Such a
system is classified as a generalized nonorthogonal coordinate system. With this
coordinate system, body-fitted grids can be generated for any geometrical
body. However, it is important to emphasize that the handiness of the gen-
eralized body-fitted coordinate system is thwarted by the complexity of the
governing differential equations once they are transformed to generalized
coordinates.
With the use of body-fitted coordinates, flows around and through arbi-
trarily complex geometries can be treated as easily as flows in simple geom-
etries. Transforming the complex physical domain, defined by the generalized
coordinates, into a simple computational domain does this. This computational
domain, also called the transformed domain, is essentially an equally spaced
Cartesian grid — it is a rectangular domain for 2-D problems or a parallel-
epiped for 3-D problems. The transformation from the physical flow domain
to the computational domain is depicted in Figure 10.40. In a 2-D case, the
7328_C010.fm Page 395 Monday, October 16, 2006 2:27 PM
j = jmax
Transformation
η (x, y, z) → (ξ, η, ζ)
ξ ∆η
η ∆ξ
y j=1
i=1 i = imax
x ξ
FIGURE 10.40
Transformation from the physical flow domain to the computational domain.
physical domain is described using Cartesian coordinates (x,y) and the com-
putational domain using generalized coordinates (ξ,η). The transformation
(x,h,z) (x,y,z) can be described as
1 ∂Q ∂E ∂ F ∂Eυ ∂ Fυ
Γ + + − − =0 (10.47)
J ∂t ∂ξ ∂η ∂ξ ∂η
ξ x ηx
J = det = ξ x η y − ξ y ηx (10.48)
ξ y ηy
1 1
E = ( ξ x E + ξ y F ), F = ( η x E + η y F ) (10.49)
J J
1 1
Eυ = (ξ x Eυ + ξ y Fυ ) , Fυ = ( ηx Eυ + ηy Fυ ) (10.50)
J J
7328_C010.fm Page 396 Monday, October 16, 2006 2:27 PM
∂u lim u( x + ∆x) − u( x)
= (10.51)
∂x ∆x → 0 ∆x
∂u u( x + ∆x) − u( x)
≈ (10.52)
∂x ∆x
7328_C010.fm Page 397 Monday, October 16, 2006 2:27 PM
∂u ∆x 2 ∂2 u ∆x 3 ∂3u ∆x 4 ∂ 4u
u( x + ∆x) = u( x) + ∆x + + + +… (10.53)
∂x 2 ∂x 2 6 ∂x 3 24 ∂x 4
Substituting Equation 10.53 into Equation 10.52, the first order derivative
can be expressed as
u( x + ∆x) − u( x) ∂u ∆x ∂2 u ∆x 2 ∂3u ∆x 3 ∂ 4u
= + T with T = + + +…
∆x ∂x 2 ∂x 2 6 ∂x 3 24 ∂x 4
(10.54)
u( x + ∆x) − u( x) ∂u
= + θ( ∆x) (10.55)
∆x ∂x
It is important to note that as the grid spacing approaches zero all the terms
in the truncation error series approach zero.
Using Taylor series expansions or polynomial fitting, one can derive dis-
crete approximations of any order of accuracy for continuous derivatives. As
an example, several discretization schemes for the first order derivative of
the function u using the notation given in Figure 10.41 are listed in Table 10.1.
The very simple implementation of the finite difference method explains
its popularity. Nonetheless, one of the main disadvantages of this method
is that the computational grid has to be sufficiently smooth in order to
maintain high order of accuracy on nonuniform meshes. Alternative discret-
ization techniques, such as Finite Volume and Finite Elements, have been
developed to overcome the drawbacks of the Finite Difference method.
Despite its limitations, the finite difference method is still the most popular
one currently used.
∆x
FIGURE 10.41
Discretized domain.
7328_C010.fm Page 398 Monday, October 16, 2006 2:27 PM
TABLE 10.1
Examples of Different Discretization Scheme of a First-Order Derivative
Schemes Formulas
Central second-order accurate formula
∂u ui +1 − ui −1
= + ϑ( ∆x 2 )
∂x 2∆x
First-order accurate, one-sided Forward
∂u ui +1 − ui
formula = + ϑ( ∆x)
∂x ∆x
Backward
∂u ui − ui −1
= + ϑ( ∆x)
∂x ∆x
Second-order one-sided Forward
∂u −3ui + 4ui +1 − ui + 2
formula = + ϑ( ∆x 2 )
∂x 2 ∆x
Backward
∂u 3ui − 4ui −1 + ui − 2
= + ϑ( ∆x 2 )
∂x 2 ∆x
∂u ∂u ∂2 u
+c −υ 2 =0 (10.56)
∂t ∂x ∂x
∂u i u −u u −2 u i + u i −1
+ c i +1 i −1 − υ i +1 =0 (10.57)
∂t 2 ∆x ∆x 2
It can be seen that if the solution is known at the time level n, the solution
at the time level n+1 can be simply evaluated using Equation 10.59. Such a
method is called explicit time-marching scheme. However, if one wants to
evaluate the spatial derivative at the time level n + 1, then Equation 10.57
becomes
uni +1 − uni uni ++11 − uni −+11 uni ++11 − 2 uni +1 + uni −+11
+c −υ =0 (10.60)
∆t 2 ∆x ∆x 2
TABLE 10.2
Examples of Blood Vessel Parameters in Dogs and Humans
Canine Parameters [6] Human Parameters [25]
Mean Wall Mean Mean Wall
Mean Diameter Thickness Peak Reynolds Diameter Thickness
(mm) (mm) Number (mm) (mm)
Artery 4 (F) 0.4(F) 1000(F) 25 (AA) 1
5 (C) 0.3(C) 3000 (MP)
17 (MP) 0.2 (MP) 4500 (AA)
Arteriole 0.05 0.02 0.09 0.03 0.006
Capillary 0.006 0.001 0.001 0.008 0.0005
Venule 0.004 0.002 0.035 0.02 0.001
Vein 10 (IVC) 0.2 (IVC) 700 (IVC) 5 0.5
Note: MP – main pulmonary, F – femoral, IVC – inferior vena cava, C – carotid, AA –
ascending aorta.
the veins, the pulsatility induced by the beating of the heart cannot be
neglected in most cases as it greatly affects the flow field. For instance,
experimental studies clearly indicate that the flow field through a fully open,
bileaflet mechanical heart valve under steady conditions is considerably
different from the flow field during the peak forward flow phase. Addition-
ally, as can be seen from Table 10.2, blood flow through the cardiovascular
system is subjected to a wide range of Reynolds numbers, indicating that
laminar, transitional, and even turbulent flows may be encountered. For
instance, blood flow through mechanical prosthetic heart valves can become
turbulent at peak forward flow rate and such turbulence is thought to play
an important role in clinical phenomena, such as platelet activation and
thrombus formation and embolization. Appropriate turbulence models are,
therefore, required to numerically evaluate such flow fields. On the other
hand, flow through capillaries is nonhomogeneous, unsteady, and is catego-
rized by a very small Reynolds number (<< 1). One of the main challenges
encountered in simulating the flow in the capillary circulation is the contin-
uous variation of blood flow in the capillaries. Laminarization and turbu-
lence are important flow phenomena, but have not yet been modeled
together. Numerical methods capable of allowing laminar and turbulent flow
to coexist in the same domain are needed. At present, variants of large-eddy
simulation (LES) techniques appear to be the most promising for predicting
the turbulence and relaminarization that occur during the cardiac cycle.
10.9.4.5 Symmetry
True geometries of blood vessels or of heart valves often show one or more
axes of symmetry. For simplicity, one may want to assume symmetry of the
flow and simulate only parts of the flow fields. However, previous studies
have shown that symmetry of the geometry does not always imply symmetry
of the flow fields. For instance, simulated flow fields through a bileaflet
mechanical heart valve showed that, despite the symmetry of the geometry,
symmetry of the flow could not be assumed and the entire cross section had
to be modeled, even at a low Reynolds number.
10.9.4.7 Validation
In vitro and in vivo studies have shown the complexity of flow in the cardio-
vascular system. Therefore, the capability of CFD codes to simulate such
complex flow needs to be tested rigorously by comparing their results with
detailed experimental measurements. There are several experimental studies
reported in the literature but, so far, only a few have been carried out with
CFD model development and validation in mind. For example, no experi-
ments have been carried out focusing on the details of the flow in realistic
mechanical heart valve geometries for Reynolds numbers over which lami-
nar flow conditions prevail. Even though such Reynolds numbers would not
be relevant from a physiological standpoint, the availability of a comprehen-
sive data set for laminar flow would greatly facilitate the validation of the
numerical method free of any turbulence modeling-related uncertainties.
Detailed experiments are also needed for Reynolds numbers in the physio-
logical range, which, unlike most previous work, should be driven by specific
needs for fine-tuning and validating advanced turbulence models. Ideally,
all numerical methods developed should be extensively validated, both qual-
itatively and quantitatively, using sophisticated experimental measurement
techniques.
10mm
FIGURE 10.42
Example of generation of computational and experimental models from raw MR images.
interpolated (Figure 10.42b). The blood vessel of interest, in this case, the
Total Cavo-Pulmonary Connection, is then segmented from the rest of the
image (Figure 10.42c). The obtained geometry can then be used to generate
both a computational mesh (Figure 10.42d) and an experimental Rapid Pro-
totype (Figure 10.42e) of the connection. Because this procedure can provide
identical experimental and computational geometries, it allows for an accu-
rate validation of the numerical tool to be made.
Previous blood flow studies have shown that extended exposure of blood
cells to high levels of shear stress leads to hemolysis, platelet activation,
and initiation of the coagulation cascade. Furthermore, aggregation of lysed
or activated blood cells and thrombus formation occur in regions of low flow
where the blood cell residence time is sufficiently elevated. Additionally,
data indicate that platelet activation depends on both shear stress levels
and exposure time while thrombus formation is a function of the residence
time. It is, therefore, essential to also assess these parameters in order to
7328_C010.fm Page 404 Monday, October 16, 2006 2:27 PM
Bibliography
Asakura, T., and Karino, T. (1990) Flow patterns and spatial distributions of
atherosclerotic lesions in human coronary arteries. Circulation Research
66:1054–1066.
Atkinson, P., and Woodcock, J. (1982) Doppler Ultrasound and Its Use in Clinical
Measurement, Academic Press, London.
Ballyk, P.D., Walsh, C., Butany, J., and Ojha, M. Compliance mismatch may promote
graft-artery intimal hyperplasia by altering suture-line stresses. J. Biomechan-
ics 1998. 31: 229–37.
Becker, G.J. Intravascular stents: General principles and status of lower-extremity
arterial applications. Circulation 1991. 83(Suppl I): I–122–I–136.
Bellhouse, B.J., and Bellhouse, F.H. Mechanism of closure of the aortic valve. Nature,
1968. 217(123): 86–7.
Bergel, D.H. (1961a) The static elastic properties of the arterial wall, J. Physiol., 156:
445–457.
Bergel, D.H. (1961b) The dynamic elastic properties of the arterial wall. J. Physiol.,
156: 458–469.
Berger, S.A., and Jou, L.A., Flows in stenotic vessels. Annual Reviews of Fluid
Mechanics, 2000. 32: 347–382.
Brown, M.M. Balloon angioplasty for extracranial carotid disease in Advances in
Vascular Surgery Vol. 4, 1996, A.D. Whittemore, Ed., Elsevier, Philadelphia.
Burton, A.C. (1954) Yearbook Publishers, Elsevier, New York.
Burton, A.C. (1971) Physiology and Biophysics of the Circulation. Year Book Medical
Publishers, Chicago.
Caro, C.G., Fitz-Gerald, J. M., and Schroter, R. C. (1971) Atheroma and arterial wall
shear. Observation, correlation and proposal of a shear dependent mass trans-
fer mechanism for atherogenesis. Proc. Royal Soc. London B. 177: 109–159.
Caro, C.G., Pedley, T.J., Schroter, R.C., and Seed, W.A. (1978) The Mechanics of the
Circulation, Oxford Medical Publications, Oxford.
Chandran, K.B. (1993) Flow dynamics in the human aorta. J. Biomech. Eng. 115: 611–615.
Cho, K.J., and Williams, D.M., (1991) Current Therapy in Vascular Surgery, 2nd ed.,
Ernst, C.B., and Stanley, J.C. Eds., B.C. Decker, Mosby Yearbook, Elsevier,
Philadelphia
Cooney, D.O. (1976) Biomedical Engineering Principles: An Introduction to Fluid, Heat
and Mass Transport Processes, Marcel Dekker, New York.
Copley, A.L., and Stainsby, G. (Eds.) (1960) Flow Properties of Blood, Pergamon Press,
Oxford
Dean, W.R. (1927) Note on the motion of fluid in a curved pipe. Phil. Mag. 20:
208–223.
DeArmond, S.J., Fusco, M.M., and Dewey, M.M. (1974) Structure of the Human Brain,
Oxford University Press, London.
405
7328_C011.fm Page 406 Monday, October 16, 2006 2:27 PM
DeBakey, M.E., Lawrie, G.M., and Glaeser, D.H. Patterns of atherosclerosis and
their surgical significance. Ann. Surg., 1985. 201(2): 115–31.
Didisheim, P. and Watson, J.T. Cardiovascular Applications in Biomaterials Science:
An Introduction to Materials in Medicine, Ratner, B.D., Ed., 1996, Elsevier,
Philadelphia
Dinsdale, A., and Moore, F. (1962) Viscosity and Its Measurement, Reinhold Publishing,
New York.
Engeler, C.E., Edlicky, J., Letourneau, J.W., Castaneda-Auniga, J.G., Hunger, W.R.,
and Amplat, D. W. Intravascular sonography in the detection of arterioscle-
rosis and evaluation vascular interventional procedures. AJR, 1991, 156:
1087–1090.
Ernst, C.B., and Stanley, J.C. (1991) Current Therapy in Vascular Surgery, 2nd ed. B.C.
Decker, Philadelphia.
Farthing, S., and Peronneau, P. Flow in the thoracic aorta. Cardiovasc. Res., 1979.
13(11): 607–20.
Fields C, Cassano A, Allen C., Meyer A.C., Rittgers, S.E., Makhoul, R.G., and Bowlin
G.L. The persistence of electrostatically seeded endothelial cells lining a small
diameter expanded polytetrafluoroethylene vascular graft. J. Biomaterials
Applications 16: 157–173, 2001
Frazin, L.J., Lanza, G., Vonesh, M., Khasho, F., Spitzzeri, C., McGee, S., Mehlman,
D., Chandran, K.B., Talano, J., and McPherson, D.D. (1990) Functional chiral
asymmetry in descending thoracic aorta. Circulation 82: 1985–1994.
Fry, D.L. (1968) Acute vascular endothelial changes associated with increased blood
velocity gradients. Circulation Research 22: 165–197.
Fung, Y.C. (1984) Biodynamics: Circulation, Springer-Verlag, New York.
Geddes, L.A. (1984) Cardiovascular Devices and Applications, John Wiley & Sons,
New York.
Ghista, D.N., and Rao, A.P. Structural mechanics of the mitral valve: stresses sus-
tained by the valve; non-traumatic determination of the stiffness of the in
vivo valve. J. Biomech., 1972. 5(3): 295–307.
Gotto, A.M. Jr., Robertson, A.L. Jr., Epstein, S.E., DeBakey, M.E., and McCollum,
C.H. III. (1977) Atherosclerosis, Upjohn, Pfizer, Inc., New York.
Gupta, S.K., (1993), Vascular Graft Monitoring, R.G. Landes Co., Austin, Texas
Guyton, A.C., and Hall, J.E. (2000) A Textbook of Medical Physiology, 10th ed., W.B.
Saunders, Philadelphia.
Gwertzman, G. Clinically tested intravascular stents and endovascular stent grafts
in Endovascular Stented Grafts for the Treatment of Vascular Diseases, Marin, M.L.,
Veith F.J., and Levine, B.A. Eds., 1995, R.G. Landes Co., Austin, Texas
Hatle, L., and Angelson, B. (1985) Doppler Ultrasound in Cardiology: Physical Principles
and Clinical Applications, Lea and Febiger, Philadelphia.
He, X., and Ku, D.N. Pulsatile flow in the human left coronary artery bifurcation:
average conditions. J. Biomech. Eng., 1996. 118(1): 74–82.
Helmlinger, G., Geiger, R.V., Schreck, S., and Nerem, R.M. Effects of pulsatile
flow in cultured vascular endothelial cell morphology. J. Biomech. Engr., 1991.
113: 123–131.
Hibbeler, R.C. (2003) Mechanical of Materials, 5th ed., Pearson Prentice Hall, Upper
Saddle River, NJ.
Holzapfel, G.A., Gasser, T.C., and Ogden, R. W. (2000) A new constitutive frame-
work for arterial wall mechanics and a comparative study of material models.
J. Elasticity 61: 1–48.
7328_C011.fm Page 407 Monday, October 16, 2006 2:27 PM
Bibliography 407
How, T.V., Rowe, C.S., Gilling-Smith, G.L., and Harris, P. Interposition vein
cuff anastomosis alters wall shear stress distribution in the recipient
artery, J. Vascular Surgery 2000. 31 : 1008–17.
Ionescu, M.I., and Ross, D.N. Heart-valve replacement with autologous fascia lata.
Lancet, 1969. 2(7616): 335–8.
Jacob, S.W., and Francone, C.A. (1974) Structure and Function in Man, 3rd ed., W.B.
Saunders Company, Philadelphia.
Kaiser, G.A., Hancock, W.D. Lukban, S.B., and Litwak, R.S. Clinical use of a new
design stented xenograft heart valve prosthesis. Surg Forum, 1969. 20: 137–8.
Keynton, R.S., Rittgers, S.E., and Shu, M.C.S. The effect of angle and flow rate upon
hemodynamics in distal vascular graft anastomosis: An in vitro model study.
J. Biomech. Eng,. 1991. 113: 458–463.
Keynton, R.S., Evancho, M.M., Sims, R.L., and Rittgers, S.E. The effect of graft
caliber upon wall shear within in vivo distal vascular anastomosis. J. Biomech.
Engr., 1999. 121: 79–88.
Keynton, R.S., Evancho, M.M., Sims, R.L., Rodway, N.V., Gobin, A., and Rittgers,
S.E. Intimal hyperplasia and wall shear in arterial bypass graft distal anas-
tomoses: An in vivo model study, J. Biomech. Engr. 2001; 123: 464–73.
Kim, W.Y., Bisgaard, T., Nielsen, S.L., Poulsen, J.K., Pedersen, E.M., Hasenkam,
J.M., and Yoganathan, A.P. Two-dimensional mitral flow velocity profiles in
pig models using epicardial Doppler echocardiography. J. Am. Coll. Cardiol.,
1994. 24(2): 532–45.
Kleinstreuer, C., Buchanan, Jr., J.R., Lei, M., and Truskey, G.A. Computational
analysis of particle-hemodynamics and prediction of the onset of arterial
diseases, in Cardiovascular Techniques, Leondes, C. Ed. 2001, CRC Press: Boca
Raton. I–1 to I–69.
Ku, D.N., Giddens, D.P., Zarins, C.K., and Glagov, S. Pulsatile flow and atheroscle-
rosis in the human carotid bifurcation. Positive correlation between plaque
location and low oscillating shear stress. Arteriosclerosis, 1985. 5(3): 293–302.
Ku, D. N. (1997) Blood flow in arteries, Ann. Rev. Fluid Mech. 29: 399–434.
Kunzelman, K.S., Cochran, R.P., Chuong, C., Ring, W.S., Verrier, E.D., and Eberhart, R.D.
Finite element analysis of the mitral valve. J. Heart Valve Dis. 1993. 2(3): 326–40.
Langille, B.L., and O’Donnell, F. (1986) Reductions in arterial diameter produced by
chronic decreases in blood flow are endothelium-dependent, Science 231: 405–407.
Leo, H.L., He, Z., Ellis, J.T., and Yoganathan, A.P. Microflow fields in the hinge
region of the CarboMedics bileaflet mechanical heart valve design. J. Thorac.
Cardiovasc Surg., 2002. 124(3): 561–74.
Li, J. K-J. (1987) Arterial System Dynamics, New York University Press, New York.
Li X.M, Evaluation of hemodynamic factors at the distal end-to-side anastomosis
of a bypass graft with different POS:DOS ratios, Ph.D. Dissertation, Dept. of
Biomedical Engineering, The University of Akron, 1998
Lim, K.O., and Boughner, D.R. Mechanical properties of human mitral valve
chordae tendineae: variation with size and strain rate. Can. J. Physiol Pharma-
col., 1975. 53(3): 330–9.
LoGerfo, F.W., Nowak, M.D., and Quist, W.C. Structural details of boundary layer
separation in a model human carotid bifurcation under steady and pulsatile
flow conditions. J. Vasc. Surg. 1985. 2: 263–269.
Lutz, R. J., Hsu, L., Menawat, A., Zrubek, J., and Edwards, K. Comparison of steady
and pulsatile flow in a double branching arterial model. J. Biomech. 1983. 16:
753–766.
7328_C011.fm Page 408 Monday, October 16, 2006 2:27 PM
Malek, A.M., Alper, S.L., and Izumo, S. (1999) Hemodynamic shear stress and its
role in atherosclerosis. JAMA. 282(21): 2035–42.
Marcus, M.L. (1983) The Coronary Circulation in Health and Disease, McGraw-Hill,
New York.
McDonald, D.A. (1974) Blood Flow in Arteries. Williams and Wilkins, Baltimore.
Merrill, E.W., Margetts, W.G., Cokelet, G.R., and Gilliland, E.R. (1965) The Casson
equation and rheology of blood near zero shear. In Proceedings of the Fourth
International Congress on Rheology (Copley, A.L. Ed.), Wiley-Interscience, NY,
135–143.
Milnor, W.R. (1989) Hemodynamics, 2nd ed., Williams and Wilkins, Baltimore Morgan,
G.W., and Kiely, J.P. (1954) Wave propagation in a viscous liquid contained in
a flexible fluid, J. Acoust. Soc. Am., 26: 323–328.
Moore, J.E., Jr., and Berry, J.L. (2002) Fluid and solid mechanical implications of
vascular stenting. Ann. Biomed. Eng. 30: 498–508.
Nanda, N.C. (1985) Doppler Echocardiography, Igaku-Shoin, New York.
Nichols, W.W., and O’Rourke, M.F. (1990) McDonald’s Blood Flow in Arteries, 3rd Ed.
Lea and Febiger, Philadelphia.
Noordergraaf, A. (1978) Circulatory Systems Dynamics, Academic Press, NY.
Ojha, M., Ethier, C.R., Johnson, K.W., and Cobbold, R.S.C. Influence of angle on
wall shear stress distribution for an end-to-side anastomosis model, J. Vasc.
Surg. 1990. 12: 747–53.
O'Rourke, M.F., Vascular impedance in studies of arterial and cardiac function.
Physiol Rev, 1982. 62(2): 570–623.
Papadaki, M., and Eskin, S.G. Effects of fluid shear stress on gene regulation of
vascular cells. Biotechnol Prog. 1997. 13(3): 209–21.
Paulsen, P.K., and Hasenkam, J.M. Three-dimensional visualization of velocity
profiles in the ascending aorta in dogs, measured with a hot-film anemometer.
J. Biomech., 1983. 16(3): 201–10.
Peterson, L.H., Jensen, R.E., and Parnell, J. (1960) Mechanical properties of arteries
in vivo. Circulation Research 8: 622–639.
Poots, K., Cobbold, R.S.C., Johnson, K.W., Appugliese, R., Kassam, M., Zuech, P.E.,
and Hummel, R.L. A new pulsatile flow visualization method using a pho-
tochromic dye with application to Doppler ultrasound. Ann. Biomedical En-
gineering, 1986. 14: 203–218.
Reul, H., Talukder, N., and Muller, E.W. Fluid mechanics of the natural mitral valve.
J. Biomech., 1981. 14(5): 361–72.
Riley, W.F., Sturges, L.D., and Morris, D.H. (1998) Mechanics of Materials, 5th ed., John
Wiley & Sons, New York.
Rittgers, S.E., and Shu, M.C.S. Doppler color-flow images from a stenosed arterial
model: Interpretation of flow patterns. J. Vasc. Surg. 1990. 12: 511–22,
Rushmer, R.F. (1976) Cardiovascular Dynamics, W. B. Saunders, Philadelphia.
Salisbury, P.F., Cross, C.E., and Rieben, P.A. Chorda Tendinea Tension. Am. J. Phys-
iol., 1963. 205: 385–92.
Sawyer, P.N., (1987) Modern Vascular Grafts, McGraw-Hill Book Co., New York.
Scott, M.J., and Vesely, I. Morphology of porcine aortic valve cusp elastin. J. Heart
Valve Dis., 1996. 5(5): 464–71.
Schlant, R C., and Alexander, R.W., Eds. (1994) Hurst’s The Heart, Arteries and Veins,
8th ed., McGraw-Hill, Health Professions Division, New York.
Seed, W. A., and Wood, N. B. (1971) Velocity patterns in the aorta. Cardiovascular
Research 5: 319–330.
7328_C011.fm Page 409 Monday, October 16, 2006 2:27 PM
Bibliography 409
Senning, A., Fascia lata replacement of aortic valves. J. Thorac. Cardiovasc Surg.,
1967. 54(4): 465–70.
Senning, A., Results following aortic valve replacement with autologous Fascia lata.
Thoraxchir. Vask. Chir., 1971. 19(4): 304–8.
Silverthorn, D.U. (2001) Human Physiology. An Integrated Approach, 2nd ed., Prentice
Hall, Upper Saddle River, NJ.
Sloth, E., Houlind, K.C., Oyre, S., Kim, W.Y., Pedersen, E.M., Jorgensen, H.S., and
Hasenkam, J.M. Three-dimensional visualization of velocity profiles in the
human main pulmonary artery with magnetic resonance phase-velocity map-
ping. Am. Heart. J., 1994. 128: 1130–8.
Texon, M. (1957) A hemodynamic concept of atherosclerosis, with particular refer-
ence to coronary occlusion. Arch. Int. Med. 99: 418–427.
Thubrikar, M., Piepgrass, W.C., Shaner, T.W., and Nolan, S.P. The design of the
normal aortic valve. Am. J. Physiol., 1981. 241(6): H795–801.
Travis, B.R., Leo, H.L., Shah, P.A., Frakes, D.H., and Yoganathan, A.P. An analysis
of turbulent shear stresses in leakage flow through a bileaflet mechanical
prostheses. J. Biomech. Eng., 2002. 124(2): 155–65.
Tsakiris, A.G., Von Bernuth, G., Rastelli, G.C., Bourgeois, M.J., Titus, J.L., and Wood,
E.H. Size and motion of the mitral valve annulus in anesthetized intact dogs.
J. Appl. Physiol., 1971. 30(5): 611–8.
Vesely, I., and Noseworthy, R. Micromechanics of the fibrosa and the ventricularis
in aortic valve leaflets. J. Biomech., 1992. 25(1): 101–13.
Vesely, I., Macris, N., Dunmore, P.J., and Boughner, D. The distribution and mor-
phology of aortic valve cusp lipids. J. Heart Valve Dis., 1994. 3(4): 451–6.
Villemarette, P.A., Kornick, A.L., Rosenberg, D.M., Holmes, J., and Hower, J.F. Jr.
Use of color flow Doppler to evaluate vascular access graft function, J. Vasc.
Tech., 1989. 13: 164–170.
White, R.A., and Cavaye, D.M. (1993), in A Text and Atlas of Arterial Imaging —
Modern and Developing Technology, Cavaye, D.M., and White, R.A., Eds., W.
W. Lippincott, Philadelphia.
Whitmore, R.L. (1968) Rheology of Circulation, Pergamon Press, New York.
Womersley, J.R. (1955a) Method for the calculation of velocity, rate of flow and
viscous drag in arteries when the pressure gradient is known, J. Physiol. 127:
553–563.
Womersley, J.R. (1955b) Oscillatory motion of a viscous liquid in a thin-walled elastic
tube. I. The linear approximation for long waves. Phil. Mag. 46: 199–221.
Womersley, J.R. (1957a) The mathematical analysis of arterial circulation in a state of
oscillatory motion. Wright Air Development Center, Technical Report WADC-
TR-56-614.
Womersley, J.R. (1957b) Oscillatory flow in arteries: the constrained elastic tube as a
model of arterial flow and pulse transmission, Phys. Med. Biol. 2: 178–187.
Yellin, E.L., Peskin, C., Yoran, C., Koenigsberg, M., Matsumoto, M., Laniado,
S., McQueen, D., Shore, D., and Frater, R.W. Mechanisms of mitral valve
motion during diastole. Am. J. Physiol, 1981. 241(3): H389–400.
Yoganathan, A.P., Cardiac Valve Prostheses, in The Biomedical Engineering Handbook,
Bronzino, J.D. Ed. 2000, CRC Press: Boca Raton, FL.
Yoganathan, A.P., Eberhardt, C.E., and Walker, P.G. Hydrodynamic performance of
the Medtronic Freestyle Aortic Root Bioprosthesis. J. Heart Valve Dis. 1994.
3(5): 571–80.
Young, D.F. (1979) Fluid mechanics of arterial stenosis. J. Biomech. Eng. 101: 157–173.
7328_C011.fm Page 410 Monday, October 16, 2006 2:27 PM
7328_Index Page 411 Tuesday, October 10, 2006 11:57 AM
Index
A Basement membrane, 87
Bernoulli equation, 26–29, 169–178
Abdominal aortic aneurysm, 172, 251, 319 aneurysms, 170–172
Advanced lesions, 107 arterial stenoses, 170–172
Adventitia, 87, 152 cardiac valve stenoses, 172–178
Adverse pressure gradients, 44 fluid dynamics, 26–29
Aliasing, 371 total vs. hydrostatic pressure measurement,
Alveolar epithelium, 97 169–170
Anastomoses, 317 Bifurcations and branches, flow through,
Anastomotic intimal hyperplasia, 324–333 236–241
Aneurysms, 170–172 Bileaflet valve, prosthetic heart valve, 299–302
flow through, 241–253 Bingham plastics, 6
formation of, 251 Björk–Shiley, 282
Angina pectoris, 93 Blood damage, prosthetic heart valve, in vitro
Angiogenin, 104 studies, 308–309
Angioplasty, 333–338 Blood vessel mechanics, 152–165
Angiotensin II, 105 cardiac muscle, material characterization
Aortic valve, 257, 259–267 of, 164–165
mechanical properties, 261–263 material behavior, 155–161
valve dynamics, 263–267 residual stress, 162–164
Aortoiliac bifurcation, 236 wall of blood vessel, structural components,
Apparent viscosity, 6 152–155
Arterial stenoses, 170–172 Boundary layer separation, 39, 44–45,
flow through, 241–253 183
Arteriovenous fistulas, 317–319 fluid dynamics, 44–45
Atherogenesis, hemodynamic theories, Bragg cell, 375
222–228 Bulk modulus, 7, 56–57, 193
low pressure, low, high wall shear stress Bundle of His, 74
theories, 223–227
oscillatory shear index, 227–228
time varying wall shear stress, 227–228 C
wall shear stress gradients, 227–228
Atherogenic, defined, 225 Caged ball valve, prosthetic heart valve,
Atheroprotective, defined, 225 298–299
Atherosclerosis, 94, 105–110 Capillary endothelial cell layer, 97
lesion growth, sequence, 107–110 Capillary viscometer, 116–123
morphology of, 106–107 Cardiac conduction, 74–77
Atherosclerotic plaque, 105, 107 Cardiac cycle, 76
Cardiac function, 77–82
Cardiac muscle, material characterization,
B 164–165
Cardiac structure, 73–74
Back-scattering mode, 375 Cardiac valves, 82–84
Baroreceptors, 104 stenoses, 172–178
411
7328_Index Page 412 Tuesday, October 10, 2006 11:57 AM
Index 413
Index 415
Index 417
Index 419